Jump to content

Bernoulli number: Difference between revisions

From Wikipedia, the free encyclopedia
Content deleted Content added
correct spelling and link to Uppsala University, not changing disputed content
→‎Efficient computation of Bernoulli numbers: remove reference which had (and still has at another url) an exact copy of same AFP story as the preceding reference, remove spaces between refs
Line 213: Line 213:
''Digits'' is to be understood as the exponent of 10 when B(n) is written as a real in normalized [[scientific notation]].
''Digits'' is to be understood as the exponent of 10 when B(n) is written as a real in normalized [[scientific notation]].


In 2009, a sixteen year-old Iraqi immigrant living in Sweden submitted a proof to [[Uppsala University]], correctly simplifying the computation of Bernoulli numbers. Mathematicians claim this proof has been discovered before and is well known in the field.
In 2009, a sixteen year-old Iraqi immigrant living in Sweden submitted a proof to [[Uppsala University]], correctly simplifying the computation of Bernoulli numbers. Mathematicians claim this proof has been discovered before and is well known in the field.<ref>[http://news.yahoo.com/s/afp/20090528/od_afp/swedeneducationoffbeatreply_20090528174259 "Iraqi teen tackles maths puzzle, but not the first: university"]</ref><ref>[http://uk.news.yahoo.com/18/20090528/twl-iraq-born-teen-cracks-maths-puzzle-3cd7efd.html "Iraq-born teen cracks maths puzzle"]</ref><ref>[http://www.foxnews.com/story/0,2933,522698,00.html?test=latestnews "Iraqi Teen Solves 300-Year-Old Math Puzzle in Four Months"]</ref>
<ref>[http://news.yahoo.com/s/afp/20090528/od_afp/swedeneducationoffbeatreply_20090528174259 "Iraqi teen tackles maths puzzle, but not the first: university"]</ref> <ref>[http://uk.news.yahoo.com/18/20090528/twl-iraq-born-teen-cracks-maths-puzzle-3cd7efd.html "Iraq-born teen cracks maths puzzle"]</ref>
<ref>[http://www.naharnet.com/domino/tn/NewsDesk.nsf/getstory?openform&AwayPolitics/$first Iraq-Born Teen Cracks Maths Puzzle]</ref>
<ref>[http://www.foxnews.com/story/0,2933,522698,00.html?test=latestnews "Iraqi Teen Solves 300-Year-Old Math Puzzle in Four Months"]</ref>


== Reconstruction of 'Summae Potestatum' ==
== Reconstruction of 'Summae Potestatum' ==

Revision as of 14:24, 1 June 2009

In mathematics, the Bernoulli numbers are a sequence of rational numbers with deep connections to number theory. They are closely related to the values of the Riemann zeta function at negative integers.

In Europe, they were first studied by Jakob Bernoulli, after whom they were named by Abraham de Moivre. In Japan, perhaps earlier, they were independently discovered by Seki Kōwa. They appear in the Taylor series expansions of the tangent and hyperbolic tangent functions, in the Euler–Maclaurin formula, and in expressions for certain values of the Riemann zeta function.

In note G of Ada Lovelace's notes on the Analytical Engine from 1842, Lovelace describes an algorithm for generating Bernoulli numbers with Babbage's machine [~ 1]. As a result, the Bernoulli numbers have the distinction of being the subject of the first computer program.

Introduction

The Bernoulli numbers Bn, sometimes denoted bn. were first discovered in connection with the closed forms of the sums

for various fixed values of n. The closed forms are always polynomials in m of degree n + 1. The coefficients of these polynomials are closely related to the Bernoulli numbers, in connection with Faulhaber's formula:

For example, taking n to be 1,

Generating functions

The Bernoulli numbers may also be defined using generating functions. Their exponential generating function is x/(ex − 1), so that:

for all values of x of absolute value less than 2π (the radius of convergence of this power series).

These definitions can be shown to be equivalent using mathematical induction. The initial condition is immediate from L'Hôpital's rule. To obtain the recurrence, multiply both sides of the equation by . Then, using the Taylor series for the exponential function,

By expanding this as a Cauchy product and rearranging slightly, one obtains

It is clear from this last equality that the coefficients in this power series satisfy the same recurrence as the Bernoulli numbers.

Other forms and conventions

One may also write

where Bn + 1(m) is the (n + 1)th-degree Bernoulli polynomial.

Bernoulli numbers may be calculated by using the following recursive formula:

for m > 0, and B0 = 1.

