This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

De Moivre and Bell polynomials

Cormac  O’Sullivan111Date: Mar 5, 2022.
   2010 Mathematics Subject Classification: 05A15, 05A16, 13F25.
   Support for this project was provided by a PSC-CUNY Award, jointly funded by The Professional Staff Congress and The City
   University of New York.
Abstract

We survey a family of polynomials that are very useful in all kinds of power series manipulations, and appearing more frequently in the literature. Applications to formal power series, generating functions and asymptotic expansions are described, and we discuss the related work of De Moivre, Arbogast and Bell.

1 Introduction

Let a1x+a2x2+a3x3+a_{1}x+a_{2}x^{2}+a_{3}x^{3}+\cdots be a formal power series without a constant term and with coefficients in {\mathbb{C}}. It could represent a function in some neighborhood of x=0x=0, but usually we don’t require convergence. The kkth power of this series may be expanded into another series

(a1x+a2x2+a3x3+)k=n𝒜n,k(a1,a2,a3,)xn(k0),\mathopen{}\mathclose{{}\left(a_{1}x+a_{2}x^{2}+a_{3}x^{3}+\cdots}\right)^{k}=\sum_{n\in{\mathbb{Z}}}{\mathcal{A}}_{n,k}(a_{1},a_{2},a_{3},\dots)x^{n}\qquad\quad(k\in{\mathbb{Z}}_{\geqslant 0}), (1.1)

and it can be seen that the new coefficients 𝒜n,k(a1,a2,a3,){\mathcal{A}}_{n,k}(a_{1},a_{2},a_{3},\dots) can only be nonzero if nkn\geqslant k, in which case they depend at most on a1,a2,,ama_{1},a_{2},\dots,a_{m} for m=nk+1m=n-k+1. The multinomial development

(a1x+a2x2++amxm)k=j1+j2++jm=k(kj1,j2,,jm)(a1x)j1(a2x2)j2(amxm)jm\mathopen{}\mathclose{{}\left(a_{1}x+a_{2}x^{2}+\cdots+a_{m}x^{m}}\right)^{k}=\sum_{j_{1}+j_{2}+\dots+j_{m}=k}\binom{k}{j_{1},j_{2},\dots,j_{m}}(a_{1}x)^{j_{1}}(a_{2}x^{2})^{j_{2}}\cdots(a_{m}x^{m})^{j_{m}}

shows that

𝒜n,k(a1,a2,a3,)=1j1+2j2++mjm=nj1+j2++jm=k(kj1,j2,,jm)a1j1a2j2amjm{\mathcal{A}}_{n,k}(a_{1},a_{2},a_{3},\dots)=\sum_{\begin{subarray}{c}1j_{1}+2j_{2}+\dots+mj_{m}=n\\ j_{1}+j_{2}+\dots+j_{m}=k\end{subarray}}\binom{k}{j_{1},j_{2},\dots,j_{m}}a_{1}^{j_{1}}a_{2}^{j_{2}}\cdots a_{m}^{j_{m}} (1.2)

where the sum is over all possible j1j_{1}, j2j_{2}, …, jm0j_{m}\in{\mathbb{Z}}_{\geqslant 0} and 00=10^{0}=1 throughout.

Thus 𝒜n,k{\mathcal{A}}_{n,k} is a simply described polynomial in a1,a2,,ank+1a_{1},a_{2},\dots,a_{n-k+1}. In the literature they are designated ‘partial ordinary Bell polynomials’, and related to the better-known partial Bell polynomials n,k{\mathcal{B}}_{n,k}. We will make the case that in fact the 𝒜n,k{\mathcal{A}}_{n,k} polynomials are the more natural and useful version. Prior to Bell, Arbogast [Arb00] was already working with them in 1800. However, their earliest appearance seems to be in a 1697 paper of De Moivre [DM97]. There he says he was curious to see if he could generalize to (1.1) his friend Mr. Newton’s work with binomials. His solution is equivalent to (1.2), though uses an interesting recurrence to describe which products a1j1a2j2amjma_{1}^{j_{1}}a_{2}^{j_{2}}\cdots a_{m}^{j_{m}} appear. This is discussed in section 6. See also [Sch68, Sect. 5.1] for historical context. We propose a new, more succinct name for these fundamental objects:

Definition 1.1.

For nn, kk\in{\mathbb{Z}} with k0k\geqslant 0, define the De Moivre polynomial 𝒜n,k(a1,a2,){\mathcal{A}}_{n,k}(a_{1},a_{2},\dots) by (1.1). If n<kn<k it is 0. If nkn\geqslant k it is given by (1.2), making a polynomial in a1,a2,,ank+1a_{1},a_{2},\dots,a_{n-k+1} of homogeneous degree kk with positive integer coefficients and number of terms equalling the number of partitions of nn with kk parts; see (6.1).

In this article we aim to provide a convenient reference for these polynomials, while also describing some new results about them. The author has found them to be of great use in giving explicit forms for asymptotic expansions, and we will see examples of this in section 7. More generally, as shown in section 3, composing, inverting and taking powers of generating functions and power series becomes easy with the help of the De Moivre polynomials, giving clear expressions for the new coefficients. We find in section 6 that some familiar generating functions take on a new look with this treatment. Many applications are discussed; we may also mention [Pit06, Chap. 1], [Cha02, Chap. 11], for example, for further uses of De Moivre\Bell polynomials in probability and statistics and for relations among symmetric polynomials.

Before reviewing the basic properties of 𝒜n,k{\mathcal{A}}_{n,k} in the next section, we complete this introduction by including more of the history of these ideas. Replacing the ordinary series in (1.1) with an exponential series, (in other words a Taylor series), defines the partial Bell polynomials:

1k!(a1x1!+a2x22!+a3x33!+)k=n=0n,k(a1,a2,a3,)xnn!(k0).\frac{1}{k!}\mathopen{}\mathclose{{}\left(a_{1}\frac{x}{1!}+a_{2}\frac{x^{2}}{2!}+a_{3}\frac{x^{3}}{3!}+\cdots}\right)^{k}=\sum_{n=0}^{\infty}{\mathcal{B}}_{n,k}(a_{1},a_{2},a_{3},\dots)\frac{x^{n}}{n!}\qquad\quad(k\in{\mathbb{Z}}_{\geqslant 0}). (1.3)

(We are following Comtet’s terminology from [Com74, Chap. 3].) Hence we have the relation

n,k(a1,a2,a3,)=n!k!𝒜n,k(a11!,a22!,a33!,).{\mathcal{B}}_{n,k}(a_{1},a_{2},a_{3},\dots)=\frac{n!}{k!}{\mathcal{A}}_{n,k}(\frac{a_{1}}{1!},\frac{a_{2}}{2!},\frac{a_{3}}{3!},\dots). (1.4)

The complete Bell polynomials are defined as

𝒴n(a1,a2,a3,):=k=0nn,k(a1,a2,a3,).{\mathcal{Y}}_{n}(a_{1},a_{2},a_{3},\dots):=\sum_{k=0}^{n}{\mathcal{B}}_{n,k}(a_{1},a_{2},a_{3},\dots).

There is also a related Bell polynomial in one variable:

n(x):=n![tn]ex(et1)=k=0nn,k(1,1,1,)xk{\mathcal{B}}_{n}(x):=n![t^{n}]e^{x(e^{t}-1)}=\sum_{k=0}^{n}{\mathcal{B}}_{n,k}(1,1,1,\dots)x^{k}

where [tn][t^{n}] indicates the coefficient of tnt^{n} in a series.

Bell introduced the complete polynomials 𝒴n{\mathcal{Y}}_{n} in [Bel34] as a wide generalization of the Hermite and Appell polynomials. They are also related to the partition polynomials he was previously considering - see [O’S, Thm. 6.6]. Riordan studied 𝒴n{\mathcal{Y}}_{n} and n,k{\mathcal{B}}_{n,k} in [Rio68], naming them Bell polynomials. Comtet emphasized that what he termed the ‘partial ordinary’ version should be used when dealing with ordinary series like a0+a1x+a2x2+a_{0}+a_{1}x+a_{2}x^{2}+\cdots, (he used the notation ^n,k\hat{{\mathcal{B}}}_{n,k} but we prefer 𝒜n,k{\mathcal{A}}_{n,k} to clearly differentiate the versions). However, all the formulas in [Com74, Chap. 3] involve the polynomials n,k{\mathcal{B}}_{n,k} and this might account for their dominance in the literature.

Research into the origins of Faà di Bruno’s formula in [Knu97, pp. 52, 481 - 483] and [Cra05, Joh02b] uncovered the work of Arbogast in [Arb00] where the formula in fact first appeared. Arbogast takes powers of polynomials in his method and gives the formula (1.2) in [Arb00, pp. 43-44]. For example, in a table on p. 29 of [Arb00], the entry

6β5ζ+15β4(2γε+δ2)+60β3γ2δ+15β2γ46\beta^{5}\zeta+15\beta^{4}(2{\gamma}\varepsilon+\delta^{2})+60\beta^{3}{\gamma}^{2}\delta+15\beta^{2}{\gamma}^{4} (1.5)

associated to x10x^{10} and D6D^{6} is provided. In our notation this is 𝒜10,6(β,γ,δ,ε,ζ,){\mathcal{A}}_{10,6}(\beta,\gamma,\delta,\varepsilon,\zeta,\dots), or more clearly,

𝒜10,6(a1,a2,a3,)=6a15a5+30a14a2a4+15a14a32+60a13a22a3+15a12a24.{\mathcal{A}}_{10,6}(a_{1},a_{2},a_{3},\dots)=6a_{1}^{5}a_{5}+30a_{1}^{4}a_{2}a_{4}+15a_{1}^{4}a_{3}^{2}+60a_{1}^{3}a_{2}^{2}a_{3}+15a_{1}^{2}a_{2}^{4}. (1.6)

We explain what he was doing after Theorem 3.3.

Abraham De Moivre (1667-1754), though perhaps best known today for a trigonometric formula, was an important pioneer in areas such as probability and statistics, roots of equations, analytic geometry and infinite series. Born in France, he moved to England with other Huguenots to escape persecution, and there became part of the circle that included Newton, Halley and Stirling. The biography [BG07] gives more details about his life, and the papers [Sch68, Cra05, Gél] describe aspects of his mathematics, some of which we will touch on.

2 Basic properties of 𝒜n,k{\mathcal{A}}_{n,k}

The results in this section follow from the generating function (1.1) as exercises. Some of these and their partial Bell polynomial analogs appear in [Rio68, pp. 188 – 192] and [Com74, Sec. 3.3]. The following equalities (2.1) – (2.12) are identities in the ring [a1,a2,]{\mathbb{Z}}[a_{1},a_{2},\dots].

For small kk,

𝒜n,0(a1,a2,)\displaystyle{\mathcal{A}}_{n,0}(a_{1},a_{2},\dots) =δn,0,\displaystyle=\delta_{n,0}, (2.1)
𝒜n,1(a1,a2,)\displaystyle{\mathcal{A}}_{n,1}(a_{1},a_{2},\dots) =an(n1),\displaystyle=a_{n}\qquad\qquad(n\geqslant 1), (2.2)
𝒜n,2(a1,a2,)\displaystyle{\mathcal{A}}_{n,2}(a_{1},a_{2},\dots) =j=1n1ajanj.\displaystyle=\sum_{j=1}^{n-1}a_{j}a_{n-j}. (2.3)

The identities (2.2), (2.3) generalize to give a symmetric formula for 𝒜n,k{\mathcal{A}}_{n,k}:

𝒜n,k(a1,a2,)=n1+n2++nk=nan1an2ank(k1),{\mathcal{A}}_{n,k}(a_{1},a_{2},\dots)=\sum_{n_{1}+n_{2}+\dots+n_{k}=n}a_{n_{1}}a_{n_{2}}\cdots a_{n_{k}}\qquad\qquad(k\geqslant 1), (2.4)

where the sum in (2.4) is over all possible n1n_{1}, n2,,nk1n_{2},\dots,n_{k}\in{\mathbb{Z}}_{\geqslant 1}. The two main recursion relations for 𝒜n,k{\mathcal{A}}_{n,k} in nn and kk are given by

𝒜n+k,k(a1,a2,a3,)\displaystyle{\mathcal{A}}_{n+k,k}(a_{1},a_{2},a_{3},\dots) =j=0k(kj)a1kj𝒜n,j(a2,a3,),\displaystyle=\sum_{j=0}^{k}\binom{k}{j}a_{1}^{k-j}{\mathcal{A}}_{n,j}(a_{2},a_{3},\dots), (2.5)
𝒜n,k+1(a1,a2,a3,)\displaystyle{\mathcal{A}}_{n,k+1}(a_{1},a_{2},a_{3},\dots) =j=kn1anj𝒜j,k(a1,a2,a3,).\displaystyle=\sum_{j=k}^{n-1}a_{n-j}{\mathcal{A}}_{j,k}(a_{1},a_{2},a_{3},\dots). (2.6)

In fact (2.6) is the =1\ell=1 case of the following identity:

j=0n𝒜j,k(a1,a2,)𝒜nj,(a1,a2,)=𝒜n,k+(a1,a2,).\sum_{j=0}^{n}{\mathcal{A}}_{j,k}(a_{1},a_{2},\dots)\cdot{\mathcal{A}}_{n-j,\ell}(a_{1},a_{2},\dots)={\mathcal{A}}_{n,k+\ell}(a_{1},a_{2},\dots). (2.7)

