Explicit formulas using partitions of integers for numbers defined by recursion

Similar documents
Generating Functions (Revised Edition)

3.3 Accumulation Sequences

Takao Komatsu School of Mathematics and Statistics, Wuhan University

Generating Functions

Chapter Generating Functions

MATH3283W LECTURE NOTES: WEEK 6 = 5 13, = 2 5, 1 13

are the q-versions of n, n! and . The falling factorial is (x) k = x(x 1)(x 2)... (x k + 1).

q-series Michael Gri for the partition function, he developed the basic idea of the q-exponential. From

Section 4.1: Sequences and Series

Integer-Valued Polynomials

SUMS OF POWERS AND BERNOULLI NUMBERS

A q-analogue OF THE GENERALIZED FACTORIAL NUMBERS

MA22S3 Summary Sheet: Ordinary Differential Equations

Generating Functions of Partitions

Notes on generating functions in automata theory

Basic counting techniques. Periklis A. Papakonstantinou Rutgers Business School

ON A KIND OF GENERALIZED ARITHMETIC-GEOMETRIC PROGRESSION

Introduction to Series and Sequences Math 121 Calculus II Spring 2015

Notes on Continued Fractions for Math 4400

Discrete Structures Lecture Sequences and Summations

The Humble Sum of Remainders Function

ON DIVISIBILITY OF SOME POWER SUMS. Tamás Lengyel Department of Mathematics, Occidental College, 1600 Campus Road, Los Angeles, USA.

Ex. Here's another one. We want to prove that the sum of the cubes of the first n natural numbers is. n = n 2 (n+1) 2 /4.

A Note about the Pochhammer Symbol

5. Sequences & Recursion

MACMAHON S PARTITION ANALYSIS IX: k-gon PARTITIONS

Linear chord diagrams with long chords

Common Core State Standards for Mathematics - High School PARRC Model Content Frameworks Mathematics Algebra 2

Chapter 11 - Sequences and Series

Miller Objectives Alignment Math

MATRIX ALGEBRA AND SYSTEMS OF EQUATIONS. + + x 1 x 2. x n 8 (4) 3 4 2

Generating Functions

arxiv: v1 [math.co] 3 Nov 2014

Some Congruences for the Partial Bell Polynomials

A LATTICE POINT ENUMERATION APPROACH TO PARTITION IDENTITIES

Moreover this binary operation satisfies the following properties

Pearson Mathematics Algebra 2 Common Core 2015

Calculus 2502A - Advanced Calculus I Fall : Local minima and maxima

1 Determinants. 1.1 Determinant

Intermediate Math Circles March 11, 2009 Sequences and Series

Sequences and Series. Copyright Cengage Learning. All rights reserved.

Algebra II/Math III Curriculum Map

Sums of Powers of Integers. Julien Doucet. and A. Saleh-Jahromi

MTH 299 In Class and Recitation Problems SUMMER 2016

Mathematics Standards for High School Algebra II

Partition of Integers into Distinct Summands with Upper Bounds. Partition of Integers into Even Summands. An Example

The Generating Functions for Pochhammer

MULTI-RESTRAINED STIRLING NUMBERS

A new approach to Bernoulli polynomials

Arithmetic properties of lacunary sums of binomial coefficients

Math 0230 Calculus 2 Lectures

Fall 2017 Test II review problems

Algebra 2 Mississippi College- and Career- Readiness Standards for Mathematics RCSD Unit 1 Data Relationships 1st Nine Weeks

DETERMINANTS. , x 2 = a 11b 2 a 21 b 1

Recurrence Relations

Determinants of Partition Matrices

= = = = = =

On the Convergence of the Summation Formulas Constructed by Using a Symbolic Operator Approach

ESSAI sur la maniére de trouver le terme général des séries recurrentes.

x arctan x = x x x x2 9 +

Introduction to Number Theory

Comparison of Virginia s College and Career Ready Mathematics Performance Expectations with the Common Core State Standards for Mathematics

