arXiv:1311.5067v5 [math.CO] 27 Jan 2021 and ϕ eso h rtadscn id ohplnma aiisapa as appear Su families order. higher Both of derivatives certain kind. D of second expansions and in first coefficients the of bers X w obyidxdfmle fhmgnosadioai oyoil in exponentia isobaric (partial) and mials the homogeneous considered: are of indeterminates families several indexed doubly Two Abstract family 09,46E25 40E99, inversio Lagrange MSC: numbers, Stirling 2010 laws, Inversion formula, Bruno Keywords: h nioeo h aad rn ofagba tcnb represe be can It algebra. B Hopf essentia the is Bruno Fa`a of solution di the X that terms Schmitt on in and antipode problem Haiman the inversion to According Lagrange the polynomials. of solution Comtet’s h offiinsof coefficients the oyoil.A o-rva xml,aSh¨mlhtp oml sde is co Schl¨omilch-type formula the a between example, expressing relations non-trivial of a cases As special polynomials. be to demonstrated are j 1 j − n nisinverse its on and eateto ahmtc n ahmtclEuain Uni Education, Mathematical and Mathematics of Department ( (2 mi address: Email hw htboth that shows ϕ B n (the ) − B n,k S 1) n,k n,k S bya neso a hc eeaie hto h trignum- Stirling the of that generalizes which law inversion an obey n, n e family new a and S j 1 (1 t eiaieo xdfunction fixed a of derivative -th n,k utvraeSiln oyoil,Bl oyoil,Fa`a di polynomials, Bell polynomials, Stirling Multivariate oevr eea xaso oml hthlsfrtewhole the for holds that formula expansion general a Moreover, . 51,0A9 59,1B3 18,1C8 31,16T30, 13N15, 11C08, 11B83, 11B73, 05A99, 05A19, 05A18, ≤ utvraeSiln Polynomials Stirling Multivariate ntrso h elpolynomials Bell the of terms in u e aps1 -44 lnbr,Germany Flensburg, D-24943 1, Campus dem Auf [email protected] k S fteFrtadScn Kind Second and First the of ≤ n,k S n n,k eea motn rpriso h trignumbers Stirling the of properties important Several . ϕ setbihdtgte ihacoe xrsinfor expression closed a with together established is ) epciey oenwlgti hdteeyon thereby shed is light new Some respectively. , and S B n,k lrdSchreiber Alfred n,k ∈ r ieeta oyoil eedn on depending polynomials differential are Z [ X 1 ,X . . . , Afe Schreiber) (Alfred ϕ npaeo h indeterminates the of place in ) n − B k +1 n,k uhthat such ] n ieversa. vice and , est fFlensburg, of versity elpolyno- Bell l X rresponding 1 − bstituting (2 n tdby nted n − 1) rived S Lie lly n,k ell 1. Introduction 1.1. Background and problem It is well-known that a close connection exists between iterated differen- tiation and Stirling numbers (see, e. g., [18, 34, 56]). Let s1(n, k) denote the signed Stirling numbers of the first kind, s2(n, k) the Stirling numbers of the second kind, and D the operator d/dx. Then, for all positive integers n, the nth iterate (xD)n can be expanded into the sum

n n k k (xD) = s2(n, k)x D . (1.1) k X=1 An expansion in the reverse direction is also known to be valid (see, e. g., [18, p. 197] or [34, p. 45]):

n n −n k D = x s1(n, k)(xD) . (1.2) k X=1 Let us first look at (1.1). The occurrence of the Stirling numbers can be explained combinatorially as follows. Observing

(xD)nf(x)= Dn(f ◦ exp)(log x) we can use the classical higher-order chain rule (named after Fa`a di Bruno; cf. [16, 18], [20, pp. 52, 481]) to calculate the nth derivative of the composite function f ◦ g:

n (n) ′ (n−k+1) (k) (f ◦ g) (x)= Bn,k(g (x),...,g (x)) · f (g(x)), (1.3) k X=1 where Bn,k ∈ Z[X1,...,Xn−k+1], 1 ≤ k ≤ n, is the (partial) exponential Bell polynomial

n! r1 r2 Bn,k(X1,...,Xn−k+1)= X X ... (1.4) r !r ! ... (1!)r1 (2!)r2 ... 1 2 r ,r ,... 1 2 1X2 the sum to be taken over all non-negative integers r1,r2,...,rn−k+1 such that r1 + r2 + ... + rn−k+1 = k and r1 +2r2 + ... +(n − k + 1)rn−k+1 = n. The coefficient in Bn,k counts the partitions of n distinct objects into k blocks

2 (subsets) with rj blocks containing exactly j objects (1 ≤ j ≤ n − k + 1). Therefore, the sum of these coefficients is equal to the number s2(n, k) of k all such partitions. So we have Bn,k(x,...,x) = s2(n, k)x . Evaluating (f ◦ exp)(n)(log x) by (1.3) then immediately gives the right-hand side of (1.1). Question. Can also (1.2) be interpreted in this way by substituting jth derivatives in place of the indeterminates Xj of some polynomial Sn,k ∈ Z[X1, . . . ,Xn−k+1], the coefficients of which add up to s1(n, k)? The main purpose of the present paper is to give a positive and compre- hensive answer to this question including recurrences, a detailed study of the inverse relationship between the polynomial families Bn,k and Sn,k, as well as fully explicit formulas (with some applications to Stirling numbers and Lagrange inversion). The issue turns out to be closely related to the problem of generalizing (1.1), that is, finding an expansion for the operator (θD)n (n ≥ 1, θ a function of x). Note that, in the case of scalar functions, (θD)f is the Lie derivative of f with respect to θ. Several authors have dealt with this problem. In [7] and [29] a polynomial family Cn,k ∈ Z[X0,X1,...,Xn−k] has been defined1 by differential recurrences and shown to comply with (θD)n = n ′ (n−k) k k=1 Cn,k(θ, θ ,...,θ )D . Comtet [7] has tabulated Cn,k up to n = 7 n and stated that Cn,k(x,...,x)= c(n, k)x , where c(n, k) := |s1(n, k)| denotes theP signless Stirling numbers of the first kind (‘cycle numbers’ according to the terminology in [19]). Since however all coefficients of Cn,k are positive, Cn,k does not appear to be a suitable companion for Bn,k with regard to the desired inversion law. Todorov [54, 55] has studied the above Lie derivation with respect to a ′ function θ of the special form θ(x) = 1/ϕ′(x), ϕ (x) =6 0. His main results in [54] ensure the existence of Sn,k ∈ Z[X1, . . . ,Xn−k+1] such that n ′ −1 n ′ (n−k+1) (k) ϕ (x) D f(x)= An,k(ϕ (x),...,ϕ (x)) · f (x), (1.5) k=1  X −(2n−1) where An,k := X1 Sn,k. While differential recurrences for An,k can read- ily be derived from (1.5) (cf. [54, Equation (27)] or a slightly modified version

1 Here we write Cn,k instead of Comtet’s An,k (cf. [7]) in order to avoid misunderstand- ings. Note that in the present paper An,k is exclusively used to denote the ‘Lie coefficients’ according to Todorov (see (1.5) below).

3 in [55, Theorem 2]), a simple representation for Sn,k — as is (1.4) for Bn,k — was still lacking up to now. Todorov [54, p. 224] erroneously believed that the somewhat cumbersome ‘explicit’ expression in [7] for the coefficients of Cn,k would directly yield the coefficients of Sn,k. Also the determinantal form ′ n presented in [54, Theorem 6] for (D/ϕ ) (and thus also for Sn,k) may only in a modest sense be regarded as explicit. Nevertheless, Todorov’s choice (θ = 1/ϕ′) eventually proves to be a crucial idea. Among other things, it reveals that An,k (and thus Sn,k) is connected with the classical Lagrange problem of computing the compositional inverse n f of a given series f(x)= n≥1(fn/n!)x , f1 =6 0. As we shall see later, the Taylor coefficients f of f(x) can be expressed simply by applying A to n P n,1 the coefficients of f as follows:

f n = An,1(f1,...,fn). (1.6)

On the other hand, Comtet [8] found an inversion formula that expresses f n in terms of the (partial) exponential Bell polynomials:

n−1 k −n−k f n = (−1) f1 Bn+k−1,k(0, f2,...,fn). (1.7) k X=0 This result has been shown by Haiman and Schmitt [14, 38] to provide es- sentially both a combinatorial representation and a cancellation-free com- putation of the antipode on a Fa`adi Bruno Hopf algebra (a topic that has received a lot of attention in quantum field theory due to its application to renormalization; cf. [23, 6, 11]). Combining (1.6) with (1.7) we obtain an expression for An,1 in terms of the Bell polynomials. This suggests looking for a similar representation for the whole family An,k, 1 ≤ k ≤ n. As a main result (Theorem 6.1), we shall prove the following substantially extended version of (1.6) & (1.7):

n−1 n−1−r 2n − 2 − r −(2n−1)+r A = (−1) X B˜ − − − − − . (1.8) n,k k − 1 1 2n 1 k r,n 1 r r k− =X1   The tilde over B indicates that X1 has been replaced by 0. From (1.8) we eventually get the desired explicit standard representation for An,k that corresponds to the one for Bn,k given in (1.4). Equation (1.8) states a somewhat intricate relationship between the fam- ilies An,k and Bn,k. A simpler connection of both expressions is the following

4 basic inversion law which generalizes the orthogonality of the Stirling num- bers (cf. Section 5):

n

An,jBj,k = δn,k (1 ≤ k ≤ n), (1.9) j k X= where δn,n = 1, δn,k = 0 if n =6 k (Kronecker symbol).

1.2. Terminology and notation

Considering (1.9) and the fact that the sum of the coefficients of An,k and of Bn,k are equal to s1(n, k) and to s2(n, k), respectively, it may be justified to call An,k and Bn,k multivariate of the first and second kind. There should be no risk of confusing them with polynomials in one variable like those introduced and named after Stirling by Nielsen [31, 32], neither with the closely related ‘Stirling polynomials’ fk(n) := s2(n + k, n) and gk(n) := c(n, n−k) Gessel and Stanley [12] have investigated as functions of n ∈ Z. A sequence r1,r2,r3,... of non-negative integers is said to be an (n, k)- partition type,0 ≤ k ≤ n, if r1 + r2 + r3 + ... = k and r1 +2r2 +3r3 + ... = n. The set of all (n, k)-partition types is denoted by P(n, k); we write P for the union of all P(n, k). In the degenerate case (k = 0) set P(n, 0) = ∅, if n> 0, and P(0, 0) = {0} otherwise. Let k ≥ 1. Since n−k +1 is the greatest j such that rj > 0, we often write (n, k)-partition types as ordered (n−k +1)-tuples (r1,...,rn−k+1). The polynomials to be considered in the sequel have the form

r1 r2 Pπ = π(r1,r2,...)X1 X2 ..., X where the sum ranges over all elements (r1,r2,...) of a full set P(n, k). As a consequence, Pπ is homogeneous of degree k and isobaric of degree n. The coefficients of Pπ may be regarded as values of a map π : P −→ Z defined by some combinatorially meaningful expression, at least in typical cases like the following:

(r1 +2r2 + ...)! ω(r1,r2,...) := order function (Lah) (1.10) r1! r2! · ...