Note that (2.5) is useful to compute 𝒜n+k,k{\mathcal{A}}_{n+k,k} if nn is small, since the terms in the sum can only be nonzero for jnj\leqslant n. For example, when k1k\geqslant 1 we have

𝒜k,k(a1,a2,)=a1k,\displaystyle{\mathcal{A}}_{k,k}(a_{1},a_{2},\dots)=a_{1}^{k}, (2.8)
𝒜k+1,k(a1,a2,)=ka1k1a2,\displaystyle{\mathcal{A}}_{k+1,k}(a_{1},a_{2},\dots)=ka_{1}^{k-1}a_{2}, (2.9)
𝒜k+2,k(a1,a2,)=(k1)a1k1a3+(k2)a1k2a22,\displaystyle{\mathcal{A}}_{k+2,k}(a_{1},a_{2},\dots)=\binom{k}{1}a_{1}^{k-1}a_{3}+\binom{k}{2}a_{1}^{k-2}a_{2}^{2}, (2.10)
𝒜k+3,k(a1,a2,)=(k1)a1k1a4+2(k2)a1k2a2a3+(k3)a1k3a23,\displaystyle{\mathcal{A}}_{k+3,k}(a_{1},a_{2},\dots)=\binom{k}{1}a_{1}^{k-1}a_{4}+2\binom{k}{2}a_{1}^{k-2}a_{2}a_{3}+\binom{k}{3}a_{1}^{k-3}a_{2}^{3}, (2.11)
𝒜k+4,k(a1,a1,)=(k1)a1k1a5+(k2)a1k2(a32+2a2a4)+3(k3)a1k3a22a3+(k4)a1k4a24.\displaystyle{\mathcal{A}}_{k+4,k}(a_{1},a_{1},\dots)=\binom{k}{1}a_{1}^{k-1}a_{5}+\binom{k}{2}a_{1}^{k-2}\mathopen{}\mathclose{{}\left(a_{3}^{2}+2a_{2}a_{4}}\right)+3\binom{k}{3}a_{1}^{k-3}a_{2}^{2}a_{3}+\binom{k}{4}a_{1}^{k-4}a_{2}^{4}. (2.12)

De Moivre continued this list as far as 𝒜k+6,k{\mathcal{A}}_{k+6,k} in [DM97, Fig. 5].

For a variable zz, the general binomial coefficients satisfy (z0):=1\binom{z}{0}:=1 and for positive integers kk,

(zk):=z(z1)(zk+1)k!,(zk)=(1)k(z+k1k).\binom{z}{k}:=\frac{z(z-1)\cdots(z-k+1)}{k!},\qquad\binom{-z}{k}=(-1)^{k}\binom{z+k-1}{k}. (2.13)
Lemma 2.1.

For nk0n\geqslant k\geqslant 0, we have the following identities in [z]{\mathbb{Q}}[z]:

𝒜n,k(1,1,1,)\displaystyle{\mathcal{A}}_{n,k}\mathopen{}\mathclose{{}\left(1,1,1,\dots}\right) =(n1nk),\displaystyle=\binom{n-1}{n-k}, (2.14)
𝒜n,k((z0),(z1),(z2),)\displaystyle{\mathcal{A}}_{n,k}\mathopen{}\mathclose{{}\left(\binom{z}{0},\binom{z}{1},\binom{z}{2},\dots}\right) =(kznk),\displaystyle=\binom{kz}{n-k}, (2.15)
𝒜n,k((z0),(z+11),(z+22),)\displaystyle{\mathcal{A}}_{n,k}\mathopen{}\mathclose{{}\left(\binom{z}{0},\binom{z+1}{1},\binom{z+2}{2},\dots}\right) =(n+kz1nk).\displaystyle=\binom{n+kz-1}{n-k}. (2.16)
Proof.

The equality (2.15) follows from the binomial theorem. Then (2.15) implies (2.16), using the right identity in (2.13). Lastly, (2.14) is a special case of (2.16). ∎

The next relations are often useful:

𝒜n,k(0,a1,a2,a3,)\displaystyle{\mathcal{A}}_{n,k}(0,a_{1},a_{2},a_{3},\dots) =𝒜nk,k(a1,a2,a3,),\displaystyle={\mathcal{A}}_{n-k,k}(a_{1},a_{2},a_{3},\dots), (2.17)
𝒜n,k(ca1,ca2,ca3,)\displaystyle{\mathcal{A}}_{n,k}(ca_{1},ca_{2},ca_{3},\dots) =ck𝒜n,k(a1,a2,a3,),\displaystyle=c^{k}{\mathcal{A}}_{n,k}(a_{1},a_{2},a_{3},\dots), (2.18)
𝒜n,k(ca1,c2a2,c3a3,)\displaystyle{\mathcal{A}}_{n,k}(ca_{1},c^{2}a_{2},c^{3}a_{3},\dots) =cn𝒜n,k(a1,a2,a3,).\displaystyle=c^{n}{\mathcal{A}}_{n,k}(a_{1},a_{2},a_{3},\dots). (2.19)

Equalities (2.14) and (2.18) easily give:

Lemma 2.2.

Suppose the aja_{j} are complex numbers satisfying |aj|Q|a_{j}|\leqslant Q. Then for nk0n\geqslant k\geqslant 0,

|𝒜n,k(a1,a2,)|(n1nk)Qk2nQk.\mathopen{}\mathclose{{}\left|{\mathcal{A}}_{n,k}(a_{1},a_{2},\dots)}\right|\leqslant\binom{n-1}{n-k}Q^{k}\leqslant 2^{n}Q^{k}.

Some further special values of 𝒜n,k{\mathcal{A}}_{n,k} are described next, related to the exponential function and the logarithm. We have

𝒜m+k,k(10!,11!,12!,)=kmm!(m,k0).{\mathcal{A}}_{m+k,k}\mathopen{}\mathclose{{}\left(\frac{1}{0!},\frac{1}{1!},\frac{1}{2!},\dots}\right)=\frac{k^{m}}{m!}\qquad(m,k\geqslant 0). (2.20)

Also, for n,k0n,k\geqslant 0,

(et1)k\displaystyle(e^{t}-1)^{k} =n=0k!n!{nk}tn\displaystyle=\sum_{n=0}^{\infty}\frac{k!}{n!}{\genfrac{\{}{\}}{0.0pt}{}{n}{k}}t^{n} \displaystyle\implies\qquad 𝒜n,k(11!,12!,13!,)\displaystyle{\mathcal{A}}_{n,k}\mathopen{}\mathclose{{}\left(\frac{1}{1!},\frac{1}{2!},\frac{1}{3!},\dots}\right) =k!n!{nk},\displaystyle=\frac{k!}{n!}{\genfrac{\{}{\}}{0.0pt}{}{n}{k}}, (2.21)
(log(1t))k\displaystyle(-\log(1-t))^{k} =n=0k!n![nk]tn\displaystyle=\sum_{n=0}^{\infty}\frac{k!}{n!}{\genfrac{[}{]}{0.0pt}{}{n}{k}}t^{n} \displaystyle\implies\qquad 𝒜n,k(11,12,13,)\displaystyle{\mathcal{A}}_{n,k}\mathopen{}\mathclose{{}\left(\frac{1}{1},\frac{1}{2},\frac{1}{3},\dots}\right) =k!n![nk].\displaystyle=\frac{k!}{n!}{\genfrac{[}{]}{0.0pt}{}{n}{k}}. (2.22)

The left identities of (2.21) and (2.22), (see [GKP94, (7.49), (7.50)]), may be taken as the definitions of the Stirling numbers, as in [Com74, p. 51], and all their properties developed from this starting point. Alternatively, the Stirling subset numbers {nk}{\genfrac{\{}{\}}{0.0pt}{}{n}{k}} count the number of ways to partition nn elements into kk nonempty subsets, and the Stirling cycle numbers [nk]{\genfrac{[}{]}{0.0pt}{}{n}{k}} count the number of ways to arrange nn elements into kk cycles. More advanced special values of 𝒜n,k{\mathcal{A}}_{n,k} are shown in [O’S, Sect. 9.3].

We notice that the De Moivre polynomials in (2.21) appear with arguments shifted from the ones in (2.20). This type of shifting will also be seen in examples in sections 6 and 7. In the simplest case, adding or removing the first coefficient a1a_{1} in 𝒜n,k{\mathcal{A}}_{n,k} has a simple effect, by the binomial theorem:

Lemma 2.3.

For k0k\geqslant 0,

𝒜n,k(a2,a3,)\displaystyle{\mathcal{A}}_{n,k}(a_{2},a_{3},\dots) =j=0k(a1)kj(kj)𝒜n+j,j(a1,a2,),\displaystyle=\sum_{j=0}^{k}(-a_{1})^{k-j}\binom{k}{j}{\mathcal{A}}_{n+j,j}(a_{1},a_{2},\dots), (2.23)
𝒜n,k(a1,a1,)\displaystyle{\mathcal{A}}_{n,k}(a_{1},a_{1},\dots) =j=0ka1kj(kj)𝒜nk,j(a2,a3,).\displaystyle=\sum_{j=0}^{k}a_{1}^{k-j}\binom{k}{j}{\mathcal{A}}_{n-k,j}(a_{2},a_{3},\dots). (2.24)

(The identity (2.24) is just (2.5) again.) Then applying (2.23) and (2.24) rr times leads to the following.

Proposition 2.4.

For r,k0r,k\geqslant 0, 𝒜n,k(ar+1,ar+2,){\mathcal{A}}_{n,k}(a_{r+1},a_{r+2},\dots) equals

j1+j2++jr+1=k(kj1,j2,,jr+1)(a1)j1(a2)j2(ar)jr𝒜n+J+rjr+1,jr+1(a1,a2,)\sum_{j_{1}+j_{2}+\dots+j_{r+1}=k}\binom{k}{j_{1},j_{2},\dots,j_{r+1}}(-a_{1})^{j_{1}}(-a_{2})^{j_{2}}\cdots(-a_{r})^{j_{r}}{\mathcal{A}}_{n+J+rj_{r+1},j_{r+1}}(a_{1},a_{2},\dots) (2.25)

and 𝒜n,k(a1,a2,){\mathcal{A}}_{n,k}(a_{1},a_{2},\dots) equals

j1+j2++jr+1=k(kj1,j2,,jr+1)a1j1a2j2arjr𝒜n+Jrk,jr+1(ar+1,ar+2,)\sum_{j_{1}+j_{2}+\dots+j_{r+1}=k}\binom{k}{j_{1},j_{2},\dots,j_{r+1}}a_{1}^{j_{1}}a_{2}^{j_{2}}\cdots a_{r}^{j_{r}}{\mathcal{A}}_{n+J-rk,j_{r+1}}(a_{r+1},a_{r+2},\dots) (2.26)

where JJ means (r1)j1+(r2)j2++1jr1(r-1)j_{1}+(r-2)j_{2}+\cdots+1j_{r-1} and the summations are over all j1j_{1}, …, jr+10j_{r+1}\in{\mathbb{Z}}_{\geqslant 0} with sum kk.

For example, taking r=nk+1r=n-k+1 and setting ar+1=ar+2==0a_{r+1}=a_{r+2}=\cdots=0 in (2.26) recovers (1.2) since the only nonzero terms have n+Jrk=jr+1=0n+J-rk=j_{r+1}=0 by (2.1).

3 Manipulating power series

Our power series coefficients aja_{j} may come from a ring RR. Though more general cases can be considered, a natural choice for RR is an integral domain containing {\mathbb{Z}} (and so of characteristic 0). Throughout this section we assume RR has these properties and hence R[[x]]R[[x]], the ring of formal power series over RR, is also an integral domain containing {\mathbb{Z}}. See [GJ83, Chap. 1], [Com74, Sect. 1.12] or [FS09, Sect. A5] for more on formal power series rings. An important point is that, at each step, a new coefficient must depend only on finitely many others, as if we were working with polynomials; see the notions of summable and admissible in [GJ83, Chap. 1]. This is why the inner series in Proposition 3.1 cannot have a constant term, for example.

3.1 Series composition

Proposition 3.1.

Suppose that f(x)=a1x+a2x2+f(x)=a_{1}x+a_{2}x^{2}+\cdots and g(x)=b0+b1x+b2x2+g(x)=b_{0}+b_{1}x+b_{2}x^{2}+\cdots are two power series in R[[x]]R[[x]]. Then g(f(x))=c0+c1x+c2x2+g(f(x))=c_{0}+c_{1}x+c_{2}x^{2}+\cdots is in R[[x]]R[[x]] with

cn=k=0nbk𝒜n,k(a1,a2,).\quad c_{n}=\sum_{k=0}^{n}b_{k}\cdot{\mathcal{A}}_{n,k}(a_{1},a_{2},\dots). (3.1)
Proof.

We have

g(f(x))=k=0bkf(x)k=k=0bkn=k𝒜n,k(a1,a2,)xn,g(f(x))=\sum_{k=0}^{\infty}b_{k}\cdot f(x)^{k}=\sum_{k=0}^{\infty}b_{k}\sum_{n=k}^{\infty}{\mathcal{A}}_{n,k}(a_{1},a_{2},\dots)x^{n},

and switching the order of summation gives (3.1). ∎

Though this result is for formal power series, it applies very widely. A commonly occurring situation has f(x)f(x) and g(x)g(x) holomorphic in neighborhoods of 0 with f(0)=0f(0)=0. Then g(f(x)g(f(x) is also holomorphic in a neighborhood of 0 and its Taylor coefficients may be computed using Proposition 3.1.