EVALUATION OF A FAMILY OF BINOMIAL DETERMINANTS

NOTES ON OCT 4, 2012

Solution to Homework 1

Links Between Sums Over Paths in Bernoulli s Triangles and the Fibonacci Numbers

Symmetric functions and the Vandermonde matrix

Upper Bounds for Partitions into k-th Powers Elementary Methods

Named numbres. Ngày 25 tháng 11 năm () Named numbres Ngày 25 tháng 11 năm / 7

AP Calculus Chapter 9: Infinite Series

5.7 Cramer's Rule 1. Using Determinants to Solve Systems Assumes the system of two equations in two unknowns

Discrete Mathematics & Mathematical Reasoning Chapter 6: Counting

Series Solutions of Differential Equations

3 Finite continued fractions

EECS 1028 M: Discrete Mathematics for Engineers

Applications. More Counting Problems. Complexity of Algorithms

Foundations of Mathematics MATH 220 FALL 2017 Lecture Notes

MAT115A-21 COMPLETE LECTURE NOTES

AMSCO Algebra 2. Number and Quantity. The Real Number System

IB Mathematics HL Year 2 Unit 11: Completion of Algebra (Core Topic 1)

SEQUENCES, MATHEMATICAL INDUCTION, AND RECURSION

Advanced Counting Techniques

THE p-adic VALUATION OF LUCAS SEQUENCES

Last Update: March 1 2, 201 0

School District of Marshfield Course Syllabus

Tropical Polynomials

EIGENVALUES AND EIGENVECTORS 3

A matrix over a field F is a rectangular array of elements from F. The symbol

Algebra 2 Standards. Essential Standards:

Sums of Products of Bernoulli Numbers*

Final Exam Review. 2. Let A = {, { }}. What is the cardinality of A? Is

A Note on Extended Binomial Coefficients

ACI-matrices all of whose completions have the same rank

Tennessee s State Mathematics Standards - Algebra II

SECTION 9.2: ARITHMETIC SEQUENCES and PARTIAL SUMS

Some Results Concerning Uniqueness of Triangle Sequences

On Linear Recursive Sequences with Coefficients in Arithmetic-Geometric Progressions

CALCULUS JIA-MING (FRANK) LIOU

2Algebraic. foundations

and the compositional inverse when it exists is A.

Transcription:

Explicit formulas using partitions of integers for numbers defined by recursion Giuseppe Fera, Vittorino Talamini, arxiv:2.440v2 [math.nt] 27 Feb 203 DCFA, Sezione di Fisica e Matematica, Università di Udine, Via delle Scienze 206, 3300 Udine, Italy INFN, Sezione di Trieste, Via Valerio 2, 3427 Trieste, Italy giuseppe.fera@uniud.it, vittorino.talamini@uniud.it Abstract In this article we obtain an explicit formula in terms of the partitions of the positive integer n to express the n-th term of a wide class of sequences of numbers defined by recursion. Our proof is based only on arithmetics. We compare our result with similar formulas obtained with different approaches already in the XIX century. Examples are given for Bernoulli, Euler and Fibonacci numbers. Introduction and main result Consider two sequences of numbers a = {a 0,a,...} and b = {b 0,b,...} such that for all integers n 0 the following condition holds: a n h b h = δ 0,n, () h=0 in which δ 0,n is if n = 0 and 0 otherwise. We stress that for n = 0, Condition () implies a 0 b 0 =. It is convenient to consider a 0 = b 0 =. If this is not the case, one may consider the sequences obtained from a and b by factorizing out the factors a 0 and b 0, that is with coefficients a i /a 0 and b i /b 0, i 0. Therefore, from now on we will suppose a 0 = b 0 =. If, with the two given sequences, one builds the formal power series a(x) = a i x i, b(x) = b i x i, (2) i=0 i=0

