Some Fun with Divergent Series

Similar documents
MOMENTS OF HYPERGEOMETRIC HURWITZ ZETA FUNCTIONS

Bernoulli Polynomials

Synopsis of Complex Analysis. Ryan D. Reece

ζ (s) = s 1 s {u} [u] ζ (s) = s 0 u 1+sdu, {u} Note how the integral runs from 0 and not 1.

Conformal maps. Lent 2019 COMPLEX METHODS G. Taylor. A star means optional and not necessarily harder.

NOTES ON RIEMANN S ZETA FUNCTION. Γ(z) = t z 1 e t dt

Topic 7 Notes Jeremy Orloff

Solutions to practice problems for the final

Taylor and Laurent Series

Part IB. Further Analysis. Year

FRACTIONAL HYPERGEOMETRIC ZETA FUNCTIONS

Complex Variables. Instructions Solve any eight of the following ten problems. Explain your reasoning in complete sentences to maximize credit.

Complex Analysis MATH 6300 Fall 2013 Homework 4

1 Discussion on multi-valued functions

Convergence of Some Divergent Series!

Part IB. Complex Analysis. Year

13 Definite integrals

Theorem [Mean Value Theorem for Harmonic Functions] Let u be harmonic on D(z 0, R). Then for any r (0, R), u(z 0 ) = 1 z z 0 r

Complex Analysis Homework 9: Solutions

Considering our result for the sum and product of analytic functions, this means that for (a 0, a 1,..., a N ) C N+1, the polynomial.

2 Write down the range of values of α (real) or β (complex) for which the following integrals converge. (i) e z2 dz where {γ : z = se iα, < s < }

Here are brief notes about topics covered in class on complex numbers, focusing on what is not covered in the textbook.

VII.8. The Riemann Zeta Function.

Math Final Exam.

INFINITE SEQUENCES AND SERIES

Math Homework 2

Part IB Complex Analysis

AP Calculus Testbank (Chapter 9) (Mr. Surowski)

The Calculus of Residues

Riemann Zeta Function and Prime Number Distribution

MA3111S COMPLEX ANALYSIS I

Math 185 Fall 2015, Sample Final Exam Solutions

Problems for MATH-6300 Complex Analysis

MA 412 Complex Analysis Final Exam

Taylor and Maclaurin Series

MTH4101 CALCULUS II REVISION NOTES. 1. COMPLEX NUMBERS (Thomas Appendix 7 + lecture notes) ax 2 + bx + c = 0. x = b ± b 2 4ac 2a. i = 1.

1 Res z k+1 (z c), 0 =

Power series and Taylor series

LAURENT SERIES AND SINGULARITIES

Mathematics 324 Riemann Zeta Function August 5, 2005

Series Solutions. 8.1 Taylor Polynomials

Solutions to Complex Analysis Prelims Ben Strasser

Complex Variables & Integral Transforms

Let s Get Series(ous)

INTRODUCTION TO COMPLEX ANALYSIS W W L CHEN

Complex Series. Chapter 20

16.4. Power Series. Introduction. Prerequisites. Learning Outcomes

MATH COMPLEX ANALYSIS. Contents

Last Update: March 1 2, 201 0

Section Taylor and Maclaurin Series

(a) To show f(z) is analytic, explicitly evaluate partials,, etc. and show. = 0. To find v, integrate u = v to get v = dy u =

MORE CONSEQUENCES OF CAUCHY S THEOREM

d 1 µ 2 Θ = 0. (4.1) consider first the case of m = 0 where there is no azimuthal dependence on the angle φ.

Infinite series, improper integrals, and Taylor series

Section 9.7 and 9.10: Taylor Polynomials and Approximations/Taylor and Maclaurin Series

1 Euler s idea: revisiting the infinitude of primes

MATH 311: COMPLEX ANALYSIS CONTOUR INTEGRALS LECTURE

THE RESIDUE THEOREM. f(z) dz = 2πi res z=z0 f(z). C

Review of Power Series

Second Midterm Exam Name: Practice Problems March 10, 2015

SOLVED PROBLEMS ON TAYLOR AND MACLAURIN SERIES

13 Maximum Modulus Principle

Qualifying Exam Complex Analysis (Math 530) January 2019

Taylor and Maclaurin Series. Copyright Cengage Learning. All rights reserved.

