Fact-checked by Grok 2 weeks ago

Diffusion equation

The diffusion equation is a fundamental (PDE) in and physics that models the of a quantity, such as concentration of a substance or , as it spreads through a medium due to random molecular motion. In its standard form in one spatial dimension, it is expressed as \frac{\partial u}{\partial t} = D \frac{\partial^2 u}{\partial x^2}, where u(x,t) represents the quantity of interest, t is time, x is position, and D > 0 is the diffusion coefficient characterizing the medium's . In higher dimensions, the equation generalizes to \frac{\partial u}{\partial t} = D \nabla^2 u, where \nabla^2 is the Laplacian operator. The equation arises from the principle of or energy combined with Fick's first law of diffusion, which states that the diffusive flux J is proportional to the negative gradient of concentration: J = -D \nabla u. This law, proposed by Adolf Fick in 1855, posits that particles move from regions of higher concentration to lower concentration to equalize differences. Applying the \frac{\partial u}{\partial t} + \nabla \cdot J = 0 yields the diffusion equation, assuming D is constant. Historically, the diffusion equation first emerged in the context of heat conduction through Joseph Fourier's work in his 1822 treatise Théorie analytique de la chaleur, where it describes heat flow as \frac{\partial T}{\partial t} = \kappa \nabla^2 T with thermal diffusivity \kappa. Fourier's formulation provided the mathematical foundation, later adapted by Fick for mass diffusion and by in 1905 to explain , linking microscopic random walks to macroscopic diffusion. This PDE is parabolic, exhibiting smoothing behavior that leads to solutions approaching uniform equilibrium over time, distinguishing it from hyperbolic wave equations or elliptic steady-state equations. The diffusion equation has broad applications across disciplines, including modeling solute transport in porous media, in ecology, movement in semiconductors, and even financial option pricing via the Black-Scholes equation, a variant incorporating drift. Numerical solutions often employ or finite element methods due to the equation's linearity and well-posedness under appropriate initial and boundary conditions, such as Dirichlet or Neumann types. Extensions include nonlinear, fractional, or reaction-diffusion variants to capture more complex phenomena like in biology.

Introduction

General statement

The diffusion equation is a that models the spread of a substance through a medium over time. In its standard linear form for isotropic with constant diffusivity, it is given by \frac{\partial u}{\partial t} = D \nabla^2 u, where u(\mathbf{x}, t) represents the concentration or of the diffusing substance as a of \mathbf{x} and time t, D > 0 is the , and \nabla^2 denotes the Laplacian operator. Here, u quantifies the amount of the substance per unit volume at a given point, evolving temporally due to random particle motion, while \mathbf{x} typically spans one, two, or three spatial dimensions depending on the context. The coefficient D characterizes the material's , determining the rate of spreading; it is a positive scalar in isotropic media where occurs equally in all directions. For more general cases, including variable or direction-dependent diffusivity, the equation takes the form \frac{\partial u}{\partial t} = \nabla \cdot (D \nabla u), where D may be a position- or concentration-dependent scalar (for isotropic but inhomogeneous media) or a tensor (for , allowing different rates along principal directions). The diffusion coefficient D has dimensions of length squared per time, such as m²/s in units, reflecting the probabilistic nature of . This formulation bears a close mathematical resemblance to the , which describes temperature evolution in a conducting medium under analogous assumptions.

Physical and mathematical context

The provides a for where particles or substances spread through random motion, leading to a net flux from regions of higher concentration to lower concentration, as conceptualized in . This physical interpretation captures processes like the mixing of solutes in fluids or in solids, where the driving force is the concentration gradient rather than external forces. Mathematically, the diffusion equation belongs to the class of parabolic partial differential (PDEs), which are second-order in spatial variables and first-order in time. This classification arises from the general theory of linear second-order PDEs, where the of the principal symbol determines the type: parabolic equations feature a repeated real root, distinguishing them from hyperbolic PDEs (like the wave equation, which propagate disturbances at finite speeds along characteristics) and elliptic PDEs (like , which describe states without and exhibit smoothing). Parabolic equations, in contrast, demonstrate infinite propagation speed but with an inherent smoothing effect, where irregularities in the initial data are rapidly damped over time. To solve the diffusion equation in a bounded domain, appropriate conditions are specified, such as Dirichlet conditions (prescribing the value of the on the ), Neumann conditions (prescribing the normal derivative or ), or periodic conditions (for repeating domains). Additionally, an is required, typically of the form u(\mathbf{x}, 0) = u_0(\mathbf{x}), which sets the distribution at the starting time. The standard formulation of the diffusion equation relies on several key assumptions: the medium is homogeneous, the diffusion coefficient [D](/page/D*) is constant and positive, and the process is linear, meaning superposition holds for solutions. These simplifications enable analytical tractability but may require extensions for heterogeneous or nonlinear scenarios.

Historical development

Origins in diffusion processes

The phenomenon of diffusion has been observed empirically for centuries, with informal descriptions of substances spreading through , such as the gradual dissemination of vapors in air or dispersing in , predating any formal mathematical treatment. These pre-mathematical models highlighted the spontaneous mixing of matter without mechanical agitation, as noted in early qualitative accounts of natural processes. In the 19th century, systematic experiments began to quantify diffusion in gases and liquids, laying the groundwork for later theoretical developments. Scottish chemist Thomas Graham conducted pioneering studies in the 1820s and 1830s, observing the rates at which gases effuse through small openings or diffuse into other gases, such as the classic demonstration of ammonia and hydrogen chloride vapors meeting to form a white ring. His work culminated in Graham's law of effusion, formulated around 1833, which states that the rate of diffusion of a gas is inversely proportional to the square root of its density, thereby linking macroscopic transport to the underlying molecular motion. Building on these observations, Adolf Fick introduced a quantitative framework in 1855 through his seminal paper "On Liquid Diffusion," where he drew an analogy to Joseph Fourier's earlier work on heat conduction. Fick's first law posits that the diffusive flux \mathbf{J} is proportional to the negative gradient of the concentration u, expressed as \mathbf{J} = -D \nabla u, with D as the diffusion coefficient; this empirical relation was derived from experiments measuring salt diffusion in aqueous solutions. A microscopic justification for diffusion emerged in 1905 with Albert Einstein's theoretical analysis of , which modeled the erratic movement of suspended particles as a driven by molecular collisions. Einstein demonstrated that this leads to a proportional to time, directly connecting it to the macroscopic diffusion equation and providing empirical validation through predictable particle trajectories in fluids.

