Fact-checked by Grok 2 weeks ago

Bernoulli polynomials

Bernoulli polynomials are a sequence of polynomials B_n(x) of degree n, defined by the \frac{t e^{xt}}{e^t - 1} = \sum_{n=0}^\infty B_n(x) \frac{t^n}{n!}, which generalize the Bernoulli numbers B_n = B_n(0) and play a fundamental role in and for expressing of integers. The Bernoulli numbers were introduced by in 1713 in his work to solve problems involving sums of consecutive powers, while the polynomials were introduced by Leonhard Euler in 1738 through the explicit . Key properties include the explicit formula B_n(x) = \sum_{k=0}^n \binom{n}{k} B_{n-k} x^k, which expresses them in terms of Bernoulli numbers, and the B_n(x+1) - B_n(x) = n x^{n-1}, useful for deriving the formula for the sum of the first m n-th powers as \sum_{k=1}^m k^n = \frac{1}{n+1} \left[ B_{n+1}(m+1) - B_{n+1} \right]. Additional notable features are their B_n(1-x) = (-1)^n B_n(x) for all non-negative integers n, the differentiation rule \frac{d}{dx} B_n(x) = n B_{n-1}(x), and their expansion B_n(x) = -\frac{n!}{(2\pi i)^n} \sum_{k \neq 0} \frac{e^{2\pi i k x}}{k^n} for non-integer x. These polynomials appear in diverse applications, including the Euler-Maclaurin formula for approximating sums by integrals, the study of zeta functions, and asymptotic expansions in .

History

Origins and Bernoulli Numbers

The Bernoulli numbers were first introduced by the Swiss mathematician in his posthumously published treatise in 1713, where he developed them as a tool for deriving explicit formulas for the of the first n positive integers, such as \sum_{k=1}^n k^p. also introduced the Bernoulli polynomials B_n(x) in this work to express these sums as \sum_{k=1}^m k^n = \frac{1}{n+1} [B_{n+1}(m+1) - B_{n+1}(0)]. 's approach involved identifying a recursive pattern in these sums, leading to a general expression that incorporated a sequence of rational coefficients now known as the Bernoulli numbers. This work built on earlier efforts by mathematicians like Johann Faulhaber, but 's systematic treatment marked a significant advance in understanding power sums through combinatorial methods. Despite their foundational role, the numbers are named after , even though subsequent key developments, including their broader applications and generalizations, were advanced by later figures such as Leonhard Euler. In the , Euler incorporated Bernoulli numbers into the Euler-Maclaurin summation formula, independently discovered around 1735 alongside , to approximate sums by integrals with correction terms derived from these numbers. This formula highlighted the numbers' utility in bridging discrete sums and continuous integrals, drawing an analogy between finite differences in discrete calculus and derivatives in continuous analysis, and built upon the earlier Bernoulli polynomials for power sums. The Bernoulli polynomials themselves appear as special cases evaluated at specific points, such as x = 0 or , preserving their recursive and combinatorial properties in the context of the Euler-Maclaurin formula.

Development and Applications

Leonhard Euler further developed the Bernoulli polynomials, providing their in 1738 and exploring them for arbitrary x in his Institutiones calculi differentialis (1755), to facilitate the summation of and applications in the Euler-Maclaurin formula, which approximates definite integrals by finite and provides asymptotic expansions for . This extension allowed for more flexible expressions in analyzing , such as \sum_{k=1}^n k^m, as polynomials in n. In the , mathematicians advanced the study of Bernoulli polynomials, with J.L. Raabe coining the term "Bernoulli polynomials" in 1851. They were used in the expansion of trigonometric series and the study of periodic functions, particularly in connection with zeta function evaluations. The saw significant developments through Gian-Carlo Rota's formalization of umbral in the , which provided a unified operator-based framework for manipulating Bernoulli polynomials alongside other like Hermite and . Rota's approach, using linear functionals on polynomial rings, revealed deep structural similarities and facilitated derivations of identities for these sequences without explicit computations. Early applications of Bernoulli polynomials extended to , where they underpin expansions in the Euler-Maclaurin for approximating integrals and sums in physics and engineering contexts. In , extensions of the von Staudt–Clausen theorem to Bernoulli polynomials, such as those for Hurwitz zeta values, determine denominators and modular properties, influencing results on L-functions and arithmetic progressions.

Definitions

Generating Function

The exponential generating function for the Bernoulli polynomials B_n(x) is \frac{t e^{x t}}{e^t - 1} = \sum_{n=0}^{\infty} B_n(x) \frac{t^n}{n!}, valid for |t| < 2\pi. This series defines the polynomials B_n(x) as the coefficients of t^n / n! in the Laurent series expansion of the left-hand side around t = 0. This generating function arises from the corresponding exponential generating function for the Bernoulli numbers, \frac{t}{e^t - 1} = \sum_{n=0}^{\infty} B_n \frac{t^n}{n!}, also valid for |t| < 2\pi, where B_n = B_n(0). Multiplying by the exponential shift factor e^{x t} incorporates the parameter x, yielding the more general form and allowing the extraction of B_n(x) for arbitrary x. This construction was first introduced by Euler in 1738. The structure of the generating function, expressed as e^{x t} g(t) with g(t) = t / (e^t - 1), identifies the Bernoulli polynomials as an Appell sequence. Consequently, they satisfy the translation formula B_n(x + y) = \sum_{k=0}^n \binom{n}{k} B_k(x) y^{n-k}, which derives from equating the generating functions for x + y and for x, then multiplying by e^{y t} and comparing coefficients via the binomial theorem.

Explicit Formula

The standard closed-form expression for the Bernoulli polynomials B_n(x) is given by B_n(x) = \sum_{k=0}^n \binom{n}{k} B_k x^{n-k}, where B_k denotes the kth Bernoulli number. This formula arises directly from the generating function for the Bernoulli polynomials, \frac{t e^{xt}}{e^t - 1} = \sum_{n=0}^\infty B_n(x) \frac{t^n}{n!}. Rewriting the left side as e^{xt} \cdot \frac{t}{e^t - 1} and expanding each factor as a power series—e^{xt} = \sum_{m=0}^\infty x^m \frac{t^m}{m!} and \frac{t}{e^t - 1} = \sum_{k=0}^\infty B_k \frac{t^k}{k!}—the coefficient of \frac{t^n}{n!} in the product is obtained via the Cauchy product of the series, which simplifies to the binomial sum above using the binomial theorem. The expression assumes the conventional choice B_1 = -\frac{1}{2} for the , which ensures B_1(x) = x - \frac{1}{2} and aligns with the generating function definition. In the alternative convention B_1^+ = +\frac{1}{2}, the corresponding polynomials B_n^+(x) satisfy B_n^+(x) = (-1)^{n+1} B_n(1 - x) for n \geq 2, altering the explicit sum to use B_k^+ instead; this impacts applications like the by changing the sign for odd-degree terms. An alternative closed-form representation expresses the Bernoulli polynomials in terms of the Hurwitz zeta function: B_n(x) = -n \zeta(1 - n, x), \quad n \geq 1, \ \operatorname{Re} x > 0, where \zeta(s, x) = \sum_{k=0}^\infty (k + x)^{-s} for \operatorname{Re} s > 1, extended analytically. This form highlights connections to and asymptotic expansions.

