• Aucun résultat trouvé

On generating series of complementary planar trees

N/A
N/A
Protected

Academic year: 2021

Partager "On generating series of complementary planar trees"

Copied!
28
0
0

Texte intégral

(1)

HAL Id: hal-00002767

https://hal.archives-ouvertes.fr/hal-00002767

Preprint submitted on 3 Sep 2004

HAL is a multi-disciplinary open access archive for the deposit and dissemination of sci- entific research documents, whether they are pub- lished or not. The documents may come from teaching and research institutions in France or

L’archive ouverte pluridisciplinaire HAL, est destinée au dépôt et à la diffusion de documents scientifiques de niveau recherche, publiés ou non, émanant des établissements d’enseignement et de recherche français ou étrangers, des laboratoires

On generating series of complementary planar trees

Roland Bacher

To cite this version:

Roland Bacher. On generating series of complementary planar trees. 2004. �hal-00002767�

(2)

ccsd-00002767, version 1 - 3 Sep 2004

On generating series of complementary planar trees

Roland Bacher

Abstract

1

: We generalize and reprove an identity of Parker and Loday.

It states that certain pairs of generating series associated to pairs of labelled rooted planar trees are mutually inverse under composition.

1 Introduction

In [1] Carlitz, Scoville and Vaughan consider finite words (in a finite alphabet A ) such that all pairs of consecutive letters belong to a fixed subset L ⊂ A × A . They show (Theorems 6.8 and 7.3 of [1]) that suitably defined pairs of signed generating series counting such words associated to L ⊂ A × A and to its complementary set L = A × A \ L are each others inverse. Their result was generalized in the first part of Parker’s thesis [5] who showed an analogous result for suitable classes of finite trees having labelled vertices.

Loday in [3], motivated by questions concerning combinatorial realisations of operads, rediscovered Parkers result toghether with a different proof, based on homological arguments.

This paper presents a further generalisation of Parkers and Lodays result.

A typical example of our identity can be described as follows: Associate to two complex matrices

M

1

=

a b c d

, M

2

=

α β γ δ

the following two systems of algebraic equations

g

1

= ( − X + ag

1

+ bg

2

)( − X + αg

1

+ βg

2

) g

2

= ( − X + cg

1

+ dg

2

)( − X + γg

1

+ δg

2

) g = − X + g

1

+ g

2

and

˜

g

1

= (X − (1 − a)˜ g

1

− (1 − b)˜ g

2

)(X − (1 − α)˜ g

1

− (1 − β)˜ g

2

)

˜

g

2

= (X − (1 − c)˜ g

1

− (1 − d)˜ g

2

)(X − (1 − γ)˜ g

1

− (1 − δ)˜ g

2

)

˜

g = − X + ˜ g

1

+ ˜ g

2

.

1Math. class: 05A15, 05C05, 06A10. Keywords: Integer sequence, generating function, inversion of power series, planar tree, spin model

(3)

Choosing continuous determinations satisfying g = g

1

= g

2

= ˜ g = ˜ g

1

= ˜ g

2

= 0 at X = 0 we get holomorphic functions g = g(X), g ˜ = ˜ g(X) for X in an open neighbourhood of 0 ∈ C . We have now

g(˜ g(X)) = X for all X in a small open disc centered at 0.

This result holds of course formally for the corresponding generating series and can be verified by computing for instance a minimal polynomial P (u, v) = P

i,j

p

i,j

u

i

v

j

for g (i.e. satisfying in particular P (g(X), X) = 0) and checking that we have P (X, g(X)) = 0. Since this identity is algebraic, ˜ the field C can be replaced by an arbitrary commutative ring.

The sequel of this paper is organized as follows: The next section states the main result in purely algebraic terms over a not necessarily commutative ring. Parkers and Lodays result corresponds to the special case of matri- ces with coefficients in { 0, 1 } . Our main result removes the restriction on the coefficients. It is also somewhat easier to state (at least in a commuta- tive setting) since it avoids combinatorial descriptions. It follows from our formulation that all involved generating functions are algebraic in a commu- tative setting and over a finite alphabet. Section 3 fixes notations concerning trees. Section 4 describes spin models on trees and recasts our main result using partition functions of spin models. Section 5 proves the main result using a spin model on grafted trees (called “graftings” in [3]). The proof avoids homological arguments and is thus in some sense more elementary (although perhaps not simpler) than the proof of [3]. Section 6 describes briefly a further generalisation involving arbitrary (not necessarily regular) finite trees which appears already in Parkers work. Section 7 is a digression generalizing the notion of grafted trees. Section 8 contains the computations for example (i) of [3]. We display the defining polynomial of the relevant (algebraic) generating function and discuss briefly its asymptotics.