Three important cases are as follows. Suppose that the integral domain RR contains {\mathbb{Q}}. Applying Proposition 3.1 with g(x)=(1+x)αg(x)=(1+x)^{\alpha}, eαxe^{\alpha x} and log(1+αx)\log(1+\alpha x) gives

[xn](1+f(x))α\displaystyle[x^{n}](1+f(x))^{\alpha} =k=0n(αk)𝒜n,k(a1,a2,),\displaystyle=\sum_{k=0}^{n}\binom{\alpha}{k}{\mathcal{A}}_{n,k}(a_{1},a_{2},\dots), (3.2)
[xn]eαf(x)\displaystyle[x^{n}]e^{\alpha f(x)} =k=0nαkk!𝒜n,k(a1,a2,),\displaystyle=\sum_{k=0}^{n}\frac{\alpha^{k}}{k!}{\mathcal{A}}_{n,k}(a_{1},a_{2},\dots), (3.3)
[xn]log(1+αf(x))\displaystyle[x^{n}]\log(1+\alpha f(x)) =k=1n(1)k1αkk𝒜n,k(a1,a2,).\displaystyle=\sum_{k=1}^{n}(-1)^{k-1}\frac{\alpha^{k}}{k}{\mathcal{A}}_{n,k}(a_{1},a_{2},\dots). (3.4)

The polynomials in a1,a2,a_{1},a_{2},\dots on the right sides of (3.2), (3.3) and (3.4) are essentially the potential, complete exponential and logarithmic polynomials, respectively, of [Com74, Sections 3.3, 3.5]. Recalling (2.13), the right sides of (3.2), (3.3) and (3.4) are also degree nn polynomials in α\alpha. Therefore the series (1+f(x))α(1+f(x))^{\alpha}, eαf(x)e^{\alpha f(x)} and log(1+αf(x))\log(1+\alpha f(x)) make sense for arbitrary α\alpha, giving elements of R[α][[x]]R[\alpha][[x]]. Further, if f(x)f(x) is holomorphic in a neighborhood of 0, then so are the compositions g(f(x))g(f(x)), with (3.2), (3.3), (3.4) (times n!n!) giving their Taylor coefficients.

For an example we will need later, consider

log(1+log(1+x)u)=n=1n(u)xn,\log\mathopen{}\mathclose{{}\left(1+\frac{\log(1+x)}{u}}\right)=\sum_{n=1}^{\infty}\ell_{n}(u)x^{n}, (3.5)

and we would like to know how n(u)\ell_{n}(u) depends on uu. Stepping through the proof of Proposition 3.1 shows

k=1(1)k+1kuklogk(1+x)\displaystyle\sum_{k=1}^{\infty}\frac{(-1)^{k+1}}{k\cdot u^{k}}\log^{k}(1+x) =k=1(1)k+1kukn=k𝒜n,k(1,12,13,)xn\displaystyle=\sum_{k=1}^{\infty}\frac{(-1)^{k+1}}{k\cdot u^{k}}\sum_{n=k}^{\infty}{\mathcal{A}}_{n,k}(1,-{\textstyle\frac{1}{2}},{\textstyle\frac{1}{3}},\dots)x^{n}
=n=1xnk=1n(1)k+1kuk𝒜n,k(1,12,13,).\displaystyle=\sum_{n=1}^{\infty}x^{n}\sum_{k=1}^{n}\frac{(-1)^{k+1}}{k\cdot u^{k}}{\mathcal{A}}_{n,k}(1,-{\textstyle\frac{1}{2}},{\textstyle\frac{1}{3}},\dots).

It is already clear that n(u)\ell_{n}(u) is a polynomial of degree nn in 1/u1/u with no constant term. With (2.18), (2.19) and (2.22) we obtain the simplification

n(u)=(1)n+1k=1n(k1)!n![nk]uk.\ell_{n}(u)=(-1)^{n+1}\sum_{k=1}^{n}\frac{(k-1)!}{n!}{\genfrac{[}{]}{0.0pt}{}{n}{k}}u^{-k}. (3.6)

A very common case that follows from (3.2) should be highlighted:

Proposition 3.2.

Let RR be an integral domain containing {\mathbb{Z}} with a0+a1x+a2x2+a_{0}+a_{1}x+a_{2}x^{2}+\cdots in R[[x]]R[[x]]. Then for mm\in{\mathbb{Z}},

(a0+a1x+a2x2+)m=n=0[k=0n(mk)a0mk𝒜n,k(a1,a2,)]xn\mathopen{}\mathclose{{}\left(a_{0}+a_{1}x+a_{2}x^{2}+\cdots}\right)^{m}=\sum_{n=0}^{\infty}\mathopen{}\mathclose{{}\left[\sum_{k=0}^{n}\binom{m}{k}a_{0}^{m-k}{\mathcal{A}}_{n,k}(a_{1},a_{2},\dots)}\right]x^{n} (3.7)

is in R[[x]]R[[x]], (provided a0a_{0} is invertible in RR if m<0m<0), and in particular, the multiplicative inverse of the series is given by

(a0+a1x+a2x2+)1=n=0[k=0n(1)ka0k1𝒜n,k(a1,a2,)]xn.\mathopen{}\mathclose{{}\left(a_{0}+a_{1}x+a_{2}x^{2}+\cdots}\right)^{-1}=\sum_{n=0}^{\infty}\mathopen{}\mathclose{{}\left[\sum_{k=0}^{n}(-1)^{k}a_{0}^{-k-1}{\mathcal{A}}_{n,k}(a_{1},a_{2},\dots)}\right]x^{n}. (3.8)

3.2 Arbogast’s formula

See [Cra05] and [Joh02b] for the early history of the next famous result which is usually named for Faà di Bruno. It is essentially equivalent to Proposition 3.1.

Theorem 3.3 (Arbogast’s formula).

For nn times differentiable functions ff and gg,

1n!dndxng(f(x))=k=0ng(k)(f(x))k!𝒜n,k(f(x)1!,f′′(x)2!,).\frac{1}{n!}\frac{d^{n}}{dx^{n}}g(f(x))=\sum_{k=0}^{n}\frac{g^{(k)}(f(x))}{k!}\cdot{\mathcal{A}}_{n,k}\mathopen{}\mathclose{{}\left(\frac{f^{\prime}(x)}{1!},\frac{f^{\prime\prime}(x)}{2!},\dots}\right). (3.9)
Proof.

Use induction to establish that (3.9) must be true for some polynomial Pn,kP_{n,k} instead of 𝒜n,k{\mathcal{A}}_{n,k}. To identify Pn,kP_{n,k} set f(x)=a1x+a2x2+f(x)=a_{1}x+a_{2}x^{2}+\cdots, g(x)=b0+b1x+b2x2+g(x)=b_{0}+b_{1}x+b_{2}x^{2}+\cdots and g(f(x))=c0+c1x+c2x2+g(f(x))=c_{0}+c_{1}x+c_{2}x^{2}+\cdots. Then evaluating at x=0x=0 finds

cn=k=0nbkPn,k(a1,a2,).c_{n}=\sum_{k=0}^{n}b_{k}\cdot P_{n,k}\mathopen{}\mathclose{{}\left(a_{1},a_{2},\dots}\right).

Comparing this with (3.1) now shows that Pn,k𝒜n,kP_{n,k}\equiv{\mathcal{A}}_{n,k}, since the coefficients aia_{i} and bjb_{j} are arbitrary. ∎

Arbogast had a similar point of view in [Arb00]. He considered f(x)=a0+a1x+a2x2+f(x)=a_{0}+a_{1}x+a_{2}x^{2}+\cdots and gave the formula

1n!dndxng(f(x))|x=0=k=0ng(k)(a0)k!𝒜n,k(a1,a2,),\frac{1}{n!}\mathopen{}\mathclose{{}\left.\frac{d^{n}}{dx^{n}}g(f(x))}\right|_{x=0}=\sum_{k=0}^{n}\frac{g^{(k)}(a_{0})}{k!}\cdot{\mathcal{A}}_{n,k}\mathopen{}\mathclose{{}\left(a_{1},a_{2},\dots}\right), (3.10)

computing (3.10) explicitly in tables for 1n101\leqslant n\leqslant 10. For example, see Figure 1 of [Cra05], reproduced from [Arb00, p. 29], where 𝒜10,k(β,γ,δ,ε,ζ,ν,θ,ι,κ,λ,){\mathcal{A}}_{10,k}(\beta,\gamma,\delta,\varepsilon,\zeta,\nu,\theta,\iota,\kappa,\lambda,\dots) is given correctly for 1k101\leqslant k\leqslant 10.

In the notation of Proposition 3.1, a further composition h(g(f(x)))h(g(f(x))) would require 𝒜n,r(c0,c1,){\mathcal{A}}_{n,r}(c_{0},c_{1},\dots). Computing this using (3.1) looks hopelessly complicated, with De Moivre polynomials inside De Moivre polynomials, but in fact a slight extension of a lemma of Charalambides, [Cha02, Lemma 11.1], gives a simple formula.

Proposition 3.4.

With the assumptions of Proposition 3.1,

𝒜n+r,r(c0,c1,)=k=0n𝒜n,k(a1,a2,)𝒜k+r,r(b0,b1,).{\mathcal{A}}_{n+r,r}(c_{0},c_{1},\dots)=\sum_{k=0}^{n}{\mathcal{A}}_{n,k}(a_{1},a_{2},\dots)\cdot{\mathcal{A}}_{k+r,r}(b_{0},b_{1},\dots). (3.11)
Proof.

Write y=f(x)y=f(x) so that

(yg(y))r=k=r𝒜k,r(b0,b1,)yk,(y\cdot g(y))^{r}=\sum_{k=r}^{\infty}{\mathcal{A}}_{k,r}(b_{0},b_{1},\dots)y^{k},

and hence

g(y)r\displaystyle g(y)^{r} =k=0𝒜k+r,r(b0,b1,)yk\displaystyle=\sum_{k=0}^{\infty}{\mathcal{A}}_{k+r,r}(b_{0},b_{1},\dots)y^{k}
=k=0𝒜k+r,r(b0,b1,)n=k𝒜n,k(a1,a2,)xn\displaystyle=\sum_{k=0}^{\infty}{\mathcal{A}}_{k+r,r}(b_{0},b_{1},\dots)\sum_{n=k}^{\infty}{\mathcal{A}}_{n,k}(a_{1},a_{2},\dots)x^{n}
=n=0xnk=0n𝒜n,k(a1,a2,)𝒜k+r,r(b0,b1,).\displaystyle=\sum_{n=0}^{\infty}x^{n}\sum_{k=0}^{n}{\mathcal{A}}_{n,k}(a_{1},a_{2},\dots)\cdot{\mathcal{A}}_{k+r,r}(b_{0},b_{1},\dots).

To finishes the proof, note that

𝒜n+r,r(c0,c1,)=[xn+r](c0x+c1x2+)r=[xn]g(y)r.{\mathcal{A}}_{n+r,r}(c_{0},c_{1},\dots)=[x^{n+r}](c_{0}x+c_{1}x^{2}+\cdots)^{r}=[x^{n}]g(y)^{r}.\qed

(Use (2.17) to simplify this result if b0=c0=0b_{0}=c_{0}=0.) Now we see that Proposition 3.1 is just the r=1r=1 case of Proposition 3.4. Also (3.11) suggests the following natural extension of Arbogast’s formula to rrth powers of a composition. It is probably contained in the large literature on Arbogast’s formula, but we have not found it.

Theorem 3.5.

Let nn and rr be nonnegative integers. For nn times differentiable functions ff and gg,

1n!dndxng(f(x))r=k=0n𝒜n,k(f(x)1!,f′′(x)2!,)𝒜k+r,r(g(f(x))0!,g(f(x))1!,).\frac{1}{n!}\frac{d^{n}}{dx^{n}}g(f(x))^{r}=\sum_{k=0}^{n}{\mathcal{A}}_{n,k}\mathopen{}\mathclose{{}\left(\frac{f^{\prime}(x)}{1!},\frac{f^{\prime\prime}(x)}{2!},\dots}\right)\cdot{\mathcal{A}}_{k+r,r}\mathopen{}\mathclose{{}\left(\frac{g(f(x))}{0!},\frac{g^{\prime}(f(x))}{1!},\dots}\right). (3.12)
Proof.

