Why Do Partitions Occur in Faà di Bruno s Chain Rule For Higher Derivatives?

Similar documents
Math 210B. Artin Rees and completions

NOTES ON POLYNOMIAL REPRESENTATIONS OF GENERAL LINEAR GROUPS

Formal power series rings, inverse limits, and I-adic completions of rings

Topics in linear algebra

fy (X(g)) Y (f)x(g) gy (X(f)) Y (g)x(f)) = fx(y (g)) + gx(y (f)) fy (X(g)) gy (X(f))

AHAHA: Preliminary results on p-adic groups and their representations.

Problem Set 6: Solutions Math 201A: Fall a n x n,

Combinatorial Hopf algebras in particle physics I

Infinite-Dimensional Triangularization

Course 311: Michaelmas Term 2005 Part III: Topics in Commutative Algebra

(a i1,1 a in,n)µ(e i1,..., e in ) i 1,...,i n. (a i1,1 a in,n)w i1,...,i n

Eilenberg-Steenrod properties. (Hatcher, 2.1, 2.3, 3.1; Conlon, 2.6, 8.1, )

Exercises on chapter 0

NONCOMMUTATIVE POLYNOMIAL EQUATIONS. Edward S. Letzter. Introduction

Theorem 5.3. Let E/F, E = F (u), be a simple field extension. Then u is algebraic if and only if E/F is finite. In this case, [E : F ] = deg f u.

THE QUANTUM DOUBLE AS A HOPF ALGEBRA

2. Intersection Multiplicities

Lecture 17: Invertible Topological Quantum Field Theories

Representations and Linear Actions

Math 121 Homework 4: Notes on Selected Problems

15. Polynomial rings Definition-Lemma Let R be a ring and let x be an indeterminate.

Review of Linear Algebra

CHAPTER V DUAL SPACES

Projection-valued measures and spectral integrals

Topological vectorspaces

Lifting Smooth Homotopies of Orbit Spaces of Proper Lie Group Actions

Some notes on linear algebra

Let V, W be two finite-dimensional vector spaces over R. We are going to define a new vector space V W with two properties:

CHAPTER 9. Embedding theorems

INTRODUCTION TO LIE ALGEBRAS. LECTURE 2.

3. The Sheaf of Regular Functions

6 Lecture 6: More constructions with Huber rings

MATH 205B NOTES 2010 COMMUTATIVE ALGEBRA 53

Definition 5.1. A vector field v on a manifold M is map M T M such that for all x M, v(x) T x M.

12. Hilbert Polynomials and Bézout s Theorem

l(y j ) = 0 for all y j (1)

Synopsis of material from EGA Chapter II, 4. Proposition (4.1.6). The canonical homomorphism ( ) is surjective [(3.2.4)].

Combinatorics in Banach space theory Lecture 12

LINEAR EQUATIONS WITH UNKNOWNS FROM A MULTIPLICATIVE GROUP IN A FUNCTION FIELD. To Professor Wolfgang Schmidt on his 75th birthday

Groupoids and Faà di Bruno Formulae for Green Functions

1 Introduction. 2 Categories. Mitchell Faulk June 22, 2014 Equivalence of Categories for Affine Varieties

Math 396. Bijectivity vs. isomorphism

Injective semigroup-algebras

B 1 = {B(x, r) x = (x 1, x 2 ) H, 0 < r < x 2 }. (a) Show that B = B 1 B 2 is a basis for a topology on X.

MATH 326: RINGS AND MODULES STEFAN GILLE

THEODORE VORONOV DIFFERENTIABLE MANIFOLDS. Fall Last updated: November 26, (Under construction.)

1 Directional Derivatives and Differentiability

D-MATH Alessio Savini. Exercise Sheet 4

Hilbert spaces. 1. Cauchy-Schwarz-Bunyakowsky inequality

Holomorphic line bundles

NOTES ON LINEAR ALGEBRA OVER INTEGRAL DOMAINS. Contents. 1. Introduction 1 2. Rank and basis 1 3. The set of linear maps 4. 1.