2 Main result

Consider a (not necessarily finite) alphabet A and a (not necessarily com- mutative) associative ring R having a unit 1. We denote by Y

A

a set of (non- commutative) variables indexed by elements α ∈ A and by X a supplemen- tary (non-commutative) variable. We denote by ˜ R[[X]] = Z[[R, Y

A

, R]] ⊗

Z

Z the ring of formal power series. An element of ˜ R[[X]] is a (generally infinite) sum of monomials of the form

± r

1

Z

1

r

2

Z

2

· · · r

l

Z

l

r

l+1

with r

i

∈ R and Z

i

∈ { X } ∪ Y

A

. Let k ≥ 2 be a natural integer and let

M

1

, . . . , M

k

be a set of matrices with rows and columns indexed by A and

(4)

coefficients M

j

(α, β) ∈ R for α, β ∈ A . For α ∈ A , let g

α

∈ R[[X]] be the ˜ power series g

α

= Y

α

X

k

+ (terms of higher order in X) which satisfies

g

α

= Y

α

(X − (M

1

V )

α

)(X − (M

2

V )

α

) · · · (X − (M

k

V )

α

)

where V is the column vector with coordinates g

β

, β ∈ A and where (M

j

V )

α

= P

β∈A

M

j

(α, β)g

β

denotes the α − th coordinate of the matrix- product M

j

V . Set g = − X + P

α∈A

g

α

.

Remark that our notation is slightly misleading: g, g

1

, . . . , g

l

are power series in X defining “functions” of X. The letter X stands of course for the power series at X = 0 of the identity function X 7−→ X. Remark also that the requirements k ≥ 2 and g

α

= Y

α

X

k

+ terms of higher order in X ensure that g

α

is well-defined: Its n − th coefficient involving X n times depends only on coefficients with indices ≤ n − k + 1 of g

β

, β ∈

m

athcalA involving at most n − k + 1 occurences of X.

Define “complementary” matrices ˜ M

1

, . . . , M ˜

k

with coefficients ˜ M

j

(α, β) = 1 − M

j

(α, β) by setting ˜ M

j

= J − M

j

where J is the all 1 matrix with rows and columns indexed by A . For α ∈ A we introduce “complementary”

functions ˜ g

α

= Y

α

( − X)

k

+ terms of higher order in X ∈ R[[X]] satisfying ˜

˜

g

α

= Y

α

( − X + ( ˜ M

1

V ˜ )

α

) · · · ( − X + ( ˜ M

k

V ˜ )

α

)

where ˜ V is the column vector with coordinates ˜ g

β

, β ∈ A and where ( ˜ M

j

V ˜ )

α

= P

β∈A

M ˜

j

(α, β)˜ g

β

. We define now ˜ g = − X + P

l

j=1

˜ g

j

.

Given two formal power series f, h ∈ R[[X]] such that every monomial ˜ of h is at least of degree 1 in X, the composition f ◦

X

h of f with h is defined as the formal power series obtained by replacing every occurence of X

k

, k = 1, 2, 3, . . . in every monomial of f by the series h

k

.

The main result of this paper can now be stated as follows:

Theorem 2.1 We have (formally)

g ◦

X

˜ g = ˜ g ◦

X

g = X .

Remark 2.2 (i) In a commutative setting with R = C the field of complex numbers, a finite alphabet A , and Y

α

∈ C for α ∈ A , the functions g

α

, g ˜

α

and thus also g and g ˜ are holomorphic determinations of algebraic functions in an open neighbourhood of 0.

(ii) The above definitions of g

α

and g ˜

α

are well-suited for iterative computations of the power-series g

α

, g ˜

α

by “bootstrapping” (see e.g. section 5.4 of [4]). Indeed, given g

α,n

, α ∈ A such that g

α

− g

α,n

= O(X

1+(n+1)(k1)

) we have

g

α

− Y

α

(t − (M

1

V

n

)

α

) · · · (t − (M

k

V

n

)

α

) = O(X

1+(n+2)(k1)

)

where V

n

is the column vector with coordinates g

α,n

, α ∈ A . An analogous

result holds of course for g ˜

1

, . . . , g ˜

l

.

(5)

(iii) Changing the grading and considering also all variables Y

α

, α ∈ A as beeing of degree 1, Theorem 2.1 holds also for k = 1 (as a special case, one obtains the main result of [1]). In this case, g ◦

X

g ˜ boils down to a simple product in R ˜ since g and ˜ g are of the form cX with c ∈ R[[(Y

α

)

α∈A

]].

