Bernoulli polynomials

From Wikipedia, the free encyclopedia
(Redirected from Bernoulli polynomial)

Bernoulli polynomials

In mathematics, the Bernoulli polynomials, named after Jacob Bernoulli, combine the Bernoulli numbers and binomial coefficients. They are used for series expansion of functions, and with the Euler–MacLaurin formula.

These polynomials occur in the study of many special functions and, in particular, the Riemann zeta function and the Hurwitz zeta function. They are an Appell sequence (i.e. a Sheffer sequence for the ordinary derivative operator). For the Bernoulli polynomials, the number of crossings of the x-axis in the unit interval does not go up with the degree. In the limit of large degree, they approach, when appropriately scaled, the sine and cosine functions.

A similar set of polynomials, based on a generating function, is the family of Euler polynomials.

Representations[edit]

The Bernoulli polynomials Bn can be defined by a generating function. They also admit a variety of derived representations.

Generating functions[edit]

The generating function for the Bernoulli polynomials is

The generating function for the Euler polynomials is

Explicit formula[edit]

for n ≥ 0, where Bk are the Bernoulli numbers, and Ek are the Euler numbers.

Representation by a differential operator[edit]

The Bernoulli polynomials are also given by

where D = d/dx is differentiation with respect to x and the fraction is expanded as a formal power series. It follows that

cf. integrals below. By the same token, the Euler polynomials are given by

Representation by an integral operator[edit]

The Bernoulli polynomials are also the unique polynomials determined by

The integral transform

on polynomials f, simply amounts to

This can be used to produce the inversion formulae below.

Integral Recurrence[edit]

In,[1][2] it is deduced and proof that the Bernoulli polynomials can be obtained by the following integral recurrence

Another explicit formula[edit]

An explicit formula for the Bernoulli polynomials is given by

That is similar to the series expression for the Hurwitz zeta function in the complex plane. Indeed, there is the relationship

where is the Hurwitz zeta function. The latter generalizes the Bernoulli polynomials, allowing for non-integer values of n .

The inner sum may be understood to be the nth forward difference of that is,

where is the forward difference operator. Thus, one may write

This formula may be derived from an identity appearing above as follows. Since the forward difference operator Δ equals

where D is differentiation with respect to x, we have, from the Mercator series,

As long as this operates on an mth-degree polynomial such as one may let n go from 0 only up to m .

An integral representation for the Bernoulli polynomials is given by the Nörlund–Rice integral, which follows from the expression as a finite difference.

An explicit formula for the Euler polynomials is given by

The above follows analogously, using the fact that

Sums of pth powers[edit]

Using either the above integral representation of or the identity , we have

(assuming 00 = 1).

The Bernoulli and Euler numbers[edit]

The Bernoulli numbers are given by

This definition gives for

An alternate convention defines the Bernoulli numbers as

The two conventions differ only when since

The Euler numbers are given by

Explicit expressions for low degrees[edit]

The first few Bernoulli polynomials are:

The first few Euler polynomials are:

Maximum and minimum[edit]

At higher n the amount of variation in between and gets large. For instance, but Lehmer (1940)[3] showed that the maximum value (Mn) of between 0 and 1 obeys

unless n is 2 modulo 4, in which case

(where is the Riemann zeta function), while the minimum (mn) obeys

unless n = 0 modulo 4 , in which case

These limits are quite close to the actual maximum and minimum, and Lehmer gives more accurate limits as well.

Differences and derivatives[edit]

The Bernoulli and Euler polynomials obey many relations from umbral calculus:

(Δ is the forward difference operator). Also,

These polynomial sequences are Appell sequences:

Translations[edit]

These identities are also equivalent to saying that these polynomial sequences are Appell sequences. (Hermite polynomials are another example.)

Symmetries[edit]

Zhi-Wei Sun and Hao Pan [4] established the following surprising symmetry relation: If r + s + t = n and x + y + z = 1, then

where

Fourier series[edit]

The Fourier series of the Bernoulli polynomials is also a Dirichlet series, given by the expansion

Note the simple large n limit to suitably scaled trigonometric functions.

This is a special case of the analogous form for the Hurwitz zeta function

This expansion is valid only for 0 ≤ x ≤ 1 when n ≥ 2 and is valid for 0 < x < 1 when n = 1.

The Fourier series of the Euler polynomials may also be calculated. Defining the functions

and

for , the Euler polynomial has the Fourier series

and

Note that the and are odd and even, respectively:

and

They are related to the Legendre chi function as

and

Inversion[edit]

The Bernoulli and Euler polynomials may be inverted to express the monomial in terms of the polynomials.

Specifically, evidently from the above section on integral operators, it follows that

and

Relation to falling factorial[edit]

The Bernoulli polynomials may be expanded in terms of the falling factorial as

where and

denotes the Stirling number of the second kind. The above may be inverted to express the falling factorial in terms of the Bernoulli polynomials:

where

denotes the Stirling number of the first kind.

Multiplication theorems[edit]

The multiplication theorems were given by Joseph Ludwig Raabe in 1851:

For a natural number m≥1,

Integrals[edit]

Two definite integrals relating the Bernoulli and Euler polynomials to the Bernoulli and Euler numbers are:[5]

Another integral formula states[6]

with the special case for

Periodic Bernoulli polynomials[edit]

A periodic Bernoulli polynomial Pn(x) is a Bernoulli polynomial evaluated at the fractional part of the argument x. These functions are used to provide the remainder term in the Euler–Maclaurin formula relating sums to integrals. The first polynomial is a sawtooth function.

Strictly these functions are not polynomials at all and more properly should be termed the periodic Bernoulli functions, and P0(x) is not even a function, being the derivative of a sawtooth and so a Dirac comb.

The following properties are of interest, valid for all :

See also[edit]

References[edit]

  1. ^ Hurtado Benavides, Miguel Ángel. (2020). De las sumas de potencias a las sucesiones de Appell y su caracterización a través de funcionales. [Tesis de maestría]. Universidad Sergio Arboleda. https://repository.usergioarboleda.edu.co/handle/11232/174
  2. ^ Sergio A. Carrillo; Miguel Hurtado. Appell and Sheffer sequences: on their characterizations through functionals and examples. Comptes Rendus. Mathématique, Tome 359 (2021) no. 2, pp. 205-217. doi : 10.5802/crmath.172. https://comptes-rendus.academie-sciences.fr/mathematique/articles/10.5802/crmath.172/
  3. ^ Lehmer, D.H. (1940). "On the maxima and minima of Bernoulli polynomials". American Mathematical Monthly. 47: 533–538.
  4. ^ Zhi-Wei Sun; Hao Pan (2006). "Identities concerning Bernoulli and Euler polynomials". Acta Arithmetica. 125 (1): 21–39. arXiv:math/0409035. Bibcode:2006AcAri.125...21S. doi:10.4064/aa125-1-3. S2CID 10841415.
  5. ^ Takashi Agoh & Karl Dilcher (2011). "Integrals of products of Bernoulli polynomials". Journal of Mathematical Analysis and Applications. 381: 10–16. doi:10.1016/j.jmaa.2011.03.061.
  6. ^ Elaissaoui, Lahoucine & Guennoun, Zine El Abidine (2017). "Evaluation of log-tangent integrals by series involving ζ(2n+1)". Integral Transforms and Special Functions. 28 (6): 460–475. arXiv:1611.01274. doi:10.1080/10652469.2017.1312366. S2CID 119132354.

External links[edit]