7. Baker-Campbell-Hausdorff formula

A linear algebra proof of the fundamental theorem of algebra

arxiv:math/ v1 [math.ag] 24 Nov 1998

A linear algebra proof of the fundamental theorem of algebra

Math 429/581 (Advanced) Group Theory. Summary of Definitions, Examples, and Theorems by Stefan Gille

Problems in Linear Algebra and Representation Theory

Locally convex spaces, the hyperplane separation theorem, and the Krein-Milman theorem

Transcendental Numbers and Hopf Algebras

NOTES ON LINEAR ODES

LECTURE 4: REPRESENTATION THEORY OF SL 2 (F) AND sl 2 (F)

10. Smooth Varieties. 82 Andreas Gathmann

Operads. Spencer Liang. March 10, 2015

Math 121 Homework 5: Notes on Selected Problems

Topological Data Analysis - Spring 2018

SPECTRAL THEORY EVAN JENKINS

CONSTRUCTING FROBENIUS ALGEBRAS

Equivalence Relations

Math 396. Quotient spaces

Institut für Mathematik, Universität Wien, Strudlhofgasse 4, A-1090 Wien, Austria.

LECTURE 3: REPRESENTATION THEORY OF SL 2 (C) AND sl 2 (C)

Chapter 1. Measure Spaces. 1.1 Algebras and σ algebras of sets Notation and preliminaries

1 Recall. Algebraic Groups Seminar Andrei Frimu Talk 4: Cartier Duality

W if p = 0; ; W ) if p 1. p times

58 CHAPTER 2. COMPUTATIONAL METHODS

Groups of Prime Power Order with Derived Subgroup of Prime Order

Descents, Peaks, and Shuffles of Permutations and Noncommutative Symmetric Functions

Math 530 Lecture Notes. Xi Chen

Two chain rules for divided differences

(df (ξ ))( v ) = v F : O ξ R. with F : O ξ O

2-DIMENSIONAL TOPOLOGICAL QUANTUM FIELD THEORIES AND FROBENIUS ALGEBRAS. Contents 1. The main theorem 1

are Banach algebras. f(x)g(x) max Example 7.4. Similarly, A = L and A = l with the pointwise multiplication

Real Analysis Notes. Thomas Goller

Overview of normed linear spaces

Introduction to Arithmetic Geometry Fall 2013 Lecture #18 11/07/2013

Spanning and Independence Properties of Finite Frames

THE CLOSED-POINT ZARISKI TOPOLOGY FOR IRREDUCIBLE REPRESENTATIONS. K. R. Goodearl and E. S. Letzter

arxiv:math/ v1 [math.fa] 26 Oct 1993

Solutions of exercise sheet 8

55 Separable Extensions

5 Structure of 2-transitive groups

Vector spaces, duals and endomorphisms

1 The Local-to-Global Lemma

where c R and the content of f is one. 1

Mobius Inversion on Partially Ordered Sets

Chapter 8. P-adic numbers. 8.1 Absolute values

The Algebra of Tensors; Tensors on a Vector Space Definition. Suppose V 1,,V k and W are vector spaces. A map. F : V 1 V k

Vector Space Basics. 1 Abstract Vector Spaces. 1. (commutativity of vector addition) u + v = v + u. 2. (associativity of vector addition)

MATH 101A: ALGEBRA I PART C: TENSOR PRODUCT AND MULTILINEAR ALGEBRA. This is the title page for the notes on tensor products and multilinear algebra.

Math212a1413 The Lebesgue integral.