Infinite Series. 1 Introduction. 2 General discussion on convergence

Summation Techniques, Padé Approximants, and Continued Fractions

Complex Homework Summer 2014

Chapter 11 - Sequences and Series

Fourier Sin and Cos Series and Least Squares Convergence

Complex Analysis Important Concepts

Part 3.3 Differentiation Taylor Polynomials

This ODE arises in many physical systems that we shall investigate. + ( + 1)u = 0. (λ + s)x λ + s + ( + 1) a λ. (s + 1)(s + 2) a 0

TAYLOR AND MACLAURIN SERIES

Chapter 6: Residue Theory. Introduction. The Residue Theorem. 6.1 The Residue Theorem. 6.2 Trigonometric Integrals Over (0, 2π) Li, Yongzhao

Topic 4 Notes Jeremy Orloff

Math 312 Fall 2013 Final Exam Solutions (2 + i)(i + 1) = (i 1)(i + 1) = 2i i2 + i. i 2 1

Complex Variables Notes for Math 703. Updated Fall Anton R. Schep

MASTERS EXAMINATION IN MATHEMATICS

Complex Variables...Review Problems (Residue Calculus Comments)...Fall Initial Draft

Math 259: Introduction to Analytic Number Theory Functions of finite order: product formula and logarithmic derivative

Math 259: Introduction to Analytic Number Theory Functions of finite order: product formula and logarithmic derivative

Complex Analysis Qualifying Exam Solutions

Complex variables lecture 6: Taylor and Laurent series

Answer Key 1973 BC 1969 BC 24. A 14. A 24. C 25. A 26. C 27. C 28. D 29. C 30. D 31. C 13. C 12. D 12. E 3. A 32. B 27. E 34. C 14. D 25. B 26.

1. Introduction Interest in this project began with curiosity about the Laplace transform of the Digamma function, e as ψ(s + 1)ds,

MTH3101 Spring 2017 HW Assignment 4: Sec. 26: #6,7; Sec. 33: #5,7; Sec. 38: #8; Sec. 40: #2 The due date for this assignment is 2/23/17.

MATH 452. SAMPLE 3 SOLUTIONS May 3, (10 pts) Let f(x + iy) = u(x, y) + iv(x, y) be an analytic function. Show that u(x, y) is harmonic.

Chapter 11. Cauchy s Integral Formula

Polynomial expansions in the Borel region

Suggested Homework Solutions

2. Complex Analytic Functions

MTH 3102 Complex Variables Final Exam May 1, :30pm-5:30pm, Skurla Hall, Room 106

JUST THE MATHS UNIT NUMBER DIFFERENTIATION APPLICATIONS 5 (Maclaurin s and Taylor s series) A.J.Hobson

First, let me recall the formula I want to prove. Again, ψ is the function. ψ(x) = n<x

THE GAMMA FUNCTION AND THE ZETA FUNCTION

EE2 Mathematics : Complex Variables

. Then g is holomorphic and bounded in U. So z 0 is a removable singularity of g. Since f(z) = w 0 + 1

C. Complex Numbers. 1. Complex arithmetic.

MATH5685 Assignment 3

Math 411, Complex Analysis Definitions, Formulas and Theorems Winter y = sinα

Transcription:

Some Fun with Divergent Series 1. Preliminary Results We begin by examining the (divergent) infinite series S 1 = 1 + 2 + 3 + 4 + 5 + 6 + = k=1 k S 2 = 1 2 + 2 2 + 3 2 + 4 2 + 5 2 + 6 2 + = k=1 k 2 (i) Define T = 1 1 + 1 1 + 1 1 + Rewrite as T = 1 1 + 1 1 + 1 1 + Add 2T = 1 + + + + + + = 1 T = ½ (ii) Define s = 1 3 + 5 7 + 9 11 + Rewrite as s = 1 3 + 5 7 + 9 11 + Add 2s = 1 2 + 2 2 + 2 2 + = 1 2T = 1 1 = s = (iii) Define T 1 = 1 2 + 3 4 + 5 6 + Rewrite as T 1 = 1 2 + 3 4 + 5 6 + Add 2T 1 = 1 1 + 1 1 + 1 1 + = T = ½ T 1 = ¼ (iv) Define T 2 = 1 2 2 2 + 3 2 4 2 + 5 2 6 2 + Rewrite as T 2 = 1 2 2 2 + 3 2 4 2 + 5 2 Add 2T 2 = 1 2 (2 2 1 2 ) + (3 2 2 2 ) (4 2 3 2 ) + (5 2 4 2 ) (6 2 5 2 ) + = 1 (2 1)(2 + 1) + (3 2)(3 + 2) (4 3)(4 + 3) + (5 4)(5 + 4) = 1 3 + 5 7 + 9 11 + = s = T 2 = Note that these results can be formalized by using summability methods, such as those attributed to Abel, Cesàro and others, which define the sum of an infinite series more generally than the usual method of determining the limit to which the partial sums of the series converge. They are designed so that values can be attached to certain series that would otherwise be considered divergent. Abel summability is described later.