ω(r1,r2,...) γ(r1,r2,...) := cycle function (Cauchy) (1.11) 1r1 2r2 · ... 5 ω(r1,r2,...) β(r1,r2,...) := subset function (Fa`adi Bruno) (1.12) (1!)r1 (2!)r2 · ... These coefficients count the number of ways a set can be partitioned into non-empty blocks according to a given partition type, that is, rj denotes the number of blocks containing j elements (j = 1, 2,...). The result depends on the meaning of ‘block’: linearly ordered subset (ω), cyclic order (γ), or unordered subset (β). It should be noticed that the corresponding polynomials Pω,Pγ,Pβ(= Bn,k) are closely related to well-known combinatorial number-families:

+ Pω(1,..., 1) = l (n, k), unsigned Lah numbers [24, 34] n P (1,..., 1) = c(n, k)= , unsigned Stirling numbers of the 1st kind γ k   n P (1,..., 1) = s (n, k)= , Stirling numbers of the 2nd kind. β 2 k   1.3. Overview This paper is organized as follows: In Section 2 a general setting is sketched that allows functions and derivations to be treated algebraically. Section 3 contains a study of the iterated Lie operator D(ϕ)−1D. An ex- pansion formula for (D(ϕ)−1D)n is established together with a differential recurrence for the resulting Lie coefficients An,k. Doing the same with re- n spect to the inverse function ϕ will yield, conversely, D expanded and Bn,k as the corresponding Lie coefficients. A by-product of Section 3 is Fa`adi Bruno’s formula and its applications to the partial Bell polynomials Bn,k to be briefly summarized in Section 4. These basic facts then lead to both inversion and recurrence relations, which we shall demonstrate and discuss in Section 5. The main task in Section 6 is to find an explicit polynomial expression for Sn,k. This is eventually achieved by means of (1.8), a proof of which makes up a central part of the section. In Section 7 we give some ap- plications to the Lagrange inversion problem and to exponential generating functions.

2. Function algebra with derivation 2.1. Basic notions Menger [26] has introduced the notion of a ‘tri-operational algebra’ of functions, which in the sequel (since 1960) stimulated to a great extent studies

6 of generalized function algebras, e. g., [10, 27, 44, 45, 51]. In what follows I will propose a variant of Menger’s original system tailored to our specific purposes of treating functions and their derivatives in a purely algebraic way. Let (F, +, ·) be a non-trivial commutative ring of characteristic zero, 0 and 1 its identity elements with respect to addition and multiplication. We will think of the elements of F as ‘functions (of one variable)’ and therefore assume that F has a third binary operation ◦ (called composition) together with an identity element ι such that the following axioms are satisfied:

(F1) f ◦ (g ◦ h)=(f ◦ g) ◦ h (F2) (f + g) ◦ h =(f ◦ h)+(g ◦ h) (F3) (f · g) ◦ h =(f ◦ h) · (g ◦ h) (F4) f ◦ ι = ι ◦ f = f (F5) 1 ◦ 0=1 (F4) is assumed to be valid for all f ∈ F; hence ι is unique. Let f be any element of F. From (F2) we conclude 0 ◦ f = 0; so we get ι =6 0 (by (F4)) and ι =6 1 (by (F5)). (F2) furthermore implies (−f) ◦ g = −(f ◦ g). The least subring of F containing 1 will in the following conveniently be identified with Z. (F5) then extends to the integers, that is, n ◦ 0= n holds for all n ∈ Z. Given a unit f in F (i. e., f is an element invertible with respect to multi- −1 plication), we write f (or 1/f) for the inverse (henceforth called reciprocal) of f. 2.1 Remark. It must be emphasized that ◦ has to be understood as a partial operation (of course, ι−1 ◦ 0 is not defined). We therefore assign truth values to formulas, especially to our postulates (F1–3), whenever the terms involved are meaningful. Let f,g ∈ F be functions such that f ◦ g = g ◦ f = ι. Then g is called the compositional inverse of f, and vice versa. It is unique and will be denoted by f. The following is obvious: ι = ι, f = f, and f ◦ g = g ◦ f.

2.1 Definition. Suppose (F, +, ·, ◦) satisfies (F1–5). We then call a mapping D : F −→ F derivation on F, and (F, +, ·, ◦,D) a function algebra with derivation, if D meets the following conditions:

(D1) D(f + g)= D(f)+ D(g)

7 (D2) D(f · g)= D(f) · g + f · D(g) (D3) D(f ◦ g)=(D(f) ◦ g) · D(g) (D4) D(ι)=1 (D5) D(f)=0=⇒ f ◦ 0= f

The classical derivation rules (D1), (D2) make F into a differential ring. Some simple facts are immediate: D(0) = D(1) = 0, D(m · f) = m · D(f) for all m ∈ Z. By an inductive argument the product rule (D2) can be generalized:

n

D(f1 ··· fn)= f1 ··· fk−1 · D(fk) · fk+1 ··· fn. (2.1) k X=1

n n−1 By putting fi = f, 1 ≤ i ≤ n, (2.1) becomes D(f ) = nf D(f). If f is a unit, this holds also for n ≤ 0. As usual, f m for m< 0 is defined by (f −1)−m. (D4) prevents D from operating trivially. In the case of a field F, (D4) can be weakend to D(f) =6 0 (for some f ∈ F), since the chain rule (D3) then gives D(f)= D(f ◦ ι)= D(f) · D(ι). Applying (D3) and (D4) to D(f ◦ f) we obtain the inversion rule 1 D(f)= . (2.2) D(f) ◦ f

In a differential ring, it is customary to define the subring K of constants as the kernel of the additive homomorphism D, that is,

K := {f ∈F| D(f)=0}.

We have Z ⊆K. Constants behave as one would expect.

2.1 Proposition. c ∈K ⇐⇒ c ◦ f = c for all f ∈F.

Proof. ⇒: Suppose D(c) = 0. Then, for any f ∈ F we have by (F1) and (D5): c ◦ f =(c ◦ 0) ◦ f = c ◦ (0 ◦ f)= c ◦ 0= c.— ⇐: Set f = 0 and apply the chain rule (D3). If f ◦ 0 exists for f ∈F, then it is obviously a constant.

8 2.2 Examples (Function algebras with derivation). In each case below, the ‘functions’ carry some argument X (indeterminate, variable) that can be substituted in the usual sense: f(g(X)) =: (f ◦ g)(X). 1. The rational function field Q(X) together with the algebraically defined derivation R 7−→ R′, R ∈ Q(X). 2. The ring R[[X]] of all power series with formal differentiation. 3. The ring of real-valued C∞ functions on an open real interval with the ordinary differential operator d/dx. 4. The field of all meromorphic functions on a given region in C with complex differentiation.

Given a polynomial P ∈K[X1, ...,Xn] and functions f1, ...,fn, we de- note by P (f1, ...,fn) the function obtained by substituting fi in place of Xi, 1 ≤ i ≤ n. Recalling the algebraic definition of ∂/∂Xi, we readily obtain by (2.1) the generalized chain rule:

n ∂P D(P (f1, . . . ,fn)) = (f1, ...,fn) · D(fk). (2.3) ∂Xk k X=1 In the case n = 1 this becomes D(P (f)) = P ′(f)·D(f). Thus, D restricted to polynomial functions turns out to act like the ordinary differential operator: D(P (ι)) = P ′(ι). Equation (2.3) also applies to the case that P is a rational function from K(X1,...,Xn). Notation. Suppose that ϕ is a fixed function, and let Di denote the ith iterate of D. In the following we will abbreviate P (D(ϕ),D2(ϕ),...,Dn(ϕ)) to P ϕ. ϕ We denote by Pn(ϕ) the set of all P and by Rn(ϕ) the set of all rational ϕ ϕ expressions P /Q , where P, Q ∈K[X1,...,Xn]. 2.3 Remark. Obviously (P + Q)ϕ = P ϕ + Qϕ and (P · Q)ϕ = P ϕ · Qϕ. For a homogeneous polynomial P of degree k and arbitrary a, b ∈ K we have P aϕ+b = akP ϕ. Another formula that is known from calculus and that holds in the differential ring (F,D) is the general Leibniz rule, which yields an explicit expression for the higher derivatives of a product:

s s! j1 jn D (f1 ··· fn)= D (f1) ··· D (fn). (2.4) j1! ··· jn! j1+···+jn=s j1, ...,jXn ≥ 0

9 We note the special case Ds(ιn) = ns · ιn−s, if s ≤ n, and Ds(ιn)=0 otherwise; the falling power ns is defined by n(n − 1) ··· (n − s + 1).

Convention. Throughout the remainder of this paper we denote by ϕ any function from F such that the compositional inverse ϕ exists, D(ϕ) and D(ϕ) are units, and equation (2.2) holds for f = ϕ.

With regard to such a ϕ, we define a mapping Dϕ : F −→F by

D(f) D (f) := . (2.5) ϕ D(ϕ)

Dϕ is the function-algebraic version of the Lie derivative that Todorov used in his paper [54]. (F,Dϕ) is a differential ring having the same constants as (F,D). However, with regard to Dϕ we have: (D3) ⇔ (D4) ⇔ D(ϕ) = 1. Therefore, Dϕ satisfies the chain rule if and only if Dϕ = D, that is, in the trivial case ϕ = ι + c, c ∈K. The following simple but useful statement concerns the relation between n n Dϕ and D (Pourchet’s formula according to [8, p. 220] and [54, p. 223–224]).

n n 2.2 Proposition. Dϕ(f)= D (f ◦ ϕ) ◦ ϕ for all n ≥ 0.

n n Proof. Verify D (f ◦ ϕ) = Dϕ(f) ◦ ϕ by induction on n. The case n = 0 is clear. For the induction step (n → n +1) use (D3):

n+1 n D (f ◦ ϕ)=(D(Dϕ(f)) ◦ ϕ)) · D(ϕ)=:(∗).

n D(Dϕ(f)) n+1 Applying (2.2) to D(ϕ) yields (∗)= D(ϕ) ◦ ϕ = Dϕ (f) ◦ ϕ. 2.2. Exponential and logarithm For some purposes it will prove convenient to have in (F, +, ·, ◦,D) be- sides the three identity elements more functions with special properties, the most important examples being the exponential (exp) and its compo- sitional inverse (log). Of course we expect exp and log to have the familiar properties known from analysis, like D(exp) = exp, D(log) · ι = 1, and Dk(log) = (−1)k−1(k − 1)! ι−k for all k ≥ 1. In such a case we call F an extended function algebra. Items 3 and 4 from Examples 2.2 are extended function algebras. In R[[x]] (item 2), however, neither D(log) nor log have counterparts. In Section 7 we shall therefore deal with log ◦(1+ι) and exp −1 instead.

10 Notation. In an extended function algebra, we prefer to write id (or x) instead of ι. If no misunderstanding is likely, we will occasionally replace exp ◦f with ef and also switch from f ◦ g to the usual notation f(g), mainly when g is a constant or the identity function.

3. Expansion of higher-order derivatives

n In this section, some basic facts regarding the expansion of Dϕ (see (1.5) and Todorov [54, 55]) as well as of Dn will be reformulated and set up for an arbitrary function algebra F with derivation D. The main idea of the presentation is to make it clear that from the beginning these results are linked by an inversion of functions. We will also examine some basic proper- ties of the multivariable polynomials involved, in particular their differential recurrences and their relationship to the Stirling numbers.

n n 3.1. Expansion formulas for Dϕ and D 3.1 Proposition. Let f be any function from F and n, k non-negative inte- gers. Then there are an,k ∈ Rn−k+1(ϕ), 0 ≤ k ≤ n, such that

n n k Dϕ(f)= an,k · D (f). k X=0

The coefficients an,k are uniquely determined by the recurrence

a + D(a ) a = n,k−1 n,k (1 ≤ k ≤ n + 1), n+1,k D(ϕ) where a0,0 =1, ai,0 =0(i> 0), and ai,j =0 (0 ≤ i < j).

n Proof. Recall that D(ϕ) is a unit and (F,Dϕ) is a differential ring. So, Dϕ(f) is defined for every n ≥ 0 and can successively be calculated by applying (D1) −1 −2 and (D2) together with the rule Dϕ(g )= −g Dϕ(g)(g a unit). The proof is then carried out by a simple induction on n, the details of which can be omitted here.