(iv) The choice of an integer k ≥ 2 (or k = 1) corresponds to the case of k − regular trees. This restriction can be removed: Section 6 describes a further generalization (already contained in [5]), corresponding to arbitrary finite trees.

3 Trees

A tree is a connected graph without cycles. A rooted tree contains a marked vertex r, called the root. In particular, a rooted tree is non-empty. A rooted tree T = { r } reduced to its root is trivial. The edges of a rooted tree are canonically oriented by requiring the root to be the unique source of the directed tree obtained by orienting all edges away from the root. We write e = (α(e), ω(e)) for the edge e oriented from α(e) to ω(e). Given an edge e = (α(e), ω(e)) we call ω(e) a son of α(e) and α(e) the father of ω(e). Each vertex v 6 = r ∈ T except the root has a unique father. The set of vertices sharing a common father is a brotherhood. A leaf is a vertex without sons. A vertex having at least one son is interior. The level of a vertex is its (combinatorial) distance to the root. A rooted tree is planar if every brotherhood is totally ordered. Every brotherhood of a locally finite rooted tree is finite . A rooted tree is thus locally finite if and only if the set of vertices of given level N is finite for all N ∈ N. A rooted tree is k − regular if every non-empty brotherhood of strictly positive level contains exactly k vertices. A rooted tree is finite if and only if the set of vertices of level N is finite for all N and empty for N large enough. Given a vertex v ∈ T , we denote by T (v) the subtree of T rooted at v which is defined by considering all vertices w ∈ T such that v belongs to the unique geodesic path joining the root r ∈ T to w (the vertices of T (v) are thus v and its descendants). We call such a subtree maximal. A principal subtree of a non-trivial tree T is a maximal subtree T (s

i

) rooted at a son s

i

of the root r ∈ T . A non-trivial k − regular tree has thus exactly k principal subtrees.

Unless otherwise stated, a (k − regular) tree will henceforth always denote a (k − regular) finite rooted planar tree.

Every tree T can be represented by a plane tree of the halfplane R × R

0

such that a totally ordered brotherhood s

1

< s

2

< . . . of level n ≥ 0 corresponds to vertices (x

1

, n), (x

2

, n), . . . , x

1

< x

2

< . . ..

A vertex v ∈ T of level n has a unique recursively defined address a(v) = a

0

a

1

. . . a

n−1

∈ { 1, 2, . . . }

n

where a

n−1

is the number of elements ≤ v in the brotherhood of v and where

(6)

a

0

. . . a

n−2

is the address of the (unique) father of v. The address of the root r ∈ T is empty. The lexicographic order on addresses orders the set of vertices of a planar rooted tree completely.

11 111

12 21 22

23

2 24

232 234

231 233

1

Figure 1: A (finite rooted planar) tree with addressed vertices.

Figure 1 illustrates the notions of this section. It displays a tree with 14 vertices

∅ < 1 < 11 < 111 < 12 < 2 < 21 < 22 < 23 < 231 < 232 < 233 < 234 < 24 , completely ordered by their addresses. The root corresponds to the lowest vertex having an empty address. The oriented edge e with vertices 23 and 232 starts at α(e) = (23) and ends at the leave ω(e) = 232. The vertex 23 is thus the father of 232. The brotherhood of 232 are the four vertices 231, 232, 233 and 234 at level 3 having α(e) = 23 as their common father.

This tree is of course not regular since the interior vertices 1 and 2 (for instance) have respectively 2 and 4 sons. The tree of Figure 1 has two prin- cipal subtrees. Two vertices with a non-empty adress belong to a common principal subtree if and only if the first letter of their adresses coincides.

4 Non-commutative spin models on rooted planar trees

Let T = (E, V ) be a (rooted planar) tree with edges E and vertices V . We denote by V

the set of interior vertices (having at least one son) of T and by E

⊂ E the subset of leafless edges (i.e. e ∈ E

if ω(e) ∈ V

). We define a spin-model (T, w) on T by considering a (not necessarily) finite set or alphabet A of spins and a weight-function

w : E

× A × A −→ R

with values in a (not necessarily commutative) ring R containing 1. Setting w

e

(α, β) = w(e, α, β), a weight function can be identified with an application

E

−→ M

A

(R)

where M

A

(R) denotes the set of matrices with coefficients in R and rows and

columns indexed by A . We call w

e

∈ M

A

(R) the weight-matrix of e ∈ E

.

(7)

We introduce furthermore (non-commutative) variables Y

α

for α ∈ A and a supplementary (non-commutative) variable X.

Given a spin-model (T, w), its complementary spin-model is defined as (T, w) where ˜ ˜ w

e

(α, β) = 1 − w

e

(α, β) for e ∈ E

and α, β ∈ A .

