fbpx
Wikipedia

Bernoulli number

Bernoulli numbers B±
n
n fraction decimal
0 1 +1.000000000
1 ±1/2 ±0.500000000
2 1/6 +0.166666666
3 0 +0.000000000
4 1/30 −0.033333333
5 0 +0.000000000
6 1/42 +0.023809523
7 0 +0.000000000
8 1/30 −0.033333333
9 0 +0.000000000
10 5/66 +0.075757575
11 0 +0.000000000
12 691/2730 −0.253113553
13 0 +0.000000000
14 7/6 +1.166666666
15 0 +0.000000000
16 3617/510 −7.092156862
17 0 +0.000000000
18 43867/798 +54.97117794
19 0 +0.000000000
20 174611/330 −529.1242424

In mathematics, the Bernoulli numbers Bn are a sequence of rational numbers which occur frequently in analysis. The Bernoulli numbers appear in (and can be defined by) the Taylor series expansions of the tangent and hyperbolic tangent functions, in Faulhaber's formula for the sum of m-th powers of the first n positive integers, in the Euler–Maclaurin formula, and in expressions for certain values of the Riemann zeta function.

The values of the first 20 Bernoulli numbers are given in the adjacent table. Two conventions are used in the literature, denoted here by and ; they differ only for n = 1, where and . For every odd n > 1, Bn = 0. For every even n > 0, Bn is negative if n is divisible by 4 and positive otherwise. The Bernoulli numbers are special values of the Bernoulli polynomials , with and .[1]

The Bernoulli numbers were discovered around the same time by the Swiss mathematician Jacob Bernoulli, after whom they are named, and independently by Japanese mathematician Seki Takakazu. Seki's discovery was posthumously published in 1712[2][3][4] in his work Katsuyō Sanpō; Bernoulli's, also posthumously, in his Ars Conjectandi of 1713. Ada Lovelace's note G on the Analytical Engine from 1842 describes an algorithm for generating Bernoulli numbers with Babbage's machine.[5] As a result, the Bernoulli numbers have the distinction of being the subject of the first published complex computer program.

Notation

The superscript ± used in this article distinguishes the two sign conventions for Bernoulli numbers. Only the n = 1 term is affected:

In the formulas below, one can switch from one sign convention to the other with the relation  , or for integer n = 2 or greater, simply ignore it.

Since Bn = 0 for all odd n > 1, and many formulas only involve even-index Bernoulli numbers, a few authors write "Bn" instead of B2n . This article does not follow that notation.

History

Early history

The Bernoulli numbers are rooted in the early history of the computation of sums of integer powers, which have been of interest to mathematicians since antiquity.

 
A page from Seki Takakazu's Katsuyō Sanpō (1712), tabulating binomial coefficients and Bernoulli numbers

Methods to calculate the sum of the first n positive integers, the sum of the squares and of the cubes of the first n positive integers were known, but there were no real 'formulas', only descriptions given entirely in words. Among the great mathematicians of antiquity to consider this problem were Pythagoras (c. 572–497 BCE, Greece), Archimedes (287–212 BCE, Italy), Aryabhata (b. 476, India), Abu Bakr al-Karaji (d. 1019, Persia) and Abu Ali al-Hasan ibn al-Hasan ibn al-Haytham (965–1039, Iraq).

During the late sixteenth and early seventeenth centuries mathematicians made significant progress. In the West Thomas Harriot (1560–1621) of England, Johann Faulhaber (1580–1635) of Germany, Pierre de Fermat (1601–1665) and fellow French mathematician Blaise Pascal (1623–1662) all played important roles.

Thomas Harriot seems to have been the first to derive and write formulas for sums of powers using symbolic notation, but even he calculated only up to the sum of the fourth powers. Johann Faulhaber gave formulas for sums of powers up to the 17th power in his 1631 Academia Algebrae, far higher than anyone before him, but he did not give a general formula.

Blaise Pascal in 1654 proved Pascal's identity relating the sums of the pth powers of the first n positive integers for p = 0, 1, 2, ..., k.

The Swiss mathematician Jakob Bernoulli (1654–1705) was the first to realize the existence of a single sequence of constants B0, B1, B2,... which provide a uniform formula for all sums of powers.[9]

The joy Bernoulli experienced when he hit upon the pattern needed to compute quickly and easily the coefficients of his formula for the sum of the cth powers for any positive integer c can be seen from his comment. He wrote:

"With the help of this table, it took me less than half of a quarter of an hour to find that the tenth powers of the first 1000 numbers being added together will yield the sum 91,409,924,241,424,243,424,241,924,242,500."

Bernoulli's result was published posthumously in Ars Conjectandi in 1713. Seki Takakazu independently discovered the Bernoulli numbers and his result was published a year earlier, also posthumously, in 1712.[2] However, Seki did not present his method as a formula based on a sequence of constants.

Bernoulli's formula for sums of powers is the most useful and generalizable formulation to date. The coefficients in Bernoulli's formula are now called Bernoulli numbers, following a suggestion of Abraham de Moivre.

Bernoulli's formula is sometimes called Faulhaber's formula after Johann Faulhaber who found remarkable ways to calculate sum of powers but never stated Bernoulli's formula. According to Knuth[9] a rigorous proof of Faulhaber's formula was first published by Carl Jacobi in 1834.[10] Knuth's in-depth study of Faulhaber's formula concludes (the nonstandard notation on the LHS is explained further on):

"Faulhaber never discovered the Bernoulli numbers; i.e., he never realized that a single sequence of constants B0, B1, B2, ... would provide a uniform
 
or
 
for all sums of powers. He never mentioned, for example, the fact that almost half of the coefficients turned out to be zero after he had converted his formulas for Σ nm from polynomials in N to polynomials in n."[11]

Reconstruction of "Summae Potestatum"

 
Jakob Bernoulli's "Summae Potestatum", 1713[a]

The Bernoulli numbers OEISA164555(n)/OEISA027642(n) were introduced by Jakob Bernoulli in the book Ars Conjectandi published posthumously in 1713 page 97. The main formula can be seen in the second half of the corresponding facsimile. 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. The expression c·c−1·c−2·c−3 means c·(c−1)·(c−2)·(c−3) – the small dots are used as grouping symbols. Using today's terminology these expressions are falling factorial powers ck. 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).[b] 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:

 

This formula suggests setting B1 = 1/2 when switching from the so-called 'archaic' enumeration which uses only the even indices 2, 4, 6... to the modern form (more on different conventions in the next paragraph). Most striking in this context is the fact that the falling factorial ck−1 has for k = 0 the value 1/c + 1.[12] Thus Bernoulli's formula can be written

 

if B1 = 1/2, recapturing the value Bernoulli gave to the coefficient at that position.

The formula for   in the first half of the quotation by Bernoulli above contains an error at the last term; it should be   instead of  .

Definitions

Many characterizations of the Bernoulli numbers have been found in the last 300 years, and each could be used to introduce these numbers. Here only three of the most useful ones are mentioned:

  • a recursive equation,
  • an explicit formula,
  • a generating function.

For the proof of the equivalence of the three approaches.[13]

Recursive definition

The Bernoulli numbers obey the sum formulas[1]

 

where   and δ denotes the Kronecker delta. Solving for   gives the recursive formulas

 

Explicit definition

In 1893 Louis Saalschütz listed a total of 38 explicit formulas for the Bernoulli numbers,[14] usually giving some reference in the older literature. One of them is (for  ):

 

Generating function

The exponential generating functions are

 

where the substitution is  .

The (ordinary) generating function

 

is an asymptotic series. It contains the trigamma function ψ1.

Bernoulli numbers and the Riemann zeta function

 
The Bernoulli numbers as given by the Riemann zeta function.

The Bernoulli numbers can be expressed in terms of the Riemann zeta function:

B+
n
= −(1 − n)
          for n ≥ 1 .

Here the argument of the zeta function is 0 or negative.

By means of the zeta functional equation and the gamma reflection formula the following relation can be obtained:[15]

  for n ≥ 1 .

Now the argument of the zeta function is positive.

It then follows from ζ → 1 (n → ∞) and Stirling's formula that

  for n → ∞ .

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[16] which require only O(p (log p)2) operations (see big O notation).

David Harvey[17] 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 + ε) and claims that this implementation is significantly faster than implementations based on other methods. Using this implementation Harvey computed Bn for n = 108. Harvey's implementation has been included in SageMath since version 3.1. Prior to that, Bernd Kellner[18] computed Bn to full precision for n = 106 in December 2002 and Oleksandr Pavlyk[19] for n = 107 with Mathematica in 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 1672 3330
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
* Digits is to be understood as the exponent of 10 when Bn is written as a real number in normalized scientific notation.

A possible algorithm for computing Bernoulli numbers in the Julia programming language is given by[14]

b = Array{Float64}(undef, n+1) b[1] = 1 b[2] = -0.5 for m=2:n for k=0:m for v=0:k b[m+1] += (-1)^v * binomial(k,v) * v^(m) / (k+1) end end end return b 

Applications of the Bernoulli numbers

Asymptotic analysis

Arguably the most important application of the Bernoulli numbers 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[20]

 

This formulation assumes the convention B
1
= −1/2
. Using the convention B+
1
= +1/2
the formula becomes

 

Here   (i.e. the zeroth-order derivative of   is just  ). Moreover, let   denote an antiderivative of  . 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 sk denotes the rising factorial power.[21]

Bernoulli numbers are also frequently used in other kinds of asymptotic expansions. The following example is the classical Poincaré-type asymptotic expansion of the digamma function ψ.

 

Sum of powers

Bernoulli numbers feature prominently in the closed form expression of the sum of the mth powers of the first n positive integers. For m, n ≥ 0 define

 

This expression can always be rewritten as a polynomial in n of degree m + 1. The coefficients of these polynomials are related to the Bernoulli numbers by Bernoulli's formula:

 

where (m + 1
k
)
denotes the binomial coefficient.

For example, taking m to be 1 gives the triangular numbers 0, 1, 3, 6, ... OEISA000217.

 

Taking m to be 2 gives the square pyramidal numbers 0, 1, 5, 14, ... OEISA000330.

 

Some authors use the alternate convention for Bernoulli numbers and state Bernoulli's formula in this way:

 

Bernoulli's formula is sometimes called Faulhaber's formula after Johann Faulhaber who also found remarkable ways to calculate sums of powers.

Faulhaber's formula was generalized by V. Guo and J. Zeng to a q-analog.[22]

Taylor series

The Bernoulli numbers appear in the Taylor series expansion of many trigonometric functions and hyperbolic functions.