Representations

Operator Forms

One representation of the Bernoulli polynomials arises from the through a . Specifically, the nth Bernoulli polynomial is given by the nth of the evaluated at zero: B_n(x) = \left. \frac{d^n}{dt^n} \left( \frac{t e^{x t}}{e^t - 1} \right) \right|_{t=0}, where the notation \left( f(t) \right)^{\underline{n}} |_{t=0} denotes this operation. This form facilitates algebraic manipulations in operational calculus and connects directly to the exponential generating function \sum_{n=0}^\infty B_n(x) \frac{t^n}{n!} = \frac{t e^{x t}}{e^t - 1}. A key connection to finite difference operators is provided by the forward difference \Delta f(x) = f(x+1) - f(x). For Bernoulli polynomials, this yields \Delta B_n(x) = n x^{n-1}, which highlights their role in discretizing derivatives and appears in summation formulas like Euler-Maclaurin. This relation holds for n \geq 1 and underscores the polynomials' utility in approximating integrals by sums. In umbral calculus, Bernoulli polynomials admit an elegant operator interpretation where B_n(x) = (x + B)^n, with B denoting the umbral variable satisfying B^k = B_k for the kth B_k = B_k(0). This formal expansion treats the polynomials as if x and B commute in the umbral algebra, enabling mnemonic derivations of identities such as translation formulas B_n(x + y) = \sum_{k=0}^n \binom{n}{k} B_k(x) y^{n-k}. The approach leverages the structure of finite operator calculus to unify various polynomial sequences.

Integral Representations

One prominent integral representation of the Bernoulli polynomials arises from their generating function \frac{t e^{xt}}{e^t - 1} = \sum_{n=0}^\infty B_n(x) \frac{t^n}{n!}, valid for |t| < 2\pi. By extracting the coefficient of t^n/n! using Cauchy's residue theorem, with a counterclockwise contour C encircling the origin and avoiding the poles of e^t - 1 at t = 2\pi i k for integers k \neq 0, the representation is B_n(x) = \frac{n!}{2\pi i} \oint_C \frac{e^{x z}}{(e^z - 1) z^{n+1}} \, dz. This contour provides a global analytic expression useful for studying properties in the complex plane. Another contour integral representation, known as the Mellin–Barnes form, expresses the Bernoulli polynomials in terms of the Gamma function through the identity \pi / \sin(\pi t) = \Gamma(t) \Gamma(1 - t): B_n(x) = \frac{1}{2\pi i} \int_{-c - i\infty}^{-c + i\infty} (x + t)^n \left( \frac{\pi}{\sin(\pi t)} \right)^2 \, dt, where the vertical contour satisfies $0 < c < 1. This form links Bernoulli polynomials to the Hurwitz zeta function via \zeta(1 - n, x) = -B_n(x)/n for positive integers n \geq 2, facilitating connections to analytic number theory. Fourier-type integral representations offer real-line expressions that serve as precursors to the full Fourier series expansion of Bernoulli polynomials on [0, 1). For even degrees, B_{2n}(x) = (-1)^{n+1} 2n \int_0^\infty \frac{\cos(2\pi x) - e^{-2\pi t}}{\cosh(2\pi t) - \cos(2\pi x)} t^{2n-1} \, dt, valid for n = 1, 2, \dots and $0 < \Re x < 1. Similarly, for odd degrees greater than 1, B_{2n+1}(x) = (-1)^{n+1} (2n+1) \int_0^\infty \frac{\sin(2\pi x t)}{\cosh(2\pi t) - \cos(2\pi x)} t^{2n} \, dt, under the same conditions. These integrals highlight the periodic nature of Bernoulli polynomials and aid in deriving their Fourier series. These integral representations are particularly valuable for asymptotic analysis, where saddle-point methods applied to the contour integrals yield expansions for large |x| or high degrees. For instance, deforming the contour in the generating function integral to pass through a saddle point provides uniform asymptotic approximations, such as B_\nu(z) \sim z^\nu \sum_{k=0}^\infty c_k(\nu) z^{-k} for large |z| with \arg z bounded away from the negative real axis, enabling precise estimates in applications like summation formulas.

Properties

Recurrence Relations

The Bernoulli polynomials satisfy several recurrence relations that facilitate their computation and analysis. A fundamental relation is the difference equation B_n(x+1) - B_n(x) = n x^{n-1}, valid for all nonnegative integers n and real x. This identity can be derived from the generating function \frac{t e^{xt}}{e^t - 1} = \sum_{n=0}^\infty B_n(x) \frac{t^n}{n!}. Substituting x+1 yields \frac{t e^{(x+1)t}}{e^t - 1} = e^t \cdot \frac{t e^{xt}}{e^t - 1}, so the difference of generating functions is (e^t - 1) \cdot \frac{t e^{xt}}{e^t - 1} = t e^{xt}. Expanding the right side as t \sum_{k=0}^\infty \frac{(xt)^k}{k!} = \sum_{n=1}^\infty n x^{n-1} \frac{t^n}{n!} and equating coefficients gives the recurrence. An integral recurrence follows from the derivative property B_n'(x) = n B_{n-1}(x), which is obtained by differentiating the generating function with respect to x. Integrating both sides yields \int B_{n-1}(x) \, dx = \frac{1}{n} B_n(x) + C. For the definite integral over an interval of length 1, \int_0^1 B_n(x + t) \, dt = \frac{B_{n+1}(x+1) - B_{n+1}(x)}{n+1}. This holds because the substitution u = x + t transforms the left side to \int_x^{x+1} B_n(u) \, du, and applying the antiderivative gives the right side. These relations enable inductive computation of the polynomials. Starting from B_0(x) = 1, the explicit summation formula B_n(x) = \sum_{k=0}^n \binom{n}{k} B_k x^{n-k}, where B_k = B_k(0) are the , allows determination of higher-degree terms using previously computed lower-degree polynomials. This sum is derived by expanding the generating function \frac{t e^{xt}}{e^t - 1} = \frac{t}{e^t - 1} \cdot e^{xt} = \left( \sum_{k=0}^\infty B_k \frac{t^k}{k!} \right) \left( \sum_{m=0}^\infty \frac{(xt)^m}{m!} \right) and collecting coefficients. The difference recurrence then verifies or extends these computations for shifted arguments.