Key contributors and milestones

Joseph Fourier's seminal work in 1822 laid the groundwork for the diffusion equation through his development of the heat equation in Théorie analytique de la chaleur, which mathematically described heat conduction as a parabolic partial differential equation and directly inspired subsequent models for mass and particle diffusion processes. Fourier's formulation demonstrated how diffusive phenomena could be captured by equations of the form \frac{\partial u}{\partial t} = \alpha \nabla^2 u, where u represents temperature or concentration, providing a template for analogous diffusion problems. In the late 19th century, Lord Rayleigh and advanced the theoretical foundations by integrating kinetic theory with , linking macroscopic to microscopic of particle collisions in gases. Rayleigh's work in the 1880s on stochastic processes rediscovered the unity between physical and probabilistic descriptions, while Boltzmann's transport equation (1872, refined through the 1890s) derived coefficients from , establishing as a consequence of random walks in equilibrium systems. Adolf Fick's publication of his laws in 1855 marked a pivotal milestone, formalizing the flux of diffusing substances proportional to the concentration gradient (J = -D \nabla c) and leading to the second law that yields the diffusion equation \frac{\partial c}{\partial t} = D \nabla^2 c. In 1905, Albert Einstein provided a rigorous probabilistic interpretation by deriving the diffusion coefficient for spherical particles as D = \frac{kT}{6\pi \eta r}, connecting Brownian motion to Fickian diffusion and enabling experimental verification of atomic theory. Early 20th-century extensions by and contemporaries applied the diffusion equation to colloidal systems, introducing boundary conditions for and while emphasizing probabilistic interpretations through Smoluchowski's 1906 theory of under external forces. John Crank's 1956 book The Mathematics of Diffusion synthesized these developments into a comprehensive reference, cataloging analytical solutions and boundary value problems for the equation across diverse geometries and conditions. From the onward, the diffusion equation was recognized as a limiting case of the Fokker-Planck equation, which generalizes it to include drift terms in processes, as formulated by Adriaan Fokker (1914) and (1917) but widely adopted in diffusion contexts during that decade for modeling inhomogeneous random walks.