Tangent
 
Cotangent
 
Hyperbolic tangent
 
Hyperbolic cotangent
 

Laurent series

The Bernoulli numbers appear in the following Laurent series:[23]

Digamma function:  

Use 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 involves Bernoulli numbers. Let ESn be the number of such exotic spheres for n ≥ 2, then

 

The Hirzebruch signature theorem for the L genus of a smooth oriented closed manifold of dimension 4n also involves Bernoulli numbers.

Connections with combinatorial numbers

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 Worpitzky numbers

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 numbers of the second kind

 

A Bernoulli number is then introduced as an inclusion–exclusion sum of Worpitzky numbers weighted by the harmonic sequence 1, 1/21/3, ...

 
B0 = 1
B1 = 1 − 1/2
B2 = 1 − 3/2 + 2/3
B3 = 1 − 7/2 + 12/36/4
B4 = 1 − 15/2 + 50/360/4 + 24/5
B5 = 1 − 31/2 + 180/3390/4 + 360/5120/6
B6 = 1 − 63/2 + 602/32100/4 + 3360/52520/6 + 720/7

This representation has B+
1
= +1/2
.

Consider the sequence sn, n ≥ 0. From Worpitzky's numbers OEISA028246, OEISA163626 applied to s0, s0, s1, s0, s1, s2, s0, s1, s2, s3, ... is identical to the Akiyama–Tanigawa transform applied to sn (see Connection with Stirling numbers of the first kind). This can be seen via the table:

Identity of
Worpitzky's representation and Akiyama–Tanigawa transform
1 0 1 0 0 1 0 0 0 1 0 0 0 0 1
1 −1 0 2 −2 0 0 3 −3 0 0 0 4 −4
1 −3 2 0 4 −10 6 0 0 9 −21 12
1 −7 12 −6 0 8 −38 54 −24
1 −15 50 −60 24

The first row represents s0, s1, s2, s3, s4.

Hence for the second fractional Euler numbers OEISA198631 (n) / OEISA006519 (n + 1):

E0 = 1
E1 = 1 − 1/2
E2 = 1 − 3/2 + 2/4
E3 = 1 − 7/2 + 12/46/8
E4 = 1 − 15/2 + 50/460/8 + 24/16
E5 = 1 − 31/2 + 180/4390/8 + 360/16120/32
E6 = 1 − 63/2 + 602/42100/8 + 3360/162520/32 + 720/64

A second formula representing the Bernoulli numbers by the Worpitzky numbers is for n ≥ 1

 

The simplified second Worpitzky's representation of the second Bernoulli numbers is:

OEISA164555 (n + 1) / OEISA027642(n + 1) = n + 1/2n + 2 − 2 × OEISA198631(n) / OEISA006519(n + 1)

which links the second Bernoulli numbers to the second fractional Euler numbers. The beginning is:

1/2, 1/6, 0, −1/30, 0, 1/42, ... = (1/2, 1/3, 3/14, 2/15, 5/62, 1/21, ...) × (1, 1/2, 0, −1/4, 0, 1/2, ...)

The numerators of the first parentheses are OEISA111701 (see Connection with Stirling numbers of the first kind).

Connection with Stirling numbers of the second kind

If S(k,m) denotes Stirling numbers of the second kind[24] then one has:

 

where jm denotes the falling factorial.

If one defines the Bernoulli polynomials Bk(j) as:[25]

 

where Bk for k = 0, 1, 2,... are the Bernoulli numbers.

Then after the following property of the binomial coefficient:

 

one has,

 

One also has the following for Bernoulli polynomials,[25]

 

The coefficient of j in (j
m + 1
)
is (−1)m/m + 1.

Comparing the coefficient of j in the two expressions of Bernoulli polynomials, one has:

 

(resulting in B1 = +1/2) which is an explicit formula for Bernoulli numbers and can be used to prove Von-Staudt Clausen theorem.[26][27][28]

Connection with Stirling numbers of the first kind

The two main formulas relating the unsigned Stirling numbers of the first kind [n
m
]
to the Bernoulli numbers (with 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 numbers, the first few of which are displayed in the following table.

Akiyama–Tanigawa number
m
n
0 1 2 3 4
0 1 1/2 1/3 1/4 1/5
1 1/2 1/3 1/4 1/5 ...
2 1/6 1/6 3/20 ... ...
3 0 1/30 ... ... ...
4 1/30 ... ... ... ...

The Akiyama–Tanigawa numbers satisfy a simple recurrence relation which can be exploited to iteratively compute the Bernoulli numbers. This leads to the algorithm shown in the section 'algorithmic description' above. See OEISA051714/OEISA051715.

An autosequence is a sequence which has its inverse binomial transform equal to the signed sequence. If the main diagonal is zeroes = OEISA000004, the autosequence is of the first kind. Example: OEISA000045, the Fibonacci numbers. If the main diagonal is the first upper diagonal multiplied by 2, it is of the second kind. Example: OEISA164555/OEISA027642, the second Bernoulli numbers (see OEISA190339). The Akiyama–Tanigawa transform applied to 2n = 1/OEISA000079 leads to OEISA198631 (n) / OEISA06519 (n + 1). Hence:

Akiyama–Tanigawa transform for the second Euler numbers
m
n
0 1 2 3 4
0 1 1/2 1/4 1/8 1/16
1 1/2 1/2 3/8 1/4 ...
2 0 1/4 3/8 ... ...
3 1/4 1/4 ... ... ...
4 0 ... ... ... ...

See OEISA209308 and OEISA227577. OEISA198631 (n) / OEISA006519 (n + 1) are the second (fractional) Euler numbers and an autosequence of the second kind.

(OEISA164555 (n + 2)/OEISA027642 (n + 2) = 1/6, 0, −1/30, 0, 1/42, ...) × ( 2n + 3 − 2/n + 2 = 3, 14/3, 15/2, 62/5, 21, ...) = OEISA198631 (n + 1)/OEISA006519 (n + 2) = 1/2, 0, −1/4, 0, 1/2, ....

Also valuable for OEISA027641 / OEISA027642 (see Connection with Worpitzky numbers).

Connection with Pascal's triangle

There are formulas connecting Pascal's triangle to Bernoulli numbers[c]

 

where   is the determinant of a n-by-n Hessenberg matrix part of Pascal's triangle whose elements are:  

Example:

 

Connection with Eulerian numbers

There are formulas connecting Eulerian numbers n
m
to Bernoulli numbers:

 

Both formulae 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.

A binary tree representation

The Stirling polynomials σn(x) are related to the Bernoulli numbers by Bn = n!σn(1). S. C. Woon described an algorithm to compute σn(1) as a binary tree:[29]

 

Woon's recursive algorithm (for n ≥ 1) starts by assigning to the root node N = [1,2]. Given a node N = [a1, a2, ..., ak] of the tree, the left child of the node is L(N) = [−a1, a2 + 1, a3, ..., ak] and the right child R(N) = [a1, 2, a2, ..., ak]. A node N = [a1, a2, ..., ak] is written as ±[a2, ..., ak] in the initial part of the tree represented above with ± denoting the sign of a1.

Given a node N the factorial of N is defined as

 

Restricted to the nodes N of a fixed tree-level n the sum of 1/N! is σn(1), thus

 

For example:

B1 = 1!(1/2!)
B2 = 2!(−1/3! + 1/2!2!)
B3 = 3!(1/4!1/2!3!1/3!2! + 1/2!2!2!)

Integral representation and continuation

The integral

 

has as special values b(2n) = B2n for n > 0.

For example, b(3) = 3/2ζ(3)π−3i and b(5) = −15/2ζ(5)π−5i. Here, ζ is the Riemann zeta function, and i is the imaginary unit. Leonhard Euler (Opera Omnia, Ser. 1, Vol. 10, p. 351) considered these numbers and calculated

 

Another similar integral representation is

 

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 odd n, Bn = En = 0 (with the exception B1), it suffices to consider the case when n is even.

 

These conversion formulas express a connection 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.[30] The magic of these numbers lies in the fact that they turn out to be rational numbers. This was first proved by Leonhard Euler in a landmark paper De summis serierum reciprocarum (On the sums of series of reciprocals) and has fascinated mathematicians ever since.[31] The first few of these numbers are

  (OEISA099612 / OEISA099617)

These are the coefficients in the expansion of sec x + tan x.

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 = 2Sn/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 (OEISA132049 / OEISA132050):

 

These rational numbers also appear in the last paragraph of Euler's paper cited above.

Consider the Akiyama–Tanigawa transform for the sequence OEISA046978 (n + 2) / OEISA016116 (n + 1):

0 1 1/2 0 1/4 1/4 1/8 0
1 1/2 1 3/4 0 5/8 3/4
2 1/2 1/2 9/4 5/2 5/8
3 −1 7/2 3/4 15/2
4 5/2 11/2 99/4
5 8 77/2
6 61/2

From the second, the numerators of the first column are the denominators of Euler's formula. The first column is −1/2 × OEISA163982.

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)!, sometimes called Euler zigzag numbers, are integers.

  (OEISA000111). See (OEISA253671).

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 simple to calculate Tn.[32]

 
Seidel's algorithm for Tn
  1. Start by putting 1 in row 0 and let k denote the number of the row currently being filled
  2. 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
  3. At the end of the row duplicate the last number.
  4. If k is even, proceed similar in the other direction.

Seidel's algorithm is in fact much more general (see the exposition of Dominique Dumont [33]) and was rediscovered several times thereafter.

Similar to Seidel's approach D. E. Knuth and T. J. Buckholtz 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'.[34]

V. I. Arnold[35] rediscovered Seidel's algorithm and later Millar, Sloane and Young popularized Seidel's algorithm under the name boustrophedon transform.

Triangular form:

1
1 1
2 2 1
2 4 5 5
16 16 14 10 5
16 32 46 56 61 61
272 272 256 224 178 122 61

Only OEISA000657, with one 1, and OEISA214267, with two 1s, are in the OEIS.

Distribution with a supplementary 1 and one 0 in the following rows:

1
0 1
−1 −1 0
0 −1 −2 −2
5 5 4 2 0
0 5 10 14 16 16
−61 −61 −56 −46 −32 −16 0

This is OEISA239005, a signed version of OEISA008280. The main andiagonal is OEISA122045. The main diagonal is OEISA155585. The central column is OEISA099023. Row sums: 1, 1, −2, −5, 16, 61.... See OEISA163747. See the array beginning with 1, 1, 0, −2, 0, 16, 0 below.