A colouring

ϕ : V

−→ A

of all interior vertices in T by elements of A is a state. Its energy f

T

(ϕ) ∈ R[X, (Y

α

)

α∈A

] is recursively defined as follows: If T = { r } is trivial then f

T

(ϕ) = X. Otherwise, we set

f

T

(ϕ) = Y

ϕ(r)

(w(1, ϕ)f

T1

1

)) (w(2, ϕ)f

T2

2

)) · · ·

where T

1

, T

2

, . . . are the principal subtrees associated to the linearly ordered sons s

1

< s

2

< . . . of the root r ∈ T , where

w(i, ϕ) =

w((r, s

i

), ϕ(r), ϕ(s

i

)) if s

i

∈ V

1 otherwise

and where f

Ti

(ϕ) is the energy of the spinmodel on the principal subtree T

i

defined by the i − th son s

i

of the root r ∈ T with weights and state obtained by restriction.

In a commutative setting, this boils down to f (ϕ) = X

♯(V\V)

Y

v∈V

Y

ϕ(v)

Y

e∈E

w

e

(ϕ(α(e)), ϕ(ω(e))) which is the familiar definition used in statistical physics.

The total sum

Z = Z(T ) = X

ϕ∈AV

f (ϕ)

of energies over all states is the partition function. We denote by Z

α

= Z

α

(T ) = X

ϕ∈AV, ϕ(r)=α

f (ϕ)

the restricted partition function obtained by computing the total energy of all states with prescribed colour ϕ(r) = α on the root. We have obviously

Z = X

α∈A

Z

α

if T is non-trivial. For T = { r } trivial, we have Z = X and Z

α

= 0 for α ∈ A .

The partition function of a tree T can be computed as follows: We

have Z (T ) = X for T = { r } the trivial tree. Otherwise, consider the

principal subtrees T

1

, T

2

, . . . associated to the k linearly ordered sons s

1

<

(8)

s

2

< . . . < s

k

of r. For α ∈ A denote by Z

α

(T

j

) the obvious restricted partition functions of the principal subtree T

j

with colour α on its root s

j

. We denote by e

j

= { r = α(e

j

), s

j

= ω(e

j

) } the oriented edge joining the root r of T to its j − th son s

j

= ω(e

j

).

The definition of the restricted partition function implies then easily the following result:

Proposition 4.1 If T 6 = { r } is non-trivial, we have Z

α

(T ) = Y

α

C

1

(α)C

2

(α) · · · C

k

(α)

where T

1

, . . . , T

k

are the principal subtrees associated to the k linearly ordered sons s

1

< s

2

< . . . < s

k

of the root r ∈ T and where

C

k

(α) = P

β∈A

w((r, s

i

), α, β)Z

β

(T

i

) if T

i

6 = { s

i

}

X otherwise .

An edge e ∈ E of a k − regular tree is of type j if the extremity ω(e) (having level m + 1) of e has an address

a

0

a

1

. . . a

m

= a

0

a

1

. . . a

m−1

j

ending with 1 ≤ j = a

m

≤ k. The endpoint ω(e) of a type j edge is thus the j − th element in its (totally ordered) brotherhood.

Given a (finite) k − regular tree T and k weight-matrices M

1

, . . . , M

k

(with indices in A × A ) we consider the spin model with spins A and weight function w

e

(α, β) = M

j

(α, β) on leafless edges of type j.

1 121

2

22 222

11 12

Figure 2: A 2 − regular tree.

Example. We consider the commutative spin model on the 2 − regular tree T of Figure 2 with A = { 1, 2 } and

M

1

=

1 1 1 2

, M

2

= (M

1

)

1

=

2 − 1

− 1 1

.

For the sake of simplicity, we set X = Y

1

= Y

2

= 1. Given an interior vertex v of T , we denote by Z

(v) the vector

Z

1

(T (v)) Z

2

(T (v))

where T(v) denotes the

(9)

maximal subtree of T with root v (obtained by considering the connected component of v in T \ a(v) where a(v) is the father of v 6 = r). Since all sons of 11, 121, 222 are leaves, we have

Z

(11) = Z

(121) = Z

(222) = 1

1

where we denote a vertex by its address. We get now Z

(12) =

1 1 1 2

Z

(121) =

1 1 1 2

1 1

= 2

3

,

Z

(22) =

2 − 1

− 1 1

Z

(222) =

2 − 1

− 1 1 1 1

= 1

0

and

Z

(2) =

2 − 1

− 1 1

Z

(22) =

2 − 1

− 1 1 1 0

= 2

− 1

. Denoting by U ∗ W the Hadamard product (“student’s vector product”) αα

ββ

of two vectors α

β

, α

β

we have then

Z

(1) =

1 1 1 2

Z

(11)

2 − 1

− 1 1

Z

(12)

=

1 1 1 2

1 1

2 − 1

− 1 1 2 3

= 2

3

∗ 1

1

= 2

3

and

Z

( ∅ ) =

1 1 1 2

Z

(1)

2 − 1

− 1 1

Z

(2)

=

1 1 1 2

2 3

2 − 1

− 1 1

2

− 1

= 5

8

∗ 5

− 3

= 25

− 24

.