Mathematical formulation

Derivation from conservation laws

The diffusion equation can be derived from the fundamental principle of mass conservation, expressed through the , combined with a constitutive relation describing the diffusive flux. This approach assumes a description of the diffusing substance, where the local concentration u(\mathbf{x}, t) represents the per unit volume at position \mathbf{x} and time t. The continuity equation for mass conservation, in the absence of sources or sinks, states that the rate of change of concentration within a volume equals the negative divergence of the flux \mathbf{J} through its surface: \frac{\partial u}{\partial t} + \nabla \cdot \mathbf{J} = 0. $$/03%3A_Diffusion/10%3A_Diffusion/10.01%3A_Continuum_Diffusion) Fick's first law provides the relation between the flux and the concentration gradient, stating that the diffusive flux is proportional to the negative gradient of the concentration, reflecting the tendency of particles to move from regions of high to low concentration: \mathbf{J} = -D \nabla u, [](https://www.researchgate.net/publication/7926018_On_the_sesquicentennial_of_Fick%27s_laws_of_diffusion) where $D > 0$ is the [diffusion](/page/Diffusion) coefficient, a material-specific constant characterizing the ease of diffusion. This law was originally proposed by Adolf Fick in 1855 as an analogy to Fourier's law of heat conduction.[](https://www.researchgate.net/publication/7926018_On_the_sesquicentennial_of_Fick%27s_laws_of_diffusion) Substituting Fick's first law into the [continuity equation](/page/Continuity_equation) yields the diffusion equation: \frac{\partial u}{\partial t} = \nabla \cdot (D \nabla u). [](https://www.comsol.com/multiphysics/diffusion-equation) For constant $D$ and in Cartesian coordinates, this simplifies to the standard form: \frac{\partial u}{\partial t} = D \nabla^2 u. To illustrate the [derivation](/page/Derivation) in one dimension, consider a thin slab of material between positions $x$ and $x + \Delta x$ with cross-sectional area $A = 1$ for simplicity. The mass in this [control volume](/page/Control_volume) is $u(x, t) \Delta x$, and its time rate of change is $\frac{\partial u}{\partial t} \Delta x$. The net mass flux out of the slab is the difference in one-dimensional flux: $J(x + \Delta x, t) - J(x, t)$. By [conservation of mass](/page/Conservation_of_mass), \frac{\partial u}{\partial t} \Delta x = - [J(x + \Delta x, t) - J(x, t)]. Dividing by $\Delta x$ and taking the limit as $\Delta x \to 0$ gives \frac{\partial u}{\partial t} = -\frac{\partial J}{\partial x}. Applying Fick's first law in one dimension, $J = -D \frac{\partial u}{\partial x}$, and assuming constant $D$, substitution results in \frac{\partial u}{\partial t} = D \frac{\partial^2 u}{\partial x^2}. This derivation relies on several key assumptions: the absence of sources or sinks (no production or consumption terms in the continuity equation), isotropic diffusion (where $D$ is a scalar rather than a tensor, implying uniform diffusion in all directions), and local thermodynamic equilibrium (justifying the linear relation in Fick's law).[](https://scholar.harvard.edu/files/schwartz/files/2-diffusion_0.pdf) ### General forms in different dimensions The diffusion equation generalizes naturally to higher spatial dimensions by replacing the second derivative in one dimension with the Laplacian operator, which accounts for diffusion in all directions. In two dimensions, the isotropic form is given by \frac{\partial u}{\partial t} = D \left( \frac{\partial^2 u}{\partial x^2} + \frac{\partial^2 u}{\partial y^2} \right), where $u(x,y,t)$ represents the concentration or temperature, and $D$ is the diffusion coefficient. This equation describes the evolution of diffusive processes on a plane, such as heat spread in a thin sheet. In polar coordinates $(r, \theta)$, the Laplacian takes the form $\nabla^2 u = \frac{1}{r} \frac{\partial}{\partial r} \left( r \frac{\partial u}{\partial r} \right) + \frac{1}{r^2} \frac{\partial^2 u}{\partial \theta^2}$, yielding \frac{\partial u}{\partial t} = D \left[ \frac{1}{r} \frac{\partial}{\partial r} \left( r \frac{\partial u}{\partial r} \right) + \frac{1}{r^2} \frac{\partial^2 u}{\partial \theta^2} \right]. This coordinate system is particularly useful for problems with radial or angular symmetry, like diffusion from a circular source.[8][26] In three dimensions, the equation becomes \frac{\partial u}{\partial t} = D \nabla^2 u, where $\nabla^2 u = \frac{\partial^2 u}{\partial x^2} + \frac{\partial^2 u}{\partial y^2} + \frac{\partial^2 u}{\partial z^2}$ is the scalar Laplacian in Cartesian coordinates. For spherical symmetry in coordinates $(r, \theta, \phi)$, the Laplacian expands to $\nabla^2 u = \frac{1}{r^2} \frac{\partial}{\partial r} \left( r^2 \frac{\partial u}{\partial r} \right) + \frac{1}{r^2 \sin \theta} \frac{\partial}{\partial \theta} \left( \sin \theta \frac{\partial u}{\partial \theta} \right) + \frac{1}{r^2 \sin^2 \theta} \frac{\partial^2 u}{\partial \phi^2}$, simplifying to the radial form $\frac{\partial u}{\partial t} = D \frac{1}{r^2} \frac{\partial}{\partial r} \left( r^2 \frac{\partial u}{\partial r} \right)$ when angular dependence is absent. This is applicable to scenarios like solute diffusion in a spherical particle.[](https://people.maths.bris.ac.uk/~macpd/apde2/chap7.pdf)[](http://web.mit.edu/8.Math/www/lectures/lec13/3.5.3.pdf) For anisotropic diffusion, where diffusivity varies by direction, $D$ becomes a symmetric positive-definite tensor $\mathbf{D}$, and the equation is \frac{\partial u}{\partial t} = \nabla \cdot (\mathbf{D} \nabla u). In matrix form for Cartesian coordinates, this expands to $\frac{\partial u}{\partial t} = \sum_{i,j=1}^n \frac{\partial}{\partial x_i} \left( D_{ij} \frac{\partial u}{\partial x_j} \right)$, capturing directional preferences in materials like composites.[](https://link.aps.org/doi/10.1103/PhysRevE.74.061103)[](https://homepage.tudelft.nl/d2b4e/numanal/marlen_afst.pdf) The standard forms assume infinite domains, but for finite or bounded regions, [boundary](/page/Boundary) conditions are essential to specify [flux](/page/Flux) at interfaces. No-[flux](/page/Flux) (Neumann) boundaries enforce $\mathbf{n} \cdot (\mathbf{D} \nabla u) = 0$, where $\mathbf{n}$ is the outward [normal](/page/Normal), preventing diffusive transport across impermeable walls and conserving total quantity.