−n 3.1 Remark. an,n = D(ϕ) for all n ≥ 0. 2n−1 One obtains by induction that the denominator in an,k is D(ϕ) .

11 2n−1 3.2 Corollary. Set sn,k := D(ϕ) an,k. Then sn,k ∈ Pn−k+1(ϕ) for (n, k) =6 (0, 0), and the following recurrence holds:

2 sn+1,k = −(2n − 1)D (ϕ)sn,k + D(ϕ) · (sn,k−1 + D(sn,k)),

−1 where s0,0 = D(ϕ) , s1,1 =1, si,0 =0(i> 0), and si,j =0 (0 ≤ i < j). It is natural to ask whether, conversely, Dn can be expanded into a linear k combination of the Dϕ (k =0, 1,...,n). The next proposition gives a positive answer. 3.3 Proposition. Let f be any function and n, k non-negative integers. Then, there are bn,k ∈ Pn−k+1(ϕ), 0 ≤ k ≤ n, such that

n n k D (f)= bn,k · Dϕ(f). k X=0 The coefficients bn,k are uniquely determined by the recurrence

bn+1,k = D(ϕ) · bn,k−1 + D(bn,k) (1 ≤ k ≤ n + 1), where b0,0 =1, bi,0 =0(i> 0), and bi,j =0 (0 ≤ i < j). Proof. We apply Proposition 3.1 to the compositional inverse ϕ, thus ob- ′ n n ′ k taining an,k ∈ Rn−k+1(ϕ) so that Dϕ(f) = k=0 an,k · D (f). Since accord- ing to Pourchet’s formula (Proposition 2.2) the left-hand side is equal to Dn(f ◦ ϕ) ◦ ϕ, we get by (F3) P

n n ′ k D (f ◦ ϕ)= (an,k ◦ ϕ) · (D (f) ◦ ϕ). (3.1) k X=0 We now replace f by f ◦ ϕ in (3.1). Then again Pourchet’s formula, applied to the second factor on the right-hand side of (3.1), yields

n n ′ k D (f)= (an,k ◦ ϕ) · Dϕ(f). k X=0 ′ ′ Now set bn,k := an,k ◦ ϕ. We then have b0,0 = a0,0 ◦ ϕ = 1 ◦ ϕ = 1, likewise bi,0 =0(i> 0), bi,j =0 (0 ≤ i < j), and by (2.2)

′ −1 ′ ′ bn+1,k = an+1,k ◦ ϕ =(D(ϕ) ◦ ϕ) · ((D(an,k) ◦ ϕ)+(an,k−1 ◦ ϕ))

12 ′ = D(ϕ) · (D(an,k) ◦ ϕ)+ D(ϕ) · bn,k−1

= D(bn,k)+ D(ϕ) · bn,k−1 (by the chain rule (D3)).

Finally, bn,k ∈ Pn−k+1(ϕ) follows from this recurrence by an inductive argu- ment.

n 3.2 Remark. bn,n = D(ϕ) for all n ≥ 0.

3.2. Fundamental properties of the coefficients

Let us now have a closer look at the coefficient functions an,k and bn,k. We start with bn,k. As it is a polynomial expression in the derivatives n−k+1 D(ϕ),...,D (ϕ), we get bn,k from a suitable polynomial family Bn,k by j ϕ substituting D (ϕ) in place of the indeterminates Xj, that is, Bn,k = bn,k. In the case of an,k it is likewise clear by Corollary 3.2 that sn,k, too, comes from ϕ 2n−1 certain polynomials Sn,k satisfying Sn,k = sn,k = D(ϕ) an,k. We therefore −(2n−1) ϕ define An,k := X1 Sn,k. Then, of course, An,k = an,k holds. Note that An,k is a Laurent polynomial, and that is especially also true for S0,0 (see Corollary 3.2). The polynomials Sn,k (An,k) and Bn,k are closely connected. 3.4 Proposition.

ϕ 2n−1 ϕ ϕ ϕ (i) Bn,k = D(ϕ) · (Sn,k ◦ ϕ), Bn,k = An,k ◦ ϕ ϕ 2n−1 ϕ ϕ ϕ (ii) Sn,k = D(ϕ) · (Bn,k ◦ ϕ), An,k = Bn,k ◦ ϕ

Proof. (i): From the proof of Proposition 3.3 we obtain

ϕ ′ ϕ Bn,k = an,k ◦ ϕ = An,k ◦ ϕ.

Now note the generalized inversion rule obtained by induction from (2.2):

D(ϕ)m ◦ ϕ = D(ϕ)−m for all integers m ≥ 0.

ϕ 2n−1 ϕ −(2n−1) ϕ Hence Sn,k ◦ ϕ =(D(ϕ) ◦ ϕ) · (An,k ◦ ϕ)= D(ϕ) · Bn,k. (ii): Replace ϕ in (i) by ϕ. From the foregoing we gather the following special values:

−n n An,n = X1 , Sn+1,n+1 = Bn,n = X1 (n ≥ 0),

13 Ai,0 = Si,0 = Bi,0 =0 (i> 0),

Ai,j = Si,j = Bi,j =0 (0 ≤ i < j). What still remains to be done is transforming the differential recurrences for an,k, sn,k, bn,k into recurrences for the corresponding polynomials An,k, Sn,k, Bn,k. Consider the derivative n−k+1 D(an,k)= D(An,k(D(ϕ),...,D (ϕ))). Applying (2.3) to the right-hand side, we obtain

n−k +1 ∂A D(a )= n,k D(ϕ),...,Dn−k+1(ϕ) · Dj+1(ϕ) n,k ∂X j=1 j X n−k+1 ϕ  ∂An,k = Xj+1 . ∂Xj j=1 ! X D(sn,k) and D(bn,k) resolve in the same manner. So, according to Proposi- tion 3.1, Corollary 3.2 and Proposition 3.3 we have the following 3.5 Proposition.

n−k+1 1 ∂An,k (i) An+1,k = An,k−1 + Xj+1 , X ∂Xj 1 j=1 ! X n−k+1 ∂Sn,k (ii) Sn+1,k = −(2n − 1)X2Sn,k + X1 Sn,k−1 + Xj+1 , ∂Xj j=1 ! X n−k+1 ∂Bn,k (iii) Bn+1,k = X1Bn,k−1 + Xj+1 . ∂Xj j=1 X It follows (by induction) from Proposition 3.5 that these polynomi- als have integral coefficients. We denote by σn,k(r1,...,rn−k+1) and r1 rn−k+1 βn,k(r1,...,rn−k+1) the coefficients of X1 ··· Xn−k+1 in Sn,k and in Bn,k, respectively, thus obtaining 3.6 Corollary.

r1 rn−k+1 (i) Sn,k = σn,k(r1,...,rn−k+1)X1 ··· Xn−k+1 , P − − − (2n X1 k,n 1) r1 rn−k+1 (ii) Bn,k = βn,k(r1,...,rn−k+1)X1 ··· Xn−k+1 . P X(n,k) 14 Proof. (i): By induction on n. For n = 1 we have the degenerate case of a (0, 0)-partition type, r1 = 0. Thus S1,1 = 1 can be achieved by choosing σ1,1(0) = 1. The induction step (n → n + 1) is carried out by examining the ∂Sn,k partition types produced by the terms X Sn,k, X Sn,k− , and X Xj in 2 1 1 1 +1 ∂Xj part (ii) of Proposition 3.5. Each of them makes ri (= n − 1) increase by 1 and makes iri (= 2n − 1 − k) increase by 2, which gives the appropriate P (2n +1 − k, n)-partition types for Sn+1,k. — (ii): Obviously due to a similar argument. P n−1 3.3 Remark. We already know that Sn,n = X1 . Taking k = n in Proposi- n n−2 tion 3.5 (ii) then yields Sn,n−1 = − 2 X1 X2 (n ≥ 2). Todorov [55] has also calculated S − and S − this way. n,n 2 n,n 3  3.4 Remark. As a consequence of Corollary 3.6, Sn,k is homogeneous of degree n − 1 and isobaric of degree 2n − 1 − k, while Bn,k is homogeneous of degree k and isobaric of degree n.

We now define integers s1(n, k) := An,k(1,..., 1) = Sn,k(1,..., 1) and s2(n, k) := Bn,k(1,..., 1), that is, s1(n, k), s2(n, k) are the sums of the coef- ficients of An,k (Sn,k) and Bn,k, respectively.

3.7 Proposition. Let n, k be integers, 0 ≤ k ≤ n. Then, s1(n, k) are the signed Stirling numbers of the first kind, and s2(n, k) are the Stirling numbers of the second kind: n n (i) s (n, k)=(−1)n−k , (ii) s (n, k)= . 1 k 2 k     Proof. (i): From the special values above we gather s1(0, 0) = 1, s1(i, 0)=0 (i> 0), and s1(i, j)=0(0 ≤ i < j). It suffices to show that s1 satisfies the recurrence s1(n +1,k)= s1(n, k − 1) − ns1(n, k), which defines the Stirling numbers of the first kind (see e. g. [34, p. 33] ). We use Corollary 3.6 (i). Consider first ∂Sn,k Xj+1 = rj σn,k(r1,...,rn−k+1) × ∂Xj P − − − (2n X1 k,n 1) r1 rj −1 rj+1+1 rn−k+1 × X1 ··· Xj Xj+1 ··· Xn−k+1 . Replacing all indeterminates by 1 and then taking the sum from j = 1 to n − k + 1 yields n−k+1 n−k+1

rj σn,k(r1,...,rn−k+1)= s1(n, k) rj. j=1 P − − − j=1 X (2n X1 k,n 1) X 15 Observing r1 + ··· + rn−k+1 = n − 1 we get by Proposition 3.5 (ii)

s1(n +1,k)= −(2n − 1)s1(n, k)+ s1(n, k − 1)+(n − 1)s1(n, k)

= s1(n, k − 1) − ns1(n, k).

(ii): The recurrence s2(n +1,k) = s2(n, k − 1) + ks2(n, k) that defines the Stirling numbers of the second kind, can be verified by a similar argument using (iii) from Proposition 3.5. 3.5 Examples. (i) We consider some special cases in an extended function algebra: exp −n log −n (1) An,k = s1(n, k) · exp (1’) Bn,k = s1(n, k) · id exp k log k (2) Bn,k = s2(n, k) · exp (2’) An,k = s2(n, k) · id .

(1’) and (2’) immediately follow by Proposition 3.4 from (1) and (2), respec- tively. It is enough to perform the calculation for (1):

exp n−k+1 An,k = An,k(D(exp),...,D (exp)) −(2n−1) = D(exp) · Sn,k(exp,..., exp) −(2n−1) n−1 = exp · exp · Sn,k(1,..., 1) (Remark3.4) −n = s1(n, k) · exp . (Def. s1, Proposition 3.7)

(ii) Since Proposition 3.7 tells us that s1(n, k) are in fact the signed Stirling numbers of the first kind, we choose ϕ = log in the expansion of Proposi- tion 3.3, which implies equation (1.2) in the form n n n log k −n k D (f)= Bn,k · Dlog(f) = id s1(n, k)(id · D) (f). k k X=1 X=1 Analogously combining (2) and (2’) with Proposition 3.1, the reader should verify also the following expansion formula that corresponds to (1.1): n n n k k (id · D) (f)= Dlog(f)= s2(n, k) · id · D (f). k=1 (iii) It may be of interest to show howX some special Stirling numbers can be directly calculated within the machinery of an extended function algebra. Let us, for instance, compute s1(n, 1). By (1) and Proposition 3.4 n exp n log n n s1(n, 1) = exp ·An,1 = exp ·(Bn,1 ◦ exp) = exp ·(D (log) ◦ exp).

