Convection–diffusion equation

Source: Wikipedia, the free encyclopedia.

The convection–diffusion equation is a combination of the

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,[1] or (generic) scalar transport equation.[2]

Equation

General

A contaminant plume moving in porous media via the Convection-Diffusive Equation(also called Advective-Diffusive Equation) (time stamps of 1 day, 85 days, 462 days and 674 days) with advection as primary transport mechanism.

The general equation in

conservative form is[3][4]
where

For example, if c is the concentration of a molecule, then R describes how the molecule can be created or destroyed by chemical reactions. R may be a function of c and of other parameters. Often there are several quantities, each with its own convection–diffusion equation, where the destruction of one quantity entails the creation of another. For example, when methane burns, it involves not only the destruction of methane and oxygen but also the creation of carbon dioxide and water vapor. Therefore, while each of these chemicals has its own convection–diffusion equation, they are coupled together and must be solved as a system of simultaneous differential equations.

  • represents gradient and ∇ ⋅ represents divergence. In this equation, c represents concentration gradient.

Understanding the current density terms involved

The convection-diffusion equation is a particular example of conservation equation. A conservation equation has the general form: Where jc is the current density term associated to the variable of interest c.

In a convection-diffusion equation, the current density of the quantity c is the sum of two terms:

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]

In this form, the convection–diffusion equation combines both

partial differential equations
.

In this case the equation can be put in the simple convective form:

where the derivative of the left hand side is the material derivative of the variable c. In non-interacting material, D=0 (for example, when temperature is close to absolute zero, dilute gas has almost zero mass diffusivity), hence the transport equation is simply the continuity equation:

Using

integral kernel
), 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[8] via time of flight method.[9]

Stationary version

The stationary convection–diffusion equation describes the

steady-state
behavior of a convective-diffusive system. In a steady state, c/t = 0, so the equation to solve becomes the second order equation:

One dimensional case

In one dimension the spatial gradient operator is simply:

so the equation to solve becomes the single-variable second order equation:

Which can be integrated one time in the space variable x to give:

Where D is not zero, this is an inhomogeneous first-order linear differential equation with variable coefficients in the variable c(x):

where the coefficients are: and:

This equation has in fact a relatively simple analytical solution (see the link above to first-order linear differential equation with variable coefficients).

On the other hand, in the positions x where D=0, the first-order diffusion term disappears and the solution becomes simply the ratio:

Derivation

The convection–diffusion equation can be derived in a straightforward way

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 j is the total
Fick's first law
: 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: The total flux (in a stationary coordinate system) is given by the sum of these two: Plugging into the continuity equation:

Complex mixing phenomena

In general, D, v, and R 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 v depends on temperature in the heat transfer formulation and reaction–diffusion pattern formation when R depends on concentration in the mass transfer formulation.

Velocity in response to a force

In some cases, the average velocity field 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[10] (not to be confused with the Einstein–Smoluchowski relation or Smoluchowski coagulation equation).

Typically, the average velocity is directly proportional to the applied force, giving the equation:[11][12] where F is the force, and ζ characterizes the friction or viscous drag. (The inverse ζ−1 is called mobility.)

Derivation of Einstein relation

When the force is associated with a

steady-state
solution to the above equation (i.e. 0 = R = c/t) is: (assuming D and ζ 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:[12]

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 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 a collection 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.[12] However, the Langevin equation is more general.[12]

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.

  • 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.[13]
  • 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:

where M is the momentum of the fluid (per unit volume) at each point (equal to the density ρ multiplied by the velocity v), μ 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 the diffusion of momentum by viscosity; 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

As carriers are generated (green:electrons and purple:holes) due to light shining at the center of an intrinsic semiconductor, they diffuse towards two ends. Electrons have higher diffusion constant than holes leading to fewer excess electrons at the center as compared to holes.

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:[14]
where

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

absolute temperature. The drift current and diffusion current
refer separately to the two terms in the expressions for J, namely:

This equation can be solved together with Poisson's equation numerically.[15]

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.

See also

References

  1. ^ . See equation (312)
  2. – via Google Books.
  3. – via Google Books.
  4. ^ a b Socolofsky, Scott A.; Jirka, Gerhard H. "Advective Diffusion Equation" (PDF). Lecture notes. Archived from the original (PDF) on June 25, 2010. Retrieved April 18, 2012.
  5. ^ Bejan A (2004). Convection Heat Transfer.
  6. ^ Bird, Stewart, Lightfoot (1960). Transport Phenomena.{{cite book}}: CS1 maint: multiple names: authors list (link)
  7. ^ Probstein R (1994). Physicochemical Hydrodynamics.
  8. .
  9. .
  10. .
  11. ^ "Smoluchowski Diffusion Equation" (PDF).
  12. ^ .
  13. .
  14. .
  15. .

Further reading