[](http://web.mit.edu/1.061/www/dream/FOUR/FOURTHEORY.PDF) When the diffusion coefficient varies with time, the equation adjusts to $\frac{\partial u}{\partial t} = \nabla \cdot (D(t) \nabla u)$, though this introduces challenges like non-self-adjoint operators that may complicate analytical solutions and require case-specific analysis for well-posedness.[](https://link.aps.org/doi/10.1103/PhysRevE.72.011107) ## Properties and theory ### Fundamental solutions and Green's functions The fundamental solution to the diffusion equation $\partial_t u = D \partial_{xx} u$ on the infinite domain $-\infty < x < \infty$ with a Dirac delta initial condition at $x=0$ is given by the Gaussian kernel G(x, t) = \frac{1}{\sqrt{4\pi D t}} \exp\left( -\frac{x^2}{4 D t} \right) for $t > 0$.[](https://www.math.ucsd.edu/~jmckerna/Teaching/18-19/Autumn/110A/l_11.pdf) This explicit form arises from solving the equation via self-similar [ansatz](/page/Ansatz) or [Fourier transform](/page/Fourier_transform) methods and satisfies the initial condition in the sense of distributions, as $\int_{-\infty}^{\infty} G(x, t) \, dx = 1$ for all $t > 0$.[](https://www.math.ucsd.edu/~jmckerna/Teaching/18-19/Autumn/110A/l_11.pdf) Physically, $G(x, t)$ describes the evolution of a point-source initial concentration of total mass 1 at the [origin](/page/Origin), with the [diffusion process](/page/Diffusion_process) spreading the mass symmetrically. The [distribution](/page/Distribution) has zero [mean](/page/Mean) and variance $2 D t$, reflecting the linear growth of spatial spread with time.[](https://neherlab.org/PoL_notebooks/2023/lecture04b_DiffusionEquation.pdf) In higher dimensions, the fundamental solution generalizes to G(\mathbf{x}, t) = \frac{1}{(4\pi D t)^{n/2}} \exp\left( -\frac{|\mathbf{x}|^2}{4 D t} \right) for $\mathbf{x} \in \mathbb{R}^n$, preserving the total [integral](/page/Integral) of 1 and exhibiting variance $2 D t$ per [dimension](/page/Dimension).[](https://www.damtp.cam.ac.uk/user/dbs26/1BMethods/Heat.pdf) The Green's function approach leverages this kernel to represent general solutions to the initial-value problem. For initial data $u(\mathbf{x}, 0) = u_0(\mathbf{x})$, the solution is the convolution u(\mathbf{x}, t) = \int_{\mathbb{R}^n} G(\mathbf{x} - \mathbf{y}, t) u_0(\mathbf{y}) , d\mathbf{y}, which linearly superposes point-source responses and ensures mass conservation if $u_0$ integrates to a finite total.[](https://www.math.ucsd.edu/~jmckerna/Teaching/18-19/Autumn/110A/l_11.pdf) This integral form highlights the smoothing effect of diffusion, transforming arbitrary initial profiles into progressively Gaussian-like distributions over time. A key property of the fundamental solution is its [self-similarity](/page/Self-similarity), stemming from the scaling invariance of the diffusion equation under transformations $x \mapsto \lambda x$, $t \mapsto \lambda^2 t$, and $u \mapsto \lambda^{-n} u$ in $[n](/page/N+)$ dimensions. The solution depends only on the similarity variable $\eta = \mathbf{x} / \sqrt{4 D t}$, yielding $G(\mathbf{x}, t) = (4 D t)^{-n/2} g(\eta)$ where $g$ is a fixed radial Gaussian profile.[](https://www.damtp.cam.ac.uk/user/dbs26/1BMethods/Heat.pdf) This reduces the PDE to [an ODE](/page/An_Ode) in $\eta$, facilitating analytical insight into the spreading dynamics. Asymptotically, as $t \to 0^+$, the kernel sharpens to a Dirac delta distribution $\delta(\mathbf{x})$, preserving the initial point-source localization. Conversely, as $t \to \infty$, the solution flattens with peak height decaying as $(4\pi D t)^{-n/2}$ and width expanding proportionally to $\sqrt{t}$, illustrating the irreversible homogenization inherent to diffusion.[](https://www.damtp.cam.ac.uk/user/dbs26/1BMethods/Heat.pdf) ### Uniqueness and maximum principles The maximum principle for solutions of the [diffusion equation](/page/Diffusion_equation), a prototypical [parabolic partial differential equation](/page/Parabolic_partial_differential_equation), asserts that the maximum value of the solution $u(x,t)$ over a bounded [domain](/page/Domain) $\Omega \subset \mathbb{R}^n$ and time interval $[0,T]$ occurs either on the initial time $t=0$ or on the parabolic boundary $\partial \Omega \times [0,T]$.[](https://www.math.ubc.ca/~gustaf/M516/L17.pdf) This principle holds for the classical [heat equation](/page/Heat_equation) $\partial_t u = \Delta u$ under suitable boundary conditions, such as Dirichlet or [Neumann](/page/Neumann), and extends to more general linear parabolic equations with bounded coefficients.[](https://projecteuclid.org/journals/pacific-journal-of-mathematics/volume-8/issue-2/Remarks-on-the-maximum-principle-for-parabolic-equations-and-its/pjm/1103040096.pdf) A proof sketch relies on the comparison [principle](/page/Principle): suppose $u$ achieves an interior maximum at $(x_0, t_0)$ with $t_0 > 0$; then at that point, $\partial_t u(x_0, t_0) \leq 0$ and $\Delta u(x_0, t_0) \leq 0$, contradicting the equation unless the maximum is constant, which propagates backward to the boundary via energy methods or the strong [maximum principle](/page/Maximum_principle).[](https://www.math.ucdavis.edu/~hunter/pdes/ch6.pdf) Uniqueness of solutions to the diffusion equation follows from energy estimates in appropriate function spaces. For the initial-boundary value problem on a bounded domain with $u_0 \in L^2(\Omega)$, if two solutions $u$ and $v$ satisfy the equation, their difference $w = u - v$ solves the homogeneous equation with zero initial and boundary data; multiplying by $w$ and integrating yields $\frac{d}{dt} \int_\Omega |w|^2 \, dx + 2 \int_\Omega |\nabla w|^2 \, dx = 0$, implying $\int_\Omega |\nabla w|^2 \, dx$ is nonincreasing and $w \equiv 0$ by Gronwall's inequality or Poincaré estimates.[](https://math.uchicago.edu/~may/REU2014/REUPapers/Ji.pdf) This holds under conditions like $L^2$ integrability of initial data and compatibility with boundary conditions, ensuring well-posedness in Hilbert spaces.[](https://www.math.cmu.edu/~rcristof/pdf/Teaching/Spring2017/Energy-uniqueness%281%29.pdf) Existence of solutions can be established via the Galerkin method for weak solutions in Sobolev spaces or through semigroup theory in Banach spaces. In the Galerkin approach, approximate solutions are projected onto finite-dimensional subspaces of $H^1_0(\Omega)$, yielding a system of ODEs whose solutions converge weakly to a solution of the variational formulation as the dimension increases, with uniform energy bounds ensuring compactness.[](https://www.math.ucdavis.edu/~hunter/notes/nonlinev.pdf) Semigroup methods generate a contraction semigroup on $L^2(\Omega)$ via the Laplacian's self-adjointness, providing mild solutions that coincide with classical ones under higher regularity.