16 Observing that s1(n, k) ∈K, we obtain

s1(n, 1) = s1(n, 1) ◦ 0 =1 · (Dn(log) ◦ 1)=(−1)n−1(n − 1)!.

I use the term multivariate Stirling polynomial (MSP) to denote both Sn,k (MSP of the first kind) and Bn,k (MSP of the second kind). Proposition 3.7 may be regarded as a good reason for this eponymy (see also my comments in Section 1 concerning notation and terminology).

3.6 Remark. The Bn,k are widely known as partial exponential Bell polyno- mials (see e. g. [5, 8], also the explicit formula (1.4)). Their complete form n is defined by Bn := k=1 Bn,k. Applying Proposition 3.5 (iii) to each term n ∂Bn of this sum gives the differential recurrence Bn+1 = X1Bn + Xj+1 , P j=1 ∂Xj which originally has been studied by Bell [2]. In [34, p. 49] the complete Bell P polynomials are tabulated up to n = 8.

3.7 Remark. In [20, p. 52 and pp. 481–483] Knuth analyzes, from a com- binatorial point of view, the coefficients of Bn,k in connection with (3.1) thus establishing a recurrence for βn,k. Set βn,k(...) = 0 for partition types ∈/ P(n, k). Then β1,1(1) = 1, and for every (r1,...,rn−k+2) ∈ P(n +1,k):

βn+1,k(r1,...,rn−k+2)= βn,k−1(r1 − 1,r2,...,rn−k+2)+ n−k+1

(rj + 1)βn,k(...,rj +1,rj+1 − 1,... ). j=1 X This could also be obtained more formally by combining Proposi- tion 3.5 (iii) with Corollary 3.6 (ii). I give here without proof also a recurrence for σn,k (to be obtained by using Proposition 3.5 (ii) and Corollary 3.6 (i)). If we agree in an analogous way to let σn,k vanish for partition types ∈/ P(2n − 1 − k, n − 1), then we have σ1,1(1) = 1, and for every (r1,...,rn−k+2) ∈ P(2n +1 − k, n):

σn+1,k(r1,...,rn−k+2) = σn,k−1(r1 − 1,r2,...,rn−k+2)

+ (r1 − 2n + 1)σn,k(r1,r2 − 1,...,rn−k+1) n−k+1

+ (rj + 1)σn,k(r1 − 1,...,rj +1,rj+1 − 1,...,rn−k+1). j=2 X 17 4. A brief summary on Bell polynomials

′ Replacing the coefficient an,k ◦ ϕ in (3.1) (cf. the proof of Proposition 3.3) ϕ by Bn,k, we obtain for n ≥ 0

n n ϕ k D (f ◦ ϕ)= Bn,k · (D (f) ◦ ϕ). (4.1) k X=0 This is, in function-algebraic notation, the well-known Fa`adi Bruno formula (1.3) for the higher derivatives of a composite function. Though it has been known for a long time, it may, from a systematic point of view, appear appropriate to briefly examine here some of the related classical results on Bn,k within our framework. Let F ∈ F[X]. We denote by [F (ϕ) | ϕ = 0 ] the result of substituting 0 for ϕ in the monomial products ϕj of F (ϕ) with j ≥ 1. Example: Let F = ϕ + X2 + 3(1 − X)D(ϕ); then [F (ϕ) | ϕ =0]=3D(ϕ).

4.1 Proposition. For 1 ≤ k ≤ n we have 1 Bϕ = Dn(ϕk) | ϕ =0 . n,k k!   Proof. By (4.1) we obtain

n n k n k ϕ j k D (ϕ )= D (ι ◦ ϕ)= Bn,j · (D (ι ) ◦ ϕ) j=0 X k ϕ j k−j = Bn,j · k ϕ j=1 X k−1 ϕ ϕ j k−1−j = k! · Bn,k + ϕ Bn,j · k ϕ . (4.2) j=1 X Taking ϕ to 0 gives the asserted.

4.2 Proposition. For 1 ≤ k ≤ n we have

k 1 k Bϕ = (−1)k−j ϕk−jDn(ϕj). n,k k! j j=1 X   18 Proof. We rewrite (4.2) in the form

k k Dn(ϕk)= j!Bϕ ϕk−j. j n,j j=1 X   Applying binomial inversion (cf. [1, p. 96-97]) to the equivalent equation

k k ϕ−kDn(ϕk)= ϕ−jj!Bϕ j n,j j=1 X   yields

k k ϕ−kk!Bϕ = (−1)k−j ϕ−jDn(ϕj). n,k j j=1 X   Division by ϕ−kk! then gives the asserted.

Now recall the definition of the subset function β in (1.12).

4.3 Proposition. We have βn,k(r1,...,rn−k+1) = β(r1,...,rn−k+1) for all (r1,...,rn−k+1) ∈ P(n, k), that is,

n! r1 rn−k+1 Bn,k = X ··· X . r1 r − +1 1 n−k+1 r1! ··· rn−k+1! · 1! ··· (n − k + 1)! n k P X(n,k) Proof. It follows from the Leibniz rule (2.4)

n! Dn(ϕk)= Dj1 (ϕ) ··· Djk (ϕ). j1! ··· jk! j1+···+jk=n j1,...,jXk ≥ 0

From this we get by Proposition 4.1

1 n! Bn,k = Xj1 ··· Xjk . (4.3) k! j1! ··· jk! j1+···+jk=n j1, ...,jXk ≥ 1

19 Denote by rm the number of j’s equal to m ∈{1,...,n − k +1}. Then, each sequence (j1,...,jk) in (4.3) is obtained from its corresponding linearly or- k! dered k-tuple i1 ≤···≤ ik by permutations. Hence (4.3) becomes r1!···rn−k+1! 1 n! Bn,k = · Xi1 ··· Xik , r1! ··· rn−k+1! i1! ··· ik! i1+···+ik=n 1≤i1X≤...≤ik

r1 r2 rn−k+1 where i1!i2! ··· ik! = 1! · 2! ··· (n + k + 1)! . This yields the asserted equation, and βn,k (the coefficient function of Bn,k according to Corollary 3.6) is shown to agree with β on P(n, k).

4.1 Remark. Some historical comments related to Fa`adi Bruno’s formula are given in [16]. One example that ‘deserves to be better known’ (Johnson), is a formula stated by G. Scott (1861) (cf. [46] and [16, p.220]). Proposition 4.1 reformulates it in function-algebraic terms. According to [54], the expression ϕ for Bn,k given in Proposition 4.2 is due to J. Bertrand [3, p. 140]). Instead of ‘a not so easy induction’ (Todorov), its verification needs merely applying bino- mial inversion to Scott’s formula. Finally, taking ϕ = exp makes Bertrand’s formula into a well-known explicit expression for the Stirling numbers of the 1 k k−j k n second kind (cf. [1, p. 97]): s2(n, k)= k! j=1(−1) j j . Three corollaries will be useful in laterP sections. 

Because of the relatively simple structure of the Bn,k the partial deriva- tives in Proposition 3.5 (iii) can be given a closed non-differential form. 4.4 Corollary.

∂Bn,k n = B − − (1 ≤ j ≤ n − k + 1). ∂X j n j,k 1 j   Proof. The assertion follows by applying ∂ to the explicit expression of ∂Xj ∂ Bn,k in Proposition 4.3. Observe that takes each (n, k)-partition type ∂Xj into a (n − j, k − 1)-partition type. The details are left to the reader. 4.5 Corollary.

k n r B = X B − − (0,X ,...,X − ). n,k r 1 n r,k r 2 n k+1 r=0 X   20 Proof. Immediate from Proposition 4.3. See also Comtet [8, p. 136]. Notation. (i) The right-hand side of the equation in Corollary 4.5 gives rise to the notation Bn,k := Bn,k(0,X2,...,Xn−k+1). We call Bn,k associated Bell polynomial (or, associated MSP of the second kind). The coefficients of B˜n,k count only partitionse with no singleton blocks. Note thate s ˜2(n, k) := B˜n,k(1,..., 1) are the well-known associated Stirling numbers of the second kind [15, 34]. (ii) We call (unsigned) Lah polynomial the expression

+ r1 rn−k+1 Ln,k := Pω = ω(r1,...,rn−k+1) X1 ...Xn−k+1 , P X(n,k) where ω is the order function in (1.10). We have

n! n − 1 L+ (1,..., 1) = =: l+(n, k). n,k k! k − 1   Let l(n, k) denote the signed Lah numbers (−1)nl+(n, k). Then

+ 1 2 n−k+1 Ln,k((−1) , (−1) ,..., (−1) )= l(n, k), which follows from the observation that r1 + r3 + r5 + ... ≡ n (mod 2) holds whenever (r1,...,rn−k+1) ∈ P(n, k).

+ 4.6 Corollary. Ln,k = Bn,k(1!X1, 2!X2,..., (n − k + 1)!Xn−k+1). Proof. Immediate from Proposition 4.3. See also Comtet [8, p. 134].

5. Inversion formulas and recurrences

We now establish some statements concerning inversion as well as recurrence relations of An,k and Bn,k. The first one is the polynomial analogue of the well-known inversion law satisfied by the Stirling numbers of the first and second kind (see (1.9)).

5.1 Theorem (Inversion Law). For all n ≥ k ≥ 1

n n

An,jBj,k = δn,k and Bn,jAj,k = δn,k. j k j k X= X= 21 5.1 Remark. Defining lower triangular matrices An := (Ai,j)1≤i,j≤n and Bn := (Bi,j)1≤i,j≤n we can rewrite the statements of the theorem as ma- trix inversion formulas, for instance, the first one: AnBn = In (identity matrix) for every n ≥ 1 (which may also be equivalently expressed by means Aϕ Bϕ I of differential terms: n n = n). In the special case X1 = X2 = ... = 1, where the Stirling numbers of the first and second kind are the entries of An and Bn, respectively, both matrices can be considered as transformation ma- trices connecting the linearly independent polynomial sequences (x1,...,xn) and (x1,...,xn) (cf. [1]). Unfortunately, there is much to suggest that this method does not work in our general case. The following proof therefore makes no use of it. Proof. We prove the first equation of Theorem 5.1. Suppose 1 ≤ k ≤ n and n denote by dn,k the sum j=k an,jbj,k. We need to show that dn,k = δn,k. This is clear for n = 1. Now we proceed by induction on n using the differential recurrences in the PropositionsP 3.1 and 3.3. First, observe that applying D to both sides of the induction hypothesis yields D(dn,k)= D(δn,k) = 0, whence n n

D(an,j)bj,k = − an,jD(bj,k). (5.1) j k j k X= X= We then have n

dn+1,k = an+1,n+1 bn+1,k + an+1,j bj,k j=k X n an,n 1 = b + (a − b + D(a )b ) D(ϕ) n+1,k D(ϕ) n,j 1 j,k n,j j,k j=k X n D(bn,k) 1 = a b − + + (a − D(ϕ)b − − + n,n n,k 1 D(ϕ) D(ϕ) n,j 1 j 1,k 1   j=k X n 1 + a − D(b − )) + D(a )b . n,j 1 j 1,k D(ϕ) n,j j,k j k X= Replacing the last sum by the right-hand side of (5.1), we obtain after a short computation: n D(bn,k) D(bn,k) d = a b − + a + a − b − − − a n+1,k n,n n,k 1 n,n D(ϕ) n,j 1 j 1,k 1 n,n D(ϕ) j k X= 22 n