and makes the Cauchy product of the two series (with formal power series one does not bother about convergence problems), then Condition () is equivalent to the equation a(x)b(x) =. Condition () is also equivalent to a recursive definition of the elements of one of the two sequences. One has, in fact: n b n = a n h b h, n > 0, b 0 =, (3) h=0 that allows to determine the n-th element of the sequence b from all the preceding ones. Many sequences of numbers are defined through Eqs. () or (3) (some examples are given in Table ), and in those cases to calculate the n-th element of the sequence one has to calculate all the elements of the sequence with index smaller than n. Formulas that allow to calculate the n-th element of the sequence only in terms of n, and not in terms of other elements of the sequence, are then useful and welcome, often more for theoretical than for practical interest. These formulas are often called closed formulas, but we prefer to call them explicit formulas. In this article we present, in Theorem, an explicit formula that allows to calculate the n-th element of a sequence of numbers satisfying Condition () in terms of the partitions of the integer n. This is not a new result, as we will see, but we will prove it in an elementary way that seems absent in the literature. Later on we will list and comment alternative ways to get the same results. Before to state our theorem we recall some basic definitions concerning partitions and compositions. Given a positive integer n, a composition c of n is an unordered set of l(c) positive integers n i that have sum n: l(c) c = {n,n 2,...,n l(c) }, n = n i. The number l(c) is called the length of the composition c. A partition p is a composition in which all the n i are ordered, usually in decreasing order. The length l(p) of a partition p is the number of elements i= 2

of p. If p is a partition obtained from the composition c, one obviously has l(p) = l(c). If p = {n,n 2,...,n l(p) }, one then has: n i n i+, i =,...,l(p). For any partition p there are many compositions formed with the same l(p) elements of p. Let µ(p) be the number of different compositions formed by the same numbers in the partition p. It is an easy exercise to verify that the number µ(p) is given by the formula: / µ(p) = l(p)! m p (n i )!, (4) n i (p) where m p (n i ) is the multiplicity of n i in p, that is the number of times the number n i appears in p, and the product at the denominator has to be done for all different numbers n i that appear in p, that is, for all numbers n i that appear in the union (p) of the elements of p. The multiplicities m p (n i ) that appear at the denominator of (4) always form a composition of the length l(p) that appears at the numerator. The second member of Eq. (4) can also be written using the multinomial coefficient: ( ) l(p) µ(p) =, m p (n ), m p (n 2 ),... where the numbers m p (n ), m p (n 2 ),... are those appearing in the denominator of the second member of Eq. (4). Let C(n) and P(n) be the set of all compositions of n and the set of all partitions of n, respectively. We are ready to state our main result. Theorem. Let {a 0,a,...} and {b 0,b,...} be sequences such that a 0 = b 0 =, and such that Condition () holds, then one may calculate the elements b n, n > 0, from the following explicit formulas involving the sequence {a 0,a,...} and the compositions or the partitions of n: b n = ( a ni ) (5) b n = c C(n) p P(n) n i c µ(p) ( a ni ) (6) 3