[](https://web.stanford.edu/class/math220b/handouts/heateqn.pdf) Solutions to the diffusion equation exhibit strong regularity properties for $t > 0$, even with rough initial data. Specifically, if the initial condition $u_0$ is merely bounded or in $L^p(\Omega)$, the solution $u(\cdot, t)$ becomes infinitely differentiable in space and Hölder continuous in time, with estimates like $\|u(\cdot, t)\|_{C^{k,\alpha}(\overline{\Omega})} \leq C t^{-\beta}$ for suitable $\beta > 0$, reflecting the parabolic smoothing effect.[](https://people.math.harvard.edu/~spicard/notes-parabolicpde.pdf) This Hölder continuity, often in the Schauder sense, follows from potential theory or parametrix constructions, ensuring classical solutions interior to the domain for positive times.[](https://www.math.ucdavis.edu/~hunter/pdes/ch6.pdf) These principles have limitations in nonlinear variants of the diffusion equation, such as those with reaction terms or degenerate diffusion coefficients, where the strong [maximum principle](/page/Maximum_principle) may fail due to interior maxima or "[needles](/page/The_Needles)" in the [solution](/page/Solution) profile.[](https://www.researchgate.net/publication/278099699_Failure_of_the_strong_maximum_principle_in_nonlinear_diffusion_Existence_of_needles) ## Solution methods ### Analytical approaches Analytical approaches to solving the diffusion equation, ∂u/∂t = D ∇²u, yield exact closed-form [solutions](/page/Solution) for well-posed initial-boundary value problems, particularly in one or higher dimensions with specific geometries. These methods, rooted in transform techniques and series expansions, transform the [partial differential equation](/page/Partial_differential_equation) (PDE) into solvable ordinary differential equations (ODEs) or [integral](/page/Integral) equations, often leveraging [symmetry](/page/Symmetry) or superposition principles. Pioneered in the context of [heat](/page/Heat) conduction, such techniques remain foundational for theoretical [analysis](/page/Analysis) in diffusion processes.[](https://www3.nd.edu/~powers/ame.20231/fourier1822.pdf) The [separation of variables](/page/Separation_of_variables) method is applicable to bounded domains with homogeneous boundary conditions, assuming a product form u(x,t) = X(x)T(t) that decouples the PDE into spatial and temporal ODEs. For the one-dimensional case on [0, L] with Dirichlet conditions u(0,t) = u(L,t) = 0, substitution yields X''/X = (1/D) T'/T = -λ, leading to the eigenvalue problem X'' + λX = 0 with eigenvalues λ_n = (nπ/L)^2 and eigenfunctions X_n(x) = [sin](/page/Sin)(nπx/L) for n = 1,2,... The temporal equation T' + D λ T = 0 gives T_n(t) = [exp](/page/Exp)(-D λ_n t), so the general solution is the superposition u(x,t) = \sum_{n=1}^\infty b_n \sin\left(\frac{n\pi x}{L}\right) \exp\left(-D \left(\frac{n\pi}{L}\right)^2 t\right), where coefficients b_n are Fourier sine series coefficients of the initial condition u(x,0). This approach extends to higher dimensions and other boundary types, such as Neumann conditions, by adjusting the eigenfunctions accordingly. Originally developed by Fourier for heat flow in solids, it provides explicit series solutions for transient diffusion in finite geometries.[](https://www3.nd.edu/~powers/ame.20231/fourier1822.pdf)[](https://www-eng.lbl.gov/~shuman/NEXT/MATERIALS&COMPONENTS/Xe_damage/Crank-The-Mathematics-of-Diffusion.pdf) For unbounded or semi-infinite domains, the Fourier transform method exploits translational invariance. The one-dimensional Fourier transform of u(x,t) is û(ω,t) = ∫_{-∞}^∞ u(x,t) e^{-i ω x} dx, transforming the diffusion equation to ∂û/∂t = -D ω² û(ω,t). With initial condition û(ω,0), the solution is û(ω,t) = û(ω,0) exp(-D ω² t), and u(x,t) is recovered via inverse transform u(x,t) = \frac{1}{2\pi} \int_{-\infty}^\infty \hat{u}(\omega,0) \exp(i \omega x - D \omega^2 t) , d\omega. This yields the fundamental Gaussian solution for an initial delta function, u(x,0) = δ(x), as u(x,t) = (4π D t)^{-1/2} exp(-x²/(4 D t)), which convolves with arbitrary initials via superposition. The method is particularly effective for infinite domains without boundaries.[](https://www-eng.lbl.gov/~shuman/NEXT/MATERIALS&COMPONENTS/Xe_damage/Crank-The-Mathematics-of-Diffusion.pdf) The [Laplace transform](/page/Laplace_transform), applied with respect to time, suits initial-boundary value problems by converting the time derivative to an algebraic term. For one dimension, the transform ũ(x,s) = ∫_0^∞ u(x,t) e^{-s t} dt satisfies s ũ(x,s) - u(x,0) = D ∂²ũ/∂x², a second-order [ODE](/page/Ode) solvable subject to transformed boundary conditions. Inversion via [contour integration](/page/Contour_integration) or tables then yields u(x,t). This technique handles non-homogeneous or time-dependent boundaries efficiently, often complementing [separation of variables](/page/Separation_of_variables) for finite times.[](https://www-eng.lbl.gov/~shuman/NEXT/MATERIALS&COMPONENTS/Xe_damage/Crank-The-Mathematics-of-Diffusion.pdf) The [method of images](/page/Method_of_images) addresses [boundary](/page/Boundary) effects in semi-infinite or finite domains by extending the infinite-domain solution with fictitious "[image](/page/Image)" sources to enforce conditions. For a reflecting (no-flux) [boundary](/page/Boundary) at x=0, an [image](/page/Image) source at -x_0 mirrors a real source at x_0, yielding u(x,t) = (4π D t)^{-1/2} [exp(-(x - x_0)^2/(4 D t)) + exp(-(x + x_0)^2/(4 D t))], satisfying ∂u/∂x|_{x=0} = 0. For absorbing boundaries (u=0), a negative [image](/page/Image) is used. Multiple images handle slabs or cylinders via infinite series or periodic extensions. This approach simplifies problems reducible to fundamental solutions.[](https://www-eng.lbl.gov/~shuman/NEXT/MATERIALS&COMPONENTS/Xe_damage/Crank-The-Mathematics-of-Diffusion.pdf) In the long-time limit t → ∞, transient terms decay, and the diffusion equation reduces to the steady-state form ∇²u = 0, known as [Laplace's equation](/page/Laplace's_equation). Solutions are harmonic functions, analytic and satisfying the mean-value property, determined solely by boundary values without sources. For example, in one dimension, u(x) = a + b x for constant flux boundaries. These limits provide equilibrium profiles essential for interpreting asymptotic behavior in confined systems.