Transcription:

Why Do Partitions Occur in Faà di Bruno s Chain Rule For Higher Derivatives? arxiv:math/0602183v1 [math.gm] 9 Feb 2006 Eliahu Levy Department of Mathematics Technion Israel Institute of Technology, Haifa 32000, Israel email: eliahu@techunix.technion.ac.il Abstract It is well known that the coefficients in Faà di Bruno s chain rule for n-th derivatives of functions of one variable can be expressed via counting of partitions. It turns out that this has a natural form as a formula for the vector case. Viewed as a purely algebraic fact, it is briefly explained in the first part of this note why a proof for this formula leads to partitions. In the rest of the note a proof for this formula is presented from first principles for the case of n-th Fréchet derivatives of mappings between Banach spaces. Again the proof explains why the formula has its form involving partitions. 0 Introduction If I is a finite set, I will denote its cardinality. N will denote the set of non-negative integers. P(I) is the power set of I. Faà di Bruno s formula (see [J] for an excellent survey and bibiliography) gives a somewhat complicatedexpressionfor(g f) (n) (x),f andg functionsofonevariable, intermsofderivatives up to order n of f at x and of g at f(x). The coefficients can be expressed using numeration of partitions of a set of n elements (see below). It might seem that this settles the problem for mappings between multi-dimensional spaces X: the existence of the n-th derivative for the composite mapping follows from 1-st derivative theorems, the n-th derivatives in the vector case being treated as iterated 1-st derivatives, and the formula for it is obtained from Faà di Bruno s formula for one variable by composing with linear mappings X R or R X. Moreover, since it is clear, by iterating the 1-st derivative chain rule, that (g f) (n) (x) is a certain polynomial in f (k) (x) and g (k) (f(x)), 1 k n, the problem is only to find its form, a purely algebraic problem, and in this way it is treated in the literature. It turns out, however, that writing the formula for mappings between vector spaces gives it a very natural form, obscured by the particulars of the 1-dimensional case. This chain-rule formula for n-th derivatives of mappings between vector spaces is constructed in terms of partitions of a set I of cardinality n, as follows: 1

2 0 INTRODUCTION (g f) (n) (x), v i = g ( π ) (f(x)), S π f ( S ) (x), i S v i (3.1) Where f : X Y, g : Y Z, X, Y, Z are vector spaces, x,v i X, I is a set with n elements and H(I) is the set of partitions of I, i.e. H(I) is the subset of P(P(I)) consisting of all disjoint collections of subsets of I whose union is I and that do not contain the empty set. Thus, for example: H( ) = { } H({1}) = {{1}} H({1,2}) = {{{1},{2}}, {{1,2}}} H({1,2,3}) = {{{1},{2},{3}}, {{1},{2,3}}, {{2},{1,3}}, {{3},{1,2}},{{1,2,3}}} For the one-dimensional case X = Y = Z = the scalars it suffices to let v i = 1 and (3) reads: (g f) (n) (x),1 = g ( π ) (f(x)), S π f ( S ) (x),1, that is: a form of Faà di Bruno s formula. (g f) (n) (x) = g ( π ) (f(x)) f ( S ) (x), (3.1 ) S π If (3.1) is viewed as a purely algebraic statement, one can deduce it in a straightforward, though abstract, manner (see 1), where the role of partitions is explained. In Theorem 1. in 3, though, (3.1) is proved from basic principles for Fréchet derivatives between Banach spaces. Once guessed, (3.1) can be proved in a straightforward manner using induction. We have preferred to present a different proof which explains directly the form of the formula. To this end the derivatives are treated via n-th iterated differences, i.e. alternating signs sums over vertices of n-dimensional parallelepipeds, rather than iteratively. The formula is a consequence of an identity (Lemma 2) involving such sums. This identity and the proof are best expressed in the language of free linear spaces and free commutative algebras, which is therefore introduced. Also, this approach to the n-th derivative allows one to require just their existence in the strict sense (a definition is given in 2) at the particular relevant points. A slight complication arises from the fact that the identity of Lemma 2 involves sums over parallelepipeds of dimensions higher than n, the order of the highest participating derivative. This is dealt with using Lemma 1. In Bourbaki, Variétés ([B]) the linear space of point distributions of order n at a point in a C (n) -manifold is defined and its properties stated. The fact that this is properly defined and is a functor may be proved using formula (3.1). Thus one may say that in spirit, (3.1) is present in [B].