Formulas (5) and (6) in Theorem differ only for the use of compositions in place of partitions. In general, the use of partitions is preferable in calculations because it involves fewer terms. In Table 2 we list explicit formulas for the general terms of some sequences of numbers in terms of partitions only, but there exist similar formulas using compositions, too. It must be said that the explicit formulas that use partitions of integers to calculate the general n-th term of a sequence of numbers are often not convenient for practical computations, because the cardinality of P(n) grows very rapidly with n. Proof of Theorem (elementary version). We want to solve the linear system expressed by Condition () in the indeterminates b,b 2,... We recall that we suppose a 0 = b 0 =. Following Scherk [7], we do this by substitution, using recursive formula (3) to solve the equation for b, then that one for b 2, and so on. At each step, the equation number n contains only the indeterminates b,...,b n, so the substitution method is easy to apply. The first few equations are not difficult to solve: b = a b 2 = a 2 a b = a 2 +a 2 b 3 = a 3 a 2 b a b 2 = a 3 +2a 2 a a 3 b 4 = a 4 a 3 b a 2 b 2 a b 3 = a 4 +2a 3 a +a 2 2 3a 2a 2 +a4 b 5 = a 5 a 4 b a 3 b 2 a 2 b 3 a b 4 = = a 5 +2a 4 a +2a 3 a 2 3a 3 a 2 3a 2 2a +4a 2 a 3 a 5... (7) At this point one notes that, at least for h =,2,3,4,5,..., b h is obtained as the sum of all products like whose indices satisfy the conditions a n a n2... a nk, k h, n +n 2 +...+n k = h, and each term is taken with the + sign if k is even, with the sign if k is odd. The k positive numbers n,...,n k form then a composition of length k of the positive integer h. We are then led to write the formula: b n = ( a ni ), n, (8) c C(n) n i c where C(n) is the set of all compositions of the integer n. Let s verify by induction that Formula (8) is true for all values of n. From (7) we see 4

that Eq. (8) is true for n 5. Supposing (8) true up to a given n, let s prove that it is true also for n substituted by n+. Starting from (3), and using (8), one finds: = a n+ b n+ = a n+ c C(h) a n+ h b h = ( a ni ) a n+ h. n i c For any composition c = {n,n 2,...,n l(c) } C(h), consider the set obtained by appending to the l(c) numbers n i in c the number n 0 = (n+) h, that appears in the index of the last factor a n+ h (so that it can be rewritten in the form a n0 ). The set c = {n 0,n,n 2,...,n l(c) } is now a composition of (n + ) of length l(c ) = l(c) +. Considering then all possible values of h resulting from the sum of the last expression, we obtain in this way all possible compositions of(n+) of length greater than. Considering the term outside the summation symbol as corresponding to the unique composition {n + } of length of (n + ), we can rewrite the last expression in the following form: b n+ = ( a ni ). n i c c C(n+) This is Formula (8) written for the integer (n+) instead of n, so our proof by induction is now complete and (8) is true for all values of n. Usually it is more convenient to use partitions in place of compositions. Using the partitions, Formula (8) becomes b n = µ(p) ( a ni ), n, p P(n) where the factor µ(p) is given in Equation (4). 2 Different approaches As already remarked, one may find results similar to those expressed by Theorem also in other papers. We will give here a short historical review and with it we will stress how a simple proof of Theorem is not available. In fact, the result expressed by Theorem is usually seen in the context of the theory of formal power series or in relation with the Taylor series expansion 5

of functions. In 825, Scherk (pages 59 66 of [7]) showed how to solve with iterated substitutions a triangular linear system like the one corresponding to Eq. (3) (that we present in Eq. (7)), and found correctly that the n-th indeterminate b n is obtained by summing up 2 n terms involving products of the sequence elements a i, i =,...,n. The number 2 n is equal to the number of compositions of n, but Scherk did not recognize the sum over the compositions of n in his result. The triangular system considered by Scherk was written in matrix form by Brioschi in 858 [2], and solved using Cramer s rule, arriving at the following expression for the n-th term of the sequence: b n = ( ) n det a 0... 0 a 2 a... 0 a 3 a 2 a... 0....... a n a n a n 2... a Brioschi then obtained the following explicit formula: b n = q +2q 2 +3q 3 +...+nq n=n ( ) q q! q!q 2! q n! aq a q 2. 2 a qn n, (9) where q = q + q 2 +... + q n, and the sum is over all sets of non-negative integers {q,q 2,...,q n } that are solutions of the diophantine equation q +2q 2 +3q 3 +...+nq n = n. (0) In his work Brioschi also reported that the same formula (9) was obtained one year before by Fergola [5]. Indeed, Fergola found Eq. (9) in 857 [5] using an explicit formula he earlier found in [4] for the expansion of the n-th derivative of the inverse function f (x) in terms of the derivatives of the function f(x) up to the n-th order, where f(x) is any differentiable function. At the beginning of his paper [5], Fergola said that Sylvester remarked the importance of partitions in connection with his work and the end of the same paper he stated without proofthatthenumberofsolutions{q,q 2,...,q n }ofthediophantineequation (0), with q i 0, is equal to the number of partitions of n. Sylvester in 87 [8] stated without proof that the set of solutions of the diophantine equation (0) are in a one to one correspondence with the set of partitions of n. This fact is crucial to show that Eq. (9) and the result 6