= an,j bj,k−1 = δn,(k−1) = δ(n+1),k. j k− =X1 We conclude from Theorem 5.1 a statement that generalizes ‘Stirling in- version’ for sequences of real numbers; cf. [1, Corollary 3.38 (ii)]. 5.2 Corollary (Inversion of sequences). Let E be an arbitrary overring of Z[X1,...,Xn−k+1], P0,P1,P2,... and Q0, Q1, Q2,... any sequences in E. Then the following conditions are equivalent:

n

(i) Pn = Bn,kQk for all n ≥ 0, k=0 Xn

(ii) Qn = An,kPk for all n ≥ 0. k X=0 5.2 Examples. (i) Theorem 5.1 implies the above mentioned special inver- sion law for the Stirling numbers s1(n, k) = An,k(1,..., 1) and s2(n, k) = Bn,k(1,..., 1) of the first and second kind (see Proposition 3.7): n

s1(n, j)s2(j, k)= δn,k (1 ≤ k ≤ n). j k X= (ii) The signed Lah numbers are known to be self-inverse:

n

l(n, j)l(j, k)= δn,k (1 ≤ k ≤ n). j k X= x In order to prove this, set ϕ(x) = − 1+x . Since ϕ is involutory (ϕ = ϕ) and ϕ(0) = 0, Proposition 3.4 (ii) yields

ϕ ϕ ϕ ϕ An,k(0) = (Bn,k ◦ ϕ)(0) = Bn,k ◦ ϕ(0) = Bn,k(0). It is easily seen that Dj(ϕ)(0) = (−1)jj! for all j ≥ 1. Thus by Corollary 4.6 ϕ we have Bn,k(0) = Bn,k(−1!, 2!, −3!, 4!,...)= l(n, k). Applying Theorem 5.1 then gives the desired result.

5.3 Theorem. For 1 ≤ k ≤ n

(i) An,k = Bn,k(A1,1,...,An−k+1,1),

(ii) Bn,k = An,k(A1,1,...,An−k+1,1).

23 ϕ ϕ Proof. (i): By Proposition 3.4 (ii) An,k = Bn,k ◦ ϕ. Corollary 3.6 yields

n−k+1 ϕ j rj Bn,k ◦ ϕ = βn,k(r1,...,rn−k+1) · (D (ϕ) ◦ ϕ) . P j=1 X(n,k) Y j ϕ ϕ Setting k =1 we get D (ϕ) ◦ ϕ = Bj,1 ◦ ϕ = Aj,1 for every j ≥ 1, hence

ϕ ϕ r1 ϕ rn−k+1 An,k = βn,k(r1,...,rn−k+1)(A1,1) ··· (An−k+1,1) P(n,k) X ϕ ϕ = Bn,k(A1,1,...,An−k+1,1) ϕ = Bn,k(A1,1,...,An−k+1,1) . (ii): Similarly by Proposition 3.4 (i). The equations of Theorem 5.3 can be rewritten as statements about the polynomials Sn,k. 5.4 Corollary.

k−1 (i) Sn,k = X1 · Bn,k(S1,1,...,Sn−k+1,1), 2k−n (ii) Bn,k = X1 · Sn,k(S1,1,...,Sn−k+1,1). Proof. We show only (ii). It follows from Theorem 5.3 (ii)

1 S S B = S 1,1 ,..., n−k+1,1 = X2n−1 × n,k (A )2n−1 n,k X1 2(n−k+1)−1 1 1,1 1 X1 ! − n k+1 n−k+1 − P (2j−1)rj rj j=1 σn,k(r1,r2,...) · (Sj,1) · X1 P − − − j=1 ! (2n X1 k,n 1) Y 2n−1 −(3n−2k−1) = X1 · X1 · Sn,k(S1,1,...,Sn−k+1,1). In the last two lines Corollary 3.6 (ii) has been used. 5.3 Remark. Though equations (i) from both Theorem 5.3 and Corollary 5.4 look like recurrences, their practical (computational) value is rather poor, insofar as they work recursively only for 2 ≤ k ≤ n. For example, one can 2 2 2 3 actually get S5,3 by evaluating X1 ·B5,3(S1,1,S2,1,S3,1) to 45X1 X2 −10X1 X3. It should be noted, however, that the very first members of each generation, Sn,1, are the most complicated, and in this case, of all things, (i) yields the empty statement Sn,1 = Bn,1(S1,1,...,Sn,1), where Bn,1 = Xn.

24 5.4 Remark. Through Corollary 5.4, the particular role of X1 becomes evi- dent. The exponents appearing here have a combinatorial meaning. Given any (r1,...,rn−k+1) ∈ P(2n − 1 − k, n − 1), one has r1 ≥ k − 1 ≥ 0. For par- tition types from P(n, k), a corresponding, however possibly negative lower bound holds: r1 ≥ 2k − n.

5.5 Example. It has already been illustrated that putting Xj = 1(j = 1, 2,... ) converts MSP relations into statements about Stirling numbers. So we may ask what in this regard Theorem 5.3 (i) is about. By Proposition 3.7 we obtain a neat identity for the signed Stirling numbers of the first kind:

s1(n, k)= Bn,k(s1(1, 1),...,s1(n − k +1, 1)). (5.2)

j−1 We know that Bn,k = Pβ (Proposition 4.3) and s1(j, 1) = (−1) (j − 1)! (Examples 3.5 (iii)). Hence a straightforward evaluation of the right-hand n−k side of (5.2) eventually yields s1(n, k) = (−1) c(n, k) with c(n, k) =

P(n,k) γ(r1,...,rn−k+1) (Cauchy’s famous enumeration of n-permutations with exactly k cycles by means of the cycle function γ, (1.11)). Compare P equation [3i] in [8, p. 135] for a signless version of (5.2). We will now establish one recurrence relation that is satisfied by both An,k and Bn,k. However, as with Theorem 5.3, the recurrence does not work for k = 1, since Aj,1 and Bj,1 (1 ≤ j ≤ n − k + 1) are needed as initial values.

5.5 Proposition. Let n, k be integers with 1 ≤ k ≤ n. Then we have

n−k +1 n − 1 (i) A = A A − − , n,k j − 1 j,1 n j,k 1 j=1 X   n−k +1 n − 1 (ii) B = B B − − . n,k j − 1 j,1 n j,k 1 j=1 X   Proof. (i) can be easily inferred from (ii): We transform B into Bϕ and apply ◦ ϕ (from the right) on both sides of the equation. Then, Proposition 3.4 (ii) yields the desired statement. — (ii): Eliminating the partial derivative in Proposition 3.5 (iii) by Corollary 4.4 leads to

n− −k 1 +1 n − 1 B = X B − − + X B − − − n,k 1 n 1,k 1 j+1 j n 1 j,k 1 j=1 X   25 n−k +1 n − 1 = X B − − . j j − 1 n j,k 1 j=1 X  

The observation Xj = Bj,1 completes the proof.

5.6 Corollary.

n−k +1 n − 1 B˜ = X B˜ − − (1 ≤ k ≤ n). n,k j − 1 j n j,k 1 j=2 X   5.6 Remark. Proposition 5.5 (ii) is stated in Charalambides [5] (together with a proof; see ibid., p. 415). In the special case k = 2 we have Bj,1Bn−j,k−1 = XjXn−j. Thus (ii) becomes a simple explicit formula for Bn,2; see also [34, p. 48]. 5.7 Remark. From Corollary 5.6 it follows by an inductive argument that ˜ ˜ n B2n−l,n = 0, if l ≥ 1, and B2n,n = (1 · 3 · 5 ··· (2n − 1))X2 .

5.8 Examples (Proposition 5.5). The substitution Xj = 1 makes (i) and (ii) into statements about Stirling numbers. Observe s2(j, 1) = 1 and s1(j, 1) = (−1)j−1(j − 1)! (Examples 3.5 (iii)). Then, a little calculation yields the following well-known identities (cf. [34, p. 42-43] and [20, p. 68]):

n n n +1 n! j n +1 n j = , = . k +1 j! k k +1 j k j k j k   X=     X=   

6. Explicit formulas for Sn,k

We now pass to the task of finding a fully explicit expression for Sn,k. At first glance it seems a viable idea to get Sn,k by elimination from the inversion law (Theorem 5.1), which is a linear system already in triangular form. In fact, one can verify by induction that every MSP of the first kind can be expressed in terms of Bell polynomials. For instance, in the leading case k =1 < n we obtain

n−2 Sn,1 = − X1 Bn,1+ n−2 r+1 (n−2)−(j1+···+jr) (6.1) (−1) X1 Bn,jr Bjr,jr−1 ··· Bj1,1. r=1 1

2 3 Cn,k,1X1 + Cn,k,2X1 + Cn,k,3X1 + ··· , whose coefficients Cn,k,r do neither contain X1 nor products of two or more Bell polynomials. The main result we are going to establish in Theorem 6.1 may be regarded as a non-trivial counterpart of Corollary 4.5; it indeed expresses all Sn,k (and consequently An,k) in terms of associated Bell polynomials. 6.1 Theorem. For n ≥ k ≥ 1

n−1 n−1−r 2n − 2 − r r S = (−1) X B˜ − − − − − . n,k k − 1 1 2n 1 k r,n 1 r r k− =X1   Proof. The proof is divided into two parts. First we will show by induction2 that there are polynomials Cn,k,r ∈ Z[X2,...,Xn−k+1] such that

n−1 r Sn,k = Cn,k,rX1 , (6.2) r k− =X1 where the Cn,k,r are uniquely determined by a certain differential recurrence. Therefore, in the second step, it remains to show that the recurrence is r satisfied by the coefficients of X1 in the asserted equation of the theorem. n−1 1. For k = n we have Sn,n = X1 (Remark 3.3), and taking Cn,n,n−1 =1 satisfies (6.2). This includes the case n = 1. Now let n ≥ 1 and suppose (6.2) holds for all k ∈{1,...,n}, where Cn,k,r ∈ Z[X2,...,Xn−k+1], k − 1 ≤ (a) (b) (c) r ≤ n − 1. Proposition 3.5 (ii) yields Sn+1,k = Tn,k − Tn,k + Tn,k with

(a) (b) Tn,k = X1Sn,k−1, Tn,k = (2n − 1)X2Sn,k, and

2The reciprocity law for Stirling polynomials, which is proved in [42], results in a new proof that is independent of Theorem 6.1 and does not use induction.