Symmetries and Translations

The Bernoulli polynomials exhibit a fundamental translation property that reflects their structure as a binomial transform of the Bernoulli numbers. Specifically, for nonnegative integers n and real numbers x, y, B_n(x + y) = \sum_{k=0}^n \binom{n}{k} B_k(x) y^{n-k}. This relation arises directly from the generating function \sum_{n=0}^\infty B_n(x) \frac{t^n}{n!} = \frac{t e^{x t}}{e^t - 1}; substituting x \to x + y yields \frac{t e^{(x+y) t}}{e^t - 1} = e^{y t} \cdot \frac{t e^{x t}}{e^t - 1}, and expanding e^{y t} = \sum_{m=0}^\infty \frac{(y t)^m}{m!} produces the binomial expansion in the coefficients of \frac{t^n}{n!}. As a member of the Appell sequence of polynomials, this translation invariance underscores the Bernoulli polynomials' role in generalizing sequences satisfying similar shift properties. A key symmetry of the Bernoulli polynomials is captured by the reflection formula B_n(1 - x) = (-1)^n B_n(x), which holds for all nonnegative integers n. This identity implies that the polynomials are even or odd functions with respect to the point x = 1/2, depending on the parity of n: for even n, B_n(1/2 + z) = B_n(1/2 - z), while for odd n, B_n(1/2 + z) = -B_n(1/2 - z). The formula applies uniformly, including for n=1 where B_1(x) = x - 1/2 satisfies B_1(1 - x) = -B_1(x), though n=1 is exceptional in that B_1(1) = 1/2 \neq -1/2 = B_1(0), unlike higher degrees where B_n(1) = B_n(0) for n \neq 1. For odd degrees n > 1, the reflection formula combines with the vanishing of the corresponding Bernoulli numbers B_n(0) = 0 to ensure antisymmetry around x=1/2 without endpoint equality at integers beyond the n=1 case. The reflection formula can be derived from the generating function by substituting x \to 1 - x, yielding \frac{t e^{(1-x) t}}{e^t - 1} = e^{t} \cdot \frac{t e^{-x t}}{e^t - 1}. Multiplying numerator and denominator in the fraction by e^{-t} gives \frac{t e^{-x t}}{1 - e^{-t}}, and recognizing that \frac{t}{1 - e^{-t}} = -\frac{t e^{t}}{e^{t} - 1} \cdot (-1) with t \to -t relates the series to \sum_{n=0}^\infty (-1)^n B_n(x) \frac{t^n}{n!}, confirming the sign alternation. This generating function approach highlights the intrinsic embedded in the structure defining the polynomials.

Differences and Derivatives

The derivative of the Bernoulli polynomial B_n(x) satisfies the relation B_n'(x) = n B_{n-1}(x) for n \geq 1. This formula arises from term-by-term of the generating function \frac{t e^{xt}}{e^t - 1} = \sum_{n=0}^\infty B_n(x) \frac{t^n}{n!}, yielding a factor of t that shifts the degree by one, and holds in the complex domain for \Re(v) > -1 and |\arg z| < \pi when extended to non-integer orders. The forward difference \Delta B_n(x) = B_n(x+1) - B_n(x) simplifies to \Delta B_n(x) = n x^{n-1} for n \geq 1. This property, which encodes the defining role of Bernoulli polynomials in , also extends to complex arguments under the same conditions as the derivative formula. It directly implies that \frac{B_n(x)}{n} serves as an antiderivative (indefinite sum) for the monomial x^{n-1} under the forward difference operator. Higher-order forward differences \Delta^k B_n(x) for $1 \leq k \leq n can be computed iteratively from the first difference, resulting in polynomials of degree n-k whose leading term is the falling factorial coefficient n^{\underline{k}} x^{n-k} = n(n-1)\cdots(n-k+1) x^{n-k}, accompanied by lower-degree terms arising from the binomial expansions in repeated applications of \Delta. The explicit form for the translation underlying these differences is B_n(x + m) = B_n(x) + n \sum_{j=0}^{m-1} (x + j)^{n-1}, allowing computation of \Delta^k B_n(x) via inclusion-exclusion on multiple translations; for example, \Delta^2 B_n(x) = n[(x+1)^{n-1} - x^{n-1}] = n(n-1)x^{n-2} + \frac{n(n-1)(n-2)}{2} x^{n-3} + \cdots. These higher differences connect Bernoulli polynomials to the falling factorial basis (x)_l = x(x-1)\cdots(x-l+1), in which the forward difference acts exactly as \Delta (x)_l = l (x)_{l-1}, mirroring differentiation on powers. Specifically, the expansion B_n(x) = B_n + \sum_{k=1}^n \binom{n}{k} S(n-1, k-1) x^k, combined with the change-of-basis formula x^k = \sum_{l=0}^k S(k,l) (x)_l (where S(k,l) are Stirling numbers of the second kind), expresses B_n(x) as a linear combination of falling factorials with coefficients involving Stirling numbers, facilitating computations in discrete settings where the falling factorial basis diagonalizes the difference operator. This structure under differences interprets Bernoulli polynomials as tools for discrete integration by parts (summation by parts). In the summation formula \sum_{j=a}^b u_j \Delta v_j = u_{b+1} v_{b+1} - u_a v_a - \sum_{j=a}^b \Delta u_j \, v_{j+1}, setting v_j = \frac{B_n(j)}{n} yields \Delta v_j = j^{n-1}, allowing sums of powers \sum j^{n-1} to be expressed using boundary terms involving Bernoulli polynomials, analogous to how x^{n-1}/(n-1) integrates (n-1) x^{n-2} in the continuous case via integration by parts. This enables closed-form evaluation of power sums and underpins applications in numerical analysis and combinatorial identities.

