Jump to content
 







Main menu
   


Navigation  



Main page
Contents
Current events
Random article
About Wikipedia
Contact us
Donate
 




Contribute  



Help
Learn to edit
Community portal
Recent changes
Upload file
 








Search  

































Create account

Log in
 









Create account
 Log in
 




Pages for logged out editors learn more  



Contributions
Talk
 



















Contents

   



(Top)
 


1 Principle  





2 Example: 1D diffusion  





3 Example: 1D diffusion with advection for steady flow, with multiple channel connections  





4 Example: 2D diffusion  





5 CrankNicolson for nonlinear problems  





6 Application in financial mathematics  





7 See also  





8 References  





9 External links  














CrankNicolson method






Deutsch
Español
فارسی
Français

Polski
Português
Українська

 

Edit links
 









Article
Talk
 

















Read
Edit
View history
 








Tools
   


Actions  



Read
Edit
View history
 




General  



What links here
Related changes
Upload file
Special pages
Permanent link
Page information
Cite this page
Get shortened URL
Download QR code
Wikidata item
 




Print/export  



Download as PDF
Printable version
 
















Appearance
   

 






From Wikipedia, the free encyclopedia
 

(Redirected from CrankNicolson)

Innumerical analysis, the Crank–Nicolson method is a finite difference method used for numerically solving the heat equation and similar partial differential equations.[1] It is a second-order method in time. It is implicit in time, can be written as an implicit Runge–Kutta method, and it is numerically stable. The method was developed by John Crank and Phyllis Nicolson in the 1940s.[2]

For diffusion equations (and many other equations), it can be shown the Crank–Nicolson method is unconditionally stable.[3] However, the approximate solutions can still contain (decaying) spurious oscillations if the ratio of time step times the thermal diffusivity to the square of space step, , is large (typically, larger than 1/2 per Von Neumann stability analysis). For this reason, whenever large time steps or high spatial resolution is necessary, the less accurate backward Euler method is often used, which is both stable and immune to oscillations.[citation needed]

Principle

[edit]
The Crank–Nicolson stencil for a 1D problem

The Crank–Nicolson method is based on the trapezoidal rule, giving second-order convergence in time. For linear equations, the trapezoidal rule is equivalent to the implicit midpoint method[citation needed]—the simplest example of a Gauss–Legendre implicit Runge–Kutta method—which also has the property of being a geometric integrator. For example, in one dimension, suppose the partial differential equationis

Letting and evaluated for and , the equation for Crank–Nicolson method is a combination of the forward Euler methodat and the backward Euler methodat (note, however, that the method itself is not simply the average of those two methods, as the backward Euler equation has an implicit dependence on the solution):

forward Euler
backward Euler
Crank–Nicolson

Note that this is an implicit method: to get the "next" value of in time, a system of algebraic equations must be solved. If the partial differential equation is nonlinear, the discretization will also be nonlinear, so that advancing in time will involve the solution of a system of nonlinear algebraic equations, though linearizations are possible. In many problems, especially linear diffusion, the algebraic problem is tridiagonal and may be efficiently solved with the tridiagonal matrix algorithm, which gives a fast direct solution, as opposed to the usual for a full matrix, in which indicates the matrix size.

Example: 1D diffusion

[edit]

The Crank–Nicolson method is often applied to diffusion problems. As an example, for linear diffusion,

applying a finite difference spatial discretization for the right-hand side, the Crank–Nicolson discretization is then

or, letting ,

Given that the terms on the right-hand side of the equation are known, this is a tridiagonal problem, so that may be efficiently solved by using the tridiagonal matrix algorithm in favor over the much more costly matrix inversion.

A quasilinear equation, such as (this is a minimalistic example and not general)

would lead to a nonlinear system of algebraic equations, which could not be easily solved as above; however, it is possible in some cases to linearize the problem by using the old value for , that is, instead of . Other times, it may be possible to estimate using an explicit method and maintain stability.

Example: 1D diffusion with advection for steady flow, with multiple channel connections

[edit]

This is a solution usually employed for many purposes when there is a contamination problem in streams or rivers under steady flow conditions, but information is given in one dimension only. Often the problem can be simplified into a 1-dimensional problem and still yield useful information.

Here we model the concentration of a solute contaminant in water. This problem is composed of three parts: the known diffusion equation ( chosen as constant), an advective component (which means that the system is evolving in space due to a velocity field), which we choose to be a constant , and a lateral interaction between longitudinal channels ():

(1)

where is the concentration of the contaminant, and subscripts and correspond to previous and next channel.

The Crank–Nicolson method (where represents position, and time) transforms each component of the PDE into the following:

(2)
(3)
(4)
(5)
(6)
(7)

Now we create the following constants to simplify the algebra:

and substitute (2), (3), (4), (5), (6), (7), , and into (1). We then put the new time terms on the left () and the present time terms on the right () to get

To model the first channel, we realize that it can only be in contact with the following channel (), so the expression is simplified to

In the same way, to model the last channel, we realize that it can only be in contact with the previous channel (), so the expression is simplified to

To solve this linear system of equations, we must now see that boundary conditions must be given first to the beginning of the channels:

: initial condition for the channel at present time step,
: initial condition for the channel at next time step,
: initial condition for the previous channel to the one analyzed at present time step,
: initial condition for the next channel to the one analyzed at present time step.