2. The Series S 1 and S 2 S 1 T 1 = 1 + 2 + 3 + 4 + 5 + 6 + 1 + 2 3 + 4 5 + 6 = + 4 + + 8 + + 12 + = 4(1 + 2 + 3 + 4 + ) = 4S 1 Hence 3S 1 = T 1 = ¼ S 1 = 1 12 Now consider S 2 T 2 = 1 2 + 2 2 + 3 2 + 4 2 + 5 2 + 6 2 + 1 2 + 2 2 3 2 + 4 2 5 2 + 6 2 = + 2 (2 1) 2 + + 2 (2 2) 2 + + 2 (2 3) 2 + = 2 3 (1 2 + 2 2 + 3 2 + 4 2 + ) = 8S 2 Hence 7S 2 = T 2 = S 2 = 3. Summary We have proved the strange and apparently nonsensical results 1 + 2 + 3 + 4 + 5 + 6 + = 1 12 1 2 + 2 2 + 3 2 + 4 2 + 5 2 + 6 2 + = In fact these results do make sense when interpreted as special values of the Riemann zeta function. The left hand sides of the equations above are no longer sums in the normal sense but rather a (misleading) notation for the zeta function evaluated at 1 and 2 respectively. We return to this later. 4. Abel Summation An infinite series Σ k= u k is convergent to the sum S when lim S n = S where S n = Σ k= u k is n the partial sum to n terms. The series is Abel summable to S when the series Σ k= u k x k is convergent to S(x) for x < 1, and lim S(x) = S. For example, x 1 1 1 + x = 1 x + x2 x 3 + + ( 1) n x n + ( x < 1) and 1/(1 + x) ½ as x 1. Formally the right-hand side becomes the divergent series s 1 when x 1 so that, by definition, the Abel sum of T is ½ which is the result in (i). n

Note, however, that 1 + x 1 + x + x 2 = 1 x2 + x 3 x 5 + x 6 x 8 + x 9 x 11 + ( x < 1) which suggests that the Abel sum of T is lim x 1 (1 + x)/(1 + x + x2 ) = ⅔, contradicting the previous result. But written out in full, the power series on the right-hand side is 1 + x 1 + x + x 2 = 1 + x x2 + x 3 + x 4 x 5 + x 6 + x 7 x 8 + x 9 + x 1 x 11 + which shows that in fact, ⅔ = 1 + 1 + 1 + 1 + 1 + 1 + 1 + + as x 1, a different series from T. Likewise 1 + x + x 2 1 + x + x 2 + x 3 = 1 x3 + x 4 x 7 + x 8 x 11 + x 12 x 15 + ( x < 1) reduces to ¾ = 1 + + 1 + 1 + + 1 + 1 + + 1 + 1 + + 1 + in the formal limit as x 1. Thus inserting zeros in a divergent series will change its Abel sum and give different values depending on how the zeros are arranged. For other examples of Abel summation, let x 1 in the following power series 1 (1 + x) 2 = 1 2x + 3x2 4x 3 + + ( 1) n (n + 1)x n + 1 x (1 + x) 2 = 1 3x + 5x2 7x 3 + + ( 1) n (2n + 1)x n + 1 x (1 + x) 3 = 1 22 x + 3 2 x 2 4 2 x 3 + + ( 1) n (n + 1) 2 x n + convergent for x < 1. The left-hand sides approach ¼, and respectively and formally the corresponding series become T 1, s and T 2, thereby confirming the results in (iii), (ii) and (iv). The Abel sum of a convergent series is the same as its ordinary sum. For example, the Abel sum of the convergent series is defined as the limit as x 1 of 1 + 1 1! + 1 2! + 1 3! + 1 4! + + 1 n! + = e 1 + x 1! + x2 2! + x3 3! + x4 4! + + xn n! + = ex which is the value e to which the original series converges in the normal sense. We can think of Abel summation as stronger method of attaching a value to a series than the usual method of taking the limit of the partial sums.