Explicit Forms

Low-Degree Expressions

The Bernoulli polynomials for low degrees provide concrete examples that illustrate their structure and utility in computations. These polynomials are monic, meaning the coefficient of the leading term x^n is always 1 for degree n, and their constant terms correspond to the B_n. The following explicit expressions for degrees 0 through 6 are derived from the standard definition and can be verified by expanding the generating function \frac{te^{xt}}{e^t - 1} = \sum_{n=0}^\infty B_n(x) \frac{t^n}{n!} up to the respective order or by applying the explicit formula B_n(x) = \sum_{k=0}^n \binom{n}{k} B_k x^{n-k}, where B_k are the . For clarity, the polynomials are presented in the table below:
Degree nB_n(x)
0$1
1x - \frac{1}{2}
2x^2 - x + \frac{1}{6}
3x^3 - \frac{3}{2} x^2 + \frac{1}{2} x
4x^4 - 2 x^3 + x^2 - \frac{1}{30}
5x^5 - \frac{5}{2} x^4 + \frac{5}{3} x^3 - \frac{1}{6} x
6x^6 - 3 x^5 + \frac{5}{2} x^4 - \frac{1}{2} x^2 + \frac{1}{42}
These low-degree forms reveal patterns such as the absence of constant terms for odd n \geq 3 (since B_n = 0 for odd n > 1) and the progressive increase in rational coefficients, which stem from the expansions in the explicit formula. Verification for specific cases, like B_2(x), can be confirmed by substituting into the and matching coefficients: the t^2/2! term yields x^2 - x + 1/6.

Maxima and Minima

The critical points of the B_n(x) occur where its vanishes, so B_n'(x) = 0. Due to the differentiation property B_n'(x) = n B_{n-1}(x), these points coincide exactly with the roots of the B_{n-1}(x) = 0. For low-degree cases, consider n=2: the B_2(x) = x^2 - x + \frac{1}{6} has B_2'(x) = 2(x - \frac{1}{2}), yielding a single critical point at x = \frac{1}{2}. At this point, B_2\left(\frac{1}{2}\right) = -\frac{1}{12}, representing the global minimum of B_2(x) on the real line. For n=4, the critical points are the roots of B_3(x) = 0 at x=0, x=\frac{1}{2}, and x=1, with local minima at x=0 and x=1 where B_4(0) = B_4(1) = -\frac{1}{30}, and a local maximum at x=\frac{1}{2} where B_4\left(\frac{1}{2}\right) = \frac{7}{240}. In general, the number of real critical points for B_n(x) equals the number of real c_{n-1} of B_{n-1}(x). Asymptotically, as the degree m \to \infty, the number of real satisfies c_m = \frac{2m}{\pi [e](/page/E!)} + \frac{\ln m}{\pi [e](/page/E!)} + O(1). These —and thus the critical points—are symmetrically distributed about x = \frac{1}{2}, with the largest asymptotically y_m \sim \frac{m}{2\pi [e](/page/E!)}. For even degrees n \equiv 2 \pmod{4}, B_n(x) decreases on [0, \frac{1}{2}] and increases on [\frac{1}{2}, 1], attaining a local minimum at x = \frac{1}{2}.

Applications

Sums of Powers

One of the key applications of Bernoulli polynomials lies in deriving closed-form expressions for finite sums of integer powers. For a positive integer p, the sum \sum_{k=1}^{m-1} k^p equals \frac{1}{p+1} \left[ B_{p+1}(m) - B_{p+1} \right], where B_{p+1} denotes the (p+1)-th Bernoulli number and B_{p+1}(m) is the corresponding Bernoulli polynomial evaluated at m. Substituting the binomial expansion of the Bernoulli polynomial, B_{p+1}(m) = \sum_{j=0}^{p+1} \binom{p+1}{j} B_j m^{p+1-j}, the formula simplifies because the j = p+1 term cancels with the subtracted Bernoulli number, yielding \sum_{k=1}^{m-1} k^p = \frac{1}{p+1} \sum_{j=0}^p \binom{p+1}{j} B_j m^{p+1-j}. This explicit form expresses the power sum as a polynomial in m of degree p+1. The derivation relies on the forward difference property of Bernoulli polynomials: \Delta B_{p+1}(x) = B_{p+1}(x+1) - B_{p+1}(x) = (p+1) x^p. Applying this, \sum_{k=1}^{m-1} (p+1) k^p = \sum_{k=1}^{m-1} \Delta B_{p+1}(k) = B_{p+1}(m) - B_{p+1}(1). For p \geq 1, B_{p+1}(1) = B_{p+1}, confirming the earlier expression. This telescoping sum interpretation highlights how the polynomials encode the cumulative differences of power functions. Jacob Bernoulli originally encountered these patterns while computing sums of powers to study figurate numbers in the context of probability and combinatorics, as detailed in his posthumous work Ars Conjectandi (1713), which introduced the Bernoulli numbers as coefficients in such formulas. A standard generalization covers the sum from 0 to n: \sum_{k=0}^n k^p = \frac{1}{p+1} \left[ B_{p+1}(n+1) - B_{p+1} \right], which for p > 0 equals the sum from 1 to n because the k=0 term is zero, with the adjustment shifting the upper limit via the polynomial evaluation.

