ART

.

The convection–diffusion equation is a combination of the diffusion and convection (advection) 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, Smoluchowski equation,[1] or (generic) scalar transport equation[2]

Equation
General

The general equation is[3][4]

\( \frac{\partial c}{\partial t} = \nabla \cdot (D \nabla c) - \nabla \cdot (\vec{v} c) + R \)

where

c is the variable of interest (species concentration for mass transfer, temperature for heat transfer),
D is the diffusivity (also called diffusion coefficient) for mass or heat transfer,
\( \vec{v} \) is the velocity,
R describes "sources" or "sinks" of the quantity c. For example, for a chemical species, R>0 means that a chemical reaction is creating more of the species, and R<0 means that a chemical reaction is destroying the species. For heat transport, R>0 might occur if there is friction generating thermal energy,
\nabla represents gradient and \( \nabla\cdot \( represents divergence.

Common simplifications

In a common situation, the diffusion coefficient is constant, there are no sources or sinks, and the velocity field describes an incompressible flow (i.e., it has zero divergence). Then the formula simplifies to:[5][6][7]

\( \frac{\partial c}{\partial t} = D \nabla^2 c - \vec{v} \cdot \nabla c. \)

In this form, the convection–diffusion equation combines both parabolic and hyperbolic partial differential equations.
Stationary version

The stationary convection–diffusion equation describes the steady-state behavior of a convective-diffusive system. In steady-state,\( \partial c/\partial t = 0, so the formula is:

\( 0 = \nabla \cdot (D \nabla c) - \nabla \cdot (\vec{v} c) + R \)

Derivation

The convection–diffusion equation can be derived in a straightforward way[4] 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:

\( \frac{\partial c}{\partial t} + \nabla\cdot\vec{j} = s, \)

where \vec{j} is the total flux and s is a net volumetric source for c. 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:

\( \vec{j}_{\text{diffusion}} = -D \, \nabla c \)

i.e., the flux of the diffusing material (relative to the bulk motion) 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:

\( \vec{j}_{\text{advective}} = \vec{v} \, c \)

The total flux (in a stationary coordinate system) is given by the sum of these two:

\( \vec{j} = \vec{j}_{\text{diffusion}} + \vec{j}_{\text{advective}} = -D \, \nabla c + \vec{v} \, c. \)

Plugging into the continuity equation:

\( \frac{\partial c}{\partial t} + \nabla\cdot \left(\vec{-D\,\nabla c + \vec{v}\, c}\right) = s. \)

Complex mixing phenomena

In general, D, \( \vec{v} \) , and s 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 \( \vec{v} \) depends on temperature in the heat transfer formulation and reaction-diffusion pattern formation when s depends on concentration in the mass transfer formulation.
Velocity in response to a force

In some cases, the average velocity field \( \vec{v} \) 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 (as in gel electrophoresis). In this situation, it is usually called the drift-diffusion equation or the Smoluchowski equation,[1] after Marian Smoluchowski who described it in 1915[8] (not to be confused with the Einstein–Smoluchowski relation or Smoluchowski coagulation equation).

Typically, the average velocity is typically directly proportional to the applied force, giving the equation:[9][10]

\( \frac{\partial c}{\partial t} = \nabla \cdot (D \nabla c) - \nabla \cdot (\zeta^{-1} \vec{F} c) + R \)

where \vec{F} is the force, and \zeta characterizes the friction or viscous drag. (The inverse \( \zeta^{-1} \) is called mobility.)
Derivation of Einstein relation
Main article: Einstein relation (kinetic theory)

When the force is associated with a potential energy \( \vec{F} = \nabla U ( \) see conservative force), a steady-state solution to the above equation (i.e. 0 = R = ∂c/∂t) is:

\( c \propto \exp( - D^{-1} \zeta^{-1} U) \)

(assuming D and \zeta are constant). In other words, there are more particles where the energy is lower. This concentration profile is expected to agree with the Boltzmann distribution (more precisely, the Gibbs measure). From this assumption, the Einstein relation can be proven: \( D \zeta = k_B T \) .[10]
As a stochastic differential equation

The convection–diffusion equation (with no sources or drains, R=0) can be viewed as a stochastic differential equation, describing random motion with diffusivity D and bias \vec{v}. For example, the equation can describe the Brownian motion of a single particle, where the variable c 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 an ensemble of infinitely many particles (as long as the particles do not interact with each other).