Note that the two series we started with do not have Abel sums. More sophisticated methods involving the Riemann zeta function are required to justify them. 5. The Riemann Zeta Function For real x > 1, the Riemann zeta function is defined by the convergent infinite series ζ(x) = 1 + 1 2 x + 1 3 x + 1 4 x + 1 5 x + = 1 n x (1) It has a singularity at x = 1 where it reduces to the harmonic series 1 + 1 2 + 1 3 + 1 4 + 1 5 + which is well-known to be divergent. If we now replace the argument of the zeta function by the complex variable z = x + iy, then the series remains convergent provided that x > 1, i.e. ζ(z) is defined in the half-plane Re z > 1. Moreover, by defining ζ(z) in a different way which reduces to (1) for Re z > 1 we can extend the domain in which it is defined to cover the entire complex plane apart from the singularity at z = 1, a process known as analytic continuation. Consider tz 1 Substitute nt = r e t 1 dt = tz 1 e t t z 1 1 e t dt = e t t z 1 e nt dt = t z 1 e nt dt e t 1 dt = (r/n) z 1 e r n 1 dr = r z 1 e r dr = ζ(z)γ(z) where Γ(z) = r z 1 e r dr is the well-known gamma function which is analytic except at the origin and when z is a negative integer, and which reduces to (z 1)! when z is a positive integer. The advantage of this expression is that it suggests how the analytic continuation of ζ(z) should proceed. We consider the contour integral in the complex w-plane, w = u + iv, f(z) = C z 1 w e w 1 dw = C n= e(z 1) log w e w 1 dw (2) where the contour C runs from along the upper side of the branch cut u <, v = of the logarithmic function, encloses the origin in a circle of vanishingly small radius, and then returns to along the lower side of the branch cut. Note that f(z) is analytic for all finite values of z because the contour integral is uniformly convergent in any finite region of the complex z-plane.

On the upper side of the branch cut we have log w = log u + iπ, and on the lower side log w = log u iπ. It is easily shown that if x = Re z > 1 the contribution to the integral around the origin vanishes as the radius of the circle tends to zero. Thus f(z) = e(z 1) (log u +iπ) e u 1 du + (log u iπ) e(z 1) e u 1 du = 2i sin zπ ( u)z 1 e u 1 du = 2i sin zπ tz 1 e t 1 du = 2i sin zπ Γ(z)ζ(z) (Re z > 1) The last step follows from (2) but the restriction on z can be relaxed in this formula because we already know that f(z) is analytic in the entire finite region of the complex z plane. Thus with the aid of a well-known identity, Γ(z)Γ(1 z) = π csc zπ, we obtain ζ(z) = iγ(1 z) 2π z 1 w e w dw (3) 1 We now determine in which regions of the complex z-plane, in addition to Re z > 1, this formula defines ζ(z) as an analytic function. We know already that the integral is analytic in the entire finite plane so the only possible singularities of ζ(z) are those of Γ(1 z), namely simple poles at z = 1, 2, 3, all of which except z = 1 can be discarded because we already know that ζ(z) is analytic for Re z > 1. Thus (3) represents an analytic continuation of ζ(z) to all finite points in the complex z plane except z = 1 where it has a simple pole. When z is an integer, the integrals along C from to the origin and back cancel each out because there is no need to construct a branch cut to accommodate a logarithmic function in this case (w m 1, m an integer, is analytic everywhere). If one did continue to write e (m 1) log w for w m 1, the resulting factor sin(m 1)π in the combined integrals would in any case ensure the vanishing of the total integral along that part of C. Thus it is only the vanishingly small circle around the origin that will contribute to the contour integral when z = m, an integer. We have already seen that when Re z > 1 (i.e. m = 2, 3, 4 ) this contribution is zero, which is as it should be in order to nullify the singularities in Γ(1 m) at those points and thereby ensuring the analyticity of ζ(z) for Re z > 1. We are left to consider the values m =, 1, 2, 3 (we showed above that m = 1 is a singularity a simple pole). Let the circular part of C around the origin have radius ε, let the integrand in (3) be expanded in a (Laurent) series of powers of w, and consider the integration around the origin of one such term a p w p where p is an integer (positive, negative or zero) and a p is a coefficient. With w = εe iθ we obtain a p w p dw = C π ia p ε p+1 π e i(p+1)θ sin(p + 1)π p+1 dθ = 2ia p ε p + 1 C (p 1) { 2πia p (p = 1) as ε. Thus only terms of type w 1 contribute to the contour integral, a property that will be familiar from the calculus of residues.