[](https://www-eng.lbl.gov/~shuman/NEXT/MATERIALS&COMPONENTS/Xe_damage/Crank-The-Mathematics-of-Diffusion.pdf) ### Numerical discretization techniques Numerical discretization techniques approximate solutions to the diffusion equation by replacing continuous [derivatives](/page/Hartshorn) with discrete differences or integrals on a computational [grid](/page/Grid), enabling solutions for [complex](/page/Complex) geometries, nonlinear variants, or irregular boundaries where analytical methods are impractical. These methods convert the PDE into a system of ordinary [differential](/page/Differential) or algebraic equations, solved iteratively, with considerations for stability, accuracy, and computational efficiency.[](https://www-eng.lbl.gov/~shuman/NEXT/MATERIALS&COMPONENTS/Xe_damage/Crank-The-Mathematics-of-Diffusion.pdf) Finite difference methods are among the most common, approximating spatial and temporal derivatives on a uniform grid. The explicit forward-time centered-space ([FTCS](/page/FTCS_scheme)) scheme discretizes the one-dimensional equation as (u_i^{n+1} - u_i^n)/Δt = D (u_{i+1}^n - 2u_i^n + u_{i-1}^n)/(Δx)^2, yielding u_i^{n+1} = u_i^n + r (u_{i+1}^n - 2u_i^n + u_{i-1}^n), where r = D Δt / (Δx)^2 ≤ 1/2 for [stability](/page/Stability). This method is straightforward but restricted by the stability condition, limiting time step size.[](https://www-eng.lbl.gov/~shuman/NEXT/MATERIALS&COMPONENTS/Xe_damage/Crank-The-Mathematics-of-Diffusion.pdf)[](https://sites.engineering.ucsb.edu/~moehlis/APC591/tutorials/tutorial5/) Implicit schemes, such as backward Euler, use future time levels for spatial differences, resulting in a [linear system](/page/Linear_system) Au^{n+1} = u^n that must be solved at each step; they are unconditionally stable but require matrix inversion. The Crank-Nicolson method combines explicit and implicit approximations, achieving second-order accuracy in time and unconditional stability, and is widely used for its balance of efficiency and reliability in multi-dimensional problems. Extensions to higher dimensions and variable coefficients involve analogous grids, with boundary conditions incorporated via ghost points or modified stencils.[](https://www-eng.lbl.gov/~shuman/NEXT/MATERIALS&COMPONENTS/Xe_damage/Crank-The-Mathematics-of-Diffusion.pdf) Other techniques include finite element methods, which use variational formulations and basis functions for irregular domains, and finite volume methods, which conserve quantities locally by integrating over control volumes. These are particularly suited for engineering applications like [heat transfer](/page/Heat_transfer) or porous media flow, often implemented in software libraries.[](https://www-eng.lbl.gov/~shuman/NEXT/MATERIALS&COMPONENTS/Xe_damage/Crank-The-Mathematics-of-Diffusion.pdf) ## Applications ### In physical sciences The diffusion equation finds one of its primary applications in modeling heat conduction in physical systems, where it manifests as the [heat equation](/page/Heat_equation). This [parabolic partial differential equation](/page/Parabolic_partial_differential_equation) describes how thermal energy propagates through materials due to [temperature](/page/Temperature) gradients, governed by Fourier's law of heat conduction, which posits that [heat flux](/page/Heat_flux) is proportional to the negative [gradient](/page/Gradient) of [temperature](/page/Temperature). The standard one-dimensional form is \frac{\partial T}{\partial t} = \alpha \frac{\partial^2 T}{\partial x^2}, where $T$ is temperature, $t$ is time, $x$ is position, and $\alpha = k / (\rho c_p)$ is the [thermal diffusivity](/page/Thermal_diffusivity), with $k$ denoting thermal conductivity, $\rho$ the [density](/page/Density), and $c_p$ the [specific heat capacity](/page/Specific_heat_capacity). This formulation was first systematically derived by [Joseph Fourier](/page/Joseph_Fourier) in his seminal 1822 treatise *Théorie analytique de la chaleur*, revolutionizing the understanding of [heat transfer](/page/Heat_transfer) as a diffusive process rather than relying on caloric [fluid](/page/Fluid) theories.[](https://www3.nd.edu/~powers/ame.20231/fourier1878.pdf) A classic example is the cooling of a hot metal object in a cooler environment, where initial high temperatures at the surface diffuse inward, leading to uniform cooling over time; solutions to the [heat equation](/page/Heat_equation) predict [exponential decay](/page/Exponential_decay) of temperature differences, essential for analyzing transient [thermal](/page/Thermal) behaviors in [solids](/page/The_Solids) like metals or ceramics. In chemistry and physics, the diffusion equation underpins the description of [molecular diffusion](/page/Molecular_diffusion), particularly through Fick's laws, which quantify the [transport](/page/Transport) of particles from regions of high to low concentration. Fick's first law states that the diffusive [flux](/page/Flux) $J$ is $J = -D \nabla c$, where $D$ is the diffusion coefficient and $c$ is concentration; combining this with [conservation of mass](/page/Conservation_of_mass) yields the second law, the diffusion equation $\partial c / \partial t = D \nabla^2 c$, originally proposed by Adolf Fick in 1855 by analogy to heat conduction.[](https://pubs.acs.org/doi/10.1021/ed041p397) This model applies to Fickian [diffusion](/page/Diffusion) in gases, liquids, and solutions, capturing phenomena such as the spread of oxygen molecules through biological tissues, where $D$ for O₂ in [water](/page/Water) is approximately $2 \times 10^{-9}$ m²/s at 25°C, facilitating [respiration](/page/Respiration) at cellular levels./Kinetics/09%3A_Diffusion) Similarly, in environmental physics, it models [pollutant](/page/Pollutant) [dispersion](/page/Dispersion) in [water](/page/Water) bodies, like the diffusion of contaminants from a spill, where low $D$ values (e.g., $10^{-10}$ to $10^{-9}$ m²/s for organic solutes) lead to gradual spreading influenced by hydrodynamic conditions.[](https://www.sciencedirect.com/topics/engineering/ficks-law) A representative example is the diffusion of [sucrose](/page/Sucrose) (table sugar) in [water](/page/Water), with $D \approx 5 \times 10^{-10}$ m²/s at 25°C, illustrating how molecular size and [solvent](/page/Solvent) interactions determine [transport](/page/Transport) rates in dilute solutions.