In mathematics, Mathieu functions, sometimes called angular Mathieu functions, are solutions of Mathieu's differential equation

where a, q are real-valued parameters. Since we may add π/2 to x to change the sign of q, it is a usual convention to set q ≥ 0.

They were first introduced by Émile Léonard Mathieu, who encountered them while studying vibrating elliptical drumheads.[1][2][3] They have applications in many fields of the physical sciences, such as optics, quantum mechanics, and general relativity. They tend to occur in problems involving periodic motion, or in the analysis of partial differential equation (PDE) boundary value problems possessing elliptic symmetry.[4]

Definition edit

Mathieu functions edit

In some usages, Mathieu function refers to solutions of the Mathieu differential equation for arbitrary values of   and  . When no confusion can arise, other authors use the term to refer specifically to  - or  -periodic solutions, which exist only for special values of   and  .[5] More precisely, for given (real)   such periodic solutions exist for an infinite number of values of  , called characteristic numbers, conventionally indexed as two separate sequences   and  , for  . The corresponding functions are denoted   and  , respectively. They are sometimes also referred to as cosine-elliptic and sine-elliptic, or Mathieu functions of the first kind.

As a result of assuming that   is real, both the characteristic numbers and associated functions are real-valued.[6]

  and   can be further classified by parity and periodicity (both with respect to  ), as follows:[5]

Function Parity Period
  even  
  even  
  odd  
  odd  

The indexing with the integer  , besides serving to arrange the characteristic numbers in ascending order, is convenient in that   and   become proportional to   and   as  . With   being an integer, this gives rise to the classification of   and   as Mathieu functions (of the first kind) of integral order. For general   and  , solutions besides these can be defined, including Mathieu functions of fractional order as well as non-periodic solutions.

Modified Mathieu functions edit

Closely related are the modified Mathieu functions, also known as radial Mathieu functions, which are solutions of Mathieu's modified differential equation

 

which can be related to the original Mathieu equation by taking  . Accordingly, the modified Mathieu functions of the first kind of integral order, denoted by   and  , are defined from[7]

 

These functions are real-valued when   is real.

Normalization edit

A common normalization,[8] which will be adopted throughout this article, is to demand

 

as well as require   and   as  .

Stability edit

The Mathieu equation has two parameters. For almost all choices of parameter, by Floquet theory (see next section), any solution either converges to zero or diverges to infinity.

Parametrize Mathieu equation as  , where  . The regions of stability and instability are separated by curves [9]

 

Floquet theory edit

Many properties of the Mathieu differential equation can be deduced from the general theory of ordinary differential equations with periodic coefficients, called Floquet theory. The central result is Floquet's theorem:

Floquet's theorem[10] — Mathieu's equation always has at least one solution   such that  , where   is a constant which depends on the parameters of the equation and may be real or complex.

It is natural to associate the characteristic numbers   with those values of   which result in  .[11] Note, however, that the theorem only guarantees the existence of at least one solution satisfying  , when Mathieu's equation in fact has two independent solutions for any given  ,  . Indeed, it turns out that with   equal to one of the characteristic numbers, Mathieu's equation has only one periodic solution (that is, with period   or  ), and this solution is one of the  ,  . The other solution is nonperiodic, denoted   and  , respectively, and referred to as a Mathieu function of the second kind.[12] This result can be formally stated as Ince's theorem:

Ince's theorem[13] — Define a basically periodic function as one satisfying  . Then, except in the trivial case  , Mathieu's equation never possesses two (independent) basically periodic solutions for the same values of   and  .

 
An example   from Floquet's theorem, with  ,  ,   (real part, red; imaginary part, green)

An equivalent statement of Floquet's theorem is that Mathieu's equation admits a complex-valued solution of form

 

where   is a complex number, the Floquet exponent (or sometimes Mathieu exponent), and   is a complex valued function periodic in   with period  . An example   is plotted to the right.

Other types of Mathieu functions edit

Second kind edit

Since Mathieu's equation is a second order differential equation, one can construct two linearly independent solutions. Floquet's theory says that if   is equal to a characteristic number, one of these solutions can be taken to be periodic, and the other nonperiodic. The periodic solution is one of the   and  , called a Mathieu function of the first kind of integral order. The nonperiodic one is denoted either   and  , respectively, and is called a Mathieu function of the second kind (of integral order). The nonperiodic solutions are unstable, that is, they diverge as  .[14]

The second solutions corresponding to the modified Mathieu functions   and   are naturally defined as   and  .