With z = m, the integrand in (3) can be expanded in terms of the Bernoulli numbers B n defined by the generating function t/(e t 1) = Σ n= (B n t n /n!). The first values are B = 1, B 1 = 1, 2 B 2 = 1, B 6 3 =, B 4 = 1. Thus for < w < 2π (to avoid the poles at w = ±2πi), we have 3 w m 1 e w 1 = ( w) wm 2 e w 1 = ( w) n wm 2 B n n! n= = w m 2 (1 + w 2 + w2 12 w4 72 + ) For m = the contributing term is the second one with coefficient ½. Thus the integral in (3) is evaluated as πi whence ζ() = ½Γ(1) = ½. For m = 1 the relevant term is the third with coefficient 1/12, the integral has the value πi/6, and ζ( 1) = Γ(2)/12 = 1/12. There is no term of the form w 1 when m = 2 so its coefficient is and the integral is also zero. Thus ζ( 2) =. Likewise, for m = 3 we deduce from the fourth term in the series that ζ( 3) = Γ(4)/72 = 3!/72 = 1/12. If we formally substitute these results in (1), the original definition of ζ(x) for real x > 1, we obtain ζ() = 1 + 1 + 1 + 1 + 1 + 1 + = 1 2 ζ( 1) = 1 + 2 + 3 + 4 + 5 + 6 + = 1 12 ζ( 2) = 1 2 + 2 2 + 3 2 + 4 2 + 5 2 + 6 2 + = ζ( 3) = 1 3 + 2 3 + 3 3 + 4 3 + 5 3 + 6 3 + = 1 12 The second and third of these series will be recognized as the sums S 1 and S 2 with which we started this discussion. Amazingly, however, the apparently absurd results we found by boldly adventurous but totally invalid manipulations of divergent series have been confirmed rigorously as the relevant values of the Riemann zeta function. They cannot be interpreted as sums of series in the usual sense, of course. The series must be regarded as alternative notations for values of the zeta function. The first and fourth results above suggest we should be able to prove by similar questionable methods that the respective series, which we can denote by S and S 3, have sums of ½ and 1/12. The first result is easy to show by following the procedure we first used to derive S 1 and S 2. Thus S T = 1 + 1 + 1 + 1 + 1 + 1 + 1 + 1 1 + 1 1 + 1 = + 2 + + 2 + + 2 + = 2(1 + 1 + 1 + 1 + 1 + 1 + ) = 2S Hence S = T = ½

To find the sum S 3 we use exactly the same recipe in a more general notation. Let S 3 = Σ k=1 k 3 and T 3 = Σ k=1 ( 1) k+1 k 3 or alternatively T 3 = Σ k=2 ( 1) k (k 1) 3. Adding the last two expressions, we have 2T 3 = 1 + Σ k=2 ( 1) k+1 [k 3 (k 1) 3 ] = 1 + Σ k=2 ( 1) k+1 [1 + 3k(k 1)] i.e. T 3 = 1/8. Now = Σ k=1 ( 1) k+1 [1 3k(k + 1)] = T 3T 2 3T 1 = 1 3 = 2 4 1 4 S 3 T 3 = Σ k=1 so that 15S 3 = T 3 = 1/8. Thus S 3 = 1/12. [k 3 ( 1) k+1 k 3 ] = Σ k=1 2(2k) 3 = 16Σ k=1 k 3 = 16S 3 6. The General Case The previous paragraph paves the way for generalizing the method in Sections 1 and 2 to a summation of the positive integers raised to the nth power the ultimate goal of this article. Let T n = Σ k=1 ( 1) k+1 k n, (n =, 1, 2, ). An alternative but equivalent way of writing this expression is T n = Σ k=2 ( 1) k (k 1) n. Adding the two together we have yielding T = ½, and for n 1, 2T n = 1 + Σ k=2 ( 1) k+1 [k n (k 1) n ] 2T n = 1 Σ k=2 ( 1) k+1 Σ n 1 ( 1) n+m ( n m )km = 1 Σ k=1 ( 1) k+1 Σ n 1 ( 1) n+m ( n m )km + Σ n 1 ( 1) n+m ( n m ) = Σ k=1 ( 1) k+1 Σ n 1 ( 1) n+m ( n m )km = Σ n 1 ( 1) n+m+1 ( n m )Σ k=1 ( 1) k+1 k m = Σ n 1 ( 1) n+m+1 ( n m )T m Together with the initial value of T = ½ from above, this is the recursion relation for T n. It can be expressed more tidily by defining V n = ( 1) n T n so that V = ½, 2V n = Σ n 1 ( n m ) (n 1) (4)