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.[10] However, the Langevin equation is more general.[10]
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.

It is formally identical to the Fokker–Planck equation for the velocity of a particle.
It is closely related to the Black–Scholes equation and other equations in financial mathematics.
It is closely related to the Navier–Stokes equations, because the flow of momentum in a fluid is mathematically similar to the flow of mass or energy. The correspondence is clearest in the case of an incompressible Newtonian fluid, in which case the Navier–Stokes equation is:

\( \frac{\partial \mathbf{M}}{\partial t} = \frac{\mu}{\rho} \nabla^2 \mathbf{M} -\mathbf{v} \cdot \nabla \mathbf{M} + (\mathbf{f}-\nabla \text{P}) \)

where M is the momentum of the fluid (per unit volume) at each point (equal to the density \( \rho \) multiplied by the velocity v), \( \mu \) is viscosity, P is fluid pressure, and f 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 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:

\( \mathbf{J}_n/(-q) = - D_n \nabla n - n \mu_n \mathbf{E} \)
\( \mathbf{J}_p/q = - D_p \nabla p + p \mu_p \mathbf{E} \)
\( \frac{\partial n}{\partial t} = -\nabla \cdot \mathbf{J}_n + R \)
\( \frac{\partial p}{\partial t} = -\nabla \cdot \mathbf{J}_p + R \)

where

  • n and p are the concentrations (densities) of electrons and holes, respectively,
  • q>0 is the elementary charge,
  • Jn and Jp are the electric currents due to electrons and holes respectively,
  • Jn/-q and Jp/q are the corresponding "particle currents" of electrons and holes respectively,
  • R represents carrier generation and recombination (R>0 for generation of electron-hole pairs, R<0 for recombination.)
  • E is the electric field vector
  • \( \mu_n \) and \( \mu_p \) are electron and hole mobility.

The diffusion coefficient and mobility are related by the Einstein relation as above:

\( D_n = \mu_n k_B T/q, \quad D_p = \mu_p k_B T/q, \)

where kB is Boltzmann constant and T is absolute temperature. The drift current and diffusion current refer separately to the two terms in the expressions for J, i.e.:

\( \mathbf{J}_{n,\text{drift}}/(-q) = - n \mu_n \mathbf{E}, \qquad \mathbf{J}_{p,\text{drift}}/q = p \mu_p \mathbf{E}
\mathbf{J}_{n,\text{diffusion}}/(-q) = - D_n \nabla n, \qquad \mathbf{J}_{p,\text{diffusion}}/q = - D_p \nabla p \)

See also

Nernst-Planck equation
Double diffusive convection
Natural convection
Buckley–Leverett equation

References

^ a b Chandrasekhar (1943). Rev. Mod. Phys. 15: 1. doi:10.1103/RevModPhys.15.1. See equation (312)
^ Computational Fluid Dynamics in Industrial Combustion by Baukal and Gershtein, p67, google books link.
^ Introduction to Climate Modelling, by Thomas Stocker, p57, google books link
^ a b Advective Diffusion Equation, lecture notes by Scott A. Socolofsky and Gerhard H. Jirka, web link
^ Bejan A (2004). Convection Heat Transfer.
^ Bird, Stewart, Lightfoot (1960). Transport Phenomena.
^ Probstein R (1994). Physicochemical Hydrodynamics.
^ M. v. Smoluchowski, Über Brownsche Molekularbewegung unter Einwirkung äußerer Kräfte und den Zusammenhang mit der verallgemeinerten Diffusionsgleichung, Ann. Phys. 353 (4. Folge 48), 1103–1112 (1915), PDF link
^ http://www.ks.uiuc.edu/~kosztin/PHYCS498NSM/LectureNotes/chp4.pdf
^ a b c d The Theory of Polymer Dynamics by Doi and Edwards, p46-52, google books link

Granville Sewell, The Numerical Solution of Ordinary and Partial Differential Equations, Academic Press (1988). ISBN 0-12-637475-9

Retrieved from "http://en.wikipedia.org/"
All text is available under the terms of the GNU Free Documentation License

Home - Hellenica World