For the last cell of the channels (), the most convenient condition becomes an adiabatic one, so

This condition is satisfied if and only if (regardless of a null value)

Let us solve this problem (in a matrix form) for the case of 3 channels and 5 nodes (including the initial boundary condition). We express this as a linear system problem:

where

Now we must realize that AA and BB should be arrays made of four different subarrays (remember that only three channels are considered for this example, but it covers the main part discussed above):

where the elements mentioned above correspond to the next arrays, and an additional 4×4 full of zeros. Please note that the sizes of AA and BB are 12×12:

The d vector here is used to hold the boundary conditions. In this example it is a 12×1 vector:

To find the concentration at any time, one must iterate the following equation:

Example: 2D diffusion

[edit]

When extending into two dimensions on a uniform Cartesian grid, the derivation is similar and the results may lead to a system of band-diagonal equations rather than tridiagonal ones. The two-dimensional heat equation

can be solved with the Crank–Nicolson discretization of

assuming that a square grid is used, so that . This equation can be simplified somewhat by rearranging terms and using the CFL number

For the Crank–Nicolson numerical scheme, a low CFL number is not required for stability, however, it is required for numerical accuracy. We can now write the scheme as

Solving such a linear system is costly. Hence an alternating-direction implicit method can be implemented to solve the numerical PDE, whereby one dimension is treated implicitly, and other dimension explicitly for half of the assigned time step and conversely for the remainder half of the time step. The benefit of this strategy is that the implicit solver only requires a tridiagonal matrix algorithm to be solved. The difference between the true Crank–Nicolson solution and ADI approximated solution has an order of accuracy of and hence can be ignored with a sufficiently small time step.[4]

Crank–Nicolson for nonlinear problems

[edit]

Because the Crank–Nicolson method is implicit, it is generally impossible to solve exactly. Instead, an iterative technique should be used to converge to the solution. One option is to use Newton's method to converge on the prediction, but this requires the computation of the Jacobian. For a high-dimensional system like those in computational fluid dynamicsornumerical relativity, it may be infeasible to compute this Jacobian.

A Jacobian-free alternative is fixed-point iteration. If is the velocity of the system, then the Crank–Nicolson prediction will be a fixed point of the map If the map iteration does not converge, the parameterized map , with , may be better behaved. In expanded form, the update formula is

where is the current guess and is the previous time-step.

Even for high-dimensional systems, iteration of this map can converge surprisingly quickly.

A numerical solution of the Navier–Stokes equations in the vorticity form. In this case was needed for fixed-point iteration of Crank–Nicolson to converge.

Application in financial mathematics

[edit]

Because a number of other phenomena can be modeled with the heat equation (often called the diffusion equation in financial mathematics), the Crank–Nicolson method has been applied to those areas as well.[5] Particularly, the Black–Scholes option pricing model's differential equation can be transformed into the heat equation, and thus numerical solutions for option pricing can be obtained with the Crank–Nicolson method.

The importance of this for finance is that option pricing problems, when extended beyond the standard assumptions (e.g. incorporating changing dividends), cannot be solved in closed form, but can be solved using this method. Note however, that for non-smooth final conditions (which happen for most financial instruments), the Crank–Nicolson method is not satisfactory as numerical oscillations are not damped. For vanilla options, this results in oscillation in the gamma value around the strike price. Therefore, special damping initialization steps are necessary (e.g., fully implicit finite difference method).

See also

[edit]

References

[edit]
  1. ^ Tuncer Cebeci (2002). Convective Heat Transfer. Springer. ISBN 0-9668461-4-1.
  • ^ Crank, J.; Nicolson, P. (1947). "A practical method for numerical evaluation of solutions of partial differential equations of the heat conduction type". Proc. Camb. Phil. Soc. 43 (1): 50–67. Bibcode:1947PCPS...43...50C. doi:10.1017/S0305004100023197. S2CID 16676040.
  • ^ Thomas, J. W. (1995). Numerical Partial Differential Equations: Finite Difference Methods. Texts in Applied Mathematics. Vol. 22. Berlin, New York: Springer-Verlag. ISBN 978-0-387-97999-1.. Example 3.3.2 shows that Crank–Nicolson is unconditionally stable when applied to .
  • ^ "Multi-Dimensional Parabolic Problems" (PDF). Computer Science Department. RPI. Retrieved 29 May 2016.
  • ^ Wilmott, P.; Howison, S.; Dewynne, J. (1995). The Mathematics of Financial Derivatives: A Student Introduction. Cambridge Univ. Press. ISBN 0-521-49789-2. The Mathematics of Financial Derivatives Wilmott.

  • [edit]
    Retrieved from "https://en.wikipedia.org/w/index.php?title=Crank–Nicolson_method&oldid=1231136353"

    Categories: 
    Mathematical finance
    Numerical differential equations
    Finite differences
    Hidden categories: 
    Articles with short description
    Short description matches Wikidata
    All articles with unsourced statements
    Articles with unsourced statements from July 2016
    Articles with unsourced statements from September 2019
     



    This page was last edited on 26 June 2024, at 17:37 (UTC).

    Text is available under the Creative Commons Attribution-ShareAlike License 4.0; additional terms may apply. By using this site, you agree to the Terms of Use and Privacy Policy. Wikipedia® is a registered trademark of the Wikimedia Foundation, Inc., a non-profit organization.



    Privacy policy

    About Wikipedia

    Disclaimers

    Contact Wikipedia

    Code of Conduct

    Developers

    Statistics

    Cookie statement

    Mobile view



    Wikimedia Foundation
    Powered by MediaWiki