The binomial expansion (k 1) n = ( 1) n (1 k) n n = Σ ( 1) n+m ( n m )km has been used in the derivation above, with ( n ) = n!/[(n m)! m!] in the usual notation. We also noted the fact m n that when k = 1 this expansion gives = Σ ( 1) n+m ( n m ) = 1 + Σ n 1 ( 1) n+m ( n m ). The first few values of T n generated by the recurrence relation are T 1 = 1 4 T 2 = 1 4 + 1 2 (2 1 )T 1 = 1 4 + 1 4 = T 3 = 1 4 1 2 (3 1 )T 1 + 1 2 (3 2 )T 2 = 1 4 3 8 + = 1 8 T 4 = 1 4 + 1 2 (4 1 )T 1 1 2 (4 2 )T 2 + 1 2 (4 3 )T 3 = 1 4 + 1 2 1 4 = T 5 = 1 4 1 2 (5 1 )T 1 + 1 2 (5 3 )T 3 + = 1 4 5 8 + 5 8 = 1 4 T 6 = 1 4 + 1 2 (6 1 )T 1 + 1 2 (6 3 )T 3 + 1 2 (6 5 )T 5 = 1 4 + 3 4 5 4 + 3 4 = T 7 = 1 1 4 2 (7)T 1 1 + 1 2 (7)T 3 3 + 1 2 (7)T 5 5 + = 1 7 + 35 21 = 17 4 8 16 8 16 T 8 = 1 4 + 1 2 (8 1 )T 1 + 1 2 (8 3 )T 3 + 1 2 (8 5 )T 5 + 1 2 (8 7 )T 7 = 1 4 + 1 7 2 + 7 17 4 = Now let S n = Σ k=1 k n = 1 n + 2 n + 3 n + 4 n + 5 n + 6 n +. Then S n T n = Σ k=1 [1 ( 1) k+1 ]k n = Σ k=2,4,6, 2k n = Σ k=1 2(2k) n = 2 n+1 Σ k=1 k n = 2 n+1 S n Thus S n = T n /(2 n+1 1) = ( 1) n+1 V n /(2 n+1 1) (5) and from the previous results for T n 1 + 1 + 1 + 1 + 1 + 1 + = S = T = 1 2 1 + 2 + 3 + 4 + 5 + 6 + = S 1 = 1 3 T 1 = 1 12 1 2 + 2 2 + 3 2 + 4 2 + 5 2 + 6 2 + = S 2 = 1 7 T 2 = 1 3 + 2 3 + 3 3 + 4 3 + 5 3 + 6 3 + = S 3 = 1 15 T 3 = 1 12 1 4 + 2 4 + 3 4 + 4 4 + 5 4 + 6 4 + = S 4 = 1 5 + 2 5 + 3 5 + 4 5 + 5 5 + 6 5 + = S 5 = 1 63 T 5 = 1 252 1 6 + 2 6 + 3 6 + 4 6 + 5 6 + 6 6 + = S 6 = 1 7 + 2 7 + 3 7 + 4 7 + 5 7 + 6 7 + = S 7 = 1 255 T 7 = 1 24 1 8 + 2 8 + 3 8 + 4 8 + 5 8 + 6 8 + = S 8 =