This yields the partition function

Z (T ) = Z

1

(T) + Z

2

(T ) = 25 − 24 = 1 for the tree of Figure 2 with respect to the weights M

1

=

1 1 1 2

and

M

2

= M

11

associated to leafless edges indicating N W (type 1), respectively

N E (type 2).

(10)

Given a spinset (alphabet) A and pairs of complementary matrices M

1

, M ˜

1

= J − M

1

, . . . , M

k

, M ˜

k

= J − M

k

(where J denotes the all 1 matrix with row and column-indices in A ), we consider the associated complementary spin models on k − regular trees with partition function Z(T ), Z ˜ (T ) as above.

Denoting by T

k

the set of all (finite, rooted, planar) k − regular trees, we introduce the signed formal generating series

Z( T

k

) = − X

T∈Tk

( − 1)

d(T)

Z (T ) and

Z( ˜ T

k

) = X

T∈Tk

( − 1)

d(T)

Z(T ˜ )

where d

(T ) = ♯(V

) denotes the number of interior vertices and d(T ) =

♯(V \ V

) the number of leaves of a k − regular tree T ∈ T

k

. Let us moreover consider the restricted generating series

Z

α

( T

k

) = − X

T∈Tk

( − 1)

d(T)

Z

α

(T ) and

Z ˜

α

( T

k

) = X

T∈Tk

( − 1)

d(T)

Z ˜

α

(T )

where T

k

= T

k

\ { r } denotes the set of all k − regular trees which are non- trivial. We have obviously

Z( T

k

) = − X + X

α∈A

Z

α

( T

k

) and Z( ˜ T

k

) = − X + X

α∈A

Z ˜

α

( T

k

) .

Proposition 4.2 We have

Z

α

( T

k

) = Y

α

(X − (M

1

W )

α

) · · · (X − (M

k

W )

α

) and

Z ˜

α

( T

k

) = Y

α

− X + ( ˜ M

1

W ˜ )

α

· · ·

− X + ( ˜ M

k

W ˜ )

α

where W (respectively W ˜ ) is the column vector with coordinates Z

α

( T

k

) (re- spectively Z ˜

α

( T

k

)) indexed by α ∈ A .

For a fixed natural integer k ≥ 1 and k (weight-)matrices M

1

, . . . , M

k

we have now:

Corollary 4.3 We have g

α

= Z

α

( T

k

), g ˜

α

= ˜ Z

α

( T

k

) for all α ∈ A and

g = Z( T

k

), ˜ g = ˜ Z( T

k

) where g

α

, ˜ g

α

, g, g ˜ are as in Theorem 2.1 and where

Z

α

( T

k

), Z ˜

α

( T

k

), Z ( T

k

), Z( ˜ T

k

) are as above.

(11)

Proof of Proposition 4.2 Given a non-trivial k − regular tree T 6 = { r } with root r, we denote by T

1

, . . . , T

k

the principal subtrees associated to the k linearly ordered sons s

1

< s

2

. . . < s

k

of r.

The generating function Z

α

( T

k

) = − X

T∈Tk

( − 1)

d(T)

Z

α

(T ) = X

T∈Tk

( − 1)

Pjd(Tj)

Z

α

(T ) (with the last equality following from d

(T ) = 1 + P

k

j=1

d

(T

j

)) factorizes now as

Z

α

( T

k

) = Y

α

X + P

β∈A

M

1

(α, β) P

T∈Tk

( − 1)

d(T)

Z

β

(T )

·

·

X + P

β∈A

M

2

(α, β) P

T∈Tk

( − 1)

d(T)

Z

β

(T )

· · ·

· · ·

X + P

β∈A

M

k

(α, β) P

T∈Tk

( − 1)

d(T)

Z

β

(T )

= Y

α

X − P

β∈A

M

1

(α, β) Z

β

( T

k

· · ·

X − P

β∈A

M

k

(α, β) Z

β

( T

k

) . Indeed, neglecting the trivial term Y

α

, the partition function Z

α

(T) of a tree

T ∈ T

k

with prescribed spin ϕ(r) = α on its root decomposes into k obvious factors corresponding to the k principal subtrees of T : A leafless edge e ∈ E

with ω(e) ∈ T

j

yields a contribution to the j − th factor of Z

α

( T

k

). Summing over all k − regular trees T ∈ T

k

yields the above expression. The term X in the j − th factor corresponds to trees T ∈ T

k

whose j − th principal subtree T

j

is trivial. This proves the first equality of Proposition 4.2.