Integrals and Euler-Maclaurin Formula

The provides a powerful for approximating the sum of a over integers by an , with correction terms involving Bernoulli numbers derived from Bernoulli polynomials. This formula bridges discrete summation and continuous integration, enabling precise asymptotic expansions for sums of the form \sum_{k=a}^{b} f(k) where f is sufficiently smooth. The standard form of the Euler–Maclaurin formula is \sum_{k=a}^{b} f(k) = \int_{a}^{b} f(x)\, dx + \frac{f(a) + f(b)}{2} + \sum_{k=1}^{m} \frac{B_{2k}}{(2k)!} \left( f^{(2k-1)}(b) - f^{(2k-1)}(a) \right) + R, where B_{2k} are the Bernoulli numbers (the constant terms in the Bernoulli polynomials B_{2k}(x)), f^{(j)} denotes the j-th of f, and R is the remainder term. This expression arises from expanding the difference between the sum and the integral using higher-order corrections that capture endpoint behaviors and oscillatory components. Historically, Leonhard Euler developed the formula in the 1730s as part of his investigations into sums of powers and the analytic continuation of the , publishing it in 1738; independently arrived at a similar result around 1742, leading to its joint attribution. Euler's approach relied on generating functions for Bernoulli numbers, which he connected to exponential series, while later refinements by in 1823 introduced the explicit remainder term. The derivation of the formula leverages integral representations of Bernoulli polynomials and their periodic extensions. Bernoulli polynomials B_n(x) admit the generating function \frac{ze^{zx}}{e^z - 1} = \sum_{n=0}^{\infty} B_n(x) \frac{z^n}{n!}, but the key step involves the periodic Bernoulli functions \tilde{B}_n(x) = B_n(\{x\}), where \{x\} = x - \lfloor x \rfloor is the fractional part, making \tilde{B}_n(x) periodic with period 1 and mean zero over each interval. To derive the formula, consider the sum as \sum_{k=a}^{b} f(k) = \int_{a}^{b} f(x) dx + \sum_{k=a}^{b} \int_{k-1/2}^{k+1/2} (f(k) - f(x)) dx, but a more direct path uses repeated on the identity \sum_{k=a}^{b} f(k) - \int_{a}^{b} f(x) dx = \int_{a}^{b} \tilde{B}_1(\{x\}) f'(x) dx, where \tilde{B}_1(x) = \{x\} - 1/2 is the sawtooth function (the periodic extension of B_1(x)). Integrating by parts yields higher terms: the j-th integration introduces \frac{\tilde{B}_j(x)}{j!} f^{(j-1)}(x), leading to the series expansion with Bernoulli polynomials evaluated at endpoints via their properties \int_{a}^{b} \tilde{B}_j'(x) dx = B_j(\{b\}) - B_j(\{a\}). This process truncates at order $2m using only even Bernoulli numbers B_{2k} due to the odd ones vanishing except for B_1. The remainder term R after m terms involves the next higher Bernoulli polynomial in integral form: R = \int_{a}^{b} \frac{\tilde{B}_{2m+2}(\{x\})}{(2m+2)!} f^{(2m+2)}(x) dx, which incorporates higher Bernoulli numbers through the Fourier series of the periodic extension, \tilde{B}_{2k}(\{x\}) = -\frac{(2k)!}{(2\pi)^{2k}} \sum_{j \neq 0} \frac{e^{2\pi i j x}}{j^{2k}} for k > 1. This remainder can be bounded using the growth of Bernoulli numbers, |B_{2k}| \sim 4\sqrt{\pi} (k/(\pi e))^{2k}, ensuring convergence for analytic f. Poisson's summation further refines it for applications in asymptotic analysis.

Multiplication Theorems

The multiplication theorems for Bernoulli polynomials provide identities that relate the value of B_n at a scaled and shifted argument to a weighted sum of values at equally spaced shifts, generalizing translation properties to multiple arguments. These theorems are fundamental algebraic tools in the theory of Bernoulli polynomials, enabling the computation of values at rational points and facilitating connections to other special functions. A central result is the , which states that for a positive h \geq 1, k with $0 \leq k < h, and n \geq 0, B_n\left( h x + \frac{k}{h} \right) = h^{n-1} \sum_{j=0}^{h-1} B_n\left( x + \frac{j}{h} \right) e^{2 \pi i j k / h}. When k = 0, this simplifies to the basic form B_n( h x ) = h^{n-1} \sum_{j=0}^{h-1} B_n\left( x + \frac{j}{h} \right). This identity holds for all real x and is attributed to early developments in the functional equations satisfied by Bernoulli polynomials. (Nörlund, 1924, as referenced in Carlitz, 1962) For the case h=2, known as the doubling formulas, the theorem yields B_n(2x) = 2^{n-1} \left[ B_n(x) + B_n\left(x + \frac{1}{2}\right) \right] and, for the shift k=1, B_n\left(2x + \frac{1}{2}\right) = 2^{n-1} \left[ B_n(x) - B_n\left(x + \frac{1}{2}\right) \right]. These special cases arise directly from substituting h=2 into the general formula and are useful for recursive computations of Bernoulli polynomial values at half-integers. The proofs of these theorems rely on the exponential generating function \frac{t e^{x t}}{e^t - 1} = \sum_{n=0}^\infty B_n(x) \frac{t^n}{n!}. For the basic case (k=0), one considers the sum \sum_{j=0}^{h-1} e^{(j/h) t} as a geometric series, leading to a relation between the generating function at h x and the summed generating functions at the shifts after scaling the variable t appropriately. The general case incorporates the roots of unity filter: the exponential terms e^{2 \pi i j k / h} act as characters to extract the shifted component from the periodic extension, ensuring the identity holds algebraically for the polynomials. This approach leverages the uniqueness of the generating function and properties of finite differences. These theorems have applications in equidistribution theory, where they aid in analyzing the discrepancy of sequences using Fourier expansions involving Bernoulli polynomials, and in evaluating lattice sums over scaled grids, such as in analytic number theory for summing powers over arithmetic progressions.

Advanced Topics

Periodic Bernoulli Polynomials

The periodic Bernoulli polynomials, denoted \tilde{B}_n(x), are defined by \tilde{B}_n(x) = B_n(\{x\}), where B_n denotes the ordinary Bernoulli polynomials and \{x\} = x - \lfloor x \rfloor is the fractional part of x. This construction ensures that \tilde{B}_n(x + 1) = \tilde{B}_n(x) for all real x, establishing periodicity with period 1. For n = 0, \tilde{B}_0(x) = 1. For n = 1, \tilde{B}_1(x) = \{x\} - \frac{1}{2}, which exhibits discontinuities at every integer, where the left-hand limit is \frac{1}{2} and the right-hand limit is -\frac{1}{2}, resulting in a jump discontinuity of size 1; some formulations assign the value 0 at integers to symmetrize the function. For n \geq 2, the functions are continuous at integers (and everywhere), since B_n(1) = B_n(0) = B_n, the nth Bernoulli number, ensuring the left- and right-hand limits match at these points. The generating function for the periodic Bernoulli polynomials is given by \frac{t e^{t \{x\}}}{e^t - 1} = \sum_{n=0}^{\infty} \tilde{B}_n(x) \frac{t^n}{n!}. This exponential generating function captures the periodic nature through its Fourier expansion, where the coefficients involve terms inversely proportional to powers related to the degree n. Although the series converges for |t| < 2\pi in the ordinary case, the periodic extension leverages the function's repetition for applications requiring bounded domains or modular arithmetic. Periodic Bernoulli polynomials find significant applications in discrepancy theory and the analysis of uniform distribution modulo 1. In discrepancy theory, they provide bounds on the deviation of point distributions from uniformity, often appearing in error estimates for the distribution function of sequences like the fractional parts of irrational rotations. For uniform distribution modulo 1, these functions quantify irregularities in sequences through integrals involving \tilde{B}_n(x), such as in Weyl's criterion or Koksma's inequality, where higher-degree terms refine approximations of how well a sequence fills the unit interval.