An alternative convention for the Bernoulli numbers is to set B1 = 1/2 rather than −1/2. If this convention is used, then all the Bernoulli numbers may be calculated by a different recursive formula without further qualification:

Dividing both sides by m + 1 then gives a form suggestive of the connection with the Riemann zeta function if the j=0 case is understood as a limit to deal with the pole at ζ(1):

The terms of this sum are the coefficients given by Faulhaber's formula for the closed form of and so this recursive definition is merely reflecting the fact that these sums evaluate to 1 when n=1 for any m. In the alternative convention, the generating function is

Values of the Bernoulli numbers

Bn = 0 for all odd n other than 1. B1 = 1/2 or −1/2 depending on the convention adopted (see below). The first few non-zero Bernoulli numbers (sequences OEISA027641 and OEISA027642 in OEIS) and some larger ones are listed below.

n N D Bn = N / D
0 1 1 +1.00000000000
1 -1 2 -0.50000000000
2 1 6 +0.16666666667
4 -1 30 -0.03333333333
6 1 42 +0.02380952381
8 -1 30 -0.03333333333
10 5 66 +0.07575757576
12 -691 2730 -0.25311355311
14 7 6 +1.16666666667
16 -3617 510 -7.09215686275
18 43867 798 +54.9711779448

Efficient computation of Bernoulli numbers

In some applications it is useful to be able to compute the Bernoulli numbers B0 through Bp − 3 modulo p, where p is a prime; for example to test whether Vandiver's conjecture holds for p, or even just to determine whether p is an irregular prime. It is not feasible to carry out such a computation using the above recursive formulae, since at least (a constant multiple of) p2 arithmetic operations would be required. Fortunately, faster methods have been developed (Buhler et al. 2001) which require only O(p (log p)2) operations (see big-O notation).

David Harvey (Harvey 2008) describes an algorithm for computing Bernoulli numbers by computing Bn modulo p for many small primes p, and then reconstructing Bn via the Chinese Remainder Theorem. Harvey writes that the asymptotic time complexity of this algorithm is O(n2 log(n)2+eps) and claims that this implementation is significantly faster than implementations based on other methods. Harvey's implementation is included in Sage since version 3.1. Using this implementation Harvey computed Bn for n = 108, which is a new record (October 2008). Prior to that Bernd Kellner (Kellner 2002) computed Bn to full precision for n = 106 on December 2002 and Oleksandr Pavlyk (Pavlyk 2008) for n = 107 with 'Mathematica' on April 2008.

Computer Year n Digits
J. Bernoulli ~1689 10 1
L. Euler 1748 30 8
J.C. Adams 1878 62 36
D.E. Knuth, T.J. Buckholtz 1967 360 478
G. Fee, S. Plouffe 1996 10000 27677
G. Fee, S. Plouffe 1996 100000 376755
B.C. Kellner 2002 1000000 4767529
O. Pavlyk 2008 10000000 57675260
D. Harvey 2008 100000000 676752569
History of the computation of Bernoulli numbers

Digits is to be understood as the exponent of 10 when B(n) is written as a real in normalized scientific notation.

In 2009, a sixteen year-old Iraqi immigrant living in Sweden submitted a proof to Uppsala University, correctly simplifying the computation of Bernoulli numbers. Mathematicians claim this proof has been discovered before and is well known in the field.[1][2][3]

Reconstruction of 'Summae Potestatum'

Jakob Bernoulli's Summae Potestatum, 1713

The Bernoulli numbers were introduced by Jakob Bernoulli in the book 'Ars Conjectandi' published posthumously in 1713. The main formula can be seen in the second half of the facsimile given above. The constant coefficients denoted A, B, C and D by Bernoulli are mapped to the notation which is now prevalent as A = B2, B = B4, C = B6, D = B8. In the expression c·c−1·c−2·c−3 the small dots are used as grouping symbols, not as signs for multiplication. Using today's terminology these expressions are falling factorial powers . The factorial notation k! as a shortcut for 1×2×...×k was not introduced until 100 years later. The integral symbol on the left hand side goes back to Gottfried Wilhelm Leibniz in 1675 who used it as a long letter S for "summa" (sum). (The Mathematics Genealogy Project [~ 2] shows Leibniz as the doctoral adviser of Jakob Bernoulli. See also the Earliest Uses of Symbols of Calculus [~ 3].) The letter n on the left hand side is not an index of summation but gives the upper limit of the range of summation which is to be understood as 1,2,...,n. Putting things together, for positive c, today a mathematician is likely to write Bernoulli's formula as

In fact this formula imperatively suggests to set B1 = 1/2 when switching from the so called 'archaic' enumeration which uses only the even indices 2,4,.. to the modern form (more on different conventions in the next paragraph). Most striking in this context is the fact that the falling factorial has for k = 0 the value 1/(c + 1). (See the explanation in [~ 4].) Thus Bernoulli's formula can and has to be written

if B1 stands for the value Bernoulli himself has given to the coefficient at that position.