[](https://ift.onlinelibrary.wiley.com/doi/pdf/10.1111/j.1365-2621.1987.tb06655.x) In nuclear physics, the diffusion equation approximates neutron transport in reactors, simplifying the more complex Boltzmann transport equation under assumptions of isotropic scattering and low absorption. The steady-state neutron diffusion equation takes the form $D \nabla^2 \phi + (k - 1) \Sigma_a \phi = 0$, where $\phi$ is the neutron flux, $D$ is the diffusion coefficient (related to mean free path and speed), $\Sigma_a$ is the macroscopic absorption cross-section, and $k$ is the effective multiplication factor determining reactor criticality ($k = 1$ for steady operation).[](https://www.sciencedirect.com/topics/engineering/neutron-diffusion-equation) This approximation, developed in the mid-20th century as part of reactor theory, enables efficient calculation of flux distributions in fissile materials like uranium-235, crucial for designing thermal reactors where neutrons diffuse through moderators like water or graphite to sustain controlled chain reactions.[](https://www.nrc.gov/docs/ml1214/ml12142a086.pdf) Electrochemical processes in physical sciences also rely on the diffusion equation, often extended via the Nernst-Planck framework to account for ion migration under [electric fields](/page/Electric_Fields) alongside [diffusion](/page/Diffusion). The Nernst-Planck equation for species flux is $J_i = -D_i \nabla c_i - \frac{z_i F D_i c_i}{RT} \nabla \phi$, where subscripts denote species, $z_i$ is charge number, $F$ is Faraday's constant, $R$ is the [gas constant](/page/Gas_constant), $T$ is [temperature](/page/Temperature), and $\phi$ is [electric potential](/page/Electric_potential); [continuity](/page/Continuity) yields a diffusion-like equation coupled with [Poisson's equation](/page/Poisson's_equation) for charge balance.[](https://www.sciencedirect.com/topics/engineering/nernst-planck-equation) This describes ion [diffusion](/page/Diffusion) in batteries, such as lithium-ion transport in electrolytes, limiting charge-discharge rates and capacity.[](https://www.mdpi.com/2313-0105/10/7/238) In [corrosion](/page/Corrosion) modeling, it simulates the [diffusion](/page/Diffusion) of corrosive ions (e.g., Cl⁻) through protective [oxide](/page/Oxide) layers on metals, where slow diffusion coefficients predict pitting [initiation](/page/Initiation) and [propagation](/page/Propagation) rates in aqueous environments.[](https://pubs.acs.org/doi/10.1021/acsmeasuresciau.2c00070) ### In engineering and computation In engineering, the diffusion equation finds extensive application in image processing through anisotropic variants that enable [noise reduction](/page/Noise_reduction) while preserving edges. The Perona-Malik model introduces a nonlinear diffusion process defined by the partial differential equation \frac{\partial I}{\partial t} = \div \left( g(|\nabla I|) \nabla I \right), where $I$ represents the image intensity, $\nabla I$ is the [image gradient](/page/Image_gradient), and $g$ is an edge-stopping function that decreases with increasing gradient magnitude to halt diffusion across strong edges.[](https://www.sci.utah.edu/~gerig/CS7960-S2010/materials/Perona-Malik/PeronaMalik-PAMI-1990.pdf) This approach, originally proposed for [scale-space](/page/Scale_space) analysis and [edge detection](/page/Edge_detection), has become a cornerstone for denoising in [computer vision](/page/Computer_vision) tasks, balancing smoothness in homogeneous regions with fidelity at boundaries.[](https://www.sci.utah.edu/~gerig/CS7960-S2010/materials/Perona-Malik/PeronaMalik-PAMI-1990.pdf) In [fluid dynamics](/page/Fluid_dynamics), the diffusion equation underpins the modeling of viscous effects within the Navier-Stokes equations, where the [term](/page/Term) $\nu \nabla^2 \mathbf{u}$ captures [momentum diffusion](/page/Momentum_diffusion) due to [viscosity](/page/Viscosity), with $\nu$ as the kinematic [viscosity](/page/Viscosity) and $\mathbf{u}$ the velocity [field](/page/Field).[](https://www.grc.nasa.gov/www/BGH/nseqs.html) This viscous diffusion [term](/page/Term) is [essential](/page/Essential) for simulating laminar flows, [boundary layer](/page/Boundary_layer) [development](/page/Development), and [turbulence](/page/Turbulence) in [engineering](/page/Engineering) designs such as pipelines, [aircraft](/page/Aircraft) wings, and [heat](/page/Heat) exchangers.[](https://www.grc.nasa.gov/www/BGH/nseqs.html) Similarly, in [materials science](/page/Materials_science), [dopant](/page/Dopant) [diffusion](/page/Diffusion) during [semiconductor](/page/Semiconductor) fabrication relies on the diffusion equation to predict impurity profiles, enabling precise control of electrical properties in devices like transistors and solar cells through thermal annealing processes.[](https://www.cityu.edu.hk/phy/appkchu/AP6120/8.PDF) The diffusion equation also manifests in finance via the Black-Scholes partial differential equation for option pricing, given by \frac{\partial V}{\partial t} + \frac{1}{2} \sigma^2 S^2 \frac{\partial^2 V}{\partial S^2} + r S \frac{\partial V}{\partial S} - r V = 0, where $V$ is the option value, $S$ the underlying asset price, $\sigma$ the volatility, and $r$ the risk-free rate; this heat-like equation models the probabilistic diffusion of asset prices under [geometric Brownian motion](/page/Geometric_Brownian_motion).[](https://www.cs.princeton.edu/courses/archive/fall09/cos323/papers/black_scholes73.pdf) In computational engineering, solving large-scale diffusion problems benefits from GPU acceleration, which parallelizes [finite difference](/page/Finite_difference) or finite element discretizations to achieve real-time simulations in scenarios like reacting flows or [heat transfer](/page/Heat_transfer), often yielding speedups of 10-100x over CPU implementations.[](https://developer.nvidia.com/gpugems/gpugems/part-vi-beyond-triangles/chapter-38-fast-fluid-dynamics-simulation-gpu) Recent advancements integrate diffusion principles into [machine learning](/page/Machine_learning) for generative [AI](/page/Ai), where score-based models approximate the reverse [diffusion process](/page/Diffusion_process) to generate data by iteratively denoising from noise distributions, as in denoising diffusion probabilistic models that have revolutionized [image](/page/Image) and video [synthesis](/page/Synthesis).[](https://proceedings.neurips.cc/paper/2020/file/4c5bcfec8584af0d967f1ab10179ca4b-Paper.pdf) These models, rooted in [stochastic](/page/Stochastic) [differential](/page/Differential) equations akin to the [diffusion equation](/page/Diffusion), enable high-fidelity sampling in applications from art generation to [drug discovery](/page/Drug_discovery).[](https://arxiv.org/abs/2011.13456)