Let G(x):=g(x)rG(x):=g(x)^{r}. Applying Proposition 3.1 finds

1n!dndxnG(f(x))=k=0n𝒜n,k(f(x)1!,f′′(x)2!,)G(k)(f(x))k!.\frac{1}{n!}\frac{d^{n}}{dx^{n}}G(f(x))=\sum_{k=0}^{n}{\mathcal{A}}_{n,k}\mathopen{}\mathclose{{}\left(\frac{f^{\prime}(x)}{1!},\frac{f^{\prime\prime}(x)}{2!},\dots}\right)\cdot\frac{G^{(k)}(f(x))}{k!}. (3.13)

With G(x)=h(g(x))G(x)=h(g(x)) for h(x):=xrh(x):=x^{r}, another application gives

G(k)(y)k!\displaystyle\frac{G^{(k)}(y)}{k!} =j=0kh(j)(g(y))j!𝒜k,j(g(y)1!,g′′(y)2!,)\displaystyle=\sum_{j=0}^{k}\frac{h^{(j)}(g(y))}{j!}\cdot{\mathcal{A}}_{k,j}\mathopen{}\mathclose{{}\left(\frac{g^{\prime}(y)}{1!},\frac{g^{\prime\prime}(y)}{2!},\dots}\right)
=j=0k(rj)g(y)rj𝒜k,j(g(y)1!,g′′(y)2!,).\displaystyle=\sum_{j=0}^{k}\binom{r}{j}g(y)^{r-j}\cdot{\mathcal{A}}_{k,j}\mathopen{}\mathclose{{}\left(\frac{g^{\prime}(y)}{1!},\frac{g^{\prime\prime}(y)}{2!},\dots}\right). (3.14)

Since (rj)=0\binom{r}{j}=0 for j>rj>r and 𝒜k,j=0{\mathcal{A}}_{k,j}=0 for j>kj>k, we may change the upper limit of summation in (3.14) from kk to rr. Then by (2.5),

G(k)(y)k!=𝒜k+r,r(g(y),g(y)1!,),\frac{G^{(k)}(y)}{k!}={\mathcal{A}}_{k+r,r}\mathopen{}\mathclose{{}\left(g(y),\frac{g^{\prime}(y)}{1!},\dots}\right),

and the desired formula follows. ∎

Lastly we mention that it is natural to generalize (1.1) to power series of more than one variable and then consider their compositions. This results in De Moivre polynomials with elaborate indexing; see the recent paper [Sch] and its contained references for details.

3.3 Compositional inverses

We call gg a compositional inverse of ff if g(f(x))=x=f(g(x))g(f(x))=x=f(g(x)). The usual proof of the following standard result becomes especially clear by employing De Moivre polynomials.

Proposition 3.6.

The series f(x)=a1x+a2x2+f(x)=a_{1}x+a_{2}x^{2}+\cdots in R[[x]]R[[x]] has a compositional inverse in R[[x]]R[[x]] if and only if a1a_{1} is invertible in RR. This inverse of f(x)f(x) is unique.

Proof.

Suppose a1a_{1} is invertible. To find the compositional inverse, we assume it has the form g(x)=b1x+b2x2+g(x)=b_{1}x+b_{2}x^{2}+\cdots and try to solve for the coefficients using (3.1). We want

1=c1=b1𝒜1,1(a1,a2,)=b1a11=c_{1}=b_{1}\cdot{\mathcal{A}}_{1,1}(a_{1},a_{2},\dots)=b_{1}a_{1}

and so b1=1/a1Rb_{1}=1/a_{1}\in R. If we assume that we have found b1b_{1}, b2,,bm1b_{2},\dots,b_{m-1} in RR already, then bmb_{m} is the solution to

0=cm=k=1m1bk𝒜m,k(a1,a2,)+bm𝒜m,m(a1,a2,).0=c_{m}=\sum_{k=1}^{m-1}b_{k}\cdot{\mathcal{A}}_{m,k}(a_{1},a_{2},\dots)+b_{m}\cdot{\mathcal{A}}_{m,m}(a_{1},a_{2},\dots). (3.15)

With (2.8), 𝒜m,m(a1,a2,)=a1m{\mathcal{A}}_{m,m}(a_{1},a_{2},\dots)=a_{1}^{m}. Since this is again invertible, we see that bmRb_{m}\in R. By induction we obtain gg in R[[x]]R[[x]] so that g(f(x))=xg(f(x))=x. Repeating this argument for gg shows that there exists hh in R[[x]]R[[x]] with h(g(x))=xh(g(x))=x. Then

f(g(x))=h(g(f(g(x))))=h(g(x))=x.f(g(x))=h(g(f(g(x))))=h(g(x))=x.

Therefore gg is a compositional inverse of ff. Uniqueness is proved in the usual way for inverses.

In the other direction, if ff has a compositional inverse g(x)=b1x+b2x2+g(x)=b_{1}x+b_{2}x^{2}+\cdots in R[[x]]R[[x]] then we must have a1b1=1a_{1}b_{1}=1 and so a1a_{1} is invertible. ∎

The relation (3.15) gives a recursive procedure to find the coefficients of the inverse. There is a more direct way to find them though, using Lagrange inversion.

Theorem 3.7 (Lagrange inversion for R[[x]]{R[[x]]}).

Suppose f(x)f(x) and g(x)g(x) are compositional inverses in R[[x]]R[[x]], with neither having a constant term. Then

m[xm]g(x)r=r[xmr](x/f(x))m(m,r1).m[x^{m}]g(x)^{r}=r[x^{m-r}](x/f(x))^{m}\qquad(m,r\in{\mathbb{Z}}_{\geqslant 1}). (3.16)

See [GJ83, Sec. 1.2] or [Ges16] for this result and generalizations. A simple proof involves formal Laurent series and their residues, meaning the coefficients of x1x^{-1}. Note that the right side of (3.16) is in R[[x]]R[[x]] by Proposition 3.2, and using its expansion gives:

Corollary 3.8.

Suppose f(x)=a1x+a2x2+f(x)=a_{1}x+a_{2}x^{2}+\cdots in R[[x]]R[[x]] has compositional inverse b1x+b2x2+b_{1}x+b_{2}x^{2}+\cdots. Then for positive integers mm, rr,

bm=1mk=0m1(mk)a1mk𝒜m1,k(a2,a3,),\displaystyle b_{m}=\frac{1}{m}\sum_{k=0}^{m-1}\binom{-m}{k}a_{1}^{-m-k}{\mathcal{A}}_{m-1,k}(a_{2},a_{3},\dots), (3.17)
𝒜m,r(b1,b2,)=rmk=0mr(mk)a1mk𝒜mr,k(a2,a3,).\displaystyle{\mathcal{A}}_{m,r}(b_{1},b_{2},\dots)=\frac{r}{m}\sum_{k=0}^{m-r}\binom{-m}{k}a_{1}^{-m-k}{\mathcal{A}}_{m-r,k}(a_{2},a_{3},\dots). (3.18)

Recall that RR does not necessarily contain {\mathbb{Q}}. The 1/m1/m terms on the right of (3.17) and (3.18) must cancel an mm factor so that these are statements in RR. We see this cancellation explicitly in section 4.

The next result follows from (3.17), in the same way that Theorem 3.3 followed from Proposition 3.1, and easily extends to derivatives of g(x)rg(x)^{r}. See [Joh02a] for another approach and references.

Corollary 3.9.

Let f(x)f(x) be an m1m\geqslant 1 times differentiable function with compositional inverse g(x)g(x). Then g(x)g(x) is mm times differentiable and

1m!dmdxmg(x)=1mk=0m1(mk)1f(g(x))m+k𝒜m1,k(f′′(g(x))2!,f′′′(g(x))3!,).\frac{1}{m!}\frac{d^{m}}{dx^{m}}g(x)=\frac{1}{m}\sum_{k=0}^{m-1}\binom{-m}{k}\frac{1}{f^{\prime}(g(x))^{m+k}}{\mathcal{A}}_{m-1,k}\mathopen{}\mathclose{{}\left(\frac{f^{\prime\prime}(g(x))}{2!},\frac{f^{\prime\prime\prime}(g(x))}{3!},\dots}\right). (3.19)

A different kind of inversion was needed in [GWLL21, Eqs. (11), (14)]:

Proposition 3.10.

Let a1,a2,a_{1},a_{2},\dots be a sequence of complex numbers and set

bn(t):=k=1ntkk!𝒜n,k(a1,a2,)(n1).b_{n}(t):=\sum_{k=1}^{n}\frac{t^{k}}{k!}{\mathcal{A}}_{n,k}(a_{1},a_{2},\dots)\qquad(n\geqslant 1).

Then these relations may be inverted:

an=1tk=1n(1)k+1k𝒜n,k(b1(t),b2(t),)(n1).a_{n}=\frac{1}{t}\sum_{k=1}^{n}\frac{(-1)^{k+1}}{k}{\mathcal{A}}_{n,k}(b_{1}(t),b_{2}(t),\dots)\qquad(n\geqslant 1). (3.20)
Proof.

By Proposition 3.1, exp(t(a1x+a2x2+))=1+b1(t)x+b2(t)x2+\exp(t(a_{1}x+a_{2}x^{2}+\cdots))=1+b_{1}(t)x+b_{2}(t)x^{2}+\cdots. Then take the log of both sides to obtain (3.20). ∎

Of course, exp(x)\exp(x) and log(x)\log(x) above may be replaced by other pairs of inverse functions; see (6.5) and (6.6) for a different pair. More complicated inversions of this type appear in [BGW19].

4 Greatest common divisor of the coefficients of 𝒜n,k(x1,x2,){\mathcal{A}}_{n,k}(x_{1},x_{2},\dots)

It follows from Corollary 3.8, when a1=1a_{1}=1 and R=R={\mathbb{Z}}, that for integers n,m,k1n,m,k\geqslant 1 we must have

nm(mk)𝒜n,k(x1,x2,)[x1,x2,].\frac{n}{m}\binom{m}{k}{\mathcal{A}}_{n,k}(x_{1},x_{2},\dots)\in{\mathbb{Z}}[x_{1},x_{2},\dots]. (4.1)

For example, with n=9n=9 and k=4k=4 this means that all the coefficients of 𝒜9,4{\mathcal{A}}_{9,4} must be divisible by 44. The next result, combined with (4.3) below, may be used to prove (4.1) directly.

Theorem 4.1.

For nk1n\geqslant k\geqslant 1, the gcd\gcd of the coefficients of 𝒜n,k{\mathcal{A}}_{n,k} is k/gcd(n,k)k/\gcd(n,k).

Proof.

Let Δ\Delta be the gcd\gcd of the coefficients of 𝒜n,k{\mathcal{A}}_{n,k}. We may assume that n>kn>k as the theorem is clearly true when n=kn=k by (2.8). Recall that the coefficients of 𝒜n,k{\mathcal{A}}_{n,k} are the multinomial coefficients

(kj1,j2,,jm)with1j1+2j2++mjm=n.\binom{k}{j_{1},j_{2},\dots,j_{m}}\quad\text{with}\quad 1j_{1}+2j_{2}+\dots+mj_{m}=n. (4.2)

Taking j1=k1j_{1}=k-1 and jnk+1=1j_{n-k+1}=1 gives the coefficient kk and so Δk\Delta\mid k. Our next goal is to show that (k/gcd(n,k))Δ(k/\gcd(n,k))\mid\Delta.

The elegant argument in [GS95] for binomial coefficients adapts nicely to multinomial coefficients as follows. It is evident that when

C:=(kj1,j2,,jm)we havejrkC=(k1j1,,jr1,,jm)C:=\binom{k}{j_{1},j_{2},\dots,j_{m}}\quad\text{we have}\quad\frac{j_{r}}{k}C=\binom{k-1}{j_{1},\dots,j_{r}-1,\dots,j_{m}}

for 1rm1\leqslant r\leqslant m with jr0j_{r}\neq 0. Therefore Cr:=jrC/kC_{r}:=j_{r}C/k is always an integer for jr0j_{r}\geqslant 0 and so

kgcd(C,C1,C2,,Cm)\displaystyle k\cdot\gcd\mathopen{}\mathclose{{}\left(C,C_{1},C_{2},\dots,C_{m}}\right) =gcd(kC,kC1,kC2,,kCm)\displaystyle=\gcd\mathopen{}\mathclose{{}\left(kC,kC_{1},kC_{2},\dots,kC_{m}}\right)
=gcd(kC,j1C,j2C,,jmC)\displaystyle=\gcd\mathopen{}\mathclose{{}\left(kC,j_{1}C,j_{2}C,\dots,j_{m}C}\right)
=Cgcd(k,j1,j2,,jm).\displaystyle=C\gcd(k,j_{1},j_{2},\dots,j_{m}).

Hence, for any multinomial coefficient,

kgcd(k,j1,j2,,jm)divides(kj1,j2,,jm).\frac{k}{\gcd(k,j_{1},j_{2},\dots,j_{m})}\quad\text{divides}\quad\binom{k}{j_{1},j_{2},\dots,j_{m}}. (4.3)

Now gcd(j1,j2,,jm)n\gcd(j_{1},j_{2},\dots,j_{m})\mid n by the right side of (4.2) and it follows that k/gcd(n,k)k/\gcd(n,k) divides all the coefficients of 𝒜n,k{\mathcal{A}}_{n,k}.

So far we have demonstrated that (k/gcd(n,k))Δ(k/\gcd(n,k))\mid\Delta and Δk\Delta\mid k. The next lemma will let us find coefficients that ensure Δ(k/gcd(n,k))\Delta\mid(k/\gcd(n,k)).

Lemma 4.2.

Let α\alpha and β\beta be positive integers. Then

gcd((αβd1),(αβd2),,(αβdr))dividesα\gcd\mathopen{}\mathclose{{}\left(\binom{\alpha\beta}{d_{1}},\binom{\alpha\beta}{d_{2}},\dots,\binom{\alpha\beta}{d_{r}}}\right)\quad\text{divides}\quad\alpha (4.4)

where d1,d2,,drd_{1},d_{2},\dots,d_{r} are all the divisors of β\beta.

Proof.