The Akiyama–Tanigawa algorithm applied to OEISA046978 (n + 1) / OEISA016116(n) yields:

1 1 1/2 0 1/4 1/4 1/8
0 1 3/2 1 0 3/4
−1 −1 3/2 4 15/4
0 −5 15/2 1
5 5 51/2
0 61
−61

1. The first column is OEISA122045. Its binomial transform leads to:

1 1 0 −2 0 16 0
0 −1 −2 2 16 −16
−1 −1 4 14 −32
0 5 10 −46
5 5 −56
0 −61
−61

The first row of this array is OEISA155585. The absolute values of the increasing antidiagonals are OEISA008280. The sum of the antidiagonals is OEISA163747 (n + 1).

2. The second column is 1 1 −1 −5 5 61 −61 −1385 1385.... Its binomial transform yields:

1 2 2 −4 −16 32 272
1 0 −6 −12 48 240
−1 −6 −6 60 192
−5 0 66 32
5 66 66
61 0
−61

The first row of this array is 1 2 2 −4 −16 32 272 544 −7936 15872 353792 −707584.... The absolute values of the second bisection are the double of the absolute values of the first bisection.

Consider the Akiyama-Tanigawa algorithm applied to OEISA046978 (n) / (OEISA158780 (n + 1) = abs(OEISA117575 (n)) + 1 = 1, 2, 2, 3/2, 1, 3/4, 3/4, 7/8, 1, 17/16, 17/16, 33/32....

1 2 2 3/2 1 3/4 3/4
−1 0 3/2 2 5/4 0
−1 −3 3/2 3 25/4
2 −3 27/2 −13
5 21 3/2
−16 45
−61

The first column whose the absolute values are OEISA000111 could be the numerator of a trigonometric function.

OEISA163747 is an autosequence of the first kind (the main diagonal is OEISA000004). The corresponding array is:

0 −1 −1 2 5 −16 −61
−1 0 3 3 −21 −45
1 3 0 −24 −24
2 −3 −24 0
−5 −21 24
−16 45
−61

The first two upper diagonals are −1 3 −24 402... = (−1)n + 1 × OEISA002832. The sum of the antidiagonals is 0 −2 0 10... = 2 × OEISA122045(n + 1).

OEISA163982 is an autosequence of the second kind, like for instance OEISA164555 / OEISA027642. Hence the array:

2 1 −1 −2 5 16 −61
−1 −2 −1 7 11 −77
−1 1 8 4 −88
2 7 −4 −92
5 −11 −88
−16 −77
−61

The main diagonal, here 2 −2 8 −92..., is the double of the first upper one, here OEISA099023. The sum of the antidiagonals is 2 0 −4 0... = 2 × OEISA155585(n + 1). OEISA163747 − OEISA163982 = 2 × OEISA122045.

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.[36][37] 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).

Related sequences

The arithmetic mean of the first and the second Bernoulli numbers are the associate Bernoulli numbers: B0 = 1, B1 = 0, B2 = 1/6, B3 = 0, B4 = −1/30, OEISA176327 / OEISA027642. Via the second row of its inverse Akiyama–Tanigawa transform OEISA177427, they lead to Balmer series OEISA061037 / OEISA061038.

The Akiyama–Tanigawa algorithm applied to OEISA060819 (n + 4) / OEISA145979 (n) leads to the Bernoulli numbers OEISA027641 / OEISA027642, OEISA164555 / OEISA027642, or OEISA176327 OEISA176289 without B1, named intrinsic Bernoulli numbers Bi(n).

1 5/6 3/4 7/10 2/3
1/6 1/6 3/20 2/15 5/42
0 1/30 1/20 2/35 5/84
1/30 1/30 3/140 1/105 0
0 1/42 1/28 4/105 1/28

Hence another link between the intrinsic Bernoulli numbers and the Balmer series via OEISA145979 (n).

OEISA145979 (n − 2) = 0, 2, 1, 6,... is a permutation of the non-negative numbers.

The terms of the first row are f(n) = 1/2 + 1/n + 2. 2, f(n) is an autosequence of the second kind. 3/2, f(n) leads by its inverse binomial transform to 3/2 −1/2 1/3 −1/4 1/5 ... = 1/2 + log 2.

Consider g(n) = 1/2 - 1 / (n+2) = 0, 1/6, 1/4, 3/10, 1/3. The Akiyama-Tanagiwa transforms gives:

0 1/6 1/4 3/10 1/3 5/14 ...
1/6 1/6 3/20 2/15 5/42 3/28 ...
0 1/30 1/20 2/35 5/84 5/84 ...
1/30 1/30 3/140 1/105 0 1/140 ...

0, g(n), is an autosequence of the second kind.

Euler OEISA198631 (n) / OEISA006519 (n + 1) without the second term (1/2) are the fractional intrinsic Euler numbers Ei(n) = 1, 0, −1/4, 0, 1/2, 0, −17/8, 0, ... The corresponding Akiyama transform is:

1 1 7/8 3/4 21/32
0 1/4 3/8 3/8 5/16
1/4 1/4 0 1/4 25/64
0 1/2 3/4 9/16 5/32
1/2 1/2 9/16 13/8 125/64

The first line is Eu(n). Eu(n) preceded by a zero is an autosequence of the first kind. It is linked to the Oresme numbers. The numerators of the second line are OEISA069834 preceded by 0. The difference table is:

0 1 1 7/8 3/4 21/32 19/32
1 0 1/8 1/8 3/32 1/16 5/128
−1 1/8 0 1/32 1/32 3/128 1/64

Arithmetical properties of the Bernoulli numbers

The Bernoulli numbers can be expressed in terms of the Riemann zeta function as Bn = −(1 − n) for integers n ≥ 0 provided for n = 0 the expression (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. For example, the Agoh–Giuga conjecture postulates that p is a prime number if and only if pBp − 1 is congruent to −1 modulo p. 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.

The Kummer theorems

The Bernoulli numbers are related to Fermat's Last Theorem (FLT) by Kummer's theorem,[38] which says:

If the odd prime p does not divide any of the numerators of the Bernoulli numbers B2, B4, ..., Bp − 3 then xp + yp + zp = 0 has no solutions in nonzero integers.

Prime numbers with this property are called regular primes. Another classical result of Kummer are the following congruences.[39]

Let p be an odd prime and b an even number such that p − 1 does not divide b. Then for any non-negative integer k
 

A generalization of these congruences goes by the name of p-adic continuity.

p-adic continuity

If b, m and n are positive integers such that m and n are not divisible by p − 1 and mn (mod pb − 1 (p − 1)), then

 

Since Bn = −(1 − n), 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 modulo 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 modulo p − 1 to a particular a ≢ 1 mod (p − 1), and so can be extended to a continuous function ζp(s) for all p-adic integers   the p-adic zeta function.

Ramanujan's congruences

The following relations, due to Ramanujan, provide a method for calculating Bernoulli numbers that is more efficient than the one given by their original recursive definition:

 

Von Staudt–Clausen theorem

The von Staudt–Clausen theorem was given by Karl Georg Christian von Staudt[40] and Thomas Clausen[41] independently in 1840. The theorem states that for every n > 0,

 

is an integer. The sum extends over all primes p for which p − 1 divides 2n.

A consequence of this is that the denominator of B2n is given by the product of all primes p for which p − 1 divides 2n. In particular, these denominators are square-free and 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 even and 2k + 1 − m > 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 the integer in question is zero. 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 Sn,m = m!{n
m
}
. 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.

A restatement of 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 numbers. In fact Marcel Riesz proved that the RH is equivalent to the following assertion:[42]

For every ε > 1/4 there exists a constant Cε > 0 (depending on ε) such that |R(x)| < Cεxε as x → ∞.

Here R(x) is the Riesz function

 

nk denotes the rising factorial power in the notation of D. E. Knuth. The numbers 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 numbers.

Generalized Bernoulli numbers

The generalized Bernoulli numbers are certain algebraic numbers, defined similarly to the Bernoulli numbers, that are related to special values of Dirichlet L-functions in the same way that Bernoulli numbers are related to special values of the Riemann zeta function.

Let χ be a Dirichlet character modulo f. The generalized Bernoulli numbers attached to χ are defined by

 

Apart from the exceptional B1,1 = 1/2, we have, for any Dirichlet character χ, that Bk,χ = 0 if χ(−1) ≠ (−1)k.

Generalizing the relation between Bernoulli numbers and values of the Riemann zeta function at non-positive integers, one has the for all integers k ≥ 1:

 

where L(s,χ) is the Dirichlet L-function of χ.[43]

Eisenstein–Kronecker number

Eisenstein–Kronecker numbers are an analogue of the generalized Bernoulli numbers for imaginary quadratic fields.[44][45] They are related to critical L-values of Hecke characters.[45]

Appendix

Assorted identities

  • Umbral calculus gives a compact form of Bernoulli's formula by using an abstract symbol B:
     

    where the symbol Bk that appears during binomial expansion of the parenthesized term is to be replaced by the Bernoulli number Bk (and B1 = +1/2). More suggestively and mnemonically, this may be written as a definite integral:

     

    Many other Bernoulli identities can be written compactly with this symbol, e.g.