Fourier Series

The periodic Bernoulli polynomials \tilde{B}_n(x), defined as the 1-periodic extension of B_n(x) for n \geq 2 (with adjustment for n=1), admit an explicit Fourier series expansion given by \tilde{B}_n(x) = -\frac{n!}{(2\pi i)^n} \sum_{k \in \mathbb{Z} \setminus \{0\}} \frac{e^{2\pi i k x}}{k^n}, \quad n \geq 2. This representation holds for all real x, where the sum excludes the k=0 term, which would diverge for n \leq 1. The formula originates from Adolf Hurwitz's 19th-century work on trigonometric expansions of periodic functions related to Bernoulli polynomials. The derivation proceeds from the generating function for Bernoulli polynomials, \frac{t e^{x t}}{e^t - 1} = \sum_{n=0}^\infty B_n(x) \frac{t^n}{n!}, which, when extended periodically, yields a generating function for \tilde{B}_n(x). Applying the Poisson summation formula to this generating function transforms the sum over integers into a dual sum involving Fourier coefficients. Specifically, the partial fraction expansion of \pi \cot(\pi z) provides the key identity \pi \cot(\pi z) = \frac{1}{z} + \sum_{k=1}^\infty \left( \frac{1}{z - k} + \frac{1}{z + k} \right), which is \mathbb{Z}-periodic and meromorphic. Substituting z = t / (2\pi i) and expanding as a Laurent series around t=0 connects the residues at integer poles to the Bernoulli terms, leading to the exponential sum after termwise integration and periodicity enforcement. This approach leverages the rapid decay of the Fourier coefficients for higher n. The series connects to polylogarithms via the decomposition \sum_{k \neq 0} \frac{e^{2\pi i k x}}{k^n} = \mathrm{Li}_n(e^{2\pi i x}) + (-1)^n \mathrm{Li}_n(e^{-2\pi i x}), where \mathrm{Li}_n(z) = \sum_{k=1}^\infty z^k / k^n is the polylogarithm function; for integer n \geq 2, this yields the periodic Bernoulli polynomial up to the scaling factor. At integer points, such as x=0, the series reduces to -n! (2 \zeta(n)) / (2\pi i)^n for even n, linking directly to Riemann zeta function values \zeta(n) = \sum_{k=1}^\infty 1/k^n, whereas for odd n > 1, the sum \sum_{k \neq 0} 1/k^n = 0 due to antisymmetry, consistent with the vanishing Bernoulli numbers B_n = 0. These relations underpin evaluations of \zeta at positive integers. The series converges absolutely and uniformly on \mathbb{R} for n \geq 2, owing to the O(1/|k|^n) decay of terms, ensuring the partial sums approximate \tilde{B}_n(x) with error bounded by the tail sum. For n=1, convergence is conditional, resembling the of the sawtooth function. Term-by-term differentiation is justified by of the differentiated series, yielding \frac{d}{dx} \tilde{B}_n(x) = n \tilde{B}_{n-1}(x) for n \geq 2, as the resulting sum for degree n-1 matches the direct expansion; this holds iteratively down to the base case.

Inversion Formulas

Inversion formulas for Bernoulli polynomials enable the recovery of input functions or sequences from their polynomial evaluations or transforms, playing a key role in and umbral algebra. These relations often leverage the structure of posets or operator inverses to express , , or solutions to differential equations in terms of Bernoulli polynomials. A prominent example arises in the of Bernoulli polynomials, where inversion in the poset of positive integers ordered by divisibility inverts the standard expansions. The even Bernoulli polynomials admit the cosine series B_{2k}(x) = 2(-1)^{k-1} \frac{(2k)!}{(2\pi)^{2k}} \sum_{n=1}^\infty \frac{\cos(2\pi n x)}{n^{2k}}, \quad x \in [0,1), \ k \geq 1. Applying inversion yields the reciprocal relation \cos(2\pi x) = (-1)^{k-1} \frac{(2\pi)^{2k}}{2(2k)!} \sum_{n=1}^\infty \frac{\mu(n) B_{2k}(\{n x\})}{n^{2k}}, \quad x \in \mathbb{R}, \ k \geq 1, where \mu is the Möbius function and \{ \cdot \} denotes the fractional part. Similarly, for odd indices, B_{2k+1}(x) = 2(-1)^{k} \frac{(2k+1)!}{(2\pi)^{2k+1}} \sum_{n=1}^\infty \frac{\sin(2\pi n x)}{n^{2k+1}}, \quad x \in [0,1), \ k \geq 0, inverts to \sin(2\pi x) = (-1)^{k} \frac{(2\pi)^{2k+1}}{2(2k+1)!} \sum_{n=1}^\infty \frac{\mu(n) B_{2k+1}(\{n x\})}{n^{2k+1}}, \quad x \in \mathbb{R}, \ k \geq 0. These formulas recover the generating from weighted sums of Bernoulli polynomials evaluated at scaled fractional parts, facilitating asymptotic estimates and properties for rational arguments. In the framework of umbral calculus, Bernoulli polynomials form a Sheffer sequence associated with the delta operator \delta = D(e^t - 1), where D is , admitting an umbral inversion via compositional inverse. The umbra B satisfies B(B + 1) = B + 1, and its compositional inverse \hat{B} fulfills B_n(\hat{B}(x)) = x^n = \hat{B}_n(B(x)) for n \geq 0. This inversion extends to relations like the expansion of (x - B)^n, which in umbral notation generates the polynomials r_n(x) satisfying \sum_{k=0}^n \binom{n}{k} B_k r_{n-k}(x) = \delta_{n,0} x^n, where \delta_{n,0} is the ; these reciprocal polynomials coincide with signed Stirling polynomials of the first kind scaled appropriately. Such umbral inversions allow expressing monomials as linear combinations of Bernoulli polynomials and vice versa, underpinning operator manipulations in theory. Regarding the , defined for a \{a_n\} by b_n = \sum_{k=0}^n \binom{n}{k} a_k, the standard inversion recovers a_n = \sum_{k=0}^n (-1)^{n-k} \binom{n}{k} b_k. When the transform involves polynomials, such as in the expansion B_n(x) = \sum_{k=0}^n \binom{n}{k} B_k x^{n-k}, inversion expresses the Bernoulli numbers B_k from evaluations at specific points or via the umbral inverse, yielding x^n = \sum_{k=0}^n \binom{n}{k} B_k \hat{B}_{n-k}(x). This recovers the original power basis from the Bernoulli-transformed , with applications in sequence manipulation and combinatorial identities. These inversion techniques find application in solving linear recurrences with coefficients, where Bernoulli polynomials provide closed-form particular s through generating functions or representations. For instance, consider the recurrence u_{n+1} - (a + b n) u_n = p(n), where p(n) is a of d; the incorporates Bernoulli polynomials to handle the linear term in the coefficients, yielding u_n = c \prod_{k=0}^{n-1} (a + b k) + \sum_{m=0}^d c_m B_m(n) adjusted via , with inversion ensuring uniqueness of coefficients from initial conditions. This approach leverages the properties of Bernoulli polynomials, analogous to Taylor expansions for continuous cases, to resolve linear systems efficiently.