Different viewpoints and conventions

The Bernoulli numbers can be regarded from four main viewpoints:

  • as standalone arithmetical objects,
  • as combinatorial objects,
  • as values of a sequence of certain polynomials,
  • as values of the Riemann zeta function.

Each of these viewpoints leads to a set of more or less different conventions.

  • Bernoulli numbers as standalone arithmetical objects.
    Associated sequence: 1/6, −1/30, 1/42, −1/30,...
    This is the viewpoint of Jakob Bernoulli. (See the cutout from his Ars Conjectandi, first edition, 1713). The Bernoulli numbers are understood as numbers, recursive in nature, invented to solve a certain arithmetical problem, the summation of powers, which is the paradigmatic application of the Bernoulli numbers. It is misleading to call this viewpoint 'archaic'. For example Jean-Pierre Serre uses it in his highly acclaimed book A Course in Arithmetic which is a standard textbook used at many universities today.
  • Bernoulli numbers as combinatorial objects.
    Associated sequence: 1, +1/2, 1/6, 0,....
    This view focuses on the connection between Stirling numbers and Bernoulli numbers and arises naturally in the calculus of finite differences. In its most general and compact form this connection is summarized by the definition of the Stirling polynomials σn(x), formula (6.52) in Concrete Mathematics by Graham, Knuth and Patashnik.
    In consequence Bn = n! σn(1) for n ≥ 0.
  • Bernoulli numbers as values of a sequence of certain polynomials.
    Assuming the Bernoulli polynomials as already introduced the Bernoulli numbers can be defined in two different ways:
    Bn = Bn(0). Associated sequence: 1, −1/2, 1/6, 0,....
    Bn = Bn(1). Associated sequence: 1, +1/2, 1/6, 0,....
    The two definitions differ only in the sign of B1. The choice Bn = Bn(0) is the convention used in the Handbook of Mathematical Functions.
  • Bernoulli numbers as values of the Riemann zeta function.
    Associated sequence: 1, +1/2, 1/6, 0,....
    The Bernoulli numbers as given by the Riemann zeta function.

    This convention agrees with the convention Bn = Bn(1) (for example J. Neukirch and M. Kaneko). The sign '+' for B1 matches the representation of the Bernoulli numbers by the Riemann zeta function. In fact the identity nζ(1−n) = (−1)n+1Bn valid for all n > 0 is then replaced by the simpler nζ(1−n) = -Bn. (See the paper of S. C. Woon.)
    (Note that in the foregoing equation for n = 0 and n = 1 the expression −nζ(1 − n) is to be understood as limx → n −xζ(1 − x).)

Application of the Bernoulli number

Arguably the most important application of the Bernoulli number in mathematics is their use in the Euler-MacLaurin formula. Assuming that f is a sufficiently often differentiable function the Euler-MacLaurin formula can be written as [~ 5]

This formulation assumes the convention B1 = -1/2. However, if one sets B1 = 1/2 then this formula can also be written as

Here f(0) = f which is a commonly used notation identifying the zero-th derivative of f with f. Moreover, let f(-1) denote an antiderivative of f. By the fundamental theorem of calculus . Thus the last formula can be further simplified to the following succinct form of the Euler-Maclaurin formula

This form is for example the source for the important Euler-MacLaurin expansion of the zeta function

Here denotes the rising factorial power.[~ 6]

Combinatorial definitions

The connection of the Bernoulli number to various kinds of combinatorial numbers is based on the classical theory of finite differences and on the combinatorial interpretation of the Bernoulli numbers as an instance of a fundamental combinatorial principle, the inclusion-exclusion principle.

Connection with the Worpitzky number

The definition to proceed with was developed by Julius Worpitzky in 1883. Besides elementary arithmetic only the factorial function n! and the power function km is employed. The signless Worpitzky numbers are defined as

They can also be expressed through the Stirling set number

A Bernoulli number is then introduced as an inclusion-exclusion sum of Worpitzky numbers weighted by the sequence 1, 1/2, 1/3,...

