Convection–diffusion equation


The convection–diffusion equation is a combination of the diffusion and convection equations, and describes physical phenomena where particles, energy, or other physical quantities are transferred inside a physical system due to two processes: diffusion and convection. Depending on context, the same equation can be called the advection–diffusion equation, drift–diffusion equation, or scalar transport equation.

Equation

General

The general equation is
where
The right-hand side of the equation is the sum of three contributions.
In a common situation, the diffusion coefficient is constant, there are no sources or sinks, and the velocity field describes an incompressible flow. Then the formula simplifies to:
In this form, the convection–diffusion equation combines both parabolic and hyperbolic partial differential equations.
In non-interacting material, , hence the transport equation is simply:
Using Fourier transform in both temporal and spatial domain, its characteristic equation can be obtained:
which gives the general solution:
where is any differentiable scalar function. This is the basis of temperature measurement for near Bose–Einstein condensate via time of flight method.

Stationary version

The stationary convection–diffusion equation describes the steady-state behavior of a convective-diffusive system. In a steady state,, so the formula is:

Derivation

The convection–diffusion equation can be derived in a straightforward way from the continuity equation, which states that the rate of change for a scalar quantity in a differential control volume is given by flow and diffusion into and out of that part of the system along with any generation or consumption inside the control volume:
where is the total flux and is a net volumetric source for. There are two sources of flux in this situation. First, diffusive flux arises due to diffusion. This is typically approximated by Fick's first law:
i.e., the flux of the diffusing material in any part of the system is proportional to the local concentration gradient. Second, when there is overall convection or flow, there is an associated flux called advective flux:
The total flux is given by the sum of these two:
Plugging into the continuity equation:

Complex mixing phenomena

In general,,, and may vary with space and time. In cases in which they depend on concentration as well, the equation becomes nonlinear, giving rise to many distinctive mixing phenomena such as Rayleigh–Bénard convection when depends on temperature in the heat transfer formulation and reaction–diffusion pattern formation when depends on concentration in the mass transfer formulation.

Velocity in response to a force

In some cases, the average velocity field exists because of a force; for example, the equation might describe the flow of ions dissolved in a liquid, with an electric field pulling the ions in some direction. In this situation, it is usually called the drift–diffusion equation or the Smoluchowski equation, after Marian Smoluchowski who described it in 1915.
Typically, the average velocity is directly proportional to the applied force, giving the equation:
where is the force, and characterizes the friction or viscous drag.

Derivation of Einstein relation

When the force is associated with a potential energy , a steady-state solution to the above equation is:
. In other words, there are more particles where the energy is lower. This concentration profile is expected to agree with the Boltzmann distribution. From this assumption, the Einstein relation can be proven:

Smoluchowski convection-diffusion equation

The Smoluchowski convective-diffusion equation is a stochastic diffusion equation with an additional convective flow-field,
In this case, the force describes the conservative interparticle interaction force between two colloidal particles or the intermolecular interaction force between two molecules in the fluid, and it is unrelated to the externally imposed flow velocity. The steady-state version of this equation is the basis to provide a description of the pair distribution function of colloidal suspensions under shear flows.
An approximate solution to the steady-state version of this equation has been found using the method of matched asymptotic expansions. This solution provides a theory for the transport-controlled reaction rate of two molecules in a shear flow, and also provides a way to extend the DLVO theory of colloidal stability to colloidal systems subject to shear flows.
The full solution to the steady-state equation, obtained using the method of matched asymptotic expansions, has been developed by Alessio Zaccone and L. Banetta to compute the pair distribution function of Lennard-Jones interacting particles in shear flow and subsequently extended to compute the pair distribution function of charge-stabilized colloidal particles in shear flows.

As a stochastic differential equation

The convection–diffusion equation can be viewed as a stochastic differential equation, describing random motion with diffusivity and bias. For example, the equation can describe the Brownian motion of a single particle, where the variable describes the probability distribution for the particle to be in a given position at a given time. The reason the equation can be used that way is because there is no mathematical difference between the probability distribution of a single particle, and the concentration profile of a collection of infinitely many particles.
The Langevin equation describes advection, diffusion, and other phenomena in an explicitly stochastic way. One of the simplest forms of the Langevin equation is when its "noise term" is Gaussian; in this case, the Langevin equation is exactly equivalent to the convection–diffusion equation. However, the Langevin equation is more general.

Numerical solution

The convection–diffusion equation can only rarely be solved with a pen and paper. More often, computers are used to numerically approximate the solution to the equation, typically using the finite element method. For more details and algorithms see: Numerical solution of the convection–diffusion equation.

Similar equations in other contexts

The convection–diffusion equation is a relatively simple equation describing flows, or alternatively, describing a stochastically-changing system. Therefore, the same or similar equation arises in many contexts unrelated to flows through space.
where is the momentum of the fluid at each point, is viscosity, is fluid pressure, and is any other body force such as gravity. In this equation, the term on the left-hand side describes the change in momentum at a given point; the first term on the right describes viscosity, which is really the diffusion of momentum; the second term on the right describes the advective flow of momentum; and the last two terms on the right describes the external and internal forces which can act as sources or sinks of momentum.

In biology

In biology, the reaction–diffusion–advection equation is used to model chemotaxis observed in bacteria, population migration, evolutionary adaptation to changing environments, and the spatiotemporal dynamics of molecular species including morphogenesis. An example is a study of VEGFC patterning in the context of lymphangiogenesis.

In semiconductor physics

In semiconductor physics, this equation is called the drift–diffusion equation. The word "drift" is related to drift current and drift velocity. The equation is normally written:
where
The diffusion coefficient and mobility are related by the Einstein relation as above:
where is the Boltzmann constant and is absolute temperature. The drift current and diffusion current refer separately to the two terms in the expressions for, namely:
This equation can be solved together with Poisson's equation numerically.
An example of results of solving the drift diffusion equation is shown on the right. When light shines on the center of semiconductor, carriers are generated in the middle and diffuse towards two ends. The drift–diffusion equation is solved in this structure and electron density distribution is displayed in the figure. One can see the gradient of carrier from center towards two ends.