Fractional order edit

Mathieu functions of fractional order can be defined as those solutions   and  ,   a non-integer, which turn into   and   as  .[7] If   is irrational, they are non-periodic; however, they remain bounded as  .

An important property of the solutions   and  , for   non-integer, is that they exist for the same value of  . In contrast, when   is an integer,   and   never occur for the same value of  . (See Ince's Theorem above.)

These classifications are summarized in the table below. The modified Mathieu function counterparts are defined similarly.

Classification of Mathieu functions[15]
Order First kind Second kind
Integral    
Integral    
Fractional

(  non-integral)

   

Explicit representation and computation edit

First kind edit

Mathieu functions of the first kind can be represented as Fourier series:[5]

 

The expansion coefficients   and   are functions of   but independent of  . By substitution into the Mathieu equation, they can be shown to obey three-term recurrence relations in the lower index. For instance, for each   one finds[16]

 

Being a second-order recurrence in the index  , one can always find two independent solutions   and   such that the general solution can be expressed as a linear combination of the two:  . Moreover, in this particular case, an asymptotic analysis[17] shows that one possible choice of fundamental solutions has the property

 

In particular,   is finite whereas   diverges. Writing  , we therefore see that in order for the Fourier series representation of   to converge,   must be chosen such that   These choices of   correspond to the characteristic numbers.

In general, however, the solution of a three-term recurrence with variable coefficients cannot be represented in a simple manner, and hence there is no simple way to determine   from the condition  . Moreover, even if the approximate value of a characteristic number is known, it cannot be used to obtain the coefficients   by numerically iterating the recurrence towards increasing  . The reason is that as long as   only approximates a characteristic number,   is not identically   and the divergent solution   eventually dominates for large enough  .

To overcome these issues, more sophisticated semi-analytical/numerical approaches are required, for instance using a continued fraction expansion,[18][5] casting the recurrence as a matrix eigenvalue problem,[19] or implementing a backwards recurrence algorithm.[17] The complexity of the three-term recurrence relation is one of the reasons there are few simple formulas and identities involving Mathieu functions.[20]

In practice, Mathieu functions and the corresponding characteristic numbers can be calculated using pre-packaged software, such as Mathematica, Maple, MATLAB, and SciPy. For small values of   and low order  , they can also be expressed perturbatively as power series of  , which can be useful in physical applications.[21]

Second kind edit

There are several ways to represent Mathieu functions of the second kind.[22] One representation is in terms of Bessel functions:[23]

 

where  , and   and   are Bessel functions of the first and second kind.

Modified functions edit

A traditional approach for numerical evaluation of the modified Mathieu functions is through Bessel function product series.[24] For large   and  , the form of the series must be chosen carefully to avoid subtraction errors.[25][26]

Properties edit

There are relatively few analytic expressions and identities involving Mathieu functions. Moreover, unlike many other special functions, the solutions of Mathieu's equation cannot in general be expressed in terms of hypergeometric functions. This can be seen by transformation of Mathieu's equation to algebraic form, using the change of variable  :

 

Since this equation has an irregular singular point at infinity, it cannot be transformed into an equation of the hypergeometric type.[20]

Qualitative behavior edit

 
Sample plots of Mathieu functions of the first kind
 
Plot of   for varying  

For small  ,   and   behave similarly to   and  . For arbitrary  , they may deviate significantly from their trigonometric counterparts; however, they remain periodic in general. Moreover, for any real  ,   and   have exactly   simple zeros in  , and as   the zeros cluster about  .[27][28]

For   and as   the modified Mathieu functions tend to behave as damped periodic functions.

In the following, the   and   factors from the Fourier expansions for   and   may be referenced (see Explicit representation and computation). They depend on   and   but are independent of  .