This representation has B1 = 1/2.

Worpitzky's representation of the Bernoulli number
B0  =  1/1
B1  =  1/1 − 1/2
B2  =  1/1 − 3/2 + 2/3
B3  =  1/1 − 7/2 + 12/3 − 6/4
B4  =  1/1 − 15/2 + 50/3 − 60/4 + 24/5
B5  =  1/1 − 31/2 + 180/3 − 390/4 + 360/5 − 120/6
B6  =  1/1 − 63/2 + 602/3 − 2100/4 + 3360/5 − 2520/6 + 720/7

Connection with the Stirling set number

A similar combinatorial representation derives from

Here the Bernoulli numbers are an inclusion-exclusion over the set of length-n words, where the sum is taken over all words of length n with k distinct letters, and normalized by k + 1. The combinatorics of this representation can be seen from:

Connection with the Stirling cycle number

Let denote the signless Stirling cycle number. The two main formulas relating these number to the Bernoulli number (B1 = 1/2) are

and the inversion of this sum (for n ≥ 0, m ≥ 0)

Here the number An,m are the rational Akiyama-Tanigawa number, the first few of which are displayed in the following table.

Akiyama-Tanigawa number
n \ m 0 1 2 3 4
0 11/21/31/41/5
1 1/21/31/41/5...
2 1/61/63/20......
3 01/30.........
4 -1/30............

These relations lead to a simple algorithm to compute the Bernoulli number. The input is the first row, A0,m = 1/(m + 1) and the output are the Bernoulli number in the first column An,0 = Bn . This transformation is shown in pseudo-code below.

Akiyama-Tanigawa algorithm for Bn
Enter integer n.

For m from 0 by 1 to n do

    A[m] ← 1/(m + 1)
    For j from m by −1 to 1 do
   
    A[j − 1] ← j × (A[j − 1] − A[j])
Return A[0]  (which is Bn).

Connection with the Eulerian number

The Eulerian number are the number of permutations of {1,2,...,n} with m ascents. They were introduced by Leonhard Euler in 1755 and in the notation of D. E. Knuth [~ 7] written as The two main formulas connecting the Eulerian number to the Bernoulli number are:

Both formulas are valid for n ≥ 0 if B1 is set to 1/2. If B1 is set to −1/2 they are valid only for n ≥ 1 and n ≥ 2 respectively.

Asymptotic approximation

Leonhard Euler expressed the Bernoulli numbers in terms of the Riemann zeta function as

It then follows from the Stirling formula that, as n goes to infinity,

Integral representation and continuation

The integral

has as special values b(2n) = B2n for n > 0. The integral might be considered as a continuation of the Bernoulli numbers to the complex plane and this was indeed suggested by Peter Luschny in 2004.[citation needed]

For example b(3) = (3/2)ζ(3)Π−3Ι and b(5) = −(15/2) ζ(5) Π −5Ι. Here ζ(n) denotes the Riemann zeta function and Ι the imaginary unit. It is remarkable that already Leonhard Euler (Opera Omnia, Ser. 1, Vol. 10, p. 351) considered these numbers and calculated

Euler's values are unsigned and real, but obviously his aim was to find a meaningful way to define the Bernoulli numbers at the odd integers n > 1.

The relation to the Euler numbers and π

The Euler numbers are a sequence of integers intimately connected with the Bernoulli numbers. Comparing the asymptotic expansions of the Bernoulli and the Euler numbers shows that the Euler numbers E2n are in magnitude approximately (2/π)(42n − 22n) times larger than the Bernoulli numbers B2n. In consequence:

This asymptotic equation reveals that π lies in the common root of both the Bernoulli and the Euler numbers. In fact π could be computed from these rational approximations.

Bernoulli numbers can be expressed through the Euler numbers and vice versa. Since for n odd BnEn = 0 (with the exception B1), it suffices to regard the case when n is even.

These conversion formulas express an inverse relation between the Bernoulli and the Euler numbers. But more important, there is a deep arithmetic root common to both kinds of numbers, which can be expressed through a more fundamental sequence of numbers, also closely tied to π. These numbers are defined for n > 1 as

and S1 = 1 by convention. The magic of these numbers lies in the fact that they turn out to be rational numbers. This was first proved by Leonhard Euler 1734 in a landmark paper `De summis serierum reciprocarum' (On the sums of series of reciprocals) and fascinated mathematicians ever since. The first few of these numbers are

The Bernoulli numbers and Euler numbers are best understood as special views of these numbers, selected from the sequence Sn and scaled for use in special applications.