     
  • 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 the following (n + 1) × (n + 1) determinant:[46]
     
    Thus the determinant is σn(1), the Stirling polynomial at x = 1.
  • For even-numbered Bernoulli numbers, B2p is given by the (p + 1) × (p + 1) determinant::[46]
     
  • Let n ≥ 1. Then (Leonhard Euler)
     
  • Let n ≥ 1. Then[47]
     
  • Let n ≥ 0. Then (Leopold Kronecker 1883)
     
  • Let n ≥ 1 and m ≥ 1. Then[48]
     
  • Let n ≥ 4 and
     
    the harmonic number. Then (H. Miki 1978)
     
  • Let n ≥ 4. Yuri Matiyasevich found (1997)
bernoulli, number, fraction, decimal0, 0000000001, 5000000002, 1666666663, 0000000004, 0333333335, 0000000006, 0238095237, 0000000008, 0333333339, 00000000010, 07575757511, 00000000012, 2730, 25311355313, 00000000014, 16666666615, 00000000016, 3617, 0921568621. Bernoulli numbers B n n fraction decimal0 1 1 0000000001 1 2 0 5000000002 1 6 0 1666666663 0 0 0000000004 1 30 0 0333333335 0 0 0000000006 1 42 0 0238095237 0 0 0000000008 1 30 0 0333333339 0 0 00000000010 5 66 0 07575757511 0 0 00000000012 691 2730 0 25311355313 0 0 00000000014 7 6 1 16666666615 0 0 00000000016 3617 510 7 09215686217 0 0 00000000018 43867 798 54 9711779419 0 0 00000000020 174611 330 529 1242424In mathematics the Bernoulli numbers Bn are a sequence of rational numbers which occur frequently in analysis The Bernoulli numbers appear in and can be defined by the Taylor series expansions of the tangent and hyperbolic tangent functions in Faulhaber s formula for the sum of m th powers of the first n positive integers in the Euler Maclaurin formula and in expressions for certain values of the Riemann zeta function The values of the first 20 Bernoulli numbers are given in the adjacent table Two conventions are used in the literature denoted here by B n displaystyle B n and B n displaystyle B n they differ only for n 1 where B 1 1 2 displaystyle B 1 1 2 and B 1 1 2 displaystyle B 1 1 2 For every odd n gt 1 Bn 0 For every even n gt 0 Bn is negative if n is divisible by 4 and positive otherwise The Bernoulli numbers are special values of the Bernoulli polynomials B n x displaystyle B n x with B n B n 0 displaystyle B n B n 0 and B n B n 1 displaystyle B n B n 1 1 The Bernoulli numbers were discovered around the same time by the Swiss mathematician Jacob Bernoulli after whom they are named and independently by Japanese mathematician Seki Takakazu Seki s discovery was posthumously published in 1712 2 3 4 in his work Katsuyō Sanpō Bernoulli s also posthumously in his Ars Conjectandi of 1713 Ada Lovelace s note G on the Analytical Engine from 1842 describes an algorithm for generating Bernoulli numbers with Babbage s machine 5 As a result the Bernoulli numbers have the distinction of being the subject of the first published complex computer program Contents 1 Notation 2 History 2 1 Early history 2 2 Reconstruction of Summae Potestatum 3 Definitions 3 1 Recursive definition 3 2 Explicit definition 3 3 Generating function 4 Bernoulli numbers and the Riemann zeta function 5 Efficient computation of Bernoulli numbers 6 Applications of the Bernoulli numbers 6 1 Asymptotic analysis 6 2 Sum of powers 6 3 Taylor series 6 4 Laurent series 6 5 Use in topology 7 Connections with combinatorial numbers 7 1 Connection with Worpitzky numbers 7 2 Connection with Stirling numbers of the second kind 7 3 Connection with Stirling numbers of the first kind 7 4 Connection with Pascal s triangle 7 5 Connection with Eulerian numbers 8 A binary tree representation 9 Integral representation and continuation 10 The relation to the Euler numbers and p 11 An algorithmic view the Seidel triangle 12 A combinatorial view alternating permutations 13 Related sequences 14 Arithmetical properties of the Bernoulli numbers 14 1 The Kummer theorems 14 2 p adic continuity 14 3 Ramanujan s congruences 14 4 Von Staudt Clausen theorem 14 5 Why do the odd Bernoulli numbers vanish 14 6 A restatement of the Riemann hypothesis 15 Generalized Bernoulli numbers 15 1 Eisenstein Kronecker number 16 Appendix 16 1 Assorted identities 17 See also 18 Notes 19 References 20 External linksNotation EditThe superscript used in this article distinguishes the two sign conventions for Bernoulli numbers Only the n 1 term is affected B n with B 1 1 2 OEIS A027641 OEIS A027642 is the sign convention prescribed by NIST and most modern textbooks 6 B n with B 1 1 2 OEIS A164555 OEIS A027642 was used in the older literature 1 and since 2022 by Donald Knuth 7 following Peter Luschny s Bernoulli Manifesto 8 In the formulas below one can switch from one sign convention to the other with the relation B n 1 n B n displaystyle B n 1 n B n or for integer n 2 or greater simply ignore it Since Bn 0 for all odd n gt 1 and many formulas only involve even index Bernoulli numbers a few authors write Bn instead of B2n This article does not follow that notation History EditEarly history Edit The Bernoulli numbers are rooted in the early history of the computation of sums of integer powers which have been of interest to mathematicians since antiquity A page from Seki Takakazu s Katsuyō Sanpō 1712 tabulating binomial coefficients and Bernoulli numbers Methods to calculate the sum of the first n positive integers the sum of the squares and of the cubes of the first n positive integers were known but there were no real formulas only descriptions given entirely in words Among the great mathematicians of antiquity to consider this problem were Pythagoras c 572 497 BCE Greece Archimedes 287 212 BCE Italy Aryabhata b 476 India Abu Bakr al Karaji d 1019 Persia and Abu Ali al Hasan ibn al Hasan ibn al Haytham 965 1039 Iraq During the late sixteenth and early seventeenth centuries mathematicians made significant progress In the West Thomas Harriot 1560 1621 of England Johann Faulhaber 1580 1635 of Germany Pierre de Fermat 1601 1665 and fellow French mathematician Blaise Pascal 1623 1662 all played important roles Thomas Harriot seems to have been the first to derive and write formulas for sums of powers using symbolic notation but even he calculated only up to the sum of the fourth powers Johann Faulhaber gave formulas for sums of powers up to the 17th power in his 1631 Academia Algebrae far higher than anyone before him but he did not give a general formula Blaise Pascal in 1654 proved Pascal s identity relating the sums of the p th powers of the first n positive integers for p 0 1 2 k The Swiss mathematician Jakob Bernoulli 1654 1705 was the first to realize the existence of a single sequence of constants B0 B1 B2 which provide a uniform formula for all sums of powers 9 The joy Bernoulli experienced when he hit upon the pattern needed to compute quickly and easily the coefficients of his formula for the sum of the c th powers for any positive integer c can be seen from his comment He wrote With the help of this table it took me less than half of a quarter of an hour to find that the tenth powers of the first 1000 numbers being added together will yield the sum 91 409 924 241 424 243 424 241 924 242 500 Bernoulli s result was published posthumously in Ars Conjectandi in 1713 Seki Takakazu independently discovered the Bernoulli numbers and his result was published a year earlier also posthumously in 1712 2 However Seki did not present his method as a formula based on a sequence of constants Bernoulli s formula for sums of powers is the most useful and generalizable formulation to date The coefficients in Bernoulli s formula are now called Bernoulli numbers following a suggestion of Abraham de Moivre Bernoulli s formula is sometimes called Faulhaber s formula after Johann Faulhaber who found remarkable ways to calculate sum of powers but never stated Bernoulli s formula According to Knuth 9 a rigorous proof of Faulhaber s formula was first published by Carl Jacobi in 1834 10 Knuth s in depth study of Faulhaber s formula concludes the nonstandard notation on the LHS is explained further on Faulhaber never discovered the Bernoulli numbers i e he never realized that a single sequence of constants B0 B1 B2 would provide a uniform n m 1 m 1 B 0 n m 1 m 1 1 B 1 n m m 1 2 B 2 n m 1 m 1 m B m n displaystyle quad sum n m frac 1 m 1 left B 0 n m 1 binom m 1 1 B 1 n m binom m 1 2 B 2 n m 1 cdots binom m 1 m B m n right or n m 1 m 1 B 0 n m 1 m 1 1 B 1 n m m 1 2 B 2 n m 1 1 m m 1 m B m n displaystyle quad sum n m frac 1 m 1 left B 0 n m 1 binom m 1 1 B 1 n m binom m 1 2 B 2 n m 1 cdots 1 m binom m 1 m B m n right dd for all sums of powers He never mentioned for example the fact that almost half of the coefficients turned out to be zero after he had converted his formulas for S nm from polynomials in N to polynomials in n 11 Reconstruction of Summae Potestatum Edit Jakob Bernoulli s Summae Potestatum 1713 a displaystyle The Bernoulli numbers OEIS A164555 n OEIS A027642 n were introduced by Jakob Bernoulli in the book Ars Conjectandi published posthumously in 1713 page 97 The main formula can be seen in the second half of the corresponding facsimile 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 The expression c c 1 c 2 c 3 means c c 1 c 2 c 3 the small dots are used as grouping symbols Using today s terminology these expressions are falling factorial