1 Where Do Partitions Come From? A Purely Algebraic Approach In this section the scalars K may be any field. We insist on avoiding division by integers, thus the scalar field may have any characteristic. Let us be abstract. The mappings will be between germs of n-manifolds at a point, in short n-germs, defined by their function algebra, namely A = A n = K[[T 1,...,T n ]] (formal power series), where the T j are indeterminates. A has a linear topology, taking as basic neighborhoods at 0 the positive integer powers of the ideal consisting of the formal power series without constant term. Smooth mappings f from the n-germ to the m-germ are defined by continuous homomorphisms φ f : A m A n. (The actual m components of f as a mapping are the functions φ f (T i ), i = 1,...,m.) The continuity of φ f means just that any coefficient of φ(f) is only a function of a finite number of coefficients of F. Note that, by the continuity, since in the linear topology of A Ti k k 0 also (φ f (T i )) k k 0, hence φ f (T i ) have no constant term indeed f(0) = 0. Let A be the dual, i.e. the K-vector space of continuous functionals ξ : A K, K given the discrete topology. Continuity again means that ξ(f) depends only on a finite set of coefficients of F. The multiplication on A induces a comultiplication : A A A by ( (ξ))(f G) := ξ(fg) F,G A. (1.1) This comultiplication, with the counit ξ ξ(1), turns A into a (coassociative and cocommutative) coalgebra over K. A contains the Dirac delta δ at 0, defined by δ(f) := the constant term in F, which satisfies (δ) = δ δ. We also have the primitive elements in A, i.e. those satisfying (ξ) = ξ δ+δ ξ, which means by (1.1) that the functional ξ is a tangent vector at 0. Indeed, they constitute a vector space over K isomorphic to K n by the isomorphism ξ (ξ(t 1 ),...,ξ(t n )). Denote this tangent space of the n-germ by V = V n. We identify a vector in V with the corresponding vector in K n. But there is more structure: one may define the convolution of two elements ξ,η of A (giving an element of A ), and the convolution of an element ξ of A and a F A (giving an element of A) by: (ξ η)(f) := ξ(t η(s F(S+T))) ξ F := (T ξ(s F(S+T))) F A. (1.2) The first convolution makes A also into an (associative and commutative) algebra, with unit δ, which acts linearly on A by the second convolution. In particular, as one easily sees, the primitive elements v V act on A as the directional partial derivatives with direction v. Indeed, one may write: F,v = v F v(f) = F (0),v v V, F A. Convolutions of elements v in V will act, by convolution with members of A, as composition of the actions of the v, i.e. by higher-order differential operators. One finds that one may also write: (v 1 v k )(F) = F (k) (0),v 1 v k v i V, F A. (1.3) 3

4 1 WHERE DO PARTITIONS COME FROM? ALGEBRAIC APPROACH Indeed: F (k) (0),v 1 v k = F (k),v 1 v k 0 = = (v 1 v n F) 0 = (v 1 v n δ)(f) = (v 1 v n )(F) Yet there is a difference between the comultiplication and the convolution multiplication: since any smooth mapping f defines a homomorphism φ f of the algebra A, it will induce a dual map on A (which we again denote by φ f ) which will preserve the comultiplication, i.e. φ f ( (ξ)) = (φ f (ξ)), (Applying a mapping such as φ f to the tensor product is understood factorwise.) Thus, the coalgebra structure is an invariant of the manifold structure, so to speak. On the other hand the convolution was defined in (1.2) using the additive structure of K n, so to speak, and a smooth mapping need preserve it only if it is linear. Nevertheless, using the fact that the convolution in A is, in a sense, the map A A A induced by the smooth mapping K n K n K n given by addition, one shows that convolution preserves the comultiplication, hence A is a bialgebra, (it turns out to be a Hopf algebra). This can be used to compute on v i, v i V, I a finite set. (Recall that δ is the unit of convolution): ( v i ) = (v i ) = (v i δ +δ v i ) = ( i S v i ) ( ) \S v i. (1.4) S I One defines (3) : A A A A by (3) := ( id) = (id ) (the latter being equal by coassociativity), and one has [ (3) (ξ) ] (F G H) = ξ(fgh). Similarly for (n),n N. For v V one has (3) v = v δ δ +δ v δ +δ δ v etc. In analogy to (1.4), one has, for v i V, I a finite set: (k) ( v i ) = I=disj. k j=1 S j k j=1 i S j v i. (1.5) Let f be a smooth mapping from the n-germ to the m-germ. Applying (1.3) for F = f j = φ f (T j ) one obtains (recall that f (k) (0) is a linear map S k (V n ) V m for the notation S k (V) see 2.1): [φ f (v 1 v k )](T j ) = [ f (k) (0),v 1 v k ] (Tj ) = [ f (k) (0),v 1 v k ] We claim that, for v i V, I a finite set (for the notation H(I) see 0): φ f ( v i ) = j v i V j = 1,...,m. (1.6). S π f ( S ) (0), i S v i. (1.7) To get (1.7), it suffices that both sides give the same value when applied to a monomial on the T j s, that is, that applying (k) to both sides one obtains elements of the k-th tensor power of A that give the same value when applied to tensors of T j s. But that follows in a