expressed by our Theorem are equivalent, so we will give a simple proof of this fact in the following Proposition. Proposition. The sets of solutions {q,q 2,...,q n } of the diophantine equation (0), with q i 0, are in a one to one correspondence with the partitions p = {n,n 2,...,n k } of n. The correspondence is obtained by taking q i = m p (n i ), if n i p, and 0 otherwise. It then follows that l(p) = n i= q i. Proof. Let {q,q 2,...,q n } be a set of non negative integers that are solutions of the diophantine equation (0), then p = {q,2q 2,...,nq n }\{0} is a partition of n. Viceversa, given a partition p = {n,n 2,...} of n, with n i and multiplicity m p (n i ), the set {q,q 2,...,q n }, where q i = m p (n i ), if n i p, and 0 otherwise, is a set of non negative integers that are solutions of the diophantine equation (0). By the discussion following Eq. (4) one then has that l(p) = n i= q i. Proposition can now be applied to prove the following Proposition. Proposition 2. Eq. (9) is equivalent to Eq. (6). Proof. PropositionimpliesthatthesuminEq. (9)isoverallpartitionspof q! n, where p = {q,2q 2,...,nq n }\{0}. The factor q!q 2! q n! in Eq. (9) is then equaltothefactorµ(p)definedineq. (4),becauseq = q +q 2 +...+q n = l(p). Moreover, the product ( ) q a q a q 2 2 a qn n in Eq. (9) can be rewritten as ( a ) q ( a 2 ) q2 ( a n ) qn, and this is equal to the product ( a n i ) in Eq. (6). Nowadays, links between the theory of formal power series and partitions are encountered in many textbooks and papers (see, for example, [3]). We present, for example, a proof of Theorem in the context of the theory of formal power series. Proof of Theorem (formal power series version). Consider the formal power series (2) satisfying b(x) = /a(x). Being a 0 =, one has a(x) = + i= a ix i, so one can write: b(x) = a(x) = ( + ) a i x i = + i= ( ) k a i x i where in the last member the geometric series expansion has been used. Using the multinomial theorem ( ) (x +x 2 +...) k k = x r r,r 2,... xr 2 2, r +r 2 +...=k 7 k= i=

wherer,r 2,...arenon-negativeintegers, toexpandtheterms( i= a ix i ) k, and factoring out the same powers of x, one may write: = + b(x) = + = + n= x n k= k= r +r 2 +...=k r +r 2 +...=k r +2r 2 +...+nr n=n ( k r,r 2,... ) ( a x) r ( a 2 x 2 ) r2 = ( ) x r +2r 2 +... k ( a ) r ( a 2 ) r2 = r,r 2,... ( ) r +r 2 +...+r n ( a ) r ( a 2 ) r2 ( a n ) rn. r,r 2,...,r n Using Proposition and Eq. (4) one has: b(x) = + x n µ(p) ( a ni ), n= p P(n) so that, for Eq. (2), one has n > 0: b n = µ(p) ( a ni ), p P(n) This expression for b n is equal to Eq. (6) in Theorem. We listed such a different methods to obtain explicit formulas using partitions of integers for the elements of sequences satisfying Condition (), to point out that no complete elementary method, not involving calculus or formal power series, is available in the literature. We think moreover that our proof is simple because it only uses some basic arithmetics and combinatorics. 3 Examples We already observed that the equation a(x)b(x) = for two formal power series is equivalent to Condition (). This applies in the particular case of two functions f(x) and /f(x) for which a Taylor series expansion exists in x = x 0, and leads to following expressions for the sequence elements in terms of partitions of integers: a n = d n n! dx nf(x), x=x0 b n = n! d n dx n f(x) = x=x0 p P(n) 8 µ(p) ( n i! d n i f(x) dx n i x=x0 ) ()