For

Z ˜

α

( T

k

) = X

T∈Tk

( − t)

d(T)

Z ˜

α

(T ) = X

T∈Tk

( − t)

Pjd(Tj)

Z ˜

α

(T) we get the analogous factorization

Z ˜

α

( T

k

) = Y

α

− X + P

β∈A

M ˜

1

(α, β) P

T∈T

( − 1)

d(T)

Z ˜

β

(T )

·

·

− X + P

β∈A

M ˜

2

(α, β) P

T∈T

( − 1)

d(T)

Z ˜

β

(T )

· · ·

· · ·

− X + P

β∈A

M ˜

k

(α, β) P

T∈T

( − 1)

d(T)

Z ˜

β

(T )

= Y

α

− X + P

β∈A

M ˜

1

(α, β) ˜ Z

β

( T

k

)

· · ·

− X + P

β∈A

M ˜

k

(α, β) ˜ Z

β

( T

k

) which proves the second equality of Proposition 4.2. 2

The proof of Corollary 4.3 is immediate: corresponding series are recur- sively defined by the same formulae and initial data.

5 Proof of Theorem 2.1

By Corollary 4.3, the formal power series g = g(X) = − X + P

α∈A

g

α

(X) and ˜ g = ˜ g(X) = − X + P

α∈A

g ˜

α

(X) involved in Theorem 2.1 are suitably

(12)

signed generating series for the partition functions of complementary spin- models defined on k − regular trees. In order to prove Theorem 2.1 we define spin-models on a set of combinatorial objects which we call grafted trees.

A suitably signed generating series of the partition functions for these spin- models equals g ◦

X

˜ g and a direct computation establishes Theorem 2.1.

5.1 A spin model on grafted trees A grafted tree is given by

(A; B

1

, B

2

, . . . , B

d(A)

)

where A is a tree with d(A) leaves and where B

1

, . . . , B

d(A)

is a sequence of d(A) trees. A grafted tree is k-regular if it involves only k − regular trees.

The skeleton of a grafted tree (A; B

1

, . . . , B

d(A)

) is the tree obtained by grafting (gluing) the root of B

1

to the smallest (leftmost) leaf of A, by grafting the root of B

2

to the second-smallest leaf of A etc. The skeleton of (A; B

1

, . . . , B

d(A)

) has thus P

d(A)

j=1

d(B

j

) leaves.

The data of a k-regular grafted tree (A; B

1

, . . . , B

d(A)

) and k weight- matrices M

1

, . . . , M

k

indexed by A×A defines a “compositional” spin model as follows: Compute first the ordinary partition function Z(A) as defined previously. Compute also the partition functions ˜ Z(B

1

), . . . , Z(B ˜

d(A)

) with respect to the complementary weight matrices ˜ M

1

= J − M

1

, . . . , M ˜

k

= J − M

k

(where we denote by J the all 1 matrix indexed by A × A ). Replace now the i − th (occurence of the) letter X in every monomial of Z(A) by Z(B ˜

i

). The resulting formal power series (or non-commutative polynomial for a finite alphabet A ) is by definition Z(A; B

1

, . . . , B

d(A)

). It is of degree P

d(A)

i=1

d(B

i

) in X.

In particular, we have

Z(A; B

1

, . . . , B

d(A)

) = X

d(VA)

Z (A)

d(A)

Y

j=1

Z(B ˜

j

)

(with d(V

A

) denoting the number of leaves in A) in a commutative setting.

Let g(X) = − X+ P

α∈A

g

α

(X), g(X) = ˜ − X+ P

α∈A

˜ g

α

(X) be the gen- erating series involved in Theorem 2.1 associated to k matrices M

1

, . . . , M

k

. Proposition 5.1 We have

g ◦

X

g(X) = ˜ − X

(A;B1,...,Bd(A))∈Gk

( − 1)

d(A)+Pd(A)j=1 d(Bj)

Z (A; B

1

, . . . , B

d(A)

)

where G

k

denotes the set of all k − regular grafted trees.