Reflections and translations edit

Due to their parity and periodicity,   and   have simple properties under reflections and translations by multiples of  :[7]

 

One can also write functions with negative   in terms of those with positive  :[5][29]

 

Moreover,

 

Orthogonality and completeness edit

Like their trigonometric counterparts   and  , the periodic Mathieu functions   and   satisfy orthogonality relations

 

Moreover, with   fixed and   treated as the eigenvalue, the Mathieu equation is of Sturm–Liouville form. This implies that the eigenfunctions   and   form a complete set, i.e. any  - or  -periodic function of   can be expanded as a series in   and  .[4]

Integral identities edit

Solutions of Mathieu's equation satisfy a class of integral identities with respect to kernels   that are solutions of

 

More precisely, if   solves Mathieu's equation with given   and  , then the integral

 

where   is a path in the complex plane, also solves Mathieu's equation with the same   and  , provided the following conditions are met:[30]

  •   solves  
  • In the regions under consideration,   exists and   is analytic
  •   has the same value at the endpoints of  

Using an appropriate change of variables, the equation for   can be transformed into the wave equation and solved. For instance, one solution is  . Examples of identities obtained in this way are[31]

 

Identities of the latter type are useful for studying asymptotic properties of the modified Mathieu functions.[32]

There also exist integral relations between functions of the first and second kind, for instance:[23]

 

valid for any complex   and real  .

Asymptotic expansions edit

The following asymptotic expansions hold for  ,  ,  , and  :[33]

 

Thus, the modified Mathieu functions decay exponentially for large real argument. Similar asymptotic expansions can be written down for   and  ; these also decay exponentially for large real argument.

For the even and odd periodic Mathieu functions   and the associated characteristic numbers   one can also derive asymptotic expansions for large  .[34] For the characteristic numbers in particular, one has with   approximately an odd integer, i.e.  

 

Observe the symmetry here in replacing   and   by   and  , which is a significant feature of the expansion. Terms of this expansion have been obtained explicitly up to and including the term of order  .[35] Here   is only approximately an odd integer because in the limit of   all minimum segments of the periodic potential   become effectively independent harmonic oscillators (hence   an odd integer). By decreasing  , tunneling through the barriers becomes possible (in physical language), leading to a splitting of the characteristic numbers   (in quantum mechanics called eigenvalues) corresponding to even and odd periodic Mathieu functions. This splitting is obtained with boundary conditions[35] (in quantum mechanics this provides the splitting of the eigenvalues into energy bands).[36] The boundary conditions are:

 

Imposing these boundary conditions on the asymptotic periodic Mathieu functions associated with the above expansion for   one obtains

 

The corresponding characteristic numbers or eigenvalues then follow by expansion, i.e.

 

Insertion of the appropriate expressions above yields the result

 

For   these are the eigenvalues associated with the even Mathieu eigenfunctions   or   (i.e. with upper, minus sign) and odd Mathieu eigenfunctions   or   (i.e. with lower, plus sign). The explicit and normalised expansions of the eigenfunctions can be found in [35] or.[36]

Similar asymptotic expansions can be obtained for the solutions of other periodic differential equations, as for Lamé functions and prolate and oblate spheroidal wave functions.

Applications edit

Mathieu's differential equations appear in a wide range of contexts in engineering, physics, and applied mathematics. Many of these applications fall into one of two general categories: 1) the analysis of partial differential equations in elliptic geometries, and 2) dynamical problems which involve forces that are periodic in either space or time. Examples within both categories are discussed below.

Partial differential equations edit

Mathieu functions arise when separation of variables in elliptic coordinates is applied to 1) the Laplace equation in 3 dimensions, and 2) the Helmholtz equation in either 2 or 3 dimensions. Since the Helmholtz equation is a prototypical equation for modeling the spatial variation of classical waves, Mathieu functions can be used to describe a variety of wave phenomena. For instance, in computational electromagnetics they can be used to analyze the scattering of electromagnetic waves off elliptic cylinders, and wave propagation in elliptic waveguides.[37] In general relativity, an exact plane wave solution to the Einstein field equation can be given in terms of Mathieu functions.

More recently, Mathieu functions have been used to solve a special case of the Smoluchowski equation, describing the steady-state statistics of self-propelled particles.[38]

The remainder of this section details the analysis for the two-dimensional Helmholtz equation.[39] In rectangular coordinates, the Helmholtz equation is

 

Elliptic coordinates are defined by

 

where  ,  , and   is a positive constant. The Helmholtz equation in these coordinates is

 

The constant   curves are confocal ellipses with focal length  ; hence, these coordinates are convenient for solving the Helmholtz equation on domains with elliptic boundaries. Separation of variables via   yields the Mathieu equations

 

where   is a separation constant.