powers ck 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 b 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 k 1 n k c n c 1 c 1 1 2 n c k 2 c B k k c k 1 n c k 1 displaystyle sum k 1 n k c frac n c 1 c 1 frac 1 2 n c sum k 2 c frac B k k c underline k 1 n c k 1 This formula suggests setting B1 1 2 when switching from the so called archaic enumeration which uses only the even indices 2 4 6 to the modern form more on different conventions in the next paragraph Most striking in this context is the fact that the falling factorial ck 1 has for k 0 the value 1 c 1 12 Thus Bernoulli s formula can be written k 1 n k c k 0 c B k k c k 1 n c k 1 displaystyle sum k 1 n k c sum k 0 c frac B k k c underline k 1 n c k 1 if B1 1 2 recapturing the value Bernoulli gave to the coefficient at that position The formula for k 1 n k 9 displaystyle textstyle sum k 1 n k 9 in the first half of the quotation by Bernoulli above contains an error at the last term it should be 3 20 n 2 displaystyle tfrac 3 20 n 2 instead of 1 12 n 2 displaystyle tfrac 1 12 n 2 Definitions EditMany characterizations of the Bernoulli numbers have been found in the last 300 years and each could be used to introduce these numbers Here only three of the most useful ones are mentioned a recursive equation an explicit formula a generating function For the proof of the equivalence of the three approaches 13 Recursive definition Edit The Bernoulli numbers obey the sum formulas 1 k 0 m m 1 k B k d m 0 k 0 m m 1 k B k m 1 displaystyle begin aligned sum k 0 m binom m 1 k B k amp delta m 0 sum k 0 m binom m 1 k B k amp m 1 end aligned where m 0 1 2 displaystyle m 0 1 2 and d denotes the Kronecker delta Solving for B m displaystyle B m mp gives the recursive formulas B m d m 0 k 0 m 1 m k B k m k 1 B m 1 k 0 m 1 m k B k m k 1 displaystyle begin aligned B m amp delta m 0 sum k 0 m 1 binom m k frac B k m k 1 B m amp 1 sum k 0 m 1 binom m k frac B k m k 1 end aligned Explicit definition Edit In 1893 Louis Saalschutz listed a total of 38 explicit formulas for the Bernoulli numbers 14 usually giving some reference in the older literature One of them is for m 1 displaystyle m geq 1 B m k 0 m v 0 k 1 v k v v m k 1 B m k 0 m v 0 k 1 v k v v 1 m k 1 displaystyle begin aligned B m amp sum k 0 m sum v 0 k 1 v binom k v frac v m k 1 B m amp sum k 0 m sum v 0 k 1 v binom k v frac v 1 m k 1 end aligned Generating function Edit The exponential generating functions are t e t 1 t 2 coth t 2 1 m 0 B m t m m t 1 e t t 2 coth t 2 1 m 0 B m t m m displaystyle begin alignedat 3 frac t e t 1 amp frac t 2 left operatorname coth frac t 2 1 right amp amp sum m 0 infty frac B m t m m frac t 1 e t amp frac t 2 left operatorname coth frac t 2 1 right amp amp sum m 0 infty frac B m t m m end alignedat where the substitution is t t displaystyle t to t The ordinary generating function z 1 ps 1 z 1 m 0 B m z m displaystyle z 1 psi 1 z 1 sum m 0 infty B m z m is an asymptotic series It contains the trigamma function ps1 Bernoulli numbers and the Riemann zeta function Edit The Bernoulli numbers as given by the Riemann zeta function The Bernoulli numbers can be expressed in terms of the Riemann zeta function B n nz 1 n for n 1 Here the argument of the zeta function is 0 or negative By means of the zeta functional equation and the gamma reflection formula the following relation can be obtained 15 B 2 n 1 n 1 2 2 n 2 p 2 n z 2 n displaystyle B 2n frac 1 n 1 2 2n 2 pi 2n zeta 2n quad for n 1 Now the argument of the zeta function is positive It then follows from z 1 n and Stirling s formula that B 2 n 4 p n n p e 2 n displaystyle B 2n sim 4 sqrt pi n left frac n pi e right 2n quad for n Efficient computation of Bernoulli numbers EditIn 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 16 which require only O p log p 2 operations see big O notation David Harvey 17 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 e and claims that this implementation is significantly faster than implementations based on other methods Using this implementation Harvey computed Bn for n 108 Harvey s implementation has been included in SageMath since version 3 1 Prior to that Bernd Kellner 18 computed Bn to full precision for n 106 in December 2002 and Oleksandr Pavlyk 19 for n 107 with Mathematica in April 2008 Computer Year n Digits J Bernoulli 1689 10 1L Euler 1748 30 8J C Adams 1878 62 36D E Knuth T J Buckholtz 1967 1672 3330G Fee S Plouffe 1996 10000 27677G Fee S Plouffe 1996 100000 376755B C Kellner 2002 1000 000 4767 529O Pavlyk 2008 10000 000 57675 260D Harvey 2008 100000 000 676752 569 Digits is to be understood as the exponent of 10 when Bn is written as a real number in normalized scientific notation dd A possible algorithm for computing Bernoulli numbers in the Julia programming language is given by 14 b Array Float64 undef n 1 b 1 1 b 2 0 5 for m 2 n for k 0 m for v 0 k b m 1 1 v binomial k v v m k 1 end end end return bApplications of the Bernoulli numbers EditAsymptotic analysis Edit Arguably the most important application of the Bernoulli numbers 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 20 k a b 1 f k a b f x d x k 1 m B k k f k 1 b f k 1 a R f m displaystyle sum k a b 1 f k int a b f x dx sum k 1 m frac B k k f k 1 b f k 1 a R f m This formulation assumes the convention B 1 1 2 Using the convention B 1 1 2 the formula becomes k a 1 b f k a b f x d x k 1 m B k k f k 1 b f k 1 a R f m displaystyle sum k a 1 b f k int a b f x dx sum k 1 m frac B k k f k 1 b f k 1 a R f m Here f 0 f displaystyle f 0 f i e the zeroth order derivative of f displaystyle f is just f displaystyle f Moreover let f 1 displaystyle f 1 denote an antiderivative of f displaystyle f By the fundamental theorem of calculus a b f x d x f 1 b f 1 a displaystyle int a b f x dx f 1 b f 1 a Thus the last formula can be further simplified to the following succinct form of the Euler Maclaurin formula k a b f k k 0 m B k k f k 1 b f k 1 a R f m displaystyle sum k a b f k sum k 0 m frac B k k f k 1 b f k 1 a R f m This form is for example the source for the important Euler Maclaurin expansion of the zeta function z s k 0 m B k k s k 1 R s m B 0 0 s 1 B 1 1 s 0 B 2 2 s 1 R s m 1 s 1 1 2 1 12 s R s m displaystyle begin aligned zeta s amp sum k 0 m frac B k k s overline k 1 R s m amp frac B 0 0 s overline 1 frac B 1 1 s overline 0 frac B 2 2 s overline 1 cdots R s m amp frac 1 s 1 frac 1 2 frac 1 12 s cdots R s m end aligned Here sk denotes the rising factorial power 21 Bernoulli numbers are also frequently used in other kinds of asymptotic expansions The following example is the classical Poincare type asymptotic expansion of the digamma function ps ps z ln z k 1 B k k z k displaystyle psi z sim ln z sum k 1 infty frac B k kz k Sum of powers Edit Main article Faulhaber s formula Bernoulli numbers feature prominently in the closed form expression of the sum of the m th powers of the first n positive integers For m n 0 define S m n k 1 n k m 1 m 2 m n m displaystyle S m n sum k 1 n k m 1 m 2 m cdots n m This expression can always be rewritten as a polynomial in n of degree m 1 The coefficients of these polynomials are related to the Bernoulli numbers by Bernoulli s formula S m n 1 m 1 k 0 m m 1 k B k n m 1 k m k 0 m B k n m 1 k k m 1 k displaystyle S m n frac 1 m 1 sum k 0 m binom m 1 k B k n m 1 k m sum k 0 m frac B k n m 1 k k m 1 k where m 1k denotes the binomial coefficient For example taking m to be 1 gives the triangular numbers 0 1 3 6 OEIS A000217 1 2 n 1 2 B 0 n 2 2 B 1 n 1 1 2 n 2 n displaystyle 1 2 cdots n frac 1 2 B 0 n 2 2B 1 n 1 tfrac 1 2 n 2 n Taking m to be 2 gives the square pyramidal numbers 0 1 5 14 OEIS A000330 1 2 2 2 n 2 1 3 B 0 n 3 3 B 1 n 2 3 B 2 n 1 1 3 n 3 3 2 n 2 1 2 n displaystyle 1 2 2 2 cdots n 2 frac 1 3 B 0 n 3 3B 1 n 2 3B 2 n 1 tfrac 1 3 left n 3 tfrac 3 2 n 2 tfrac 1 2 n right Some authors use the alternate convention for Bernoulli numbers and state Bernoulli s formula in this way S m n 1 m 1 k 0 m 1 k m 1 k B k n m 1 k displaystyle S m n frac 1 m 1 sum k 0 m 1 k binom m 1 k B k n m 1 k Bernoulli s formula is sometimes called Faulhaber s formula after Johann Faulhaber who also found remarkable ways to calculate sums of powers Faulhaber s formula was generalized by V Guo and J Zeng to a q analog 22 Taylor series Edit The Bernoulli numbers appear in the Taylor series expansion of many trigonometric functions and hyperbolic functions Tangent tan x n 1 1 n 1 2 2 n 2 2 n 1 B 2 n 2 n x 2 n 1 x lt p 2 displaystyle begin aligned tan x amp sum n 1 infty frac 1 n 1 2 2n 2 2n 1 B 2n 2n x 2n 1 amp left x right amp lt frac pi 2 end aligned Cotangent cot x 1 x n 0 1 n B 2 n 2 x 2 n 2 n 0 lt x lt p displaystyle begin aligned cot x amp frac 1 x sum n 0 infty frac 1 n B 2n 2x 2n 2n amp qquad 0 lt x lt pi end aligned Hyperbolic tangent tanh x n 1 2 2 n 2 2 n 1 B 2 n 2 n x 2 n 1 x lt p 2 displaystyle begin aligned tanh x amp sum n 1 infty frac 2 2n 2 2n 1 B 2n 2n x 2n 1 amp x amp lt frac pi 2 end aligned Hyperbolic cotangent coth x 1 x n 0 B 2 n 2 x 2 n 2 n 0 lt x lt p displaystyle begin aligned coth x amp frac 1 x sum n 0 infty frac B 2n 2x 2n 2n amp qquad qquad 0 lt x lt pi end aligned Laurent series Edit The Bernoulli numbers appear in the following Laurent series 