27 n−k+1 (c) ∂Sn,k Tn,k = X1Xj+1 . ∂Xj j=1 X By the induction hypothesis

n−1 n (a) r r Tn,k = X1 Cn,k−1,rX1 = Cn,k−1,r−1X1 , r k− r k− =X2 =X1 n−1 (b) r Tn,k = (2n − 1)X2Cn,k,rX1 , r k− =X1 n− 1 ∂ T (c) = X X (C Xr) n,k 1 2 ∂X n,k,r 1 r k− 1 =X1 n−k+1 n−1 ∂ r + X1Xj+1 (Cn,k,rX1 ) ∂Xj j=2 r k− X =X1 n−1 n n−k+1 r ∂Cn,k,r−1 r = rX2Cn,k,rX1 + Xj+1 X1 . ∂Xj r k− r k j=2 ! =X1 X= X

For the desired expansion of Sn+1,k, we now have to find polynomials Cn+1,k,r ∈ Z[X2,...,Xn−k+2] that satisfy the recurrence

n−k+1 ∂Cn,k,r−1 Cn+1,k,r = Cn,k−1,r−1 − (2n − 1 − r)X2Cn,k,r + Xj+1 (6.3) ∂Xj j=2 X together with Cn,k,r = 1, if n = k = r + 1, and Cn,k,r = 0, if k = 0 or k ≤ r = n − 1 or r

n−1−r 2n − 2 − r C := (−1) B˜ − − − − − . (6.4) n,k,r k − 1 2n 1 k r,n 1 r  

It suffices to show that Cn,k,r meets the above conditions. First we check the initial values. Since B˜2n−1−k−r,n−1−r = B˜2(n−1−r)−l,n−1−r where l = k − 1 − r, we get by Remark 5.7: Cn,k,r = 0 for l ≥ 1, that is, for r

28 Next we substitute (6.4) into (6.3). We start with the last summand on the right-hand side of (6.3) by first evaluating the partial derivatives:

∂Cn,k,r−1 n−r 2n − 1 − r ∂ =(−1) B˜ − − − . ∂X k − 1 ∂X 2n k r,n r j   j Corollary 4.4 now yields for j ≥ 2

∂ 2n − k − r B˜ − − − = B˜ − − − − − . ∂X 2n k r,n r j 2n k r j,n r 1 j   Substituting the remaining C-terms into (6.3), we obtain after some straight- forward calculation, in particular cancelling (−1)n−r,

2n − r 2n − 1 − r − B˜ − − − = k − 1 k − 2 2n+1 k r,n r     2n − 2 − r (2n − 1 − r) X B˜ − − − − − k − 1 2 2n 1 k r,n 1 r   n−k 2n − 1 − r +1 2n − k − r + X B˜ − − − − − . (6.5) k − 1 j j+1 2n k r j,n 1 r j=2   X   Using the identity

2n − 2 − r 2n − 1 − r (2n − 1 − r) = (2n − k − r) , k − 1 k − 1     equation (6.5) is equivalently reduced to

n−k +1 2n − k − r B˜ − − − = X B˜ − − − − − , 2n+1 k r,n r j j+1 2n k r j,n 1 r j=1 X   which is the statement of Corollary 5.6. This completes the proof. 6.1 Remark. (i) Taking k = 1 on the right-hand side of the equation in The- orem 6.1, we obtain Comtet’s expansion (1.7) (see [8, p. 151]). It should, however, be pointed out that (1.7) was intended for a solution of the La- grange inversion problem (see Section 7 below). In this context, the idea of n a connection between f n in (1.7) and the Lie derivatives (θD) studied by the same author [7] did not come into play.

29 (ii) Sylvester’s note On Reciprocants (1886) deals with the task ‘to express the successive derivatives of x in regard to y in terms of those of y in regard to x’. The solution is given in the form of the polynomial terms Sn,1,1 ≤ n ≤ 7, probably one of the earliest times in the mathematical literature (cf. [53, p. 38] and [52]).

Theorem 6.1 enables us to find closed expressions for the coefficients σn,k in Corollary 3.6 (i), thus leading to an explicit formula for Sn,k, which corre- sponds to that for the Bell polynomials in Proposition 4.3.

6.2 Proposition. For every (r1,...,rn−k+1) ∈ P(2n − 1 − k, n − 1) we have

n−1−r1 2n − 2 − r1 σ (r ,...,r − )=(−1) × n,k 1 n k+1 k − 1   × β2n−1−k−r1,n−1−r1 (0,r2,...) n−1−r1 =(−1) β(k − 1,r2,...,rn−k+1).

Proof. We first consider the coefficients of the associated Bell poly- nomials. B˜2n−1−k−r,n−1−r is a sum over all partition types from P(2n − 1 − k − r, n − 1 − r), whose number r1 of one-element blocks is zero. Let us abbreviate to P˜0 the set of these ‘associated’ partition types. Then, according to Proposition 4.3 we can write

B˜2n−1−k−r,n−1−r = β2n−1−k−r,n−1−r(0,r2,...,rn−k+1)× ˜ XP0 r2 rn−k+1 × X2 ...Xn−k+1 .

Observe now that any partition type (r1,...,rn−k+1) ∈ P(2n−1−k, n−1) can be obtained by putting some r = r1 ∈ {k − 1,...,n − 1} in the first place of (0,r2,...,rn−k+1) ∈ P˜0 (cf. Remark 5.4). This means that the double n−1 summation r=k−1 P˜0 in Theorem 6.1 can be replaced by P(2n−1−k,n−1). Comparing the coefficients of Xr1 ·...·Xrn−k+1 in the resulting expression for P P 1 n−k+1 P Sn,k and in that of Corollary 3.6 (i) then finally yields the desired σn,k. We now add to the list (1.10–1.12) in Section 1 a new mapping σ : P −→ Z, that might be called Stirling function, defined by

(2n − 2 − r )! σ(r ,r ,r ,...)=(−1)n−1−r1 1 . (6.6) 1 2 3 r2 r3 (k − 1)!r2!r3! · ... · (2!) (3!) · ...

30 Proposition 6.2 shows that σn,k and σ agree on P(2n − 1 − k, n − 1).— Another succint expression of the relationship between β and σ is the follow- ing identity, which holds for all (2n − 1 − k, n − 1)-partition types. 6.3 Corollary.

2n − 1 − k 2n − 2 − r σ(r ,r ,...)=(−1)n−1−r1 1 β(r ,r ,...). r 1 2 k − 1 1 2  1    One could equivalently rewrite this equation as well for (n, k)-partition types as follows: n 2k − r σ(r ,r ,...)=(−1)k−r1 1 β(r ,r ,...). r 1 2 2k − n 1 2  1   Here, however, the assumption 2k ≥ n must be satisfied (see Remark 5.4).

6.2 Example. The explicit form of σn,k can be used to express s2(n, k) as a sum over P(2n − 1 − k, n − 1). Suppose F is an extended function algebra. Then, log from (2’) in Examples 3.5 (i) we obtain s2(n, k)= An,k ◦ 1. Corollary 3.6 (i) implies

log −(2n−1) 1 r1 2 r2 An,k = D(log) σn,k(r1,r2,...)D (log) D (log) · ... X On the right-hand side we have Dj(log) = (−1)j−1(j − 1)! · id−j for 1 ≤ j ≤ n − k + 1, and a straightforward calculation using (6.6) yields

2n−2 r1−(k−1) k−1 s2(n, k)= (−1) 2n−2 · γ(r1,...,rn−k+1). (6.7) P − − − r1  (2n X1 k,n 1)  6.3 Remark. It looks like Corollary 6.3 expresses a kind of reciprocity law for the coefficient functions β and σ. However, it remains open whether there 3 is a connection with the ‘duality law’ for the Stirling numbers : s2(n, k) = c(−k, −n) (cf. [12, p.25] and [19, p.412]). In view of the fact that many MSP relations carry over more or less verbatim to Stirling numbers (see, e. g., Remark 4.1, Theorem 5.1, Corollary 5.2, Example 5.8), we could, in the reverse direction, try to establish a reciprocity law for the corresponding

3My paper [42] answers this question in the affirmative.

31 polynomials: Bn,k = Z−k,−n. In the case n ≥ k ≥ 0 we define Zn,k := Pγ (see (1.11), the sum to be taken over P(n, k)). The meaning of both Bn,k and Cn,k can now be extended to arbitrary n ∈ Z with the aid of

k n n−k−j B − = X B˜ , n,n k k + j 1 k+j,j j=0 X   which follows from Corollary 4.5 and also applies to the Cn,n−k (compare the corresponding statements concerning associative Stirling numbers in [15]; see also [5, ch. 8, exercises 10, 12]). Again we actually have s2(n, k) = Bn,k(1,..., 1) = C−k,−n(1,..., 1) = c(−k, −n). However, Bn,k = Z−k,−n in general turns out to be false. −8 3 −7 −6 Counterexample: B5,2 =6 C−2,−5 = 105X1 X2 −120X1 X2X3 +30X1 X4, whereas B5,2(1, 1, 1, 1) = Z−2,−5(1, 1, 1, 1) = 15. Equation (i) of the theorem to follow is the MSP version of a well-known formula (cf. Example 6.4 below) expressing the signed Stirling numbers of the first kind in terms of the Stirling numbers of the second kind. 6.4 Theorem (Schl¨omilch type formulas). For n ≥ k ≥ 1

n− 1 2n − 2 − r 2n − k (i) A = (−1)n−1−r × n,k k − 1 r +1 − k r=k−1    X r−2n+1 × X1 B2n−1−k−r,n−1−r, n− 1 2n − 2 − r 2n − k (ii) B = (−1)n−1−r × n,k k − 1 r +1 − k r=k−1    X 2n−1−r × X1 A2n−1−k−r,n−1−r. Proof. For the purpose of formal convenience we adhere to the convention Bi,j = 0 (and consequently B˜i,j = 0), if j < 0. Then, the equation of Corollary 4.5 can be rewritten in the form n n n−j B = X B˜ − − , n,k j 1 j,k (n j) j=0 X   whence by binomial inversion

k j n j B˜ = (−1) X B − − . (6.8) n,k j 1 n j,k j j=0 X   32 The upper limit n has been replaced here by k, since we have Bn−j,k−j = 0 for j>k. Substituting (6.8) into the equation of Theorem 6.1 gives

−(2n−1) An,k = X1 Sn,k n− n− −r 1 1 2n − 2 − r 2n − 1 − k − r = (−1)n−1−r−j × k − 1 j r=k−1 j=0    X X r−2n+1+j × X1 B2n−1−k−r−j,n−1−r−j. Define now a new index s = r + j so that k − 1 ≤ s ≤ n − 1. Interchanging the order of summation then leads to

n−1 n−1−s s−2n+1 An,k = (−1) X1 B2n−1−k−s,n−1−s ×··· s=k−1 X s 2n − 2 − r 2n − 1 − k − r ···× . k − 1 s − r r k− =X1    (∗)

The sum (∗) can| be simplified by{z using elementary properties} of the binomial 2n−2−r 2n−2−s coefficients. First check that its summand is equal to 2n−2−s k−1 . Then, after some calculation, (∗) is reduced to   s 2n − 2 − s 2n − 2 − r 2n − 2 − s 2n − k (∗)= = . k − 1 2n − 2 − s k − 1 s +1 − k r k−  =X1      Finally, renaming the index s to r gives the assertion (i). Now we derive (ii) from (i). Using Proposition 3.4 and applying P 7−→ P ϕ ϕ to both sides of (i) makes the left-hand side into Bn,k ◦ ϕ, while the term r−2n+1 r−2n+1 ϕ X1 B2n−1−k−r,n−1−r on the right becomes D(ϕ) B2n−1−k−r,n−1−r. Next, we apply ◦ϕ (from the right) to both sides of (i). It follows

n− 1 2n − 2 − r 2n − k Bϕ = (−1)n−1−r × n,k k − 1 r +1 − k r k− =X1    r−2n+1 ϕ × (D(ϕ) ◦ ϕ) · (B2n−1−k−r,n−1−r ◦ ϕ).

r−2n+1 2n−1−r 2n−1−r ϕ Observe now that D(ϕ) ◦ ϕ = D(ϕ) =(X1 ) and, again by ϕ ϕ Proposition 3.4, B2n−1−k−r,n−1−r ◦ ϕ = A2n−1−k−r,n−1−r. This completes the

33 proof. (The argument amounts to directly applying Theorem 5.3 (ii), that is, replacing each Xj with Aj,1 on both sides of (i).)