The expression [n even] has the value 1 if n is even and 0 otherwise (Iverson bracket).

These identities show that the quotient of Bernoulli and Euler numbers at the beginning of this section is just the special case of Rn = 2 Sn / Sn+1 when n is even. The Rn are rational approximations to π and two successive terms always enclose the true value of π. Beginning with n = 1 the sequence starts

These rational numbers also appear in the last paragraph of Euler's paper cited above. But it was only in September 2007 that this classical sequence found its way into the Encyclopedia of Integer Sequences (A132049).

An algorithmic view: the Seidel triangle

The sequence Sn has another unexpected yet important property: The denominators of Sn divide the factorial (n − 1)!. In other words: the numbers Tn = Sn(n − 1)! are integers.

Thus the above representations of the Bernoulli and Euler numbers can be rewritten in terms of this sequence as

These identities make it easy to compute the Bernoulli and Euler numbers: the Euler numbers En are given immediately by T2n + 1 and the Bernoulli numbers B2n are obtained from T2n by some easy shifting, avoiding rational arithmetic.

What remains is to find a convenient way to compute the numbers Tn. However, already in 1877 Philipp Ludwig von Seidel published an ingenious algorithm which makes it extremely simple to calculate Tn.

Seidel's algorithm for Tn

[begin] Start by putting 1 in row 0 and let k denote the number of the row currently being filled. If k is odd, then put the number on the left end of the row k − 1 in the first position of the row k, and fill the row from the left to the right, with every entry being the sum of the number to the left and the number to the upper. At the end of the row duplicate the last number. If k is even, proceed similar in the other direction. [end]

Seidel's algorithm is in fact much more general (see the exposition of Dominique Dumont (1981)) and was rediscovered several times thereafter.

Similar to Seidel's approach D. E. Knuth and T. J. Buckholtz (1967) gave a recurrence equation for the numbers T2n and recommended this method for computing B2n and E2n ‘on electronic computers using only simple operations on integers’.

V. I. Arnold rediscovered Seidel's algorithm in 1991 and later Millar, Sloane and Young popularized Seidel's algorithm under the name boustrophedon transform.

A combinatorial view: alternating permutations

Around 1880, three years after the publication of Seidel's algorithm, Désiré André proved a now classic result of combinatorial analysis. Looking at the first terms of the Taylor expansion of the trigonometric functions tan x and sec x André made a startling discovery.

The coefficients are the Euler numbers of odd and even index, respectively. In consequence the ordinary expansion of tan x + sec x has as coefficients the rational numbers Sn.

André then succeeded by means of a recurrence argument to show that the alternating permutations of odd size are enumerated by the Euler numbers of odd index (also called tangent numbers) and the alternating permutations of even size by the Euler numbers of even index (also called secant numbers).

Generalizations by polynomials

The Bernoulli polynomials can be regarded as generalizations of the Bernoulli numbers the same as the Euler polynomials are generalizations of the Euler numbers.

Arithmetical properties of the Bernoulli numbers

The Bernoulli numbers can be expressed in terms of the Riemann zeta function as Bn = − nζ(1 − n) for integers n ≥ 0 provided for n = 0 and n = 1 the expression − nζ(1 − n) is understood as the limiting value and the convention B1 = 1/2 is used. This intimately relates them to the values of the zeta function at negative integers. As such, they could be expected to have and do have deep arithmetical properties, a fact discovered by Kummer in his work on Fermat's last theorem.

Divisibility properties of the Bernoulli numbers are related to the ideal class groups of cyclotomic fields by a theorem of Kummer and its strengthening in the Herbrand-Ribet theorem, and to class numbers of real quadratic fields by Ankeny-Artin-Chowla. We also have a relationship to algebraic K-theory; if cn is the numerator of Bn/2n, then the order of is −c2n if n is even, and 2c2n if n is odd.

The Agoh-Giuga conjecture postulates that p is a prime number if and only if pBp−1 is congruent to −1 mod p.

Von Staudt–Clausen theorem

The von Staudt–Clausen theorem was given by Karl Georg Christian von Staudt and Thomas Clausen independently in 1840. It describes the arithmetical structure of the Bernoulli numbers.