For pp any prime, let νp\nu_{p} denote the pp-adic valuation. Set x:=νp(α)x:=\nu_{p}(\alpha), y:=νp(β)y:=\nu_{p}(\beta) so that α=αpx\alpha=\alpha^{\prime}p^{x} and β=βpy\beta=\beta^{\prime}p^{y} for α\alpha^{\prime}, β\beta^{\prime} prime to pp. We want to establish

νp((αβpy))=x,\nu_{p}\mathopen{}\mathclose{{}\left(\binom{\alpha\beta}{p^{y}}}\right)=x, (4.5)

since that will show that the largest power of pp dividing the left side of (4.4) is the power of pp that divides α\alpha.

Use the notation sp(n)s_{p}(n) to mean the sum of the base pp digits of nn. Legendre’s well-known formula implies, as in [SMA09, Eq. (1.6)],

νp((αβpy))=sp(py)+sp(αβpy)sp(αβ)p1.\nu_{p}\mathopen{}\mathclose{{}\left(\binom{\alpha\beta}{p^{y}}}\right)=\frac{s_{p}(p^{y})+s_{p}(\alpha\beta-p^{y})-s_{p}(\alpha\beta)}{p-1}.

The numerator on the right equals

sp(py)+sp((αβpx1)py)sp(αβpx+y)=1+sp(αβpx1)sp(αβ).s_{p}(p^{y})+s_{p}((\alpha^{\prime}\beta^{\prime}p^{x}-1)p^{y})-s_{p}(\alpha^{\prime}\beta^{\prime}p^{x+y})=1+s_{p}(\alpha^{\prime}\beta^{\prime}p^{x}-1)-s_{p}(\alpha^{\prime}\beta^{\prime}).

If αβ\alpha^{\prime}\beta^{\prime} has the base pp representation c1c2ckc_{1}c_{2}\cdots c_{k} then ck0c_{k}\neq 0 and αβpx\alpha^{\prime}\beta^{\prime}p^{x} has the same digits with xx zeros added on the right. Clearly αβpx1\alpha^{\prime}\beta^{\prime}p^{x}-1 will have the representation

c1c2(ck1)(p1)(p1),c_{1}c_{2}\cdots(c_{k}-1)(p-1)\cdots(p-1),

so that sp(αβpx1)=sp(αβ)1+x(p1)s_{p}(\alpha^{\prime}\beta^{\prime}p^{x}-1)=s_{p}(\alpha^{\prime}\beta^{\prime})-1+x(p-1). Then (4.5) follows and the lemma is proved. ∎

Now let dd be any divisor of gcd(n,k)\gcd(n,k). We consider coefficients in (4.2) where only two numbers j1j_{1} and jrj_{r} are nonzero, choosing r:=(nk)/d+1>1r:=(n-k)/d+1>1. Check that j1=kdj_{1}=k-d and jr=dj_{r}=d gives 1j1+rjr=n1j_{1}+rj_{r}=n and so

(kd)fordgcd(n,k)\binom{k}{d}\quad\text{for}\quad d\mid\gcd(n,k) (4.6)

is a coefficient of 𝒜n,k{\mathcal{A}}_{n,k}. Apply Lemma 4.2 with α=k/gcd(n,k)\alpha=k/\gcd(n,k) and β=gcd(n,k)\beta=\gcd(n,k) to see that the gcd\gcd of the coefficients of the form (4.6) must divide k/gcd(n,k)k/\gcd(n,k). Therefore Δ(k/gcd(n,k))\Delta\mid(k/\gcd(n,k)) as we wanted. This finishes the proof of Theorem 4.1. ∎

By comparison, the gcd\gcds of the coefficients of the partial Bell polynomials are much larger. It may be shown for example that k+3,k{\mathcal{B}}_{k+3,k} has gcd\gcd equalling (k+34)\binom{k+3}{4}.

5 Determinant formulas

Define the n×nn\times n matrix

n(t):=(a1t1a2ta1t1antan1ta1t),\mathcal{M}_{n}(t):=\mathopen{}\mathclose{{}\left(\begin{matrix}a_{1}t&1&&\\ a_{2}t&a_{1}t&1&\\ \vdots&&\ddots&\\ a_{n}t&a_{n-1}t&\dots&a_{1}t\end{matrix}}\right),

with ones above the main diagonal and zeros above those. The next result is a slightly simplified version of [EJ, Thm. 3.1].

Proposition 5.1.

We have

k=0n(1)n+ktk𝒜n,k(a1,a2,)=detn(t).\sum_{k=0}^{n}(-1)^{n+k}t^{k}{\mathcal{A}}_{n,k}(a_{1},a_{2},\dots)=\det\mathcal{M}_{n}(t). (5.1)
Proof.

By (2.18), tk𝒜n,k(a1,a2,)=𝒜n,k(a1t,a2t,)t^{k}{\mathcal{A}}_{n,k}(a_{1},a_{2},\dots)={\mathcal{A}}_{n,k}(a_{1}t,a_{2}t,\dots) and so it is enough to prove (5.1) for t=1t=1. An induction proof is possible since repeatedly expanding the determinant along the top row gives

detn(1)=a1detn1(1)a2detn2(1)++(1)n1an.\det\mathcal{M}_{n}(1)=a_{1}\det\mathcal{M}_{n-1}(1)-a_{2}\det\mathcal{M}_{n-2}(1)+\cdots+(-1)^{n-1}a_{n}.

A more illuminating proof uses (3.8). For un:=k=0n(1)k𝒜n,k(a1,a2,)u_{n}:=\sum_{k=0}^{n}(-1)^{k}{\mathcal{A}}_{n,k}(a_{1},a_{2},\dots) we obtain

(1+a1x+a2x2+)(1+u1x+u2x2+)=1.(1+a_{1}x+a_{2}x^{2}+\cdots)(1+u_{1}x+u_{2}x^{2}+\cdots)=1.

Therefore

(1a11a2a11an1an2a11)(u1u2u3un)=(a1a2a3an)\mathopen{}\mathclose{{}\left(\begin{matrix}1&&&&\\ a_{1}&1&&&\\ a_{2}&a_{1}&1&&\\ \vdots&&\ddots&&\\ a_{n-1}&a_{n-2}&\dots&a_{1}&1\end{matrix}}\right)\mathopen{}\mathclose{{}\left(\begin{matrix}u_{1}\\ u_{2}\\ u_{3}\\ \vdots\\ u_{n}\end{matrix}}\right)=\mathopen{}\mathclose{{}\left(\begin{matrix}-a_{1}\\ -a_{2}\\ -a_{3}\\ \vdots\\ -a_{n}\end{matrix}}\right) (5.2)

and, by Cramer’s rule, un=detUn/detUu_{n}=\det U_{n}/\det U where UU is the matrix on the left of (5.2) and UnU_{n} is UU with its nnth column replaced by the column on the right side of (5.2). Moving this nnth column of UnU_{n} to the left side and changing its sign introduces a (1)n(-1)^{n} factor. ∎

The 𝒜n,k{\mathcal{A}}_{n,k} polynomials can be isolated in (5.1) since clearly

𝒜n,k(a1,a2,)=(1)n+k[tk]detn(t).{\mathcal{A}}_{n,k}(a_{1},a_{2},\dots)=(-1)^{n+k}[t^{k}]\det\mathcal{M}_{n}(t).

We may also give exponential and logarithmic versions of Proposition 5.1. Define the matrices

𝒩n(t):=(a1t1a2ta1t2a3ta2ta1t3antan1ta2ta1t),𝒪n(t):=(a1t12a2ta1t13a3ta2ta1t1nantan1ta2ta1t),\mathcal{N}_{n}(t):=\mathopen{}\mathclose{{}\left(\begin{matrix}a_{1}t&1&&&\\ a_{2}t&a_{1}t&2&&\\ a_{3}t&a_{2}t&a_{1}t&3&\\ \vdots&&&\ddots&\\ a_{n}t&a_{n-1}t&\dots&a_{2}t&a_{1}t\end{matrix}}\right),\qquad\mathcal{O}_{n}(t):=\mathopen{}\mathclose{{}\left(\begin{matrix}a_{1}t&1&&&\\ 2a_{2}t&a_{1}t&1&&\\ 3a_{3}t&a_{2}t&a_{1}t&1&\\ \vdots&&&\ddots&\\ na_{n}t&a_{n-1}t&\dots&a_{2}t&a_{1}t\end{matrix}}\right),

which are the same as n(t)\mathcal{M}_{n}(t) except for multiplying by a factor jj on row jj, above the main diagonal in 𝒩n(t)\mathcal{N}_{n}(t) and in the first column for 𝒪n(t)\mathcal{O}_{n}(t).

Proposition 5.2.

We have

k=0n(1)n+ktkk!𝒜n,k(a11,a22,)\displaystyle\sum_{k=0}^{n}(-1)^{n+k}\frac{t^{k}}{k!}{\mathcal{A}}_{n,k}\mathopen{}\mathclose{{}\left(\frac{a_{1}}{1},\frac{a_{2}}{2},\dots}\right) =1n!det𝒩n(t),\displaystyle=\frac{1}{n!}\det\mathcal{N}_{n}(t), (5.3)
k=0n(1)n+ktkk𝒜n,k(a1,a2,)\displaystyle\sum_{k=0}^{n}(-1)^{n+k}\frac{t^{k}}{k}{\mathcal{A}}_{n,k}\mathopen{}\mathclose{{}\left(a_{1},a_{2},\dots}\right) =1ndet𝒪n(t).\displaystyle=\frac{1}{n}\det\mathcal{O}_{n}(t). (5.4)
Proof.

To prove (5.3) we use (3.3) and it is convenient to set αj:=aj/j\alpha_{j}:=-a_{j}/j and f(x)=α1x+α2x2+f(x)=\alpha_{1}x+\alpha_{2}x^{2}+\cdots. For vn:=k=0n1k!𝒜n,k(α1,α2,)v_{n}:=\sum_{k=0}^{n}\frac{1}{k!}{\mathcal{A}}_{n,k}(\alpha_{1},\alpha_{2},\dots) we obtain

ef(x)=n=0vnxn,ef(x)f(x)=n=1nvnxn1.e^{f(x)}=\sum_{n=0}^{\infty}v_{n}x^{n},\qquad e^{f(x)}f^{\prime}(x)=\sum_{n=1}^{\infty}n\cdot v_{n}x^{n-1}.

It follows that

(1+v1x+v2x2+)(α1+2α2x+3α3x2+)=(v1+2v2x+3v3x2+),(1+v_{1}x+v_{2}x^{2}+\cdots)(\alpha_{1}+2\alpha_{2}x+3\alpha_{3}x^{2}+\cdots)=(v_{1}+2v_{2}x+3v_{3}x^{2}+\cdots),

and hence

(1a12a2a13an1an2a1n)(v1v2v3vn)=(a1a2a3an).\mathopen{}\mathclose{{}\left(\begin{matrix}1&&&&\\ a_{1}&2&&&\\ a_{2}&a_{1}&3&&\\ \vdots&&\ddots&&\\ a_{n-1}&a_{n-2}&\dots&a_{1}&n\end{matrix}}\right)\mathopen{}\mathclose{{}\left(\begin{matrix}v_{1}\\ v_{2}\\ v_{3}\\ \vdots\\ v_{n}\end{matrix}}\right)=\mathopen{}\mathclose{{}\left(\begin{matrix}-a_{1}\\ -a_{2}\\ -a_{3}\\ \vdots\\ -a_{n}\end{matrix}}\right). (5.5)

The proof of (5.3) is now completed using the same steps as in Proposition 5.1. The proof of (5.4) is similar, based on (3.4). ∎

6 Generating functions

Some generating functions that lead to interesting De Moivre expressions are shown next. See for example [Com74, Sect. 1.14] and [GKP94, Chapters 6, 7] for more on generating functions.

6.1 Partitions

A partition of a positive integer nn is a non-increasing sequence of positive integers that sum to nn. If pk(n)p_{k}(n) denotes the number of partitions of nn with at most kk parts, then clearly

pk(n)=1j1+2j2++njn=nj1+j2++jnk1p_{k}(n)=\sum_{\begin{subarray}{c}1j_{1}+2j_{2}+\dots+nj_{n}=n\\ j_{1}+j_{2}+\dots+j_{n}\leqslant k\end{subarray}}1 (6.1)

where the sum is over all possible j1j_{1}, j2j_{2}, …, jn0j_{n}\in{\mathbb{Z}}_{\geqslant 0} and jrj_{r} indicates the number of parts of size rr. Comparing (6.1) with (1.1) confirms the statement in Definition 1.1 that the number of terms in 𝒜n,k{\mathcal{A}}_{n,k} is the number of partitions of nn with exactly kk parts. This is pk(n)pk1(n)p_{k}(n)-p_{k-1}(n) and equals pk(nk)p_{k}(n-k) as seen from the generating function

n=0pk(n)qn=j=1k11qj.\sum_{n=0}^{\infty}p_{k}(n)q^{n}=\prod_{j=1}^{k}\frac{1}{1-q^{j}}. (6.2)

Perhaps the most elegant formulas for pk(n)p_{k}(n) use Sylvester’s theory of waves from 1857 and quasipolynomials, as described in [O’S18]. With (3.8) we can produce the less enlightening

p4(n)=k=0n𝒜n,k(1,1,0,0,2,0,0,1,1,1,),p_{4}(n)=\sum_{k=0}^{n}{\mathcal{A}}_{n,k}(1,1,0,0,-2,0,0,1,1,-1,\dots),

for example, where the rest of the sequence consists of zeros.