23 Digamma function ps z ln z k 1 B k k z k displaystyle psi z ln z sum k 1 infty frac B k kz k Use in topology Edit The Kervaire Milnor formula for the order of the cyclic group of diffeomorphism classes of exotic 4n 1 spheres which bound parallelizable manifolds involves Bernoulli numbers Let ESn be the number of such exotic spheres for n 2 then ES n 2 2 n 2 2 4 n 3 Numerator B 4 n 4 n displaystyle textit ES n 2 2n 2 2 4n 3 operatorname Numerator left frac B 4n 4n right The Hirzebruch signature theorem for the L genus of a smooth oriented closed manifold of dimension 4n also involves Bernoulli numbers Connections with combinatorial numbers EditThe 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 Worpitzky numbers Edit 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 W n k v 0 k 1 v k v 1 n k v k v displaystyle W n k sum v 0 k 1 v k v 1 n frac k v k v They can also be expressed through the Stirling numbers of the second kind W n k k n 1 k 1 displaystyle W n k k left n 1 atop k 1 right A Bernoulli number is then introduced as an inclusion exclusion sum of Worpitzky numbers weighted by the harmonic sequence 1 1 2 1 3 B n k 0 n 1 k W n k k 1 k 0 n 1 k 1 v 0 k 1 v v 1 n k v displaystyle B n sum k 0 n 1 k frac W n k k 1 sum k 0 n frac 1 k 1 sum v 0 k 1 v v 1 n k choose v B0 1 B1 1 1 2 B2 1 3 2 2 3 B3 1 7 2 12 3 6 4 B4 1 15 2 50 3 60 4 24 5 B5 1 31 2 180 3 390 4 360 5 120 6 B6 1 63 2 602 3 2100 4 3360 5 2520 6 720 7This representation has B 1 1 2 Consider the sequence sn n 0 From Worpitzky s numbers OEIS A028246 OEIS A163626 applied to s0 s0 s1 s0 s1 s2 s0 s1 s2 s3 is identical to the Akiyama Tanigawa transform applied to sn see Connection with Stirling numbers of the first kind This can be seen via the table Identity ofWorpitzky s representation and Akiyama Tanigawa transform 1 0 1 0 0 1 0 0 0 1 0 0 0 0 11 1 0 2 2 0 0 3 3 0 0 0 4 41 3 2 0 4 10 6 0 0 9 21 121 7 12 6 0 8 38 54 241 15 50 60 24The first row represents s0 s1 s2 s3 s4 Hence for the second fractional Euler numbers OEIS A198631 n OEIS A006519 n 1 E0 1 E1 1 1 2 E2 1 3 2 2 4 E3 1 7 2 12 4 6 8 E4 1 15 2 50 4 60 8 24 16 E5 1 31 2 180 4 390 8 360 16 120 32 E6 1 63 2 602 4 2100 8 3360 16 2520 32 720 64A second formula representing the Bernoulli numbers by the Worpitzky numbers is for n 1 B n n 2 n 1 2 k 0 n 1 2 k W n 1 k displaystyle B n frac n 2 n 1 2 sum k 0 n 1 2 k W n 1 k The simplified second Worpitzky s representation of the second Bernoulli numbers is OEIS A164555 n 1 OEIS A027642 n 1 n 1 2n 2 2 OEIS A198631 n OEIS A006519 n 1 which links the second Bernoulli numbers to the second fractional Euler numbers The beginning is 1 2 1 6 0 1 30 0 1 42 1 2 1 3 3 14 2 15 5 62 1 21 1 1 2 0 1 4 0 1 2 The numerators of the first parentheses are OEIS A111701 see Connection with Stirling numbers of the first kind Connection with Stirling numbers of the second kind Edit If S k m denotes Stirling numbers of the second kind 24 then one has j k m 0 k j m S k m displaystyle j k sum m 0 k j underline m S k m where jm denotes the falling factorial If one defines the Bernoulli polynomials Bk j as 25 B k j k m 0 k 1 j m 1 S k 1 m m B k displaystyle B k j k sum m 0 k 1 binom j m 1 S k 1 m m B k where Bk for k 0 1 2 are the Bernoulli numbers Then after the following property of the binomial coefficient j m j 1 m 1 j m 1 displaystyle binom j m binom j 1 m 1 binom j m 1 one has j k B k 1 j 1 B k 1 j k 1 displaystyle j k frac B k 1 j 1 B k 1 j k 1 One also has the following for Bernoulli polynomials 25 B k j n 0 k k n B n j k n displaystyle B k j sum n 0 k binom k n B n j k n The coefficient of j in jm 1 is 1 m m 1 Comparing the coefficient of j in the two expressions of Bernoulli polynomials one has B k m 0 k 1 m m m 1 S k m displaystyle B k sum m 0 k 1 m frac m m 1 S k m resulting in B1 1 2 which is an explicit formula for Bernoulli numbers and can be used to prove Von Staudt Clausen theorem 26 27 28 Connection with Stirling numbers of the first kind Edit The two main formulas relating the unsigned Stirling numbers of the first kind nm to the Bernoulli numbers with B1 1 2 are 1 m k 0 m 1 k m 1 k 1 B k 1 m 1 displaystyle frac 1 m sum k 0 m 1 k left m 1 atop k 1 right B k frac 1 m 1 and the inversion of this sum for n 0 m 0 1 m k 0 m 1 k m 1 k 1 B n k A n m displaystyle frac 1 m sum k 0 m 1 k left m 1 atop k 1 right B n k A n m Here the number An m are the rational Akiyama Tanigawa numbers the first few of which are displayed in the following table Akiyama Tanigawa number mn 0 1 2 3 40 1 1 2 1 3 1 4 1 51 1 2 1 3 1 4 1 5 2 1 6 1 6 3 20 3 0 1 30 4 1 30 The Akiyama Tanigawa numbers satisfy a simple recurrence relation which can be exploited to iteratively compute the Bernoulli numbers This leads to the algorithm shown in the section algorithmic description above See OEIS A051714 OEIS A051715 An autosequence is a sequence which has its inverse binomial transform equal to the signed sequence If the main diagonal is zeroes OEIS A000004 the autosequence is of the first kind Example OEIS A000045 the Fibonacci numbers If the main diagonal is the first upper diagonal multiplied by 2 it is of the second kind Example OEIS A164555 OEIS A027642 the second Bernoulli numbers see OEIS A190339 The Akiyama Tanigawa transform applied to 2 n 1 OEIS A000079 leads to OEIS A198631 n OEIS A06519 n 1 Hence Akiyama Tanigawa transform for the second Euler numbers mn 0 1 2 3 40 1 1 2 1 4 1 8 1 161 1 2 1 2 3 8 1 4 2 0 1 4 3 8 3 1 4 1 4 4 0 See OEIS A209308 and OEIS A227577 OEIS A198631 n OEIS A006519 n 1 are the second fractional Euler numbers and an autosequence of the second kind OEIS A164555 n 2 OEIS A027642 n 2 1 6 0 1 30 0 1 42 2n 3 2 n 2 3 14 3 15 2 62 5 21 OEIS A198631 n 1 OEIS A006519 n 2 1 2 0 1 4 0 1 2 Also valuable for OEIS A027641 OEIS A027642 see Connection with Worpitzky numbers Connection with Pascal s triangle Edit There are formulas connecting Pascal s triangle to Bernoulli numbers c B n A n n 1 displaystyle B n frac A n n 1 where A n displaystyle A n is the determinant of a n by n Hessenberg matrix part of Pascal s triangle whose elements are a i k 0 if k gt 1 i i 1 k 1 otherwise displaystyle a i k begin cases 0 amp text if k gt 1 i i 1 choose k 1 amp text otherwise end cases Example B 6 det 1 2 0 0 0 0 1 3 3 0 0 0 1 4 6 4 0 0 1 5 10 10 5 0 1 6 15 20 15 6 1 7 21 35 35 21 7 120 5040 1 42 displaystyle B 6 frac det begin pmatrix 1 amp 2 amp 0 amp 0 amp 0 amp 0 1 amp 3 amp 3 amp 0 amp 0 amp 0 1 amp 4 amp 6 amp 4 amp 0 amp 0 1 amp 5 amp 10 amp 10 amp 5 amp 0 1 amp 6 amp 15 amp 20 amp 15 amp 6 1 amp 7 amp 21 amp 35 amp 35 amp 21 end pmatrix 7 frac 120 5040 frac 1 42 Connection with Eulerian numbers Edit There are formulas connecting Eulerian numbers nm to Bernoulli numbers m 0 n 1 m n m 2 n 1 2 n 1 1 B n 1 n 1 m 0 n 1 m n m n m 1 n 1 B n displaystyle begin aligned sum m 0 n 1 m left langle n atop m right rangle amp 2 n 1 2 n 1 1 frac B n 1 n 1 sum m 0 n 1 m left langle n atop m right rangle binom n m 1 amp n 1 B n end aligned Both formulae 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 A binary tree representation EditThe Stirling polynomials sn x are related to the Bernoulli numbers by Bn n sn 1 S C Woon described an algorithm to compute sn 1 as a binary tree 29 Woon s recursive algorithm for n 1 starts by assigning to the root node N 1 2 Given a node N a1 a2 ak of the tree the left child of the node is L N a1 a2 1 a3 ak and the right child R N a1 2 a2 ak A node N a1 a2 ak is written as a2 ak in the initial part of the tree represented above with denoting the sign of a1 Given a node N the factorial of N is defined as N a 1 k 2 length N a k displaystyle N a 1 prod k 2 operatorname length N a k Restricted to the nodes N of a fixed tree level n the sum of 1 N is sn 1 thus B n tree level n N node of n N displaystyle B n sum stackrel N text node of text tree level n frac n N For example B1 1 1 2 B2 2 1 3 1 2 2 B3 3 1 4 1 2 3 1 3 2 1 2 2 2 Integral representation and continuation EditThe integral b s 2 e s i p 2 0 s t s 1 e 2 p t d t t s 2 s 1 z s p s i s 2 s z s 2 p i s displaystyle b s 2e si pi 2 int 0 infty frac st s 1 e 2 pi t frac dt t frac s 2 s 1 frac zeta s pi s i s frac 2s zeta s 2 pi i s has as special values b 2n B2n for n gt 0 For example b 3 3 2 z 3 p 3i and b 5 15 2 z 5 p 5i Here z is the Riemann zeta function and i is the imaginary unit Leonhard Euler Opera Omnia Ser 1 Vol 10 p 351 considered these numbers and calculated p 3 2 p 3 1 1 2 3 1 3 3 0 0581522 q 15 2 p 5 1 1 2 5 1 3 5 0 0254132 displaystyle begin aligned p amp frac 3 2 pi 3 left 1 frac 1 2 3 frac 1 3 3 cdots right 0 0581522 ldots q amp frac 15 2 pi 5 left 1 frac 1 2 5 frac 1 3 5 cdots right 0 0254132 ldots end aligned Another similar integral representation is b s e s i p 2 2 s 1 0 s t s sinh p t d t t 2 e s i p 2 2 s 1 0 e p t s t s 1 e 2 p t d t t displaystyle b s frac e si pi 2 2 s 1 int 0 infty frac st s sinh pi t frac dt t frac 2e si pi 2 2 s 1 int 0 infty frac e pi t st s 1 e 2 pi t frac dt t The relation to the Euler numbers and p EditThe 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 p 42n 22n times larger than the Bernoulli numbers B2n In consequence p 2 2 2 n 4 2 n B 2 n E 2 n displaystyle pi sim 2 2 2n 4 2n frac B 2n E 2n This asymptotic equation reveals that