References

  1. [1]
    [PDF] Stability of Finite Difference Schemes on the Diffusion Equation with ...
    The diffusion equation is one of the most fundamental partial differential equa- tions, with widespread applications for analyzing heat and mass transport ...
  2. [2]
    [PDF] Lecture 5.1: Fourier's law and the diffusion equation
    Assuming that the diffusion coefficient is constant, the diffusion equation becomes ut = c2uxx , c2 = −D.
  3. [3]
    [PDF] Chapter 2 – Diffusion Equation Part 1 - Benoit Cushman-Roisin
    The diffusion equation is a linear one, and a solution can, therefore, be obtained by adding several other solutions. An elementary solution ('building block').
  4. [4]
    Diffusion
    J = -D grad C , or in one dimension, J = -D dC/dx . D is a diffusion coefficient, which we usually assume to be a constant. Diffusion fluxes can cause changes ...
  5. [5]
    [PDF] Diffusion as Biology's Null Hypothesis for Dynamics - The Garcia Lab
    The goal of our thinking is to determine what amounts to an “equation of motion” that tells how the concentration field changes in both space and time. Fick's ...
  6. [6]
    [PDF] Lecture - WHOI GFD
    century of debate, Einstein definitively explained this phenomenon (6,7). 1.1 Einstein's derivation of the diffusion equation. Our interest here is in ...
  7. [7]
    [PDF] arXiv:2302.03333v1 [math.AP] 7 Feb 2023
    Feb 7, 2023 · As a basic and important mathematical model, the diffusion equation has been used to describe many physical, biological, chemical, economic, ...<|control11|><|separator|>
  8. [8]
    [PDF] The Diffusion Equation
    λ > 0: X(x) = C1 cos(√λx)+C2 sin(√λx). Substituting of the boundary conditions leads to the following equations for the constants C1 and C2: X(0) = C1 = ...
  9. [9]
    Generalized diffusion equation for anisotropic anomalous diffusion
    Dec 6, 2006 · [17] proposed a generalized diffusion equation for isotropic diffusion based on a probability balance and a gradient law for the diffusive flux.
  10. [10]
    DIFFUSION COEFFICIENT - Thermopedia
    The dimension of D in the SI system is a square meter per second. The diffusion coefficient is a physical constant dependent on molecule size and other ...
  11. [11]
    What are the units for D? - MathBench - University of Maryland
    The units of D are length2/time, and usually reported in cm2/sec. Why length2/time? It's possible to show that these are the necessary units using "dimensional ...
  12. [12]
  13. [13]
    The Mathematics of Diffusion - John Crank - Oxford University Press
    Free delivery 25-day returnsThis new edition preserves the general character of the book in providing a collection of solutions of the equations of diffusion.
  14. [14]
    The dichotomous history of diffusion - Physics Today
    At the beginning of the 19th century, the mathematics of both were established by Joseph Fourier and Pierre Simon Laplace. In 1807 Fourier submitted a monograph ...
  15. [15]
    Thomas Graham
    Thomas Graham studied gas diffusion and effusion, discovering that the rate is inversely proportional to the square root of density or molecular weight.Missing: 1830s | Show results with:1830s
  16. [16]
    Heroes and Highlights in the History of Diffusion - ResearchGate
    Aug 7, 2025 · This paper is devoted to some major landmarks and eminent pioneers of diffusion from the nineteenth and twentieth century.
  17. [17]
    On the sesquicentennial of Fick's laws of diffusion - Nature
    Apr 1, 2005 · References · Fick, A. Poggendorff's Annalen der Physik und Chemie 94, 59–86 (1855). · Fick, A. Phil. Mag. J. Sci. · Reprinted in J. Membr. Sci. 100 ...Missing: original | Show results with:original
  18. [18]
    Adolf Eugen Fick (1829-1901) - The Man Behind the Cardiac Output ...
    Oct 15, 2020 · In 1855, he proposed Fick's laws on gas diffusion. In 1870, he devised Fick's principle, which allows the measurement of cardiac output and ...
  19. [19]
    [PDF] the brownian movement - DAMTP
    The coefficient of diffusion of the suspended sub- stance therefore depends (except for universal constants and absolute temperature) only on the coefficient ...
  20. [20]
  21. [21]
  22. [22]
  23. [23]
    On the sesquicentennial of Fick's laws of diffusion - ResearchGate
    Aug 6, 2025 · On the sesquicentennial of Fick's laws of diffusion · 1. Fick, A. Poggendorff's Annalen der Physik und Chemie. 94, 59–86 (1855). · 2. Fick, A.
  24. [24]
    Diffusion Equation: Fick's Laws of Diffusion - COMSOL
    Jan 14, 2015 · Fick's second law of diffusion is a linear equation with the dependent variable being the concentration of the chemical species under consideration.Missing: 1855 | Show results with:1855
  25. [25]
    [PDF] Lecture 2: Diffusion
    The derivation of the heat equation is identical to the derivation of Fick's second law, with conservation of energy replacing conservation of particle number.
  26. [26]
    [PDF] The Laplacian Operator in Polar Coordinates
    The Laplacian Operator in Polar Coordinates. Our goal is to study the heat, wave and Laplace's equation in (1) polar coordinates in the plane.
  27. [27]
    [PDF] Chapter 7 PDEs in Three Dimensions - University of Bristol
    The three key eqns introduced in Chapter 2 were: (i) utt = c2∇2u, the wave equation,. (ii) ut = D∇2u, the diffusion equation,. (iii) ∇2u = 0, Laplace's equation ...
  28. [28]
    [PDF] 3.5.3 Diffusion within a sphere - MIT
    May 3, 2022 · As an application of spherical coordinates, let us consider the diffusion of a scalar density field n(!r, t) within a spherical volume of radius ...
  29. [29]
    [PDF] Linear and anisotropic diffusion in image processing
    J = −D ·∇u. (3.1). This equation gives the relation between the diffusion flux J and the concentration gradient ∇u. D is the diffusion coefficient. Since ...
  30. [30]
    [PDF] 4. Boundary Conditions - MIT
    For no flux, the advective and diffusive fluxes must exactly balance. If the boundary is solid, then the velocity normal to it is zero, and the constraint is ...
  31. [31]
    Time-fractional diffusion equation with time dependent diffusion ...
    Jul 18, 2005 · In summary, we have investigated the time-fractional diffusion equation with the time dependent diffusion coefficient D ( t ) = D α , γ t γ .
  32. [32]
    [PDF] 11. The diffusion equation on the line - UCSD Math
    The diffusion equation on the line. Our goal is to solve as explicitly as possible the diffusion equation on the whole line ut = kuxx.
  33. [33]
    [PDF] The diffusion equation
    Nov 29, 2023 · In this case, the solution is. This curve is a Gaussian curve with mean and variance. • The center of the distribution translates at constant ...
  34. [34]
    [PDF] 4 The Heat Equation - DAMTP
    The heat kernel is a Gaussian centred on x0. The rms width (standard deviation) of the Gaussian is %K(t − t0) while the height of the peak at x = x0 is 1/%4πK( ...
  35. [35]
    [PDF] IIID: Parabolic maximum principle - UBC Math
    Here we extend the maximum principle we proved earlier for the heat equation to more general parabolic equations. Let Ω ⇢ Rn be open, bounded and connected, ...
  36. [36]
    REMARKS ON THE MAXIMUM PRINCIPLE FOR PARABOLIC ...
    In [3] Nirenberg has proved maximum principles, both weak and strong, for parabolic equations. In § 1 of this paper we give a generalization of his strong ...
  37. [37]
    [PDF] Chapter 6: Parabolic equations - UC Davis Math
    elliptic PDEs, parabolic PDEs have strong smoothing properties. For example, there are parabolic versions of the maximum principle and Harnack's inequality,.
  38. [38]
    [PDF] discussion of the heat equation - Department of Mathematics
    Abstract. This paper discusses the heat equation from multiple perspectives. It begins with the derivation of the heat equation. Then it shows how to find.
  39. [39]
    [PDF] Energy and uniqueness
    The aim of this note is to show you a strategy in order to derive a uniqueness result for a PDEs problem by using the energy of the problem. Example: the heat ...
  40. [40]
    [PDF] Nonlinear Evolution Equations1 - UC Davis Math
    Then we formulate the definition of weak solutions of the diffusion equation and use the Galerkin method to prove the existence and uniqueness of weak solutions ...
  41. [41]
    [PDF] 2 Heat Equation
    Now differentiating with respect to x1, we have ut(x1,t) = kuxx(x1,t). Or, ut = kuxx. This is known as the diffusion equation. 2.1.2 Heat Flow. We ...
  42. [42]
    [PDF] Notes on Hölder Estimates for Parabolic PDE
    Jun 17, 2019 · These are lecture notes on parabolic differential equations, with a focus on estimates in Hölder spaces. The two main goals of our dis- cussion ...
  43. [43]
    (PDF) Failure of the strong maximum principle in nonlinear diffusion ...
    Aug 7, 2025 · SMP can have problems even for linear equations if the coefficients are not regular. Thus, Brezis and Ponce (2003) discuss the validity of the ...
  44. [44]
    [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 ...
  45. [45]
    [PDF] THE MATHEMATICS OF DIFFUSION - eng . lbl . gov
    The book contains a collection of mathematical solutions of the differential equations of diffusion and methods of obtaining them. They are discussed.
  46. [46]
    [PDF] The analytical theory of heat
    JOSEPH FOURIER. TRANSLATED, WITH NOTES,. BY. ALEXANDER FREEMAN, M.A.,. FJIlLLOW ... Equation of the M 0tJement of Heat. SECTION I. EQUATION 01' TD VABmD ...
  47. [47]
    The origin and present status of Fick's diffusion law - ACS Publications
    Journal of Chemical Education. Cite this: J. Chem. Educ. 1964, 41, 7 ...
  48. [48]
    Fick's Law - an overview | ScienceDirect Topics
    Fick's Law of diffusion was described in Section 7.3 involving a first-order differential equation that describes change in concentration as a function of time ...
  49. [49]
    Determination of Mass Diffusivity of Simple Sugars in Water by the ...
    The diffusivities of sucrose and glucose in water were de- termined to be 0.50 and 0.66 (m%ec X 109) at 25°C with activation energies of 36.3 and 31.6 Id/g-mol ...
  50. [50]
    Neutron Diffusion Equation - an overview | ScienceDirect Topics
    The neutron diffusion equation refers to a mathematical formulation that describes the distribution and flow of neutrons within a medium, applicable in both ...
  51. [51]
    [PDF] Module 6: Neutron Diffusion Dr. John H. Bickel
    • Stacey notes in Nuclear Reactor Physics (p. 49):. • “Diffusion theory is a strictly valid mathematical description of neutron flux….when assumptions used ...
  52. [52]
    Nernst-Planck Equation - an overview | ScienceDirect Topics
    According to Nernst, ions start drifting inside an electrolyte solution under the influence of external applied electric field as governed by the Nernst–Planck ...
  53. [53]
    Diffusion-Equation-Based Electrical Modeling for High-Power ...
    Jul 3, 2024 · In this work, a novel electrical model based on the solid-phase diffusion equation is proposed to capture the unique electrochemical phenomena ...
  54. [54]
    Electrochemical Impedance Spectroscopy A Tutorial
    This tutorial provides the theoretical background, the principles, and applications of Electrochemical Impedance Spectroscopy (EIS) in various research and ...
  55. [55]
    [PDF] Scale-space and edge detection using anisotropic diffusion
    Abstract The scale-space technique introduced by Witkin involves generating coarser resolution images by convolving the original image with a Gaussian kernel.
  56. [56]
    Navier-Stokes Equations
    The equations were derived independently by G.G. Stokes, in England, and M. Navier, in France, in the early 1800's. The equations are extensions of the Euler ...Missing: original | Show results with:original
  57. [57]
    [PDF] Chapter 8: Diffusion
    Diffusion and ion implantation are the two key processes to introduce a controlled amount of dopants into semiconductors and to alter the conductivity.
  58. [58]
    [PDF] Fischer Black and Myron Scholes Source: The Journal of Political Eco
    Author(s): Fischer Black and Myron Scholes. Source: The Journal of Political Economy, Vol. 81, No. 3 (May - Jun., 1973), pp. 637-654. Published by: The ...
  59. [59]
    Chapter 38. Fast Fluid Dynamics Simulation on the GPU
    This chapter describes a method for fast, stable fluid simulation that runs entirely on the GPU. It introduces fluid dynamics and the associated mathematics.
  60. [60]
    [PDF] Denoising Diffusion Probabilistic Models
    We show that diffusion models actually are capable of generating high quality samples, sometimes better than the published results on other types of generative ...
  61. [61]
    Score-Based Generative Modeling through Stochastic Differential ...
    Nov 26, 2020 · This paper introduces a stochastic differential equation (SDE) that transforms data to a prior distribution by injecting noise, and back by ...