It seems highly probable from these calculations that T n vanishes when n is even (likewise with V n and S n by definition). To prove this consider the generating function g(t) = 1 e t + 1 = a m where the coefficients a m are to be determined. Defining b j = 1 (j ), b = 2, so that we can write e t + 1 = Σ j= b j t j /j!, we deduce from the equation above 1 = j= b j a m t j+m j! m! = n= n t m m! b n ma m t n = tn n m! (n m)! n! b n ma m ( n m ) The second form of the double summation above is found by noting that all the points (j, m) in the first quadrant (covered by the first double summation) can be counted alternatively along each sloping line defined by j + m = n and then summing over all such lines from n = to. As m runs through its values from to n along each line defined by a particular value of n, the value of the index j is defined by j = n m which has been substituted in the second double summation. Since b n m = 2 when m = n and 1 otherwise, we may rewrite the last equation as n= 1 = 2a + tn n 1 n! [2a n + a m ( n m ) ] Equating coefficients of the powers of t we obtain n 1 a = ½, 2a n + a m ( n m ) = (n 1) This is the same recurrence relation as (4) showing that a n = V n and hence that g(t) is the generating function for V n = ( 1) n T n, i.e. g(t) = t m m! Define f(t) = g(t) ½ = Σ m=1 t m /m!. Then = 1 2 + m=1 t m m! (6) f(t) = 1 e t + 1 1 2 = 1 et 2(1 + e t ) = e t 1 2(e t + 1) = f( t) Thus f(t) is an odd function so that all coefficients of even powers of t in the expansion of f(t) above must vanish, i.e. V 2n = (n 1). It follows that T 2n and S 2n also vanish for n 1. Moreover, since only the odd-indexed terms V n are non-vanishing for n 1, we see from (5) that V = ½, V n = (2 n+1 1)S n (n 1) (7) since ( 1) n+1 = 1 when n is odd. It follows from the definition of g that

tg(t) = t 2 + m=1 t m+1 (2m+1 1)S m = t m! 2 + S (2t) n t n n 1 (n 1)! We now define B n = ns n 1 (n 2), B 1 = ½ with B undefined for the time being. Expressed in terms of B n the equation above becomes tg(t) = n=2 B n [ tn n! (2t)n t n ] = B n! n n! n= B n n= (2t) n The change in the lower limit of the first summation is permissible because of the way B 1 has been defined and in the second summation because B cancels out in the subtraction of the two series. If h(t) = Σ n= B n t n /n! is the generating function for B n, the last equation can be expressed in terms of this generating function by tg(t) = h(t) h(2t). Now tg(t) = t e t + 1 = t(et 1) e 2t 1 = t(et + 1) 2t e 2t = t 1 e t 1 2t e 2t 1 which shows that h(t) = t/(e t 1). Clearly h() = 1 thereby fixing the value of B as 1. Now h(t) is the generating function for the well-known Bernoulli numbers introduced in Section 5 in connection the Riemann zeta function. Thus the quantities B n defined here are actually the same familiar Bernoulli numbers. Thus with the aid of (5) and (7) we can now express the original recursive quantities T n in terms of Bernoulli numbers by the formula T = B 1 = ½, T n = (2 n+1 1)B n+1 /(n + 1) (n 1). Likewise, by (5) the sums of powers of the natural numbers series by 7. The Generating Function g(t) S = B 1, S n = B n+1 (n + 1) (n 1). In the last section I inferred the form of the generating function (6) indirectly and then verified that it did indeed yield the correct recursion relation for V n. After completing this article, however, I realised I could derive the generating function directly by adapting a clever technique used by Tao 1 to obtain the generating function for the Bernoulli numbers. We rewrite (4) as n 1 n! (n m)! and note that with D m d m /dx m It follows that m! + 2n! V n = (n 1) (8) n! n! x n m D m x n (m n) = {(n m)! (m > n) n!

so that (8) can be expressed in the form n! (n m)! (m < n) [D m x n ] x=1 + [D m x n ] x= = { 2n! (m = n) (m > n) m! {[D mx n ] x=1 + [D m x n ] x= } = (n 1). When n = the left-hand side is 2V = 1 since V = ½. Thus the case n = can be incorporated in the equation by writing it in the form m! {[D mx n ] x=1 + [D m x n ] x= } = [x n ] x= (n ). We can build on this result through linearity. Thus m! {[D m(a n x n + a k x k )] x=1 + [D m (a n x n + a k x k )] x= } = [a n x n + a k x k ] x= and in general, if P (m) (x) denotes the m th n derivative of the polynomial P(x) = Σ a m x m, then m! {P(m) (1) + P (m) ()} = P(). (9) This result can be extended to include convergent power series of the form P(x) = Σ a m x m. Let a m = t m /m!. Then we have P(x) = Σ (tx) m /m! = e tx, with values P (m) (1) = t m e t, P (m) () = t m and P() = 1. Thus from (9) which yields the generating function in agreement with (6). tm g(t) = m! {e t + 1} = 1 tm m! = 1 e t + 1 1 https://terrytao.wordpress.com/21/4/1/the-euler-maclaurin-formula-bernoulli-numbers-the-zeta-function-and-real-variableanalytic-continuation/