Entry Name a n b n Bernoulli numbers (n+)! n! B n 2 Even Bernoulli numbers 2 (2n+2)! 2n (2n)! B 2n 3 Euler numbers 4 Even Euler numbers +( ) n 2n! (2n)! n! E n (2n)! E 2n 5 Fibonacci numbers ( ) n 2 F n 6 Even Fibonacci numbers n F 2n Table : Some sequences satisfying Condition (). Many sequences of numbers can be defined as the coefficients of the Taylor series expansions of convenient generating functions g(x). In all these cases, the sequence elements can be written in terms of partitions of integers using Eq. () for the reciprocal f(x) = /g(x) of the generating function. We report in Table a few sequences of numbers satisfying Condition (). In the Table one should consider n > 0, because for n = 0 not all the listed entries are equal to, as we have supposed. The corresponding expansions in terms of partitions of integers are reported in Table 2. These expansions are usually found using Eq. (), but in the following we will prove some of them using a purely arithmetic method. The aim of the cited papers [5, 2] was to give an explicit formula for the Bernoulli and Euler numbers. For this reason we list Entries 4 in Table. The proof of Condition () for Entries, 3 and 4 is classical and can easily be found in the literature (for example in [6], starting from Eq. (3), p. 238, and Eq. () p. 256). We only report below a proof of Condition () for Entry 2. Explicit formulas for the Fibonacci numbers in terms of partitions of integers are less known. Therefore, we report below a proof of Condition () for Entries 5 and 6. 9

Proof of Condition () for Entry 2. We need Condition () for Entry of Table, that is: B h = 0. (2) (n+ h)! h! h=0 It is well-known that B 2n+ = 0, n > 0. A purely arithmetic proof of this fact is outlined in [6] p. 238. It is based on a formula published by J. Faulhaber in 63 and proved by B. Pascal in 654, as reported in [], Theorem 3. (i). Using the fact that B 2n+ = 0, n > 0, it is convenient in Eq. (2) to separate the sums over the odd and over the even indices. We can write Eq. (2) for n = 2m and for n = 2m+. When n = 2m, Eq. (2) gives: 0 = 2m h=0 B h (2m+ h)! h! = m k=0 (2m+ 2k)! B 2k (2k)! + B (2m)!, that is: B (2m)! = m k=0 When n = 2m+, Eq. (2) gives: 2m+2 2k (2m+2 2k)! B 2k (2k)! (3) 0 = 2m+ h=0 B h (2m+2 h)! h! = m k=0 (2m+2 2k)! B 2k (2k)! + B (2m+)!, that is: B (2m)! = m k=0 2m+ (2m+2 2k)! The difference between Eqs. (4) and (3) gives: 0 = m ( k=0 B 2k (2k)! 2m+ (2m+2 2k)! 2m+2 2k (2m+2 2k)! ) B2k (2k)! = (4) = m k=0 2k (2m+2 2k)! B 2k (2k)! A standard proof uses the generating function for the Bernoulli numbers g B (x) = x e x = x h B h h! h=0 and the observation that g B (x)+x/2 is an even function. 0