5 straightforward (though somewhat clumsy) manner from (1.5) and (1.6) and the fact that φ f preserves the comultiplication. Now (3.1) follows from (1.7) and (1.6): indeed, if f and g are smooth mappings, v i V, I a finite set and j = 1,...,n where n is the dimension of the image germ of g: [ (g f) ( I ) (0), v i ] = φ g = = [φ g(φ f ( v i ))](T j ) = j i S π f ( S ) (0), v (T i S j ) = g ( π ) (0), S π f ( S ) (0), v i S i 2 Preliminaries to the Banach Space Approach Let X, Y, Z denote linear spaces over a field (the scalar field) from 3 on Banach spaces unless otherwise stated. The choice of R or C as scalar field is insignificant and will not be mentioned. j 2.1 Symmetric Tensors We consider the symmetric power S n (X) of X, defined as the quotient space of the n-fold (algebraic) tensor product n X with respect to its subspace spanned by the elements x 1 x 2... x n x σ(1) x σ(2)... x σ(n) where x i X and σ is a permutation of {1,2,...,n}. The image of a tensor x 1... x n in S n (X) by the quotient map will still be denoted by x 1... x n. Thus, (x 1,...,x n ) x 1... x n S n (X) is multilinear and symmetric and one may freely employ the notation x i for a finite family (x i ). In particular, S 0 (X) may be identified with the scalar field. In the direct sum S(X) := 0 S n (X), is an associative and commutative multiplication. S(X) is called the symmetric algebra of X. When X and Y are Banach spaces, the bounded linear operators T : S n (X) Y, i.e. those satisfying T := sup T, x 1... x n < x i 1 form, with the above norm, a Banach space, to be denoted by L (n) (X,Y). Of course, this space may be identified with the Banach space of bounded symmetric n-multilinear operators from X n to Y.