Let p(n)p(n) denote the number of unrestricted partitions of nn. Their generating function is (6.2) with k=k=\infty. As we saw above, the number of terms in 𝒜m+k,k{\mathcal{A}}_{m+k,k} is pk(m)p_{k}(m). This equals p(m)p(m) for kmk\geqslant m and the terms correspond to solutions of 1j2+2j3++mjm+1=m1j_{2}+2j_{3}+\cdots+mj_{m+1}=m in (1.2). De Moivre’s description in [DM97] essentially gives what must be one of the earliest recursive constructions of the partitions of an integer mm: they are 11 followed by the partitions of m1m-1, then 22 followed by the partitions of m2m-2 with smallest part at least 22, then 33 followed by the partitions of m3m-3 with smallest part at least 33, and so on as far as m/2\lfloor m/2\rfloor. Lastly include mm.

To find a first expression for p(n)p(n) we may use Euler’s pentagonal number theorem,

j=1(1qj)=r(1)rqr(3r1)/2=m=0cmqm,\prod_{j=1}^{\infty}(1-q^{j})=\sum_{r\in{\mathbb{Z}}}(-1)^{r}q^{r(3r-1)/2}=\sum_{m=0}^{\infty}c_{m}q^{m},

with cm=(1)rc_{m}=(-1)^{r} if m=r(3r1)/2m=r(3r-1)/2 for some rr\in{\mathbb{Z}} and cm=0c_{m}=0 otherwise. Therefore, employing (3.8) again,

p(n)=k=0n𝒜n,k(c1,c2,c3,)=k=0n𝒜n,k(1,1,0,0,1,0,1,).p(n)=\sum_{k=0}^{n}{\mathcal{A}}_{n,k}(-c_{1},-c_{2},-c_{3},\dots)=\sum_{k=0}^{n}{\mathcal{A}}_{n,k}(1,1,0,0,-1,0,-1,\dots).

Ramanujan’s tau function has the related generating function

n=1τ(n)qn=qj=1(1qj)24\sum_{n=1}^{\infty}\tau(n)q^{n}=q\prod_{j=1}^{\infty}(1-q^{j})^{24} (6.3)

where the right side of (6.3) is a modular form when q=e2πizq=e^{2\pi iz}, (a weight 1212 cusp form for SL2()\mathrm{SL}_{2}({\mathbb{Z}})). Then easily,

τ(n)=𝒜n+23,24(c0,c1,c2,)=𝒜n+23,24(1,1,1,0,0,1,0,)\tau(n)={\mathcal{A}}_{n+23,24}(c_{0},c_{1},c_{2},\dots)={\mathcal{A}}_{n+23,24}(1,-1,-1,0,0,1,0,\dots) (6.4)

and Theorem 4.1 implies that τ(n)\tau(n) is divisible by 24/gcd(n+23,24)24/\gcd(n+23,24). The congruence properties of τ(n)\tau(n) and p(n)p(n) have been of great interest, and Lehmer’s 1947 question as to whether τ(n)\tau(n) is ever zero remains unanswered. By (3.2) we also have the relations

τ(n)\displaystyle\tau(n) =k=0n1(24k)𝒜n1,k(p(1),p(2),p(3),),\displaystyle=\sum_{k=0}^{n-1}\binom{-24}{k}{\mathcal{A}}_{n-1,k}(p(1),p(2),p(3),\dots), (6.5)
p(n)\displaystyle p(n) =k=0n(1/24k)𝒜n,k(τ(2),τ(3),τ(4),).\displaystyle=\sum_{k=0}^{n}\binom{-1/24}{k}{\mathcal{A}}_{n,k}(\tau(2),\tau(3),\tau(4),\dots). (6.6)

Set σ(n):=d|nd\sigma(n):=\sum_{d|n}d, the sum of the divisors of nn. For another example, take the log of the right side of (6.2) when k=k=\infty, expand this into a series and then exponentiate to produce

p(n)=k=0n1k!𝒜n,k(σ(1)1,σ(2)2,σ(3)3,),p(n)=\sum_{k=0}^{n}\frac{1}{k!}{\mathcal{A}}_{n,k}\mathopen{}\mathclose{{}\left(\frac{\sigma(1)}{1},\frac{\sigma(2)}{2},\frac{\sigma(3)}{3},\dots}\right), (6.7)

which may also be inverted. The identity (6.7) comes from [Rio68, p. 185], and for further identities along these lines see [Jha].

The authors in [BKO04, Thm. 3] develop a universal recursion for the Fourier coefficients of modular forms. The required polynomials are in fact De Moivre polynomials, and as a special case (see their remark after the theorem) we find:

τ(n+1)=24σ(n)n+k=2n(1)kk𝒜n,k(τ(2),τ(3),)(n1).\tau(n+1)=-\frac{24\sigma(n)}{n}+\sum_{k=2}^{n}\frac{(-1)^{k}}{k}{\mathcal{A}}_{n,k}(\tau(2),\tau(3),\dots)\qquad(n\geqslant 1). (6.8)

The form of (6.8) now suggests it has an easy proof: take the log of both sides of (6.3) and expand each series.

6.2 Orthogonal polynomials

Bell in [Bel34] was initially interested in generalizing the Hermite polynomials Hn(x)H_{n}(x). They may be expressed as n![tn]exp(2xtt2)n![t^{n}]\exp(2xt-t^{2}) and in Bell’s original polynomials equal 𝒴n(2x,2,0,0,){\mathcal{Y}}_{n}(2x,-2,0,0,\dots). In our notation

Hn(x)=k=0nn!k!𝒜n,k(2x,1,0,0,).H_{n}(x)=\sum_{k=0}^{n}\frac{n!}{k!}{\mathcal{A}}_{n,k}(2x,-1,0,0,\dots). (6.9)

Some other classical orthogonal polynomials have similar descriptions, as seen in [Cha02, pp. 449 - 452]. For example, the Gegenbauer polynomials may be defined with Cn(λ)(x):=[tn](12tx+t2)λC^{(\lambda)}_{n}(x):=[t^{n}](1-2tx+t^{2})^{-\lambda} giving

Cn(λ)(x)=k=0n(k+λ1k)𝒜n,k(2x,1,0,0,)C^{(\lambda)}_{n}(x)=\sum_{k=0}^{n}\binom{k+\lambda-1}{k}{\mathcal{A}}_{n,k}(2x,-1,0,0,\dots) (6.10)

by (3.2). Special cases are the Legendre polynomials Pn(x)P_{n}(x) with λ=1/2\lambda=1/2, the Chebyshev polynomials of the second kind Un(x)U_{n}(x) with λ=1\lambda=1, and the usual Chebyshev polynomials Tn(x)T_{n}(x) which may found as a limit when λ0\lambda\to 0. More explicitly, use Tn(x)=[tn](1tx)/(12tx+t2)T_{n}(x)=[t^{n}](1-tx)/(1-2tx+t^{2}) to see that

Tn(x)=k=0n[𝒜n,k(2x,1,0,0,)x𝒜n1,k(2x,1,0,0,)].T_{n}(x)=\sum_{k=0}^{n}\Bigl{[}{\mathcal{A}}_{n,k}(2x,-1,0,0,\dots)-x\cdot{\mathcal{A}}_{n-1,k}(2x,-1,0,0,\dots)\Bigr{]}. (6.11)

The Fibonacci numbers FnF_{n} equal [tn]t/(1tt2)[t^{n}]t/(1-t-t^{2}) and fit the same pattern with

Fn+1=k=0n𝒜n,k(1,1,0,0,).F_{n+1}=\sum_{k=0}^{n}{\mathcal{A}}_{n,k}(1,1,0,0,\dots). (6.12)

The identity

𝒜n,k(x,y,0,0,)=(knk)x2knynk(nk){\mathcal{A}}_{n,k}(x,y,0,0,\dots)=\binom{k}{n-k}x^{2k-n}y^{n-k}\qquad(n\geqslant k) (6.13)

may also be employed in (6.9) – (6.12). For example, using (6.13) in (6.11) yields

Tn(x)=k=1n(1)nkn2k(knk)(2x)2kn(n1).T_{n}(x)=\sum_{k=1}^{n}(-1)^{n-k}\frac{n}{2k}\binom{k}{n-k}(2x)^{2k-n}\qquad(n\geqslant 1). (6.14)

In view of the usual definition of TnT_{n} by Tn(cosθ)=cosnθT_{n}(\cos\theta)=\cos n\theta, and De Moivre’s formula for the nnth power of a complex number, it is not surprising that the polynomial (6.14) already appears in the work of De Moivre, predating Chebyshev’s introduction of Tn(x)T_{n}(x) by over 100 years. Consult [Sch68, p. 246] and [Gir] for further details.

6.3 Bernoulli numbers and polynomials

The Bernoulli numbers are usually defined by Bn:=n![tn]t/(et1)B_{n}:=n![t^{n}]t/(e^{t}-1). Another application of (3.8) yields

Bn=n!k=0n𝒜n,k(12!,13!,14!,).B_{n}=n!\sum_{k=0}^{n}{\mathcal{A}}_{n,k}\mathopen{}\mathclose{{}\left(-\frac{1}{2!},-\frac{1}{3!},-\frac{1}{4!},\dots}\right). (6.15)

Stern gave further similar expressions for BnB_{n} in [Ste43], employing the elaborate notation C𝑘𝔭n{}_{\mathfrak{p}}^{n}\overset{k}{C^{\prime}} for 𝒜n,k{\mathcal{A}}_{n,k}. Using the well-known power series for tan\tan and arctan\arctan in Corollary 3.8 also gives the alternative

2n+2(2n+21)Bn+2n+2=n!k=0n(n+kk)𝒜n,k(0,13,0,15,0,).2^{n+2}(2^{n+2}-1)\frac{B_{n+2}}{n+2}=n!\sum_{k=0}^{n}\binom{n+k}{k}{\mathcal{A}}_{n,k}\mathopen{}\mathclose{{}\left(0,-\frac{1}{3},0,-\frac{1}{5},0,\dots}\right). (6.16)

Extending (6.15) to the Bernoulli polynomials Bn(x):=n![tn]tetx/(et1)B_{n}(x):=n![t^{n}]te^{tx}/(e^{t}-1) results in

Bn(x)=n!k=0n𝒜n,k((x)2(1x)22!,(x)3(1x)33!,).B_{n}(x)=n!\sum_{k=0}^{n}{\mathcal{A}}_{n,k}\mathopen{}\mathclose{{}\left(\frac{(-x)^{2}-(1-x)^{2}}{2!},\frac{(-x)^{3}-(1-x)^{3}}{3!},\dots}\right). (6.17)

It may be seen from (6.17) that Bn(x)B_{n}(x) is a polynomial in xx of degree nn and satisfies Bn(1x)=(1)nBn(x)B_{n}(1-x)=(-1)^{n}B_{n}(x) by (2.19).

The Nörlund polynomials Bn(x):=n![tn](t/(et1))xB_{n}^{(x)}:=n![t^{n}](t/(e^{t}-1))^{x} generalize the Bernoulli numbers in another direction. Equation (3.2) finds

Bn(x)=n!k=0n(xk)𝒜n,k(12!,13!,14!,),B_{n}^{(x)}=n!\sum_{k=0}^{n}\binom{-x}{k}{\mathcal{A}}_{n,k}\mathopen{}\mathclose{{}\left(\frac{1}{2!},\frac{1}{3!},\frac{1}{4!},\dots}\right), (6.18)

making it more obvious by (2.13) that Bn(x)B_{n}^{(x)} is also a degree nn polynomial in xx. Another expression is proved in [ST88, Eq. 15]:

Bn(x)=k=0n(xk)(x+nnk)(n+kk)1{n+kk}.B_{n}^{(x)}=\sum_{k=0}^{n}\binom{-x}{k}\binom{x+n}{n-k}\binom{n+k}{k}^{-1}{\genfrac{\{}{\}}{0.0pt}{}{n+k}{k}}. (6.19)

6.4 Cyclotomic polynomials

Ramanujan sums and cyclotomic polynomials may be expressed in terms of the Möbius function μ\mu with

rj(m)=d(m,j)μ(m/d)d,Φn(x)=dn(1xd)μ(n/d)(n2),r_{j}(m)=\sum_{d\mid(m,j)}\mu(m/d)\cdot d,\qquad\Phi_{n}(x)=\prod_{d\mid n}\mathopen{}\mathclose{{}\left(1-x^{d}}\right)^{\mu(n/d)}\quad(n\geqslant 2),

respectively. The authors in [HPM21, Thm. 4.2] write Lehmer’s 1966 formula for Φn(x)\Phi_{n}(x) in an appealing way using Bell polynomials. It is even more transparent without the unnecessary factorials:

Φn(x)=m=0[k=0m1k!𝒜m,k(r1(n)1,r2(n)2,r3(n)3,)]xm,\Phi_{n}(x)=\sum_{m=0}^{\infty}\mathopen{}\mathclose{{}\left[\sum_{k=0}^{m}\frac{1}{k!}{\mathcal{A}}_{m,k}\mathopen{}\mathclose{{}\left(-\frac{r_{1}(n)}{1},-\frac{r_{2}(n)}{2},-\frac{r_{3}(n)}{3},\dots}\right)}\right]x^{m}, (6.20)

for n2n\geqslant 2, and the easy proof is similar to that of (6.7).