As a specific physical example, the Helmholtz equation can be interpreted as describing normal modes of an elastic membrane under uniform tension. In this case, the following physical conditions are imposed:[40]

  • Periodicity with respect to  , i.e.  
  • Continuity of displacement across the interfocal line:  
  • Continuity of derivative across the interfocal line:  

For given  , this restricts the solutions to those of the form   and  , where  . This is the same as restricting allowable values of  , for given  . Restrictions on   then arise due to imposition of physical conditions on some bounding surface, such as an elliptic boundary defined by  . For instance, clamping the membrane at   imposes  , which in turn requires

 

These conditions define the normal modes of the system.

Dynamical problems edit

In dynamical problems with periodically varying forces, the equation of motion sometimes takes the form of Mathieu's equation. In such cases, knowledge of the general properties of Mathieu's equation— particularly with regard to stability of the solutions—can be essential for understanding qualitative features of the physical dynamics.[41] A classic example along these lines is the inverted pendulum.[42] Other examples are

Quantum mechanics edit

Mathieu functions play a role in certain quantum mechanical systems, particularly those with spatially periodic potentials such as the quantum pendulum and crystalline lattices.

The modified Mathieu equation also arises when describing the quantum mechanics of singular potentials. For the particular singular potential   the radial Schrödinger equation

 

can be converted into the equation

 

The transformation is achieved with the following substitutions

 

By solving the Schrödinger equation (for this particular potential) in terms of solutions of the modified Mathieu equation, scattering properties such as the S-matrix and the absorptivity can be obtained.[44]

See also edit

Notes edit

  1. ^ Mathieu (1868).
  2. ^ Morse and Feshbach (1953).
  3. ^ Brimacombe, Corless and Zamir (2021)
  4. ^ a b Gutiérrez-Vega (2015).
  5. ^ a b c d e Arscott (1964), chapter III
  6. ^ Arscott (1964) 43–44
  7. ^ a b c McLachlan (1947), chapter II.
  8. ^ Arscott (1964); Iyanaga (1980); Gradshteyn (2007); This is also the normalization used by the computer algebra system Maple.
  9. ^ Butikov, Eugene I. (April 2018). "Analytical expressions for stability regions in the Ince–Strutt diagram of Mathieu equation". American Journal of Physics. 86 (4): 257–267. Bibcode:2018AmJPh..86..257B. doi:10.1119/1.5021895. ISSN 0002-9505.
  10. ^ Arscott (1964), p. 29.
  11. ^ It is not true, in general, that a   periodic function has the property  . However, this turns out to be true for functions which are solutions of Mathieu's equation.
  12. ^ McLachlan (1951), pp. 141-157, 372
  13. ^ Arscott (1964), p. 34
  14. ^ McLachlan (1947), p. 144
  15. ^ McLachlan (1947), p. 372
  16. ^ McLachlan (1947), p. 28
  17. ^ a b Wimp (1984), pp. 83-84
  18. ^ McLachlan (1947)
  19. ^ Chaos-Cador and Ley-Koo (2001)
  20. ^ a b Temme (2015), p. 234
  21. ^ Müller-Kirsten (2012), pp. 420-428
  22. ^ Meixner and Schäfke (1954); McLachlan (1947)
  23. ^ a b Malits (2010)
  24. ^ Jin and Zhang (1996)
  25. ^ Van Buren and Boisvert (2007)
  26. ^ Bibby and Peterson (2013)
  27. ^ Meixner and Schäfke (1954), p.134
  28. ^ McLachlan (1947), pp. 234–235
  29. ^ Gradshteyn (2007), p. 953
  30. ^ Arscott (1964), pp. 40-41
  31. ^ Gradshteyn (2007), pp. 763–765
  32. ^ Arscott (1964), p. 86
  33. ^ McLachlan (1947), chapter XI
  34. ^ McLachlan (1947), p. 237; Dingle and Müller (1962); Müller (1962); Dingle and Müller(1964)
  35. ^ a b c Dingle and Müller (1962)
  36. ^ a b Müller-Kirsten (2012)
  37. ^ Bibby and Peterson (2013); Barakat (1963); Sebak and Shafai (1991); Kretzschmar (1970)
  38. ^ Solon et al (2015)
  39. ^ see Willatzen and Voon (2011), pp. 61–65
  40. ^ McLachlan (1947), pp. 294–297
  41. ^ a b Meixner and Schäfke (1954), pp. 324–343
  42. ^ Ruby (1996)
  43. ^ March (1997)
  44. ^ Müller-Kirsten (2006)

References edit

External links edit