6.4 Example. Specializing Theorem 6.4 (i) by taking all indeterminates to 1, we immediately obtain Schl¨omilch’s formula for the Stirling numbers of the first kind (see e. g. [5]):

n− 1 2n − 2 − r 2n − k s (n, k)= (−1)n−1−r × (6.9) 1 k − 1 r +1 − k r k− =X1    × s2(2n − 1 − k − r, n − 1 − r).

Likewise we get from Theorem 6.1 a slightly shorter formula of a similar type:

n− 1 2n − 2 − r s (n, k)= (−1)n−1−r s˜ (2n − 1 − k − r, n − 1 − r). (6.10) 1 k − 1 2 r k− =X1  

(6.10) runs with smaller numbers than does (6.9). Compare, for instance, the computation of s1(7, 4) = −84 · 90+56 · 150 − 35 · 45 = −735 by Schl¨omilch’s formula (6.9) with that of s1(7, 4) = −84 · 15 + 56 · 10 − 35 · 1 = −735 by (6.10). However, associated Stirling numbers of the second kind do not have quite simple explicit representations (cf. [15], in particular equation (3.11), ibid.). Finally it should be noted that the statement (ii) of Theorem 6.4 enables the Stirling numbers of the second kind to be represented using the Stirling numbers of the first kind. This result is due to Gould [13].

7. Remarks on Lagrange inversion

Let ϕ be an analytic function (bijective in the real or complex domain), which is given in the form of a power series

xn ϕ(x)= f n n! n≥1 X with zero constant term and f1 =6 0. Then, the compositional inverse ϕ is unique, and one may ask for the coefficients f n in the series expansion

34 n ϕ(x) = n≥1 f nx /n!. One possible answer to this is Lagrange’s famous inversion formula (cf. [8] and [50] for proofs and further references): P d n−1 x n f = . (7.1) n dx ϕ(x)      x=0

This innocent looking expression, however, provides in most cases an all but simple method of computing. Many attempts have therefore been made to obtain alternative and more efficient expressions (see, e. g., [22], [54]). Try- ing to express f n as a function of f1, f2,...,fn is a quite natural approach. Morse and Feshbach [30, p. 412], for example, employed the Residue The- orem to show that f n can be represented as a polynomial expression over all partitions of n − 1. Comtet [8] derived the remarkable result that the n−1 k −n−k right-hand side of (7.1) is equal to k=0 (−1) f1 Bn+k−1,k(0, f2,...,fn). By Theorem 6.1 the latter can easily be seen to agree with An,1(f1,...,fn), that is, we have for all n ≥ 1: P

f n = An,1(f1,...,fn). (7.2)

In the following we will deal with some few function-algebraic aspects of Lagrange inversion. It turns out that (7.2) can be proved without using (7.1). We will also see that inverting a function ϕ is nothing else n n than switching from Dι (ϕ)(0) to Dϕ(ι)(0) in the corresponding series ex- pansions. A few examples should briefly illustrate the computational aspects.

Let F be an extended function algebra, ϕ ∈F. Throughout this section 4 we assume f0 := ϕ ◦ 0 = 0, and f1 := D(ϕ) ◦ 0 to be a unit. Since the notion of convergence has no place in F, we make use of formal power series. 7.1 Definition. We say that ϕ is an exponential generating function of the sequence of constants f0, f1, f2,... ∈ K (symbolically written ϕ(x) = n n 5 n≥0 fnx /n!), if D (ϕ)(0) = fn for every n ≥ 0. P The exponential generating function of the constant sequence 1, 1, 1,... xn ex := exp(x) := n! n≥0 X

4In the sequel we use the customary notation g(0) instead of g ◦ 0, g ∈F 5 In the case f0 = 0 we take the lower summation limit to 1.

35 obviously has the properties that one would expect from an exponential (as has been indicated in Section 2.2). Our basic statement on inversion is now the following 7.1 Proposition. xn xn (i) ϕ(x)= Dn (ϕ)(0) = Bϕ (0) , id n! n,1 n! n≥1 n≥1 X X xn xn (ii) ϕ(x)= Dn(id)(0) = Aϕ (0) . ϕ n! n,1 n! n≥1 n≥1 X X

Proof. (i): Clearly Did = D. Hence for all n ≥ 1

ϕ ϕ n n Bn,1(0) = (Xn) (0) = D (ϕ)(0) = Did(ϕ)(0). (ii): By Proposition 3.4 (ii) we have

ϕ ϕ n An,1(0) = (Bn,1 ◦ ϕ)(0) = (D (ϕ) ◦ ϕ)(0). Thus Pourchet’s formula yields

n n n n Dϕ(id)(0) = (D (ϕ) ◦ ϕ)(0) = D (ϕ)(ϕ(0)) = D (ϕ)(0).

7.2 Corollary. For every n ≥ 1

f n = An,1(f1,...,fn) 1 (−1)n−1−r1 · (2n − 2 − r )! = · 1 f r1 f r2 ··· f rn . 2n−1 r2 r 1 2 n f r2! ··· rn! · (2!) ··· (n!) n 1 P − − (2nX2,n 1) ϕ Proof. Proposition 7.1 (ii) yields f n = An,1(0) = An,1(f1,...,fn), where fj = Dj(ϕ)(0) (j =1,...,n). Then, applying Corollary 3.6 (i) and (6.6) for k =1 gives the asserted.

n Except for some simple cases, the higher Lie derivatives Dϕ(id) turn out to be considerably less complex than the Lagrangian terms Dn−1((ι/ϕ)n). Nonetheless, the most advantage is presumably to be gained from applying j Theorem 6.1 (with k = 1) or Corollary 7.2 to the coefficients fj = D (ϕ)(0) (j =1, 2,...,n). Let us consider three examples.

36 7.1 Example. Let ϕ(x) = xe−x. As is well-known (see, e. g., [50, p. 23]), the inverse ϕ is exponential generating function of the sequence r(n) (n = 1, 2, 3,...) of numbers of rooted (labeled) trees on n vertices. The function ϕ seems to be tailored to the application of (7.1), for we have Dn−1((id/ϕ)n)(x) = Dn−1(exp ◦(n · id))(x) = nn−1enx and thus readily by n−1 nx n−1 (7.1): r(n)= f n = n e |x=0 = n . Compared with this neat calculation, applying An,1 to the coefficients j j−1 fj = D (ϕ)(0) = (−1) j (j = 1,...,n) gives less satisfactory results. Corollary 7.2 yields (2n − 2 − r )! 1 (−1)r1 1 · . (7.3) r2 r3 rn r2!r3! ··· rn! 1! 2! ··· (n − 1)! P − − (2nX2,n 1) Here a combinatorial argument is needed to see that (7.3) equals nn−1. 7.2 Example. Let ϕ(x) = ex − 1. While formula (7.1) fails to yield sim- ple expressions, the Taylor series expansion of ϕ = log ◦(1 + id) is immedi- ϕ ately obtained by Proposition 7.1 (ii). We have An,1(0) = An,1(1,..., 1) = n−1 s1(n, 1) = (−1) (n − 1)! (see Examples 3.5 (iii)), and hence log(1 + x) = n−1 n n≥1(−1) x /n.

7.3P Example. Let ϕ(x) =1+2x − ex. According to Stanley [50, p. 13], ϕ is exponential generating function of the sequence t(n) (n = 1, 2, 3,...), where t(n) denotes the number of total partitions of {1,...,n}; cf. the fourth problem posed by E. Schr¨oder (1870) (‘arbitrary set bracketings’) [50, p. 178]. When applied to ϕ, the Lagrange formula does not ‘seem to yield anything interesting’ (Stanley). Let us therefore have a look at what can be achieved with the help of the MSPs. We have f1 =1, fj = −1 (j ≥ 2), and hence by Corollaries 6.3 and 7.2

2n − 2 −1 t(n)= An,1(1, −1,..., −1) = β(r1,...,rn). r1 P − − (2nX2,n 1)   The same can be expressed as well in terms of associated Stirling numbers of the second kind by applying Theorem 6.1:

n−1

t(n)= s˜2(2n − 2 − r, n − 1 − r). r=0 X Regarding the latter formula, the reader is referred to Comtet [8, p. 224].

37 n Alternatively, we can use the fact that f n = Dϕ(id)(0) (cf. Proposi- tion 7.1 (ii)). This leads to a recursive solution. The repeated application of the Lie derivation Dϕ inductively results in a representation of the form

n x −(2n−1) Dϕ(id)(x)=(2 − e ) Tn(x), where n−1 kx T1(x)=1, Tn(x)= bn−1,ke for n ≥ 2, k X=0 with non-negative integers bn−1,k. It follows for all n ≥ 1

t(n)= Tn(0) = bn−1,0 + bn−1,1 + ··· + bn−1,n−1.

kx n+1 n Equating now the coefficients of e in Dϕ (id)(x) = Dϕ(Dϕ(id))(x) gives the recurrence

bn,k = (2n − k)bn−1,k−1 +2kbn−1,k (1 ≤ k ≤ n),

bi,0 = δi,0, bi,j =0 (0 ≤ i < j). Some special values:

n−1 n−1 n bn,1 =2 , bn,2 =2 (2 − n − 1), bn,n = n!.

We obtain t(1) = 1, t(2) = 0 + 1 = 1, t(3) = 0+2+2 = 4, t(4) = 0+22 +22(23 − 3 − 1) + 3! = 26.

Our last statement concerns exponential generating functions of the form exp ◦(t · ϕ), t ∈K. 7.3 Proposition.

n xn (i) etϕ(x) = Bϕ (0) tk , n,k n! n≥0 k ! X X=0 n xn (ii) etϕ(x) = Aϕ (0) tk . n,k n! n≥0 k ! X X=0 Proof. (i): By Fa`adi Bruno’s formula (4.1) and Remark 2.3

n n n tϕ tϕ k tϕ ϕ k D (e )= Bn,k · (D (exp) ◦ (tϕ)) = e Bn,k t , k k X=0 X=0 38 hence n n tϕ tϕ(0) ϕ k D (e )(0) = e Bn,k(0) · (t ◦ 0) . k X=0 Since etϕ(0) = et·0 = e0 = 1 and t ◦ 0= t, we are done. Note ϕ(0) = 0; then (ii) follows from (i) by Proposition 3.4. We conclude with a well-known example. 7.4 Example. As in Example 7.2, take ϕ = exp −1. From (2) in Exam- ϕ ples 3.5 (i) we have Bn,k(0) = s2(n, k). Then by Proposition 7.3 (i) n xn exp(t(ex − 1)) = s (n, k) tk . 2 n! n≥0 k ! X X=0 Put t = 1. This shows that exp(ex − 1) is exponential generating function of n the sequence b(n) := k=0 s2(n, k) (see [50, p. 13]). ϕ We also have An,k(0) = s1(n, k) (from (1) in Examples 3.5 (i)). Since ϕ = log ◦(1 + ι), Proposition 7.3 (ii)P yields

n xn exp(t log(1 + x)) = s (n, k) tk . (7.4) 1 n! n≥0 k ! X X=0 Thus (1 + x)t turns out to be the exponential generating function of the n k sequence k=0 s1(n, k) t (n =1, 2, 3,...) (see, e. g., [5, p. 281]). P 8. Concluding remarks The preceding work was primarily intended as an attempt to introduce, within a general function-algebraic setting, the notion of MSP of the first and second kind (the latter being identical to that of partial Bell polyno- mial). The investigation was focussed on establishing the inverse relationship as well as other fundamental properties of the two polynomial families. The resulting picture shows that the MSPs may be understood as a kind of strong generalizations of the corresponding Stirling numbers. Supplementary to this the reader is referred to a package [39] for Math- ® ematica providing functions that generate the expressions Sn,k, An,k, and Bn,k together with a substitution mechanism. No attempt was made here to develop a combinatorial interpretation of the coefficient function σ : P −→ Z (in Sn,k and An,k). Based on [37],

39 Haiman and Schmitt [14] have offered a satisfactory explanation of Comtet’s expansion (1.7) (cf. Remark 6.1) from an incidence algebra point of view (using colored partitions of finite colored sets). So, it might be worth-while examining whether this idea will also apply to the general case formulated in Theorem 6.1. The ‘ban on one-element blocks’ observed in the case k =1 [14, p. 180] may be seen as a sign in this direction, since it is as well a striking feature of the whole family Sn,k (see, e. g., Remark 5.4, Proposition 6.2, and (6.6)).