and this is equivalent to Condition () for Entry 2 of Table. (A factor 2 has been added in Entry 2 of Table just to have a 0 = ). Proof of Condition () for Entries 5 and 6. We first recall the standard definition of the Fibonacci numbers: F = F 2 =, F n = F n +F n 2, n > 2. We then prove by induction that the following formulas are true n : F 2n = F 2h, (5) F 2n+ = + F 2h, (6) n F 2n = n+ F 2h (n h). (7) Eqs. (5), (6) and (7) are true for n =. Let s suppose they are true for a certain n 2 and let s prove they are true also for n substituted by n+. One has: F 2(n+) = F 2n+2 = F 2n+ +F 2n = F 2n+ + so Eq. (5) is true for all n. One has: F 2(n+)+ = F 2n+3 = F 2n+ +F 2n+2 = + n+ F 2h = F 2h, n+ F 2h +F 2n+2 = + F 2h, so Eq. (6) is true for all n. Using this result one has: F 2(n+) = F 2n+2 = F 2n+ +F 2n = + = n++ F 2h + n F 2h +n+ F 2h (n h) = F 2h (n h) = n++ F 2h (n+ h), and this proves that Eq. (7) is true for all n. Let s now form the sequence b, with elements b n = F 2n, n > 0, and b 0 =, and the sequence a, with a n = n, n > 0, and a 0 =. Then Eq. (7) has the form of Eq. (3) and is clearly equivalent to Condition () reported in

Entry b n B n! n = p P(n) µ(p) ( ) (n i +)! 2 2n B (2n)! 2n = p P(n) µ(p) ( ) 2 (2n i +2)! 3 E n! n = p P(n) µ(p) ( ) +( )n i 2n i! 4 E (2n)! 2n = p P(n) µ(p) ( ) (2n i )! 5 F n = p P(n) µ(p) ( ( )n i 2 ) 6 F 2n = p P(n) µ(p) n i Table 2: Explicit formulas using partitions for the sequences in Table. Entry 6 of Table. Eqs. (5) and (6) can be summarized by the following: F n = ( )n 2 n + F h ( ) n h 2. (8) Let s now form the sequence b, with elements b n = F n, n > 0, and b 0 =, and the sequence a, with a n = ( )n, n > 0, and a 2 0 =. Then Eq. (8) has the form of Eq. (3) and is clearly equivalent to Condition () reported in Entry 5 of Table. Theorem can now be applied to all the sequences listed in Table above and allows to write the explicit formulas using partitions reported in Table 2. Finally, we present two explicit calculations using P(4). One has: P(4) = {{4}, {3,}, {2,2}, {2,,}, {,,,}}, so from the first and sixth rows of Table 2 one finds: [! B 4 = 4!! + 2! 5!!! ( )2 + 2! 4!2! 2! ( )2 + 3! 3!3!!2! ( )3 3!2!2! + 4! ] 4! ( ) 4 = 2!2!2!2! 30, 2

F 8 = [ ]! 2! 2! 3! 4! 4+ 3 + 2 2+ 2 +!!! 2!!2! 4! = 2. References [] Alan F. Beardon, Sums of Powers of Integers, Amer. Math. Monthly, 03 (996) 20 23. [2] Francesco Brioschi, Sulle funzioni Bernoulliane ed Euleriane, Annali di Matematica Pura ed Applicata (858), 260 263 (in Italian). [3] Louis Comtet, Advanced Combinatorics, Reidel, Dordrecht, 974. [4] Emanuele Fergola, Dell espressione di una derivata qualunque di una funzione in termini delle derivate della funzione inversa, Memorie della Reale Accademia delle Scienze di Napoli, Vol. (856), 200 206 (in Italian). [5] Emanuele Fergola, Sopra lo sviluppo della funzione /(ce x ), e sopra una nuova espressione dei numeri di Bernoulli, Memorie della Reale Accademia delle Scienze di Napoli, Vol. 2 (857), 35 324 (in Italian). [6] Édouard Lucas, Théorie des nombres, Gauthier-Villars, Paris, 89. [7] Heinrich F. Scherk, Mathematische Abhandlungen, G. Reimer, Berlin (825) (in German). (Available at http://hdl.handle.net/2027/hvd.hnxxhh). [8] James Joseph Sylvester, A note on the theory of a point in partitions, Edinburgh British Associaton Report (87), 23 25. 3