(13)

Proof. This follows at once from Corollary 4.3 and the definition of

spin-models on grafted trees. 2

Given a k − regular tree T ∈ T

k

we denote by S (T ) the set of all k − regular grafted trees with skeleton T . Elements of S (T) are in bijection with k − regular rooted subtrees of T containing the root r ∈ T : The subtree A ⊂ T of a k − regular grafted tree (A; B

1

, . . . , B

d(A)

) ∈ S (T ) with skeleton T clearly contains r. Since the subtree B

j

⊂ T is the maximal subtree of T rooted in the j − th leaf v

j

of A, the pair of trees A ⊂ T defines the grafted tree (A; B

1

, . . . , B

d(A)

) completely.

We denote by S

(T ) = S \ { ( { r } ; T ) } the set of all k − regular grafted trees (A; B

1

, . . . , B

d(A)

) with skeleton T and non-trivial A 6 = { r } .

For α ∈ A and (A; B

1

, . . . , B

d(A)

) ∈ S

(T ) we define Z

α

(A; B

1

, . . . , B

d(A)

) in the obvious way by replacing the i − th occurence of the letter X in Z

α

(A) with ˜ Z(B

i

) for i = 1, . . . , d(A).

Proposition 5.2 We have for a non-trivial k − regular tree T ∈ T

k

and for all α ∈ A

X

(A;B1,...,Bd(A))∈S(T)

( − 1)

d(A)

Z

α

(A; B

1

, . . . , B

d(A)

) = − Z ˜

α

(T) .

Proof. The proof is by induction on the number d(T) of leaves in T . If all k principal subtrees associated to the linearly ordered sons s

1

< . . . < s

k

of the root r ∈ T are trivial, we have S

(T ) = (T ; { s

1

} , . . . , { s

k

} ) and

P

(A;B1,...,Bd(A))∈S(T)

( − 1)

d(A)

Z

α

(A; B

1

, . . . , B

d(A)

)

= ( − 1)

d(T)

Z

α

(T; { s

1

} , . . . , { s

k

} ) = − Y

α

.

Since we have ˜ Z

α

(T ) = Y

α

, Proposition 5.2 follows in this case.

Otherwise, we have X

(A;B1,...,Bd(A))∈S(T)

( − 1)

d(A)

Z

α

(A; B

1

, . . . , B

d(A)

) = − Y

α

C

1

C

2

· · · C

k

where C

i

= X if the principal subtree T

i

= { s

i

} is trivial and

C

i

= ˜ Z(T

i

) + X

β∈A

M

i

(α, β) X

(A;B1,...,Bd(A))∈S(Ti)

Z

β

(A; B

1

, . . . , B

d(A)

) if T

i

6 = { s

i

} . The contribution of a non-trivial principal subtree T

i

6 = { s

i

} is thus by induction

C

i

= Z ˜ (T

i

) − P

β∈A

M

i

(α, β) ˜ Z

β

(T

i

)

= P

β∈A

Z ˜

β

(T

i

) − M

i

(α, β) ˜ Z

β

(T

i

)

= P

β∈A

(1 − M

i

(α, β)) ˜ Z

β

(T

i

)

= P

β∈A

M ˜

i

(α, β) ˜ Z

β

(T

i

)

(14)

and

− Y

α

C

1

C

2

· · · C

k

= − Z ˜

α

(T ) . 2

Corollary 5.3 We have for a non-trivial k − regular tree T ∈ T

k

X

(A;B1,...,Bd(A))∈S(T)

( − 1)

d(A)

Z(A; B

1

, . . . , B

d(A)

) = 0 .

Proof. Sum the equality of Proposition 5.2 over α ∈ A and use Z( { r } ; T ) = ˜ Z(T) = X

α∈A

Z ˜

α

(T) . 2 Proof of Theorem 2.1. By Proposition 5.1 we have g ◦

X

˜ g = − X

T∈Tk

( − 1)

d(T)

X

(A;B1,...,Bd(A))∈S(T)

( − 1)

d(A)

Z(A; B

1

, . . . , B

d(A)

) .

It follows from Corollary 5.3 that only the trivial tree T = { r } con- tributes to the right hand side. The contribution of T = { r } amounts to

− ( − 1)

d({r})

Z( { r } ; { r } ) = − ( − 1) · X = X

which proves Theorem 2.1. 2

6 Further generalizations

We give first a generalization of Theorem 2.1 associated to not necessar- ily regular trees. The special case of this generalization where all weight- matrices have coefficients in { 0, 1 } appears already in [5].

We state then an even more general version involving trees with vertices labelled by a set T of types and formal power-series in R[[(X

τ

)

τ∈T

, (Y

α

)

α∈A

]].