Relation to Falling Factorial

The monomial powers x^n can be expressed in the falling factorial basis via the relation x^n = \sum_{k=0}^n S(n,k) (x)^{\underline{k}}, where S(n,k) denotes the Stirling numbers of the second kind, which count the number of ways to partition a set of n objects into k nonempty unlabeled subsets, and (x)^{\underline{k}} = x(x-1)\cdots(x-k+1) is the falling factorial. This change of basis is fundamental in , as the falling factorial aligns naturally with the forward difference operator \Delta f(x) = f(x+1) - f(x), satisfying \Delta (x)^{\underline{k}} = k (x)^{\underline{k-1}}, analogous to of powers. Bernoulli polynomials, being elements of the , admit a unique in this basis: B_n(x) = B_n + \sum_{k=1}^n \frac{n}{k} S(n-1,k-1) (x)^{\underline{k}}, where B_n = B_n(0) is the nth . This representation, derived from the umbral calculus framework, facilitates connections between the analytic properties of Bernoulli polynomials and discrete structures. Inverse relations also exist, allowing falling factorials to be expressed in terms of Bernoulli polynomials through appropriate linear combinations involving . In the context of finite differences, the Bernoulli polynomials play the role of discrete antiderivatives for monomials. Specifically, the defining relation B_n(x+1) - B_n(x) = n x^{n-1} implies \Delta \left[ \frac{B_n(x)}{n} \right] = x^{n-1}, for n \geq 1. This positions \frac{B_n(x)}{n} as the indefinite sum (discrete analog of the integral) of x^{n-1}, mirroring the continuous case where \int x^{n-1} \, dx = \frac{x^n}{n}. Such properties underpin summation formulas like Faulhaber's formula for sums of powers. This basis change has applications in combinatorics, particularly for counting problems with restrictions, such as selecting and ordering subsets without repetition—injections from a k-set to an x-set, enumerated by (x)^{\underline{k}}. The Stirling numbers provide combinatorial interpretations for the coefficients, enabling the expression of Bernoulli-related quantities (e.g., in power sum evaluations) in terms of partition counts and restricted selections, useful in enumerative combinatorics and algorithm analysis.