6 2 PRELIMINARIES TO THE BANACH SPACE APPROACH 2.2 The n-th strict (Fréchet) derivative It may be defined, instead of iteratively, via n-th order differences (or, in other words, alternating signs sums over vertices of parallelepipeds), as follows: Definition 1 Let X, Y be Banach spaces, U X, f : U Y, x U o (the interior of U). Let n N. We say that f has a (strict) n-th (Fréchet) derivative at x, the derivative being the element f (n) of L (n) (X,Y), if f (n) satisfies, for a set I with I = n: ( ( 1) I S f v + ) v i = f (n) (x), ( ) v i +o v x,vi 0 v i S I i S (2.1) Such an f (n) is necessarily unique, if it exists. Defined this way, f (n) (x) may exist even if previous derivatives fail to exist in a neighborhood of x or even at x itself (for example any non-continuous linear operator f has f (n) 0, n 2). Note that f (0) (x) = y means that f is continuous at x and f(x) = y. The proof of the following fact is standard: Fact: Suppose f (n) (x) exists for every x in an open U X. Then: (i) f (n) : U L (n) (X,Y) is continuous (ii) If v U, v i X are such that the parallelepiped { v + ti v i : t = (t 1,...,t n ) [0,1] n} U then ( I = n): ( ( 1) I S f v + ) ( v i = f (n) v + ) t i v i, v S I t [0,1] n i dt dt being the n-dimensional Lebesgue measure. (iii) (consequence of (ii) ): If x U, m N then f (n) has a (strict) m-th derivative at x iff f has a (strict) m+n -th derivative at x and then: f (n+m) (x), 1 i m+n v i = (f (n) ) (m) (x), 1 i m v i, v m+1 i m+n i. (iii) implies that the above definition of f (n) (x) coincides with the iterative definition whenever f (m), m < n exist in a neighborhood of x.

7 3 The Chain Rule for the n-th Fréchet Derivative Theorem 1. Let n N, f : U X Y, g : V Y Z (X,Y,Z Banach spaces) be such that f(u) V. Let x U o (interior of U), be such that f (k), 0 k n exist at x, f(x) V and g (k), 0 k n exist at f(x). Then (g f) (n) (x) exists and we have the formula ( I = n): (g f) (n) (x), v i = g ( π ) (f(x)), S π f ( S ) (x), v i S i (3.1) All derivatives are in the sense of Definition 1. Proof. Let us remark first, that if it is assumed that the derivatives of order < n exist in neighborhoods of x and f(x) and one employs the iterative definition, then (3.1) may be proved by induction in a straightforward manner (assuming one guesses the formula in advance). We shall rather give a proof that assumes existence of the derivatives only at x and f(x) using Definition 1, and moreover explains why (3.1) has its form. Before we proceed with the proof, let us introduce the following auxiliary concepts and notations: Assume, forthemoment, thatx isjustaset. DenotebyEX thefreelinearspacewithbasis denoted by (Ex) x X. In this way, E is a functor: for any sets X, Y and function f : X Y we have a linear function f E : EX EY defined by f E (Ex) = E (f(x)). In case Y is a linear space, one has also a linear f L : EX Y defined by f L (Ex) = f(x). For a linear space X, EX is a commutative algebra where Ex Ey = E(x+y), i.e. X is the group-algebra of the additive group X. It has the unit element 1 = E0. For linear f : X Y, f E is an algebra homomorphism preserving unit element. E may be iterated: for any set X we have the commutative algebra EEX. For the case that X,Y,Z are linear spaces and f : X Y, g : Y Z any functions, we have the following formulas: Note that f LE is always an algebra homomorphism. f(x) = f L (Ex) (3.2) Ef(x) = Ef L (Ex) = f LE (EEx) (3.2 ) gf(x) = g L (Ef(x)) = g L f LE (EEx) (3.2 ) Let us rewrite the left-hand side expression of (2.1) (definition 1) using these notations: S I ( 1) I S f ( v + i Sv i ) = S I ( 1) I S f L (E ( v + i S v i )) = f L( S I ( 1) I S E v i SEv i ) = f L (E v (Ev i 1)). Thus (2.1) takes the form: ( f L E v ) (Ev i 1) = f (n) (x), ( ) v i +o v x,vi 0 v i (3.3)

