Brillouin and Langevin functions

The Brillouin and Langevin functions are a pair of special functions that appear when studying an idealized paramagnetic material in statistical mechanics.

Brillouin function

The Brillouin function[1][2] is a special function defined by the following equation:

B_J(x) = \frac{2J + 1}{2J} \coth \left ( \frac{2J + 1}{2J} x \right )
                - \frac{1}{2J} \coth \left ( \frac{1}{2J} x \right )

The function is usually applied (see below) in the context where x is a real variable and J is a positive integer or half-integer. In this case, the function varies from -1 to 1, approaching +1 as x \to +\infty and -1 as x \to -\infty.

The function is best known for arising in the calculation of the magnetization of an ideal paramagnet. In particular, it describes the dependency of the magnetization M on the applied magnetic field B and the total angular momentum quantum number J of the microscopic magnetic moments of the material. The magnetization is given by:[1]

M = N g \mu_B J \cdot B_J(x)

where

x = \frac{g \mu_B J B}{k_B T}

Note that in the SI system of units B given in Tesla stands for the magnetic field, B=\mu_0 H, where H is the auxiliary magnetic field given in A/m and \mu_0 is the permeability of vacuum.

Langevin function

Langevin function (blue line), compared with \tanh(x/3) (magenta line).

In the classical limit, the moments can be continuously aligned in the field and J can assume all values (J \to \infty). The Brillouin function is then simplified into the Langevin function, named after Paul Langevin:

L(x) = \coth(x) - \frac{1}{x}

For small values of x, the Langevin function can be approximated by a truncation of its Taylor series:


   L(x) = \tfrac{1}{3} x - \tfrac{1}{45} x^3 + \tfrac{2}{945} x^5 - \tfrac{1}{4725} x^7 + \dots

An alternative better behaved approximation can be derived from the Lambert's continued fraction expansion of tanh(x):


L(x) = \frac{x}{3+\tfrac{x^2}{5+\tfrac{x^2}{7+\tfrac{x^2}{9+\ldots}}}}

For small enough x, both approximations are numerically better than a direct evaluation of the actual analytical expression, since the latter suffers from Loss of significance.

The inverse Langevin function L−1(x) is defined on the open interval (−1, 1). For small values of x, it can be approximated by a truncation of its Taylor series[3]


   L^{-1}(x) = 3 x + \tfrac{9}{5} x^3 + \tfrac{297}{175} x^5 + \tfrac{1539}{875} x^7 + \dots

and by the Padé approximant


   L^{-1}(x) = 3x \frac{35-12x^2}{35-33x^2} + O(x^7).

[[File:Cohen and Jedynak approximations.gif|thumb|Graphs of relative error for x ∈ [0, 1) for Cohen and Jedynak approximations]]

Since this function has no closed form, it is useful to have approximations valid for arbitrary values of x. One popular approximation, valid on the whole range (−1, 1), has been published by A. Cohen:[4]


   L^{-1}(x) \approx x \frac{3-x^2}{1-x^2}.

This has a maximum relative error of 4.9% at the vicinity of x = ±0.8. Greater accuracy can be achieved by using the formula given by R. Jedynak:[5]


   L^{-1}(x) \approx x \frac{3.0-2.6x+0.7x^2}{(1-x)(1+0.1x)},

valid for x ≥ 0. The maximal relative error for this approximation is 1.5% at the vicinity of x = 0.85.

Interesting and comprehensive studies of the well-known approximation formulas of the inverse Langevin function can be found in the paper written by Jedynak.[5]

High-temperature limit

When x \ll 1 i.e. when \mu_B B / k_B T is small, the expression of the magnetization can be approximated by the Curie's law:

M = C \cdot \frac{B}{T}

where C = \frac{N g^2 J(J+1) \mu_B^2}{3k_B} is a constant. One can note that g\sqrt{J(J+1)} is the effective number of Bohr magnetons.

High-field limit

When x\to\infty, the Brillouin function goes to 1. The magnetization saturates with the magnetic moments completely aligned with the applied field:

M = N g \mu_B J

References

  1. 1 2 3 C. Kittel, Introduction to Solid State Physics (8th ed.), pages 303-4 ISBN 978-0-471-41526-8
  2. Darby, M.I. (1967). "Tables of the Brillouin function and of the related function for the spontaneous magnetization". Brit. J. Appl. Phys. 18 (10): 1415–1417. Bibcode:1967BJAP...18.1415D. doi:10.1088/0508-3443/18/10/307.
  3. Johal, A. S.; Dunstan, D. J. (2007). "Energy functions for rubber from microscopic potentials". Journal of Applied Physics 101 (8): 084917. Bibcode:2007JAP...101h4917J. doi:10.1063/1.2723870.
  4. Cohen, A. (1991). "A Padé approximant to the inverse Langevin function". Rheologica Acta 30 (3): 270–273. doi:10.1007/BF00366640.
  5. 1 2 Jedynak, R. (2015). "Approximation of the inverse Langevin function revisited". Rheologica Acta 54 (1): 29–39. doi:10.1007/s00397-014-0802-2.
This article is issued from Wikipedia - version of the Wednesday, January 21, 2015. The text is available under the Creative Commons Attribution/Share Alike but additional terms may apply for the media files.