p lies in the common root of both the Bernoulli and the Euler numbers In fact p could be computed from these rational approximations Bernoulli numbers can be expressed through the Euler numbers and vice versa Since for odd n Bn En 0 with the exception B1 it suffices to consider the case when n is even B n k 0 n 1 n 1 k n 4 n 2 n E k n 2 4 6 E n k 1 n n k 1 2 k 4 k k B k n 2 4 6 displaystyle begin aligned B n amp sum k 0 n 1 binom n 1 k frac n 4 n 2 n E k amp n amp 2 4 6 ldots 6pt E n amp sum k 1 n binom n k 1 frac 2 k 4 k k B k amp n amp 2 4 6 ldots end aligned These conversion formulas express a connection 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 p These numbers are defined for n gt 1 as S n 2 2 p n k 4 k 1 n k 0 1 1 2 2 displaystyle S n 2 left frac 2 pi right n sum k infty infty 4k 1 n qquad k 0 1 1 2 2 ldots and S1 1 by convention 30 The magic of these numbers lies in the fact that they turn out to be rational numbers This was first proved by Leonhard Euler in a landmark paper De summis serierum reciprocarum On the sums of series of reciprocals and has fascinated mathematicians ever since 31 The first few of these numbers are S n 1 1 1 2 1 3 5 24 2 15 61 720 17 315 277 8064 62 2835 displaystyle S n 1 1 frac 1 2 frac 1 3 frac 5 24 frac 2 15 frac 61 720 frac 17 315 frac 277 8064 frac 62 2835 ldots OEIS A099612 OEIS A099617 These are the coefficients in the expansion of sec x tan x 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 B n 1 n 2 n even n 2 n 4 n S n n 2 3 E n 1 n 2 n even n S n 1 n 0 1 displaystyle begin aligned B n amp 1 left lfloor frac n 2 right rfloor n text even frac n 2 n 4 n S n amp n amp 2 3 ldots E n amp 1 left lfloor frac n 2 right rfloor n text even n S n 1 amp n amp 0 1 ldots end aligned 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 2Sn Sn 1 when n is even The Rn are rational approximations to p and two successive terms always enclose the true value of p Beginning with n 1 the sequence starts OEIS A132049 OEIS A132050 2 4 3 16 5 25 8 192 61 427 136 4352 1385 12465 3968 158720 50521 p displaystyle 2 4 3 frac 16 5 frac 25 8 frac 192 61 frac 427 136 frac 4352 1385 frac 12465 3968 frac 158720 50521 ldots quad longrightarrow pi These rational numbers also appear in the last paragraph of Euler s paper cited above Consider the Akiyama Tanigawa transform for the sequence OEIS A046978 n 2 OEIS A016116 n 1 0 1 1 2 0 1 4 1 4 1 8 01 1 2 1 3 4 0 5 8 3 42 1 2 1 2 9 4 5 2 5 83 1 7 2 3 4 15 24 5 2 11 2 99 45 8 77 26 61 2From the second the numerators of the first column are the denominators of Euler s formula The first column is 1 2 OEIS A163982 An algorithmic view the Seidel triangle EditThe 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 sometimes called Euler zigzag numbers are integers T n 1 1 1 2 5 16 61 272 1385 7936 50521 353792 n 0 1 2 3 displaystyle T n 1 1 1 2 5 16 61 272 1385 7936 50521 353792 ldots quad n 0 1 2 3 ldots OEIS A000111 See OEIS A253671 Thus the above representations of the Bernoulli and Euler numbers can be rewritten in terms of this sequence as B n 1 n 2 n even n 2 n 4 n T n 1 n 2 3 E n 1 n 2 n even T n 1 n 0 1 displaystyle begin aligned B n amp 1 left lfloor frac n 2 right rfloor n text even frac n 2 n 4 n T n 1 amp n amp 2 3 ldots E n amp 1 left lfloor frac n 2 right rfloor n text even T n 1 amp n amp 0 1 ldots end aligned 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 simple to calculate Tn 32 1 1 1 2 2 1 2 4 5 5 16 16 14 10 5 displaystyle begin array crrrcc amp amp color red 1 amp amp amp amp rightarrow amp color blue 1 amp color red 1 amp amp color red 2 amp color blue 2 amp color blue 1 amp leftarrow rightarrow amp color blue 2 amp color blue 4 amp color blue 5 amp color red 5 color red 16 amp color blue 16 amp color blue 14 amp color blue 10 amp color blue 5 amp leftarrow end array Seidel s algorithm for Tn 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 Seidel s algorithm is in fact much more general see the exposition of Dominique Dumont 33 and was rediscovered several times thereafter Similar to Seidel s approach D E Knuth and T J Buckholtz 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 34 V I Arnold 35 rediscovered Seidel s algorithm and later Millar Sloane and Young popularized Seidel s algorithm under the name boustrophedon transform Triangular form 11 12 2 12 4 5 516 16 14 10 516 32 46 56 61 61272 272 256 224 178 122 61Only OEIS A000657 with one 1 and OEIS A214267 with two 1s are in the OEIS Distribution with a supplementary 1 and one 0 in the following rows 10 1 1 1 00 1 2 25 5 4 2 00 5 10 14 16 16 61 61 56 46 32 16 0This is OEIS A239005 a signed version of OEIS A008280 The main andiagonal is OEIS A122045 The main diagonal is OEIS A155585 The central column is OEIS A099023 Row sums 1 1 2 5 16 61 See OEIS A163747 See the array beginning with 1 1 0 2 0 16 0 below The Akiyama Tanigawa algorithm applied to OEIS A046978 n 1 OEIS A016116 n yields 1 1 1 2 0 1 4 1 4 1 80 1 3 2 1 0 3 4 1 1 3 2 4 15 40 5 15 2 15 5 51 20 61 611 The first column is OEIS A122045 Its binomial transform leads to 1 1 0 2 0 16 00 1 2 2 16 16 1 1 4 14 320 5 10 465 5 560 61 61The first row of this array is OEIS A155585 The absolute values of the increasing antidiagonals are OEIS A008280 The sum of the antidiagonals is OEIS A163747 n 1 2 The second column is 1 1 1 5 5 61 61 1385 1385 Its binomial transform yields 1 2 2 4 16 32 2721 0 6 12 48 240 1 6 6 60 192 5 0 66 325 66 6661 0 61The first row of this array is 1 2 2 4 16 32 272 544 7936 15872 353792 707584 The absolute values of the second bisection are the double of the absolute values of the first bisection Consider the Akiyama Tanigawa algorithm applied to OEIS A046978 n OEIS A158780 n 1 abs OEIS A117575 n 1 1 2 2 3 2 1 3 4 3 4 7 8 1 17 16 17 16 33 32 1 2 2 3 2 1 3 4 3 4 1 0 3 2 2 5 4 0 1 3 3 2 3 25 42 3 27 2 135 21 3 2 16 45 61The first column whose the absolute values are OEIS A000111 could be the numerator of a trigonometric function OEIS A163747 is an autosequence of the first kind the main diagonal is OEIS A000004 The corresponding array is 0 1 1 2 5 16 61 1 0 3 3 21 451 3 0 24 242 3 24 0 5 21 24 16 45 61The first two upper diagonals are 1 3 24 402 1 n 1 OEIS A002832 The sum of the antidiagonals is 0 2 0 10 2 OEIS A122045 n 1 OEIS A163982 is an autosequence of the second kind like for instance OEIS A164555 OEIS A027642 Hence the array 2 1 1 2 5 16 61 1 2 1 7 11 77 1 1 8 4 882 7 4 925 11 88 16 77 61The main diagonal here 2 2 8 92 is the double of the first upper one here OEIS A099023 The sum of the antidiagonals is 2 0 4 0 2 OEIS A155585 n 1 OEIS A163747 OEIS A163982 2 OEIS A122045 A combinatorial view alternating permutations EditMain article Alternating permutations Around 1880 three years after the publication of Seidel s algorithm Desire Andre proved a now classic result of combinatorial analysis 36 37 Looking at the first terms of the Taylor expansion of the trigonometric functions tan x and sec x Andre made a startling discovery tan x x 2 x 3 3 16 x 5 5 272 x 7 7 7936 x 9 9 sec x 1 x 2 2 5 x 4 4 61 x 6 6 1385 x 8 8 50521 x 10 10 displaystyle begin aligned tan x amp x frac 2x 3 3 frac 16x 5 5 frac 272x 7 7 frac 7936x 9 9 cdots 6pt sec x amp 1 frac x 2 2 frac 5x 4 4 frac 61x 6 6 frac 1385x 8 8 frac 50521x 10 10 cdots end aligned 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 tan x sec x 1 x 1 2 x 2 1 3 x 3 5 24 x 4 2 15 x 5 61 720 x 6 displaystyle tan x sec x 1 x tfrac 1 2 x 2 tfrac 1 3 x 3 tfrac 5 24 x 4 tfrac 2 15 x 5 tfrac 61 720 x 6 cdots Andre 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 Related sequences EditThe arithmetic mean of the first and the second Bernoulli numbers are the associate Bernoulli numbers B0 1 B1 0 B2 1 6 B3 0 B4 1 30 OEIS A176327 OEIS A027642 Via the second row of its inverse Akiyama Tanigawa transform OEIS A177427 they lead to Balmer series OEIS A061037 OEIS A061038 The Akiyama Tanigawa algorithm applied to OEIS A060819 n 4 OEIS A145979 n leads to the Bernoulli numbers OEIS A027641 OEIS A027642 OEIS A164555 OEIS A027642 or OEIS A176327 OEIS A176289 without B1 named intrinsic Bernoulli numbers Bi n 1 5 6 3 4 7 10 2 31 6 1 6 3 20 2 15 5 420 1 30 1 20 2 35 5 84 1 30 1 30 3 140 1 105 00 1 42 1 28 4 105 1 28Hence another link between the intrinsic Bernoulli numbers and the Balmer series via OEIS A145979 n OEIS A145979 n 2 0 2 1 6 is a permutation of the non negative numbers The terms of the first row are f n 1 2 1 n 2 2 f n is an autosequence of the second kind 3 2 f n leads by its inverse binomial transform to 3 2 1 2 1 3 1 4 1 5 1 2 log 2 Consider g n 1 2 1 n 2 0 1 6 1 4 3 10 1 3 The Akiyama Tanagiwa transforms gives 0 1 6 1 4 3 10 1 3 5 14 1 6 1 6 3 20 2 15 5 42 3 28 0 1 30 1 20 2 35 5 84 5 84 1 30 1 30 3 140 1 105 0 1 140 0 g n is an autosequence of the second kind Euler OEIS A198631 n OEIS A006519 n 1 without the second term 1 2 are the fractional intrinsic Euler numbers Ei n 1 0 1 4 0 1 2 0 17 8 0 The corresponding Akiyama transform is 1 1 7 8 3 4 21 320 1 4 3 8 3 8 5 16 1 4 1 4 0 1 4 25 640 1 2 3 4 9 16 5 321 2 1 2 9 16 13 8 125 64The first line is Eu n Eu n preceded by a zero is an autosequence of the first kind It is linked to the Oresme numbers The numerators of the second line are OEIS A069834 