References

  1. [1]
    Bernoulli Polynomial -- from Wolfram MathWorld
    There are two definitions of Bernoulli polynomials in use. The nth Bernoulli polynomial is denoted here by B_n(x) (Abramowitz and Stegun 1972), ...Missing: properties | Show results with:properties
  2. [2]
    The Bernoulli Number Page
    This page gives an introduction to the Bernoulli numbers and polynomials, as well as to the Euler numbers.
  3. [3]
    [PDF] The Bernoulli Numbers: A Brief Primer - Whitman College
    May 10, 2019 · The Bernoulli polynomials are a generalization of the Bernoulli numbers. They have a variety of interesting properties, and will feature in our ...
  4. [4]
    The Origin of the Bernoulli Numbers: Mathematics in Basel and Edo ...
    Jul 23, 2021 · The Bernoulli numbers were named after the Swiss mathematician Jacob Bernoulli (1654–1705; Figure 1), whose posthumous book Ars Conjectandi ...
  5. [5]
    Sums of Powers of Positive Integers - Jakob Bernoulli (1654-1705 ...
    Bernoulli derived symbolic formulas for the sums of positive integer powers using the method conjectured above for Fermat, then noted a pattern that would make ...Missing: Jacob | Show results with:Jacob
  6. [6]
    The Powers Sums, Bernoulli Numbers, Bernoulli Polynomials ...
    Then in 1713 in his posthumous Ars conjectandi, Jacob Bernoulli [6] , mentioning Faulhaber, published the lists of ten first ∑n m . It is plausible that ...<|control11|><|separator|>
  7. [7]
    The Independent Derivations by Leonhard Euler and Colin ... - jstor
    in their own right, were given by James (Jakob) I Bernoulli (1654-1705) in. Ars Conjectandi published posthumously in 1713. Whether by the mid 1730s. Euler ...
  8. [8]
    [PDF] Bernoulli numbers and the Euler-Maclaurin summation formula
    In this note, I shall motivate the origin of the Euler-Maclaurin summation formula. I will also explain why the coefficients on the right hand side of this.
  9. [9]
    [PDF] Euler-Maclaurin Formula 1 Introduction - People | MIT CSAIL
    Both used iterative method of obtaining Bernoulli's numbers bi, but Maclaurin's approach was mainly based on geometric structure while Euler used purely ...
  10. [10]
    Bernoulli polynomials - Encyclopedia of Mathematics
    May 29, 2020 · Euler [1] was the first to study Bernoulli polynomials for arbitrary values of x. The term "Bernoulli polynomials" was introduced by J.L. Raabe ...
  11. [11]
    [PDF] Bernoulli and Euler polynomials
    In 1713 J. Bernoulli introduced the Bernoulli numbers, and used them to express. Sk(n) as a polynomial in n with degree k + 1. Later Euler ...
  12. [12]
    On the von Staudt–Clausen's theorem related to q-Frobenius–Euler ...
    Some generalizations and basic (or q-) extensions of the Bernoulli, Euler and Genocchi polynomials. Appl. Math. Inf. Sci., 5 (2011), pp. 390-444. View in ...
  13. [13]
    The umbral calculus - ScienceDirect
    On the properties of the Δn0n class of numbers and of others analogous to them, as investigated by means of representative notation.
  14. [14]
    [PDF] Applications of the classical umbral calculus - Brandeis
    We describe applications of the classical umbral calculus to bilinear generat- ing functions for polynomial sequences, identities for Bernoulli and related ...
  15. [15]
  16. [16]
    Theory of The Generalized Bernoulli-Hurwitz Numbers for ... - arXiv
    Jun 15, 2004 · This paper gives new generalization of Bernoulli and Hurwitz numbers for higher genus cases. They satisfy completely von Staudt-Clausen type theorem.Missing: polynomials | Show results with:polynomials
  17. [17]
    DLMF: §24.2 Definitions and Generating Functions ‣ Properties ...
    Symbols: B n ⁡ ( x ) : Bernoulli polynomials, E n ⁡ ( x ) : Euler polynomials, n : integer and x : real or complex
  18. [18]
    DLMF: §24.4 Basic Properties ‣ Properties ‣ Chapter 24 Bernoulli and Euler Polynomials
    ### Summary of Addition/Translation Formula for Bernoulli Polynomials
  19. [19]
    [PDF] The Gamma function - The University of British Columbia
    Dec 26, 2023 · >0 is spanned by the differential operator D = xd/dx, and the ... These are the Bernoulli polynomials. They determine in turn functions ψn ...
  20. [20]
    24.7 Integral Representations ‣ Properties ‣ Chapter 24 Bernoulli ...
    Bernoulli numbers, Euler numbers, integral representation, integral representations ... Bernoulli polynomials, Euler polynomials, integral representations; Notes ...
  21. [21]
    [PDF] Bernoulli Polynomials Old and New
    There are several reductions for this general definition. - Whenµ= 1 we have the Bernoulli polynomials Bn(z). - When z = 0 we have the generalized Bernoulli ...
  22. [22]
    [PDF] 18.785 Number Theory Fall 2015 Problem Set #8 Due
    Nov 16, 2015 · these properties uniquely determine the Bernoulli polynomials. (b) Prove that Bn(x + 1) − Bn(x) = nxn−1 and. Bn(x + y) = n. X k=0 n k. Bk(x) ...
  23. [23]
    [PDF] Bernoulli Numbers and their Applications - DSpace@MIT
    Bernoulli Numbers are a set of numbers that is created by restricting the Bernoulli polyno- mials to x = 0 and will formally proceed to define.Missing: convention | Show results with:convention
  24. [24]
    [PDF] New Formulas Involving Bernoulli and Stirling Numbers of Both Kinds
    Dec 23, 2024 · The Bernoulli numbers were first introduced by Jacob Bernoulli ... Kouba, Lecture notes, Bernoulli polynomials and numbers, preprint, 2016.
  25. [25]
    On the Maxima and Minima of Bernoulli Polynomials
    Apr 11, 2018 · (1940). On the Maxima and Minima of Bernoulli Polynomials. The American Mathematical Monthly: Vol. 47, No. 8, pp. 533-538.
  26. [26]
    [PDF] The Bernoulli numbers, power sums, and zeta values
    The Bernoulli numbers and zeta values by contour integration. For Re(s) > 1 compute that. Γ(s)n−s = Z ∞ t=0 e−nt(nt)s d(nt) nt. · n−s = Z ∞ t=0 e−ntts dt t. , n ...
  27. [27]
    [PDF] Bernoulli numbers and the Euler–Maclaurin summation formula
    We are now in a position to formulate the Euler–Maclaurin summation formula. Theorem B.5 (Euler–Maclaurin) Suppose that K is a positive integer and that f has ...Missing: history | Show results with:history
  28. [28]
    [PDF] 5. The Euler-Maclaurin Summation Formula
    Mar 10, 2011 · This formula shows how a sum can be approximated by an integral and gives an exact error term. 5.2. Theorem (Euler-Maclaurin I). Let x0 be a ...
  29. [29]
    Some Generalized Multiplication Formulas for the Bernoulli ... - EuDML
    Carlitz, L.. "Some Generalized Multiplication Formulas for the Bernoulli Polynomials and Related Functions.." Monatshefte für Mathematik 66 (1962): 1-8.Missing: Monatsh. | Show results with:Monatsh.
  30. [30]
    [PDF] Characterization of the Bernoulli polynomials via the Raabe ... - arXiv
    Mar 25, 2023 · By the Raabe multiplication theorem, we know that the polynomials λBn(X) (λ ∈ C) are all solutions of Equation (En,a). To prove the converse, we ...
  31. [31]
    [PDF] THE BERNOULLI PERIODIC FUNCTIONS - Miguel A. Lerma's
    We study slightly modified versions of the Bernoulli periodic functions with nicer structural properties, and use them to give a very simple proof of the Euler- ...
  32. [32]
    [PDF] Distribution of Sequences: A Sampler
    Jan 18, 2018 · ... periodic Bernoulli polynomials Bh(x), h = 1, 2,... . (II) B.G.Sloss and W.F.Blyth (1993) replaced the class of continuous functions by the ...
  33. [33]
    [PDF] A series representation of the cotangent
    a generating function for the Bernoulli numbers. Again for |z| < 1, πz cot ... + 2ζ(k)=0, and Euler's formula follows immediately. Since π cot πz is Z-periodic it ...
  34. [34]
    [PDF] Fourier expansions of polynomials and values of ζ
    Piecewise-polynomial functions have left and right derivatives everywhere, so the. Fourier series for Bk(x) with k > 1 will converge to Bk(x - [[x]]) for all x.
  35. [35]
    [PDF] arXiv:2004.10441v1 [math-ph] 22 Apr 2020
    Apr 22, 2020 · Bernoulli polynomials as its Fourier series and taking cuts, which includes both the Euler-Maclaurin summation formula and the Poission ...
  36. [36]
    [PDF] The Möbius inversion formula for Fourier series applied to Bernoulli ...
    Abstract. Hurwitz found the Fourier expansion of the Bernoulli polynomials over a century ago. In general, Fourier analysis can be fruitfully employed to ...Missing: \neq \frac
  37. [37]
  38. [38]
    [PDF] a73 integers 21 (2021) linear recurrences for bernoulli polynomials ...
    Sep 7, 2021 · This paper demonstrates a method to produce linear recurrence relations for Bernoulli polynomials involving two kinds of sums, using a Mellin- ...
  39. [39]
  40. [40]
    Falling Factorial -- from Wolfram MathWorld
    Is also known as the binomial polynomial, lower factorial, falling factorial power (Graham et al. ... Concrete Mathematics: A Foundation for Computer Science, 2nd ...Missing: bernoulli | Show results with:bernoulli