7 Asymptotic expansions

In this final section we discuss examples where the De Moivre polynomials play a useful role in describing the behavior of functions and sequences as a parameter goes to infinity.

7.1 Partition asymptotics

Hardy and Ramanujan famously gave the first asymptotic expansion for the partition function p(n)p(n) in 1918. Rademacher later showed how a small alteration turned this into a rapidly converging series: [Rad73, Eq. (128.1)]. A simpler, though less accurate, expansion for p(n)p(n) takes the form

p(n)=143neπ2n/3(1+C1n1/2+C2n2/2++CR1n(R1)/2+O(1nR/2))p(n)=\frac{1}{4\sqrt{3}n}e^{\pi\sqrt{2n/3}}\mathopen{}\mathclose{{}\left(1+\frac{{C}_{1}}{n^{1/2}}+\frac{{C}_{2}}{n^{2/2}}+\cdots+\frac{{C}_{R-1}}{n^{(R-1)/2}}+O\mathopen{}\mathclose{{}\left(\frac{1}{n^{R/2}}}\right)}\right) (7.1)

showing the main term along with smaller corrections. The constants Cr{C}_{r} are quite complicated, but we may give them a reasonable description. First define

αj(x)\displaystyle\alpha_{j}(x) :={24j/2 for j even24(1j)/21x(j/2(j1)/2) for j odd,\displaystyle:=\begin{cases}24^{-j/2}&\text{ for $j$ even}\\ \displaystyle-24^{(1-j)/2}\frac{1}{x}\binom{j/2}{(j-1)/2}&\text{ for $j$ odd},\end{cases} (7.2)
β(x)\displaystyle\beta_{\ell}(x) :=/2m(24)mx2m(2m)!𝒜m,2m((1/21),(1/22),(1/23),).\displaystyle:=\sum_{\ell/2\leqslant m\leqslant\ell}(-24)^{-m}\frac{x^{2m-\ell}}{(2m-\ell)!}{\mathcal{A}}_{m,2m-\ell}\mathopen{}\mathclose{{}\left(\binom{1/2}{1},\binom{1/2}{2},\binom{1/2}{3},\dots}\right). (7.3)
Proposition 7.1.

Fix R1R\geqslant 1. Let Cr{C}_{r} be defined by j=0rαj(x)βrj(x)\sum_{j=0}^{r}\alpha_{j}(x)\cdot\beta_{r-j}(x) with x=π2/3x=\pi\sqrt{2/3}. For these values, (7.1) is true as nn\to\infty with an implied constant depending only on RR.

Proof.

Using the first term in his expansion, Rademacher shows

p(n)=143nκn(11xnκn)exnκn(1+O(exn/2))p(n)=\frac{1}{4\sqrt{3}n\cdot\kappa_{n}}\mathopen{}\mathclose{{}\left(1-\frac{1}{x\sqrt{n\cdot\kappa_{n}}}}\right)e^{x\sqrt{n\cdot\kappa_{n}}}\mathopen{}\mathclose{{}\left(1+O(e^{-x\sqrt{n}/2})}\right) (7.4)

for κn:=11/(24n)\kappa_{n}:=1-1/(24n) in [Rad73, p. 278]. Put z:=1/(24n)z:=-1/(24n) and write this main term as

exn43n11+z(11xn(1+z))exn(1+z1).\frac{e^{x\sqrt{n}}}{4\sqrt{3}n}\cdot\frac{1}{1+z}\mathopen{}\mathclose{{}\left(1-\frac{1}{x\sqrt{n(1+z)}}}\right)\cdot e^{x\sqrt{n}(\sqrt{1+z}-1)}.

Treating zz as an independent complex variable, the middle factors are analytic for |z|<1|z|<1 and have a Taylor expansion. Using the usual bounds on the remainder shows

11+z(11xn(1+z))=j=0J1aj(n)zj+O(|z|J)(|z|1/2).\frac{1}{1+z}\mathopen{}\mathclose{{}\left(1-\frac{1}{x\sqrt{n(1+z)}}}\right)=\sum_{j=0}^{J-1}a_{j}(\sqrt{n})\cdot z^{j}+O(|z|^{J})\qquad(|z|\leqslant 1/2).

By the binomial theorem we find aj(n)=(1)j(3/2j)/(xn)a_{j}(\sqrt{n})=(-1)^{j}-\binom{-3/2}{j}/(x\sqrt{n}) and hence, with (7.2),

1κn(11xnκn)=j=0J1αj(x)nj/2+O(nJ/2).\frac{1}{\kappa_{n}}\mathopen{}\mathclose{{}\left(1-\frac{1}{x\sqrt{n\cdot\kappa_{n}}}}\right)=\sum_{j=0}^{J-1}\alpha_{j}(x)\cdot n^{-j/2}+O(n^{-J/2}). (7.5)

Similarly to (7.5),

exn(1+z1)==0L1b(n)z+O(|z|L)(|z|1/n).e^{x\sqrt{n}(\sqrt{1+z}-1)}=\sum_{\ell=0}^{L-1}b_{\ell}(\sqrt{n})\cdot z^{\ell}+O(|z|^{L})\qquad(|z|\leqslant 1/\sqrt{n}).

Then the expansions

exn(1+z1)\displaystyle e^{x\sqrt{n}(\sqrt{1+z}-1)} =exp(xnj=1(1/2j)zj)\displaystyle=\exp\biggl{(}x\sqrt{n}\sum_{j=1}^{\infty}\binom{1/2}{j}z^{j}\biggr{)}
=k=0(xn)kk!m=k𝒜m,k((1/21),(1/22),(1/23),)zm\displaystyle=\sum_{k=0}^{\infty}\frac{(x\sqrt{n})^{k}}{k!}\sum_{m=k}^{\infty}{\mathcal{A}}_{m,k}\mathopen{}\mathclose{{}\left(\binom{1/2}{1},\binom{1/2}{2},\binom{1/2}{3},\dots}\right)z^{m}

imply that

b(n)=k=0(xn)kk!𝒜,k((1/21),(1/22),(1/23),).b_{\ell}(\sqrt{n})=\sum_{k=0}^{\ell}\frac{(x\sqrt{n})^{k}}{k!}{\mathcal{A}}_{\ell,k}\mathopen{}\mathclose{{}\left(\binom{1/2}{1},\binom{1/2}{2},\binom{1/2}{3},\dots}\right).

After rearranging we finally obtain, with (7.3),

exn(κn1)==0L1β(x)n/2+O(nL/2).e^{x\sqrt{n}(\sqrt{\kappa_{n}}-1)}=\sum_{\ell=0}^{L-1}\beta_{\ell}(x)\cdot n^{-\ell/2}+O(n^{-L/2}). (7.6)

Use (7.5) and (7.6) in (7.4) to complete the proof. ∎

For example,

C1=72+π2246π0.443288,C2=432+π269120.0639279.{C}_{1}=-\frac{72+\pi^{2}}{24\sqrt{6}\pi}\approx-0.443288,\qquad{C}_{2}=\frac{432+\pi^{2}}{6912}\approx 0.0639279.

We see from (7.2), (7.3) that CrC_{r} is 1/π1/\pi times a polynomial in π\pi of degree r+1r+1, and so we cannot expect much simplification. Of course (7.4) is more accurate than (7.1), and including more terms of Rademacher’s series is better still. The numbers p(n)p(n) are examples of Fourier coefficients of weakly holomorphic modular forms and Rademacher type series exist for all of these. In the next examples though, and many other situations, expansions of the form (7.1) are the best available.

7.2 Laplace’s method

A simple example of Laplace’s method, (and where it overlaps with the saddle-point method), is the following. Suppose that f(z)f(z) and g(z)g(z) are holomorphic functions on a domain containing the interval [1,1][-1,1] and real-valued on this interval. The integral

I(n):=11enf(z)g(z)𝑑zI(n):=\int_{-1}^{1}e^{n\cdot f(z)}g(z)\,dz (7.7)

can be accurately estimated for large nn\in{\mathbb{R}} if f(z)f^{\prime}(z) has a simple zero at z=0z=0 (called the saddle-point) and f(z)<f(0)f(z)<f(0) for all non-zero z[1,1]z\in[-1,1]. In that case, the integral becomes highly concentrated at 0 and as nn\to\infty,

I(n)=enf(0)(r=0R1Γ(r+1/2)Ψ2r(f,g)nr+1/2+O(1nR+1/2))I(n)=e^{n\cdot f(0)}\mathopen{}\mathclose{{}\left(\sum_{r=0}^{R-1}\frac{{\Gamma}(r+1/2)\cdot\Psi_{2r}(f,g)}{n^{r+1/2}}+O\mathopen{}\mathclose{{}\left(\frac{1}{n^{R+1/2}}}\right)}\right) (7.8)

for certain constants Ψ2r(f,g)\Psi_{2r}(f,g). If we write the expansions of ff and gg at 0 as

f(z)f(0)=m=0amzm+2,g(z)=m=0bmzmf(z)-f(0)=-\sum_{m=0}^{\infty}a_{m}z^{m+2},\qquad g(z)=\sum_{m=0}^{\infty}b_{m}z^{m}

then we have

Ψs(f,g)=a0(s+1)/2m=0sbsmk=0m((s+1)/2k)𝒜m,k(a1a0,a2a0,).\Psi_{s}(f,g)=a_{0}^{-(s+1)/2}\sum_{m=0}^{s}b_{s-m}\sum_{k=0}^{m}\binom{-(s+1)/2}{k}{\mathcal{A}}_{m,k}\mathopen{}\mathclose{{}\left(\frac{a_{1}}{a_{0}},\frac{a_{2}}{a_{0}},\dots}\right). (7.9)

Laplace found the main term of (7.8) with r=0r=0 and Ψ0(f,g)=a01/2b0\Psi_{0}(f,g)=a_{0}^{-1/2}b_{0} in the 18th century. Perron in [Per17] proved (7.8) rigorously, giving a formula for the constants Ψ2r(f,g)\Psi_{2r}(f,g) in terms of a derivative. It is then easy to obtain (7.9) with (3.2). See [Nem13] for a discussion of this, and the weaker conditions on ff and gg that are possible in Laplace’s method. The general forms of (7.8) and (7.9) in Perron’s saddle-point method are also described in Corollary 5.1 and Proposition 7.2 of [O’S19].

For an important application to the gamma function write

Γ(n+1)=0ettn𝑑t=nn+11en(log(1+z)1z)𝑑z.{\Gamma}(n+1)=\int_{0}^{\infty}e^{-t}t^{n}\,dt=n^{n+1}\int_{-1}^{\infty}e^{n(\log(1+z)-1-z)}\,dz.
Corollary 7.2 (Stirling’s approximation).

As real nn\to\infty

Γ(n+1)=2πn(ne)n(1+γ1n+γ2n2++γknk+O(1nk+1)).{\Gamma}(n+1)=\sqrt{2\pi n}\mathopen{}\mathclose{{}\left(\frac{n}{e}}\right)^{n}\mathopen{}\mathclose{{}\left(1+\frac{{\gamma}_{1}}{n}+\frac{{\gamma}_{2}}{n^{2}}+\cdots+\frac{{\gamma}_{k}}{n^{k}}+O\mathopen{}\mathclose{{}\left(\frac{1}{n^{k+1}}}\right)}\right). (7.10)

From (7.9) after simplifying, we obtain the coefficients

γm=j=02m(2m+2j1)!!(1)jj!𝒜2m,j(13,14,15,),{\gamma}_{m}=\sum_{j=0}^{2m}\frac{(2m+2j-1)!!}{(-1)^{j}j!}{\mathcal{A}}_{2m,j}\mathopen{}\mathclose{{}\left(\frac{1}{3},\frac{1}{4},\frac{1}{5},\cdots}\right), (7.11)

as in [Per17, Sect. 5], [O’S19, Sect. 8.1], where (2k1)!!(2k-1)!! means the product of the odd numbers less than 2k2k and equals (2k)!/(2kk!)(2k)!/(2^{k}k!) for k0k\geqslant 0. Stirling’s series for logΓ(n)\log{\Gamma}(n), see [AAR99, Eq. (1.4.5)], is

logΓ(n+1)=log(2πn(ne)n)+B221n+B443n3++B2k2k(2k1)n2k1+O(1n2k+1),\log{\Gamma}(n+1)=\log\mathopen{}\mathclose{{}\left(\sqrt{2\pi n}\mathopen{}\mathclose{{}\left(\frac{n}{e}}\right)^{n}}\right)+\frac{B_{2}}{2\cdot 1\cdot n}+\frac{B_{4}}{4\cdot 3\cdot n^{3}}+\cdots+\frac{B_{2k}}{2k(2k-1)n^{2k-1}}+O\mathopen{}\mathclose{{}\left(\frac{1}{n^{2k+1}}}\right), (7.12)

and may be exponentiated to produce another expression for γm{\gamma}_{m}:

γm=j=0m1j!𝒜m,j(B221,0,B443,0,).{\gamma}_{m}=\sum_{j=0}^{m}\frac{1}{j!}{\mathcal{A}}_{m,j}\mathopen{}\mathclose{{}\left(\frac{B_{2}}{2\cdot 1},0,\frac{B_{4}}{4\cdot 3},0,\cdots}\right). (7.13)

Replacing the Bernoulli numbers in (7.13) with Riemann zeta values, [AAR99, Eq. (1.3.4)], gives the equivalent equality

γm=j=0m1j!𝒜m,j(ζ(1)1,ζ(2)2,ζ(3)3,),{\gamma}_{m}=\sum_{j=0}^{m}\frac{1}{j!}{\mathcal{A}}_{m,j}\mathopen{}\mathclose{{}\left(\frac{\zeta(-1)}{-1},\frac{\zeta(-2)}{-2},\frac{\zeta(-3)}{-3},\cdots}\right), (7.14)

and this has a pleasing symmetry with

Γ(z+1)=m=0[j=0m1j!𝒜m,j(γ,ζ(2)2,ζ(3)3,)](z)m,{\Gamma}(z+1)=\sum_{m=0}^{\infty}\mathopen{}\mathclose{{}\left[\sum_{j=0}^{m}\frac{1}{j!}{\mathcal{A}}_{m,j}\mathopen{}\mathclose{{}\left({\gamma},\frac{\zeta(2)}{2},\frac{\zeta(3)}{3},\cdots}\right)}\right](-z)^{m}, (7.15)

where γ{\gamma} is Euler’s constant. We have convergence in (7.15) for |z|<1|z|<1 and it is proved by taking the log of the Weierstrass product for 1/Γ(z)1/{\Gamma}(z), expanding this into a series, and then exponentiating. Further formulas for γm{\gamma}_{m} appear in [Com74, p. 267] and [Nem13, Sect. 3]. See [Gél], [Sch68, Sect. 5.3] for the history of Stirling’s series. The common form (7.12) we use today, involving Bernoulli numbers, is due to De Moivre who simplified Stirling’s treatment.