40 MSP of the 1st kind MSP of the 2nd kind

S1,1 =1 B1,1 = X1

S2,1 = −X2 B2,1 = X2 2 S2,2 = X1 B2,2 = X1 2 S3,1 =3X2 − X1X3 B3,1 = X3

S3,2 = −3X1X2 B3,2 =3X1X2 2 3 S3,3 = X1 B3,3 = X1 3 2 S4,1 = −15X2 + 10X1X2X3 − X1 X4 B4,1 = X4 2 2 2 S4,2 = 15X1X2 − 4X1 X3 B4,2 =4X1X3 +3X2 2 2 S4,3 = −6X1 X2 B4,3 =6X1 X2 3 4 S4,4 = X1 B4,4 = X1 4 2 2 2 S5,1 = 105X2 − 105X1X2 X3 + 10X1 X3 B5,1 = X5 2 3 +15X1 X2X4 − X1 X5 3 2 3 S5,2 = −105X1X2 + 60X1 X2X3 − 5X1 X4 B5,2 =5X1X4 + 10X2X3 2 2 3 2 3 S5,3 = 45X1 X2 − 10X1 X3 B5,3 = 15X1X2 + 10X1 X3 3 3 S5,4 = −10X1 X2 B5,4 = 10X1 X2 4 5 S5,5 = X1 B5,5 = X1 5 3 2 2 S6,1 = −945X2 + 1260X1X2 X3 − 280X1 X2X3 B6,1 = X6 2 2 3 − 210X1 X2 X4 + 35X1 X3X4 3 4 + 21X1 X2X5 − X1 X6 4 2 2 3 2 2 S6,2 = 945X1X2 − 840X1 X2 X3 + 70X1 X3 B6,2 = 10X3 + 15X2X4 +6X1X5 3 4 + 105X1 X2X4 − 6X1 X5 2 3 3 4 3 2 S6,3 = −420X1 X2 + 210X1 X2X3 − 15X1 X4 B6,3 = 15X2 + 60X1X2X3 + 15X1 X4 3 2 4 2 2 3 S6,4 = 105X1 X2 − 20X1 X3 B6,4 = 45X1 X2 + 20X1 X3 4 4 S6,5 = −15X1 X2 B6,5 = 15X1 X2 5 6 S6,6 = X1 B6,6 = X1

Table 1: Generations 1 – 6 of the multivariate Stirling polynomials of the first kind (Sn,k) and of the second kind (Bn,k: partial exponential Bell polynomials).

41 References

[1] M. Aigner: Combinatorial Theory (Grundlehren der mathematischen Wissenschaften, 234), Springer-Verlag: Berlin – New York (1979).

[2] E. T. Bell: Exponential polynomials. Annals of Mathematics 35 (1934), 258–277.

[3] J. Bertrand: Trait´ede calcul diff´erentiel et de calcul int´egral (Pre- mi`ere partie – Calcul diff´erentiel). Gauthier-Villars, Paris (1864).

[4] R.P. Brent and H.T. Kung: Fast algorithms for manipulating for- mal power series. J. Assoc. Comput. Mach 25 (1978), 581–595.

[5] Ch. A. Charalambides: Enumerative Combinatorics. Chapman & Hall/CRC, Boca Raton 2002.

[6] A. Connes and D. Kreimer: Hopf algebras, renormalization and noncommutative geometry. Commun. Math. Phys. 199 (1998), 203–242; hep-th/9808042.

[7] L. Comtet: Une formule explicite pour les puissances successives de l’op´erateur de d´erivations de Lie. C. R. Acad. Sci. 276 (1973), 165–168.

[8] ————: Advanced Combinatorics, rev. and enlarged edition. Reidel, Dordrecht (Holland) 1974.

[9] A. D. D. Craik: Prehistory of Fa`adi Bruno’s Formula. Amer. Math. Monthly 112 (2005), 119–130.

[10] W. A. Dudek and V. S. Trokhimenko: Algebras of Multiplace Func- tions. De Gruyter, Berlin/Boston (2012).

[11] H. Figueroa and J. M. Gracia-Bond´ıa: Combinatorial Hopf alge- bras in quantum field theory I. Review in Mathematical Physics 17 (8) (2005), 881–976.

[12] I. Gessel and R. P. Stanley: Stirling polynomials. J. Comb. Theory (A) 24 (1978), 24–33.

[13] H. W. Gould: representation problems. Proc. Amer. Math. Soc. 11 (1960), 447–451.

42 [14] M. Haiman and W. R. Schmitt: Incidence algebra antipodes and Lagrange inversion in one and several variables. J. Comb. Theory (A) 50 (1989), 172–185.

[15] F. T. Howard: Associated Stirling numbers. The Fibonacci Quarterly 18 (1980), 303–315.

[16] W. P. Johnson: The curious history of Fa`adi Bruno’s formula. Amer. Math. Monthly 109 (2002), 217–234.

[17] ————: Combinatorics of higher derivatives of inverses. Amer. Math. Monthly 109 (2002), 273–277.

[18] C. Jordan: Calculus of Finite Differences, 1st ed. Budapest (1939); 2nd ed., Repr., Chelsea Publ. Co., Inc., New York (1950).

[19] D. E. Knuth: Two notes on notation. Amer. Math. Monthly 99 (1992), 403–422.

[20] ————: The Art of Computer Programming, 3rd ed., Vol. 1: Funda- mental Algorithms. Addison Wesley Longman (1997).

[21] E. R. Kolchin: Differential algebra and algebraic groups. New York and London, 1973.

[22] Ch. Krattenthaler: Operator methods and Lagrange inversion: A unified approach to Lagrange formulas. Trans. Amer. Math. Soc. 305 (1988), 431–465.

[23] D. Kreimer: On the Hopf algebra structure of perturbative field the- ory. Adv. Theor. Math. Phys. 2.2 (1998), 303–334; q-alg/9707029.

[24] I. Lah: Eine neue Art von Zahlen, ihre Eigenschaften und Anwendung in der mathematischen Statistik. Mitt.-Bl. Math. Statistik. 7 (1955), 203–212.

[25] T. Mansour and M. Schork: Commutation Relations, Normal Or- dering, and Stirling Numbers. CRC Press, Boca Raton (2016).

[26] K. Menger: Algebra of Analysis, no. 3 in Notre Dame Mathematical Lectures. Notre Dame, Indiana (1944).

43 [27] ————: The algebra of functions: past, present, future. Rend. Mat. Appl., V. Ser. 20 (1961), 409–430.

[28] M. Mihoubi: Polynˆomes multivari´es de Bell et polynˆomes de type bi- nomial. Th`ese de Doctorat, Universit´ede Sciences et de la Technologie Houari Boumedienne (2008).

[29] Z.˘ Mijajlovic´ and Z. Markovic´: Some recurrence formulas related to the differential operator θd. Facta Universitatis (NIS)˘ 13 (1998), 7– 17.

[30] P.M. Morse and H. Feshbach: Methods of Theoretical Physics, Vol. I. McGraw-Hill Book Co., Inc., New York (1953).

[31] N. Nielsen: Recherches sur les polynˆomes et les nombres de Stirling. Annali di Matematica pura ed applicata (series 3) 10 (1904), 287–318.

[32] ————: Uber¨ die Stirlingschen Polynome und die Gammafunktion. Monatshefte f¨ur Mathematik und Physik 16 (1) (1905), 135–140.

[33] J. Quaintance and H. W. Gould: Combinatorial Identities for Stir- ling Numbers. The Unpublished Notes of H. W. Gould. World Scientific 2016.

[34] J. Riordan: Introduction to Combinatorial Analysis. New York, 1958, reprint Dover Publ., Inc., Mineola, New York 2002.

[35] ————: Combinatorial Identities. New York 1968.

[36] O. Schlomilch¨ : Recherches sur le co´efficients des facult´es analytiques. Crelle’s Journal f¨ur die reine und angewandte Mathematik 44 (1852), 344–355.

[37] W. R. Schmitt: Antipodes and Incidence Coalgebras, Ph. D. thesis, MIT, 1986.

[38] ————: Incidence Hopf Algebras. J. Pure Appl. Algebra 96 (3) (1994), 299–330.

[39] A. Schreiber: A Mathematica Package for the Multivariate Stirling Polynomials (Nov 2013). http://www.gefilde.de/ashome/software /msp/stirling.html

44 [40] ————: Didaktische Schriften zur Elementarmathematik. Logos Ver- lag, Berlin 2014.

[41] ————: Multivariate Stirling polynomials of the first and second kind. Discrete Math. 338 (2015), 2462–2484.

[42] ————: Inverse relations and reciprocity laws involving partial Bell polynomials and related extensions. Enumer. Combin. Appl. 1:1 (2021), Article S2R3. — Preprint: arXiv:2009.09201.

[43] E. Schroder¨ : Vier combinatorische Probleme. Ztschr. f¨ur Math. Phys. 15 (1870), 361–376.

[44] B. Schweizer and A. Sklar: The algebra of functions I–III. Math. Ann. 139 (1960), 366–382; 143 (1961), 440–447; 161 (1965), 171–196.

[45] ————: Function systems. Math. Ann. 172 (1967), 1–16.

[46] G. Scott: Formulae of successive differentiation. Quarterly J. Pure Appl. Math. 4 (1861), 77–92.

[47] A. Sklar: Commentary on the Algebra of Analysis and Algebra of Functions, in Karl Menger Selecta Mathematica, vol. 2 (ed. by Bert Schweizer, Abe Sklar et. al.), Springer, Wien 2003, 111–126.

[48] R. P. Stanley: Combinatorial reciprocity theorems. Adv. Math. 14 (1974), 194–253.

[49] ————: Enumerative Combinatorics, Vol.1, Wadsworth & Brooks/Cole, Monterey 1986.

[50] ————: Enumerative Combinatorics, Vol. 2, Cambridge University Press 1999.

[51] I. Suciu: Function Algebras. Noordhoff International Publishing, 1975.

[52] J. J. Sylvester: Note on Burman’s Law for the inversion of the inde- pendent variable. Philosophical Magazine, 4th series 8 (1854), 535–540. — The Collected Mathematical Papers of James Joseph Sylvester II, Cambridge 1908, pp.44–49.

45 [53] ————: On reciprocants. Messenger of Mathematics 15 (1886), 88– 92. — The Collected Mathematical Papers of James Joseph Sylvester IV, Cambridge 1912, pp.255–258.

[54] P. G. Todorov: New explicit formulas for the n-th derivative of com- posite functions (german). Pacific Journ. of Math. 92/1 (1981), 217– 236.

[55] ————: New differential recurrence relations for Bell polynomials and Lie coefficients. Comptes rendus de l’Acad´emie bulgare des Sciences 38/1 (1985), 43–45.

[56] L. Toscano: Sulla iterazione dell’operatore xd. Rendiconti di matem- atica e delle sui applicazioni V (8) (1949), 337–350.

[57] E. T. Whittaker: On the reversion of series. Gaz. Mat. Lisboa 12 (50) (1951), 1.

46