The von Staudt–Clausen theorem has two parts. The first one describes how the denominators of the Bernoulli numbers can be computed. Paraphrasing the words of Clausen it can be stated as:

“The denominator of the 2nth Bernoulli number can be found as follows: Add to all divisors of 2n, 1, 2, aa', ..., 2n the unity, which gives the sequence 2, 3, a + 1, a' + 1, ..., 2n + 1. Select from this sequence only the prime numbers 2, 3, pp', etc. and build their product.”

Clausen's algorithm translates almost verbatim to a modern computer algebra program, which looks similar to the pseudocode on the left hand site of the following table. On the right hand side the computation is traced for the input n = 88. It shows that the denominator of B88 is 61410. 

Clausen's algorithm for the denominator of Bn
Clausen: function(integer n)   |   n = 88
S = divisors(n);   |   {1, 2, 4, 8, 11, 22, 44, 88}
S = map(kk + 1, S);   |   {2, 3, 5, 9, 12, 23, 45, 89}
S = select(isprime, S);   |   {2, 3, 5, 23, 89}
return product(S);   |   61410

The second part of the von Staudt–Clausen theorem is a very remarkable representation of the Bernoulli numbers. This representation is given for the first few nonzero Bernoulli numbers in the next table.

Von Staudt–Clausen representation of Bn
B0  =  1
B1  =  1 − 1/2
B2  =  1 − 1/2 − 1/3
B4  =  1 − 1/2 − 1/3 − 1/5
B6  =  1 − 1/2 − 1/3 − 1/7
B8  =  1 − 1/2 − 1/3 − 1/5
B10  =  1 − 1/2 − 1/3 − 1/11

The theorem affirms the existence of an integer In such that

The sum is over the primes p for which p − 1 divides n. These are the same primes which are employed in the Clausen algorithm. The proposition holds true for all n ≥ 0, not only for even n. I1 = 2 and for odd n > 1, In = 1.

Consequences of the von Staudt–Clausen theorem are: the denominators of the Bernoulli numbers are square-free and for n ≥ 2 divisible by 6.

Why do the odd Bernoulli numbers vanish?

The sum

can be evaluated for negative values of the index n. Doing so will show that it is an odd function for even values of k, which implies that the sum has only terms of odd index. This and the formula for the Bernoulli sum imply that B2k+1−m is 0 for m odd and greater than 1; and that the term for B1 is cancelled by the subtraction. The von Staudt-Clausen theorem combined with Worpitzky's representation also gives a combinatorial answer to this question (valid for n > 1).

From the von Staudt-Clausen theorem it is known that for odd n > 1 the number 2Bn is an integer. This seems trivial if one knows beforehand that in this case Bn = 0. However, by applying Worpitzky's representation one gets

as a sum of integers, which is not trivial. Here a combinatorial fact comes to surface which explains the vanishing of the Bernoulli numbers at odd index. Let Sn,m be the number of surjective maps from {1, 2, ..., n} to {1, 2, ..., m}, then . The last equation can only hold if

This equation can be proved by induction. The first two examples of this equation are

n = 4 :  2 + 8 = 7 + 3,  n = 6:  2 + 120 + 144 = 31 + 195 + 40.

Thus the Bernoulli numbers vanish at odd index because some non-obvious combinatorial identities are embodied in the Bernoulli numbers.

p-adic continuity

An especially important congruence property of the Bernoulli numbers can be characterized as a p-adic continuity property. If b, m and n are positive integers such that m and n are not divisible by p − 1 and , then

Since , this can also be written

where u = 1 − m and v = 1 − n, so that u and v are nonpositive and not congruent to 1 mod p − 1. This tells us that the Riemann zeta function, with 1 − ps taken out of the Euler product formula, is continuous in the p-adic numbers on odd negative integers congruent mod p − 1 to a particular , and so can be extended to a continuous function for all p-adic integers the p-adic Zeta function.

Bernoulli numbers and the Riemann hypothesis

The connection between the Bernoulli numbers and the Riemann zeta function is strong enough to provide an alternate formulation of the Riemann hypothesis (RH) which uses only the Bernoulli number. In fact in 1916 M. Riesz proved that the RH is equivalent to the following assertion:

For every ε > 1/4 there exists a constant Cε > 0 (depending on ε) such that |R(x)| < Cε xε as x → ∞. Here R(x) is the Riesz function

denotes the rising factorial power in the notation of D. E. Knuth. The number βn = Bn/n occur frequently in the study of the zeta function and are significant because βn is a p-integer for primes p where p − 1 does not divide n. The βn are called divided Bernoulli number.

