Jump to content

Fokker–Planck equation: Difference between revisions

From Wikipedia, the free encyclopedia
Content deleted Content added
m Running Unicodifying and General Cleanup filers using AWB
Kaslanidi (talk | contribs)
Line 110: Line 110:
==External links==
==External links==
*[http://jeff560.tripod.com/f.html Fokker–Planck equation] on the [http://jeff560.tripod.com/mathword.html Earliest known uses of some of the words of mathematics]
*[http://jeff560.tripod.com/f.html Fokker–Planck equation] on the [http://jeff560.tripod.com/mathword.html Earliest known uses of some of the words of mathematics]
*[http://www.opentradingsystem.com/quantNotes/Backward_Kolmogorov_s_equation_.html Backward Kolmogorov's equation] Derivation of backward Kolmogorov's equation for various situations (accumulation of payoff, stopping time and so fourth).


[[Category:Stochastic processes]]
[[Category:Stochastic processes]]

Revision as of 18:58, 30 December 2009

A solution to the one-dimensional Fokker–Planck equation, with both the drift and the diffusion term. The initial condition is a Dirac delta function in x = 1, and the distribution drifts towards x = 0.

The Fokker–Planck equation describes the time evolution of the probability density function of the position of a particle, and can be generalized to other observables as well.[1] It is named after Adriaan Fokker and Max Planck and is also known as the Kolmogorov forward equation. The first use of the Fokker–Planck equation was for the statistical description of Brownian motion of a particle in a fluid. The first consistent microscopic derivation of the Fokker–Planck equation in the single scheme of classical and quantum mechanics was performed[2] by Nikolay Bogoliubov and Nikolay Krylov.[3]

In one spatial dimension x, the Fokker–Planck equation for a process with drift D1(x,t) and diffusion D2(x,t) is

More generally, the time-dependent probability distribution may depend on a set of macrovariables . The general form of the Fokker–Planck equation is then

where is the drift vector and the diffusion tensor; the latter results from the presence of the stochastic force.

Relationship with stochastic differential equations

The Fokker–Planck equation can be used for computing the probability densities of stochastic differential equations. Consider the Itō stochastic differential equation

where is the state and is a standard M-dimensional Wiener process. If the initial distribution is , then the probability density of the state is given by the Fokker–Planck equation with the drift and diffusion terms

Similarly, a Fokker–Planck equation can be derived for Stratonovich stochastic differential equations. In this case, noise-induced drift terms appear if the noise strength is state-dependent.

Examples

A standard scalar Wiener process is generated by the stochastic differential equation

Here the drift term is zero and the diffusion coefficient is 1/2; thus the corresponding Fokker–Planck equation is

that is the simplest form of diffusion equation. If the initial condition is , the solution is

Computational considerations

Brownian motion follows the Langevin equation, which can be solved for many different stochastic forcings with results being averaged (the Monte Carlo method, canonical ensemble in molecular dynamics). However, instead of this computationally intensive approach, one can use the Fokker–Planck equation and consider , that is, the probability of the particle having a velocity in the interval when it starts its motion with at time 0.

Solution

Being a partial differential equation, the Fokker–Planck equation can be solved analytically only in special cases. A formal analogy of the Fokker–Planck equation with the Schrodinger equation allows the use of advanced operator techniques known from quantum mechanics for its solution in a number of cases. In many applications, one is only interested in the steady-state probability distribution , which can be found from . The computation of mean first passage times and splitting probabilities can be reduced to the solution of an ordinary differential equation which is intimately related to the Fokker–Planck equation.

Particular cases with known solution and inversion

In mathematical finance for volatility smile modeling of options via local volatility, one has the problem of deriving a diffusion coefficient consistent with a probability density obtained from market option quotes. The problem is therefore an inversion of the Fokker Planck-equation: Given the density f(x,t) of the option underlying X deduced from the option market, one aims at finding the local volatility consistent with f. This is an inverse problem that has been solved in general by Dupire (1994, 1997) with a non-parametric solution. Brigo and Mercurio (2002, 2003) propose a solution in parametric form via a particular local volatility consistent with a solution of the Fokker-Plank equation given by a mixture model. More information is available also in Fengler (2008), Gatheral (2008) and Musiela and Rutkowski (2008).

Fokker-Planck equation and Path Integral

Every Fokker–Planck equation is equivalent to a path integral. The path integral formulation is an excellent[citation needed] starting point for the application of field theory methods.[4] This is used, for instance, in critical dynamics. A derivation of the path integral is possible in the same way as in quantum mechanics, simply because the FP equation is formally equivalent to the Schrödinger equation. Here are the steps for a FP equation with one variable x. Write the FP equation in the form

Integrate over a time interval ,

Insert the Fourier integral

for the -function,

This equation expresses as functional of . Iterating times and performing the limit gives a path integral with Lagrangian

The variables conjugate to are called "response variables".[5] As in quantum mechanics it now would be possible to perform the path integrals. But in the FP case this normally is not a good idea[citation needed] - the perturbation theory is simpler[5] with response variables.

See also

Notes and References

  1. ^ Leo P. Kadanoff (2000). Statistical Physics: statics, dynamics and renormalization. World Scientific. ISBN 9810237642.
  2. ^ N. N. Bogolyubov (jr) and D. P. Sankovich (1994). "N. N. Bogolyubov and statistical mechanics". Russian Math. Surveys 49(5): 19—49.
  3. ^ N. N. Bogoliubov and N. M. Krylov (1939). Fokker-Planck equations generated in perturbation theory by a method based on the spectral properties of a perturbed Hamiltonian. Zapiski Kafedry Fiziki Akademii Nauk Ukrainian SSR 4: 81-157 (in Ukrainian).
  4. ^ Zinn-Justin, Jean (1996). Quantum field theory and critical phenomena. Oxford: Clarendon Press. ISBN 0-19-851882-X.
  5. ^ a b Janssen, H. K. (1976). "On a Lagrangean for Classical Field Dynamics and Renormalization Group Calculation of Dynamical Critical Properties". Z. Physik. B23: 377–380.

Further reading

  • Bruno Dupire (1994) Pricing with a Smile. Risk Magazine, January, 18-20.
  • Dupire, B. (1997). Pricing and Hedging with Smiles. Mathematics of Derivative Securities. Edited by M.A.H. Dempster and S.R. Pliska, Cambridge University Press, Cambridge, 103-111.
  • Damiano Brigo, Fabio Mercurio, Lognormal-mixture dynamics and calibration to market volatility smiles, International Journal of Theoretical and Applied Finance, 2002, Vol: 5, Pages: 427 - 446
  • Brigo, D, Mercurio, F, Sartorelli, G, Alternative asset-price dynamics and volatility smile, QUANT FINANC, 2003, Vol: 3, Pages: 173 - 183, ISSN: 1469-7688
  • Fengler, M. R. (2008). Semiparametric Modeling of Implied Volatility, 2005, Springer Verlag.
  • Crispin W. Gardiner, "Handbook of Stochastic Methods", 3rd edition (paperback), Springer, ISBN 3-540-20882-8.
  • Jim Gatheral (2008). The Volatility Surface. Wiley and Sons.
  • Marek Musiela, Marek Rutkowski. Martingale Methods in Financial Modelling, 2008, 2nd Edition, Springer-Verlag.
  • Hannes Risken, "The Fokker–Planck Equation: Methods of Solutions and Applications", 2nd edition, Springer Series in Synergetics, Springer, ISBN 3-540-61530-X.

External links