preceded by 0 The difference table is 0 1 1 7 8 3 4 21 32 19 321 0 1 8 1 8 3 32 1 16 5 128 1 1 8 0 1 32 1 32 3 128 1 64Arithmetical properties of the Bernoulli numbers EditThe Bernoulli numbers can be expressed in terms of the Riemann zeta function as Bn nz 1 n for integers n 0 provided for n 0 the expression nz 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 For example the Agoh Giuga conjecture postulates that p is a prime number if and only if pBp 1 is congruent to 1 modulo p 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 The Kummer theorems Edit The Bernoulli numbers are related to Fermat s Last Theorem FLT by Kummer s theorem 38 which says If the odd prime p does not divide any of the numerators of the Bernoulli numbers B2 B4 Bp 3 then xp yp zp 0 has no solutions in nonzero integers Prime numbers with this property are called regular primes Another classical result of Kummer are the following congruences 39 Main article Kummer s congruence Let p be an odd prime and b an even number such that p 1 does not divide b Then for any non negative integer kB k p 1 b k p 1 b B b b mod p displaystyle frac B k p 1 b k p 1 b equiv frac B b b pmod p dd A generalization of these congruences goes by the name of p adic continuity p adic continuity Edit If b m and n are positive integers such that m and n are not divisible by p 1 and m n mod pb 1 p 1 then 1 p m 1 B m m 1 p n 1 B n n mod p b displaystyle 1 p m 1 frac B m m equiv 1 p n 1 frac B n n pmod p b Since Bn nz 1 n this can also be written 1 p u z u 1 p v z v mod p b displaystyle left 1 p u right zeta u equiv left 1 p v right zeta v pmod p b where u 1 m and v 1 n so that u and v are nonpositive and not congruent to 1 modulo p 1 This tells us that the Riemann zeta function with 1 p s taken out of the Euler product formula is continuous in the p adic numbers on odd negative integers congruent modulo p 1 to a particular a 1 mod p 1 and so can be extended to a continuous function zp s for all p adic integers Z p displaystyle mathbb Z p the p adic zeta function Ramanujan s congruences Edit The following relations due to Ramanujan provide a method for calculating Bernoulli numbers that is more efficient than the one given by their original recursive definition m 3 m B m m 3 3 j 1 m 6 m 3 m 6 j B m 6 j if m 0 mod 6 m 3 3 j 1 m 2 6 m 3 m 6 j B m 6 j if m 2 mod 6 m 3 6 j 1 m 4 6 m 3 m 6 j B m 6 j if m 4 mod 6 displaystyle binom m 3 m B m begin cases frac m 3 3 sum limits j 1 frac m 6 binom m 3 m 6j B m 6j amp text if m equiv 0 pmod 6 frac m 3 3 sum limits j 1 frac m 2 6 binom m 3 m 6j B m 6j amp text if m equiv 2 pmod 6 frac m 3 6 sum limits j 1 frac m 4 6 binom m 3 m 6j B m 6j amp text if m equiv 4 pmod 6 end cases Von Staudt Clausen theorem Edit Main article Von Staudt Clausen theorem The von Staudt Clausen theorem was given by Karl Georg Christian von Staudt 40 and Thomas Clausen 41 independently in 1840 The theorem states that for every n gt 0 B 2 n p 1 2 n 1 p displaystyle B 2n sum p 1 mid 2n frac 1 p is an integer The sum extends over all primes p for which p 1 divides 2n A consequence of this is that the denominator of B2n is given by the product of all primes p for which p 1 divides 2n In particular these denominators are square free and divisible by 6 Why do the odd Bernoulli numbers vanish Edit The sum f k n i 0 n i k n k 2 displaystyle varphi k n sum i 0 n i k frac n k 2 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 even and 2k 1 m gt 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 gt 1 From the von Staudt Clausen theorem it is known that for odd n gt 1 the number 2Bn is an integer This seems trivial if one knows beforehand that the integer in question is zero However by applying Worpitzky s representation one gets 2 B n m 0 n 1 m 2 m 1 m n 1 m 1 0 n gt 1 is odd displaystyle 2B n sum m 0 n 1 m frac 2 m 1 m left n 1 atop m 1 right 0 quad n gt 1 text is odd 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 Sn m m nm The last equation can only hold if odd m 1 n 1 2 m 2 S n m even m 2 n 2 m 2 S n m n gt 2 is even displaystyle sum text odd m 1 n 1 frac 2 m 2 S n m sum text even m 2 n frac 2 m 2 S n m quad n gt 2 text is even 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 A restatement of the Riemann hypothesis Edit 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 numbers In fact Marcel Riesz proved that the RH is equivalent to the following assertion 42 For every e gt 1 4 there exists a constant Ce gt 0 depending on e such that R x lt Cexe as x Here R x is the Riesz function R x 2 k 1 k k x k 2 p 2 k B 2 k 2 k 2 k 1 k k x k 2 p 2 k b 2 k displaystyle R x 2 sum k 1 infty frac k overline k x k 2 pi 2k left frac B 2k 2k right 2 sum k 1 infty frac k overline k x k 2 pi 2k beta 2k nk denotes the rising factorial power in the notation of D E Knuth The numbers bn Bn n occur frequently in the study of the zeta function and are significant because bn is a p integer for primes p where p 1 does not divide n The bn are called divided Bernoulli numbers Generalized Bernoulli numbers EditThe generalized Bernoulli numbers are certain algebraic numbers defined similarly to the Bernoulli numbers that are related to special values of Dirichlet L functions in the same way that Bernoulli numbers are related to special values of the Riemann zeta function Let x be a Dirichlet character modulo f The generalized Bernoulli numbers attached to x are defined by a 1 f x a t e a t e f t 1 k 0 B k x t k k displaystyle sum a 1 f chi a frac te at e ft 1 sum k 0 infty B k chi frac t k k Apart from the exceptional B1 1 1 2 we have for any Dirichlet character x that Bk x 0 if x 1 1 k Generalizing the relation between Bernoulli numbers and values of the Riemann zeta function at non positive integers one has the for all integers k 1 L 1 k x B k x k displaystyle L 1 k chi frac B k chi k where L s x is the Dirichlet L function of x 43 Eisenstein Kronecker number Edit Main article Eisenstein Kronecker number Eisenstein Kronecker numbers are an analogue of the generalized Bernoulli numbers for imaginary quadratic fields 44 45 They are related to critical L values of Hecke characters 45 Appendix EditAssorted identities Edit Umbral calculus gives a compact form of Bernoulli s formula by using an abstract symbol B S m n 1 m 1 B n m 1 B m 1 displaystyle S m n frac 1 m 1 mathbf B n m 1 B m 1 where the symbol Bk that appears during binomial expansion of the parenthesized term is to be replaced by the Bernoulli number Bk and B1 1 2 More suggestively and mnemonically this may be written as a definite integral S m n 0 n B x m d x displaystyle S m n int 0 n mathbf B x m dx Many other Bernoulli identities can be written compactly with this symbol e g 1 2 B m 2 2 m B m displaystyle 1 2 mathbf B m 2 2 m B m Let n be non negative and even z n 1 n 2 1 B n 2 p n 2 n displaystyle zeta n frac 1 frac n 2 1 B n 2 pi n 2 n The n th cumulant of the uniform probability distribution on the interval 1 0 is Bn n Let n 1 n and n 1 Then Bn is the following n 1 n 1 determinant 46 B n n 1 0 0 1 2 1 0 0 n n 1 1 0 n 1 n 2 0 n 1 0 0 1 1 2 1 0 0 1 n 1 n 1 1 0 1 n 1 1 n 1 2 0 displaystyle begin aligned B n amp n begin vmatrix 1 amp 0 amp cdots amp 0 amp 1 2 amp 1 amp cdots amp 0 amp 0 vdots amp vdots amp amp vdots amp vdots n amp n 1 amp cdots amp 1 amp 0 n 1 amp n amp cdots amp 2 amp 0 end vmatrix 8pt amp n begin vmatrix 1 amp 0 amp cdots amp 0 amp 1 frac 1 2 amp 1 amp cdots amp 0 amp 0 vdots amp vdots amp amp vdots amp vdots frac 1 n amp frac 1 n 1 amp cdots amp 1 amp 0 frac 1 n 1 amp frac 1 n amp cdots amp frac 1 2 amp 0 end vmatrix end aligned Thus the determinant is sn 1 the Stirling polynomial at x 1 For even numbered Bernoulli numbers B2p is given by the p 1 p 1 determinant 46 B 2 p 2 p 2 2 p 2 1 0 0 0 1 1 3 1 0 0 0 1 5 1 3 1 0 0 1 2 p 1 1 2 p 1 1 2 p 3 1 3 0 displaystyle B 2p frac 2p 2 2p 2 begin vmatrix 1 amp 0 amp 0 amp cdots amp 0 amp 1 frac 1 3 amp 1 amp 0 amp cdots amp 0 amp 0 frac 1 5 amp frac 1 3 amp 1 amp cdots amp 0 amp 0 vdots amp vdots amp vdots amp amp vdots amp vdots frac 1 2p 1 amp frac 1 2p 1 amp frac 1 2p 3 amp cdots amp frac 1 3 amp 0 end vmatrix Let n 1 Then Leonhard Euler 1 n k 1 n n k B k B n k B n 1 B n displaystyle frac 1 n sum k 1 n binom n k B k B n k B n 1 B n Let n 1 Then 47 k 0 n n 1 k n k 1 B n k 0 displaystyle sum k 0 n binom n 1 k n k 1 B n k 0 Let n 0 Then Leopold Kronecker 1883 B n k 1 n 1 1 k k n 1 k j 1 k j n displaystyle B n sum k 1 n 1 frac 1 k k binom n 1 k sum j 1 k j n Let n 1 and m 1 Then 48 1 m r 0 m m r B n r 1 n s 0 n n s B m s displaystyle 1 m sum r 0 m binom m r B n r 1 n sum s 0 n binom n s B m s Let n 4 and H n k 1 n k 1 displaystyle H n sum k 1 n k 1 the harmonic number Then H Miki 1978 n 2 k 2 n 2 B n k n k B k k k 2 n 2 n k B n k n k B k H n B n displaystyle frac n 2 sum k 2 n 2 frac B n k n k frac B k k sum k 2 n 2 binom n k frac B n k n k B k H n B n Let n 4 Yuri Matiyasevich found 1997 n 2 k 2 n 2 B k B n k 2 l 2 n 2 n 2 l B l B n l n n 1 B n displaystyle n 2 sum k 2 n 2 B k B n k 2 sum l 2 n 2 binom n 2 l B l B n l n n 1 B n span, wikipedia, wiki, book, books, library,

article

, read, download, free, free download, mp3, video, mp4, 3gp, jpg, jpeg, gif, png, picture, music, song, movie, book, game, games.