Ramanujan's congruences

The following relations, due to Ramanujan, provide a more efficient method for calculating Bernoulli numbers:

Use of Bernoulli numbers in topology

The Kervaire-Milnor formula for the order of the cyclic group of diffeomorphism classes of exotic (4n − 1)-spheres which bound parallelizable manifolds for involves Bernoulli numbers; if B(n) is the numerator of B4n/n, then

is the number of such exotic spheres. (The formula in the topological literature differs because topologists use a different convention for naming Bernoulli numbers; this article uses the number theorists' convention.)

Assorted identities

  • Let n be non-negative and even
  • The nth cumulant of the uniform probability distribution on the interval [−1, 0] is Bn/n.
  • Let n¡ = 1/n! and n ≥ 1. Then Bn is n! times the determinant of the following matrix.
    Bn = 
    1
    ───
    n¡
    0 0 0 ... 0
    0 0 ... 0
    0 ... 0
    ... ... ... ... ... ... ...
    (n − 2)¡ ... ... 0
    (n − 1)¡ (n − 2)¡ ... ...
    n¡ (n − 1)¡ (n − 2)¡ ... ...

    Thus the determinant is σn(1), the Stirling polynomial at x = 1.

  • Let n ≥ 1.
  • Let n ≥ 1. Then (von Ettingshausen 1827)
  • Let n ≥ 1 and m ≥ 1. Then (Carlitz 1968)
  • Let n ≥ 4 and
    the harmonic number. Then
  • Let n ≥ 4. Yuri Matiyasevich found (1997)
  • Let n ≥ 1
    This is an identity by Faber-Pandharipande-Zagier-Gessel. Choose x = 0 or x = 1 to get a Bernoulli number identity according to your favourite convention.
  • The next formula is true for n ≥ 0 if B1 = B1(1) = 1/2, but only for n ≥ 1 if B1 = B1(0) = -1/2.
  • Let n ≥ 0 and [b] = 1 if b is true, 0 otherwise.
    and