The reciprocal of the gamma function has a similar asymptotic expansion:

1Γ(n+1)=12πn(en)n(1γ1n+γ2n2+(1)kγknk+O(1nk+1)),\frac{1}{{\Gamma}(n+1)}=\frac{1}{\sqrt{2\pi n}}\mathopen{}\mathclose{{}\left(\frac{e}{n}}\right)^{n}\mathopen{}\mathclose{{}\left(1-\frac{{\gamma}_{1}}{n}+\frac{{\gamma}_{2}}{n^{2}}-\cdots+(-1)^{k}\frac{{\gamma}_{k}}{n^{k}}+O\mathopen{}\mathclose{{}\left(\frac{1}{n^{k+1}}}\right)}\right), (7.16)

and the fact that the same coefficients appear is a consequence of (7.12) containing only odd powers. Of course, since Γ(n+1)=n!{\Gamma}(n+1)=n! for n0n\in{\mathbb{Z}}_{\geqslant 0}, the results (7.10) and (7.16) may be used to estimate binomial coefficients or other factorial expressions as their entries go to infinity. See [Boy94] for a generalization of Corollary 7.2 with nn replaced by zz\in{\mathbb{C}} and |z||z|\to\infty with |argz|<π/2|\arg z|<\pi/2.

7.3 A further example

The integral

Iα(n):=1(logz)neαz𝑑z(α>0),I_{\alpha}(n):=\int_{1}^{\infty}(\log z)^{n}e^{-\alpha z}\,dz\qquad(\alpha>0), (7.17)

requires more involved methods than (7.7) to find its behavior as nn\to\infty. Since loglogz\log\log z does not have a local maximum, we look for a saddle-point z0z_{0} for the whole integrand. To locate it, recall the Lambert WW function. For x0x\geqslant 0 it may be defined as the inverse to xxexx\mapsto xe^{x} so that W(xex)=xW(xe^{x})=x. Then W(x)W(x) is non-negative and increasing for x0x\geqslant 0, satisfying W(x)logxW(x)\leqslant\log x when xex\geqslant e. An easy calculation finds

z0=eW(n/α)=n/αW(n/α).z_{0}=e^{W(n/\alpha)}=\frac{n/\alpha}{W(n/\alpha)}.

In this case we may still use Laplace’s method according to the general procedures in [FS09, Sect. B6]. Part of the series (3.5) has to be exponentiated and the polynomials 𝒜n,k{\mathcal{A}}_{n,k} keep track of all the components. Recalling (3.6), define the rational functions

ar(v):=j=02r(2r+2j1)!!j!(v2v+1)j+r𝒜2r,j(3(v),4(v),).a_{r}(v):=\sum_{j=0}^{2r}\frac{(2r+2j-1)!!}{j!}\mathopen{}\mathclose{{}\left(\frac{v^{2}}{v+1}}\right)^{j+r}{\mathcal{A}}_{2r,j}(\ell_{3}(v),\ell_{4}(v),\dots).

The next result is proved in [O’S21, Sect. 4].

Theorem 7.3.

Suppose α>0\alpha>0 and set u:=W(n/α)u:=W(n/\alpha). Then as nn\to\infty we have

Iα(n)=2πu(1+u)n(ue1/u)n(1+r=1R1ar(u)nr+O((logn)RnR))I_{\alpha}(n)=\frac{\sqrt{2\pi}u}{\sqrt{(1+u)n}}\mathopen{}\mathclose{{}\left(\frac{u}{e^{1/u}}}\right)^{n}\mathopen{}\mathclose{{}\left(1+\sum_{r=1}^{R-1}\frac{a_{r}(u)}{n^{r}}+O\mathopen{}\mathclose{{}\left(\frac{(\log n)^{R}}{n^{R}}}\right)}\right) (7.18)

where the implied constant depends only on R1R\geqslant 1 and α\alpha. Also ar(u)(logn)ra_{r}(u)\ll(\log n)^{r}.

A more general result allows suitable functions f(z)f(z) to be included in the integrand in (7.17) and this is used to prove the asymptotic expansion of the nnth Taylor coefficient of a normalized version of the Riemann zeta function ζ(s)\zeta(s) at the central symmetric point s=1/2s=1/2 as nn\to\infty. See [O’S21, Thm. 1.5], giving an explicit version of [GORZ19, Thm. 9] by expressing the expansion coefficients in terms of De Moivre polynomials.

References

  • [AAR99] George E. Andrews, Richard Askey, and Ranjan Roy. Special functions, volume 71 of Encyclopedia of Mathematics and its Applications. Cambridge University Press, Cambridge, 1999.
  • [Arb00] L. F. A. Arbogast. Du calcul des dérivations. Levrault, Strasbourg, 1800.
  • [Bel34] E. T. Bell. Exponential polynomials. Ann. of Math. (2), 35(2):258–277, 1934.
  • [BG07] David R. Bellhouse and Christian Genest. Maty’s biography of Abraham De Moivre, translated, annotated and augmented. Statist. Sci., 22(1):109–136, 2007.
  • [BGW19] Daniel Birmajer, Juan B. Gil, and Michael D. Weiner. A family of Bell transformations. Discrete Math., 342(1):38–54, 2019.
  • [BKO04] Jan H. Bruinier, Winfried Kohnen, and Ken Ono. The arithmetic of the values of modular functions and the divisors of modular forms. Compos. Math., 140(3):552–566, 2004.
  • [Boy94] W. G. C. Boyd. Gamma function asymptotics by an extension of the method of steepest descents. Proc. Roy. Soc. London Ser. A, 447(1931):609–630, 1994.
  • [Cha02] Charalambos A. Charalambides. Enumerative combinatorics. CRC Press Series on Discrete Mathematics and its Applications. Chapman & Hall/CRC, Boca Raton, FL, 2002.
  • [Com74] Louis Comtet. Advanced combinatorics. D. Reidel Publishing Co., Dordrecht, enlarged edition, 1974. The art of finite and infinite expansions.
  • [Cra05] Alex D. D. Craik. Prehistory of Faà di Bruno’s formula. Amer. Math. Monthly, 112(2):119–130, 2005.
  • [DM97] Abraham De Moivre. A method of raising an infinite multinomial to any given power, or extracting any given root of the same. Philos. Trans. R. Soc. London, 19(230):619––625, 1697. Also in Miscellanea analytica de seriebus et quadraturis, J. Tonson & J. Watts, London, 1730.
  • [EJ] Mark Elin and Fiana Jacobzon. Families of inverse functions: coefficient bodies and the Fekete–Szegö problem. arXiv:2012.07153.
  • [FS09] Philippe Flajolet and Robert Sedgewick. Analytic combinatorics. Cambridge University Press, Cambridge, 2009.
  • [Gél] Jacques Gélinas. Original proofs of Stirling’s series for log(n!). arXiv: 1701.06689.
  • [Ges16] Ira M. Gessel. Lagrange inversion. J. Combin. Theory Ser. A, 144:212–249, 2016.
  • [Gir] Kurt Girstmair. Chebyshev polynomials and Galois groups of De Moivre polynomials. arXiv:2007.14183.
  • [GJ83] I. P. Goulden and D. M. Jackson. Combinatorial enumeration. A Wiley-Interscience Publication. John Wiley & Sons, Inc., New York, 1983. With a foreword by Gian-Carlo Rota, Wiley-Interscience Series in Discrete Mathematics.
  • [GKP94] Ronald L. Graham, Donald E. Knuth, and Oren Patashnik. Concrete mathematics. Addison-Wesley Publishing Company, Reading, MA, second edition, 1994. A foundation for computer science.
  • [GORZ19] Michael Griffin, Ken Ono, Larry Rolen, and Don Zagier. Jensen polynomials for the Riemann zeta function and other sequences. Proc. Natl. Acad. Sci. USA, 116(23):11103–11110, 2019.
  • [GS95] H. W. Gould and Paula Schlesinger. Extensions of the Hermite G.C.D. theorems for binomial coefficients. Fibonacci Quart., 33(5):386–391, 1995.
  • [GWLL21] Leonardo Rydin Gorjäo, Dirk Witthaut, Klaus Lehnertz, and Pedro G. Lind. Arbitrary-order finite-time corrections for the Kramers-Moyal operator. Entropy, 23(3):Paper No. 517, pp. 15, 2021.
  • [HPM21] Andrés Herrera-Poyatos and Pieter Moree. Coefficients and higher order derivatives of cyclotomic polynomials: old and new. Expo. Math., 39(3):309–343, 2021.
  • [Jha] Sumit Kumar Jha. Formulas for the number of kk-colored partitions and the number of plane partitions of nn in terms of the Bell polynomials. arXiv:2009.04889.
  • [Joh02a] Warren P. Johnson. Combinatorics of higher derivatives of inverses. Amer. Math. Monthly, 109(3):273–277, 2002.
  • [Joh02b] Warren P. Johnson. The curious history of Faà di Bruno’s formula. Amer. Math. Monthly, 109(3):217–234, 2002.
  • [Knu97] Donald E. Knuth. The art of computer programming. Vol. 1. Addison-Wesley, Reading, MA, 1997. Fundamental algorithms, Third edition.
  • [Nem13] Gergö Nemes. An explicit formula for the coefficients in Laplace’s method. Constr. Approx., 38(3):471–487, 2013.
  • [O’S] Cormac O’Sullivan. Symmetric functions and a natural framework for combinatorial and number theoretic sequences. arXiv.
  • [O’S18] Cormac O’Sullivan. Partitions and Sylvester waves. Ramanujan J., 47(2):339–381, 2018.
  • [O’S19] Cormac O’Sullivan. Revisiting the saddle-point method of Perron. Pacific J. Math., 298(1):157–199, 2019.
  • [O’S21] Cormac O’Sullivan. Zeros of Jensen polynomials and asymptotics for the Riemann xi function. Res. Math. Sci., 8(3):Paper No. 46, 27, 2021.
  • [Per17] Oskar Perron. Über die näherungsweise Berechnung von Funktionen großer Zahlen. Sitzungsber. Bayr. Akad. Wissensch. (Münch. Ber.), pages 191–219, 1917.
  • [Pit06] J. Pitman. Combinatorial stochastic processes, volume 1875 of Lecture Notes in Mathematics. Springer-Verlag, Berlin, 2006. Lectures from the 32nd Summer School on Probability Theory held in Saint-Flour, July 7–24, 2002.
  • [Rad73] Hans Rademacher. Topics in analytic number theory. Springer-Verlag, New York, 1973. Edited by E. Grosswald, J. Lehner and M. Newman, Die Grundlehren der mathematischen Wissenschaften, Band 169.
  • [Rio68] John Riordan. Combinatorial identities. John Wiley & Sons, Inc., New York-London-Sydney, 1968.
  • [Sch] Aidan Schumann. Multivariate Bell polynomials and derivatives of composed functions. arXiv: 1903.03899.
  • [Sch68] Ivo Schneider. Der Mathematiker Abraham de Moivre (1667–1754). Arch. History Exact Sci., 5(3-4):177–317, 1968.
  • [SMA09] Armin Straub, Victor H. Moll, and Tewodros Amdeberhan. The pp-adic valuation of kk-central binomial coefficients. Acta Arith., 140(1):31–42, 2009.
  • [ST88] H. M. Srivastava and Pavel G. Todorov. An explicit formula for the generalized Bernoulli polynomials. J. Math. Anal. Appl., 130(2):509–513, 1988.
  • [Ste43] Moritz A. Stern. Ueber die Coëfficienten der Secantenreihe. J. Reine Angew. Math., 26:88–91, 1843.

Dept. of Math, The CUNY Graduate Center, 365 Fifth Avenue, New York, NY 10016-4309, U.S.A.

E-mail address: [email protected]