8 3 THE CHAIN RULE FOR THE N-TH FRÉCHET DERIVATIVE Lemma 1 If f (n) exists but one takes I > n, then for any J I, J = n, the left hand-side of (3.3) is o v x,vi 0( i J v i ). Proof of Lemma 1. f L (E v (Ev i 1)) = f L( E v \J (Ev i 1) i J (Ev i 1) ) = = f L( E v S I\J( 1) I\J S i SEv i i J (Ev i 1) ) = = S I\J( 1) I\J S f L (E ( v + i Sv i ) i J (Ev i 1)) By (3.3) (note that J = n), the last expression is ( 1) I\J S f (n) (x), ) v i J i +o v x,vi 0( v i S I\J i J And the lemma follows from the fact that the first term vanishes, since because I \J > 0. QED S I\J ( 1) I\J S = (1 1) I\J = 0 Now (3.1) will follow, using (3.3), from the following identity which will be proved later: Lemma 2 For any finite set I and vectors v,v i in a linear space X: ( ( 1) I S EE v+ ) v i = S I i S α P(I), α=i, / αee v ( [ E E v ] ) i 1) 1 A α i A(Ev ( ) Proceeding with the proof of (3.1), one has, by (3.2), (3.2 ) and (3.2 ), using ( ): S I ( 1) I S g(f ( v + v i )) = = g L f LE ( S I ( 1) I S EE ( v + i S v i ) ) = = g L f LE ( α P(I), α=i, / αee v A α(e [E v i A(Ev i 1)] 1) ) = = α P(I), α=i, / αg L( Ef( v) A α( Ef L [E v i A(Ev i 1)] 1 )) = α P(I), α=i, / αd α. D α is an expression of the form appearing in the left-hand side of (3.3), with f replaced by g, I by α and the vectors v and v i by w = f( v) and w A = f L ((E v i A(Ev i 1)). Since f is continuous at x (having a strict 0-derivative), w f(x) as v x. Also, for A I, w A = f L ((E v i A(Ev i 1)) = f A (x), i Av i +o( i A v i ) = O( i A v i )

9 which tends to 0 as v x, v i 0. Now, since g is assumed having strict derivatives at f(x) up to order n, (3.3) and Lemma 1 applied to g give: D α = g L (E w A α(ew A 1)) = { g = ( α ) (f(x)), A αw A +o( A α w A ) = O( A α w A ) if α n o( A J w A ) for any J α with J = n, if α > n (All O s and o s are for v x, v i 0). In order to deal with the second case in (3.4), note that we have: Lemma 3 If α = I, then β α such that β = I and β n. Proof Just pick for each i I an A α such that i A and let β be the set of A s picked. QED (3.4) Thus, if α = I and α > n we can always find a J α with J = n, J = I and D α = o( A J w A ) = o( v i ). If α = I and α n but α is not a partition, then D α = O( A α w A ) = o( v i ), since the A s cover I with redundancy. So we are left with the D α s for α H(I), for which we have: D α = g ( α ) (f(x)), A αw A +o( w A w A ) = = g ( α ) (f(x)), [ A α f ( A ) (x), i Av i +o( i A v i ) ] +o( v i ) = = g ( α ) (f(x)), A α f ( A ) (x), i Av i +o( v i ) which implies (3.1). The only thing left to be done is the: Proof of Lemma 2 (i.e. of ( )): S I ( 1) I S EE ( v+ i S v i ) = = S I ( 1) I S E (E v i SEv i ) = = S I ( 1) I S E [E v i S(1+(Ev i 1))] = i A(Ev i 1)] = = S I ( 1) I S E [E v A S = S I ( 1) I S E [ E v+ A S,A E v i A(Ev i 1) ] = = S I ( 1) I S EE v A S,A E (E v i A(Ev i 1)) = = S I ( 1) I S EE v A S,A [1+(E [E v i A(Ev i 1)] 1)] = = S I ( 1) I S EE v α P(S), / α A α(e [E v i A(Ev i 1)] 1) = = α P(I), / α( S I, α S( 1) I S ) EE v A α(e [E v i A(Ev i 1)] 1) and ( ) follows from the fact that for fixed α QED S I, α S ( 1) I S = (1 1) I α = This completes the proof of Theorem 1. { 1 if α = I 0 otherwise.

10 REFERENCES References [B] Bourbaki N. Variétés Différentielles et Analytiques, Fascicule de résultats. Diffusion C.C.L.S. Paris, 1983. [J] Johnson, Warren P. The curious history of Faà di Bruno s formula, Amer. Math. Monthly, 109, March 2002.