See also

References

  1. ^ Note G in the Menabrea reference
  2. ^ Mathematics Genealogy Project
  3. ^ Earliest Uses of Symbols of Calculus
  4. ^ Concrete Mathematics, (2.51)
  5. ^ Concrete Mathematics, (9.67).
  6. ^ Concrete Mathematics, (2.44) and (2.52)
  7. ^ Concrete Mathematics, Table 254 Euler's triangle.
  • Abramowitz, M.; Stegun, C. A. (1972), "§23.1: Bernoulli and Euler Polynomials and the Euler-Maclaurin Formula", Handbook of Mathematical Functions with Formulas, Graphs, and Mathematical Tables (9th printing ed.), New York: Dover, pp. 804–806.
  • André, D. (1879), "Développements de sec x et tan x.", Comptes Rendus Acad. Sci., 88: 965–967.
  • André, D. (1881), "Mémoire sur les permutations alternées", J. Math., 7: 167–184.
  • Arlettaz, D. (1998), "Die Bernoulli-Zahlen: eine Beziehung zwischen Topologie und Gruppentheorie", Math. Semesterber, 45: 61–75, doi:10.1007/s005910050037.
  • Arnold, V. I. (1991), "Bernoulli-Euler updown numbers associated with function singularities, their combinatorics and arithmetics", Duke Math. J., 63: 537–555.
  • Ayoub, A. (1981), "Euler and the Zeta Function", Amer. Math. Monthly, 74: 1067–1086.
  • Buhler, J.; Crandall, R.; Ernvall, R.; Metsankyla, T.; Shokrollahi, M. (2001), "Irregular Primes and Cyclotomic Invariants to 12 Million", Journal of Symbolic Computation, 31 (1–2): 89–96, doi:10.1006/jsco.1999.1011 {{citation}}: Italic or bold markup not allowed in: |journal= (help).
  • Carlitz, L. (1968), "Bernoulli Numbers.", Fib. Quart., 6: 71–85.
  • Clausen, Thomas (1840), "Lehrsatz aus einer Abhandlung über die Bernoullischen Zahlen", Astr. Nachr., 17: 351–352.
  • Conway, John; Guy (1996), The Book of Numbers, Springer-Verlag {{citation}}: Cite has empty unknown parameter: |unused_data= (help); Unknown parameter |firshttp://en.wikipedia.org/w/index.php?title= ignored (help).
  • Dumont, D.; Viennot, G. (1980), "A combinatorial interpretation of Seidel generation of Genocchi numbers", Ann. Discrete Math., 6: 77–87, doi:10.1016/S0167-5060(08)70696-4.
  • Dumont, D. (1981), "Matrices d'Euler-Seidel", Séminaire Lotharingien de Combinatoire
  • Entringer, R. C. (1966), "A combinatorial interpretation of the Euler and Bernoulli numbers", Nieuw. Arch. V. Wiskunde, 14: 241–6.
  • von Ettingshausen, A. (1827), Vorlesungen über die höhere Mathematik, vol. Bd. 1, Vienna: Carl Gerold.
  • Euler, Leonhard (1735), "De summis serierum reciprocarum", Opera Omnia, I.14, E 41, : 73–86{{citation}}: CS1 maint: extra punctuation (link); On the sums of series of reciprocals, arXiv:math/0506415v2 (math.HO).
  • Fee, G.; Plouffe, S. (2007), An efficient algorithm for the computation of Bernoulli numbers. arXiv:math/0702300v2 (math.NT)
  • Graham, R. L.; Knuth, D. E.; Patashnik, O. (1989), Concrete Mathematics, Addison-Wesley.
  • Harvey, David (2008), A multimodular algorithm for computing Bernoulli numbers arXiv:0807.1347v2 math.NT
  • Jordan, Charles (1950), Calculus of Finite Differences, New York: Chelsea Publ. Co..
  • Kaneko, M. (2000), "The Akiyama-Tanigawa algorithm for Bernoulli numbers", Journal of Integer Sequences, 12.
  • Kellner, Bernd (2002), Program Calcbn - A program for calculating Bernoulli numbers.
  • Knuth, D. E.; Buckholtz, T. J. (1967), "Computation of Tangent, Euler, and Bernoulli Numbers", Mathematics of Computation, 21: 663–688, doi:10.2307/2005010.
  • Luschny, Peter (2007), An inclusion of the Bernoulli numbers
  • Menabrea, L. F., "Sketch of the Analytic Engine invented by Charles Babbage, with notes upon the Memoir by the Translator Ada Augusta, Countess of Lovelace." Bibliothèque Universelle de Genève, October 1842, No. 82. http://www.fourmilab.ch/babbage/sketch.html
  • Milnor, John W.; Stasheff, James D. (1974), "Appendix B: Bernoulli Numbers", Characteristic Classes, Annals of Mathematics Studies, vol. 76, Princeton University Press and University of Tokyo Press, pp. 281–287.
  • Neukirch, Jürgen (1999). Algebraische Zahlentheorie. Grundlehren der mathematischen Wissenschaften. Vol. 322. Berlin: Springer-Verlag. ISBN 978-3-540-65399-8. MR 1697859. Zbl 0956.11021.
  • Pavlyk, Oleksandr (2008), Today We Broke the Bernoulli Record: From the Analytical Engine to Mathematica, Wolfram Blog.
  • Riesz, M. (1916), "Sur l'hypothèse de Riemann", Acta Mathematica, 40: 185–90, doi:10.1007/BF02418544.
  • Seidel, L. (1877), "Über eine einfache Entstehungsweise der Bernoullischen Zahlen und einiger verwandten Reihen", Sitzungsber. Münch. Akad., 4: 157–187.
  • Slavutskii, Ilya Sh. (1995), "Staudt and arithmetical properties of Bernoulli numbers", Historia Scientiarum, 2: 69–74.
  • von Staudt, K. G. Ch. (1840), "Beweis eines Lehrsatzes, die Bernoullischen Zahlen betreffend", Journal für die reine und angewandte Mathematik, 21: 372–374.
  • von Staudt, K. G. Ch. (1845), "De numeris Bernoullianis, commentationem alteram", Erlangen.
  • Sun, Zhi-Wei (2005/2006), Some curious results on Bernoulli and Euler polynomials {{citation}}: Check date values in: |year= (help).
  • Woon, S. C. (1998), Generalization of a relation between the Riemann zeta function and Bernoulli numbers
  • Worpitzky, J. (1883), "Studien über die Bernoullischen und Eulerschen Zahlen", Journal für die reine und angewandte Mathematik, 94: 203–232.