6.1 Trees with vertices of arbitrary degrees

Consider a subset K ⊂ N

>0

of strictly positive integers and a partition A = ∪

k∈K

A

k

of the alphabet A into non-empty parts indexed by K . For each k ∈ K choose k matrices M

1k

, M

2k

, . . . , M

kk

with rows indexed by A

k

and columns indexed by A . For i = 1, . . . , k, set ˜ M

ik

= J − M

ik

where J is the all 1 matrix with indices in A

k

× A .

For α ∈ A

k

consider the uniquely defined (non-commutative) formal power series g

α

, g ˜

α

consisting only of monomials of degree at least 1 in X, involving at least one variable of (Y

β

)

β∈A

and such that

g

α

= Y

α

(X − (M

1k

V )

α

)(X − (M

2k

V )

α

) · · · (X − (M

kk

V )

α

)

(15)

and

˜

g

α

= Y

α

( − X + ( ˜ M

1k

V ˜ )

α

) · · · ( − X + ( ˜ M

kk

V ˜ )

α

)

where V , respectively ˜ V , is the column vector with coordinates g

β

, respec- tively ˜ g

β

, for β ∈ A .

Set g = − X + P

α∈A

g

α

and ˜ g = − X + P

α∈A

g ˜

α

. Theorem 6.1 We have

g ◦

X

˜ g = ˜ g ◦

X

g = X .

The proof is an adaption of the proof of Theorem 2.1: Consider trees whose internal vertices have degrees in K and colour internal vertices of degree k ∈ K by elements in A

k

.

Example (Inversion of power series). In this example, we work over a commutative ring R and set Y

α

= 1. Choose an integer l ≥ 1 and set K = { 2, 3, 4, . . . } (this ensures existence of all relevant power-series). We consider trees with internal vertices of degree ≥ 2 having spins in a finite alphabet A containing l elements.

For 1 ≤ j ≤ k, 2 ≤ k choose elements β

k,j

∈ R and consider the diagonal weight function

w

e

(ϕ(α(e)), ϕ(ω(e)) = β

k,j

δ

ϕ(α(e)),ϕ(ω(e))

=

β

k,j

if ϕ(α(e)) = ϕ(ω(e)) 0 otherwise

where the leafless edge e joins a vertex α(e) of degree k to its j − th son ω(e).

Introduce the formal power series g

, ˜ g

∈ R[[X]] without constant term which satisfy

g

= P

k=2

Q

k

j=1

(X − β

k,j

g

)

˜

g

= P

k=2

Q

k

j=1

( − X + (l − β

k,j

) ˜ g

) .

Set g(X) = − X + l g

and ˜ g(X) = − X + l ˜ g

. Theorem 6.1, applied to the present situation shows that we have

g(˜ g(X)) = ˜ g(g(X)) = X .

Since this formula holds for any l ∈ N , it extends to an arbitrary value of l ∈ R which can thus be considered as a parameter.

This formula provides a mean (other than the celebrated Lagrange in- version formula) for computing the compositional inverse of a formal power series h(X) = P

k=1

γ

k

X

k

with γ

1

∈ R

invertible: Set f = h( − X/γ

1

) =

− X + P

k=2

β

k

t

k

. If β

2

6 = 0 set l = β

2

and choose constants β

k,j

such that g(X) = f(X) with g(X) defined as above.

Each such choice can be used to compute the compositional inverse ˜ g =

f

1

of f . The compositional inverse of the initial series h(X) is then given

by − γ

1

g(X). The case ˜ β

2

= 0 can be handled similarly by allowing the

trees to have internal vertices of degree 1.

(16)

6.2 Vertices of different types

This still more general version of Theorem 2.1 is perhaps easier to formulate in a combinatorial way:

Consider a set T of vertex-types indexing the non-empty parts of a par- tition A = ∪

τ∈T

A

τ

. Consider also a set E of edge-types together with weight-functions

w

ǫ

: A × A −→ R indexed by edge-types ǫ ∈ E .

A labelled tree is a finite rooted planar tree L with leaves labelled by elements of T , internal vertices labelled by pairs (τ, α ∈ A

τ

) ∈ T × A and edges labelled by elements of E . A vertex v ∈ L is of type τ ∈ T if it is either a leaf labelled τ or an internal vertex labelled (τ, α ∈ A

τ

). A labelled tree L is of type τ ∈ T if its root vertex is of type τ .

The energy f(L) ∈ R[(X

τ

)

τ∈T

, (Y

α

)

α∈A

] of a labelled tree L of type τ ∈ T is defined in the obvious way: f (L) = X

τ

if L is trivial (reduced to its labelled root) and

f (L) = Y

α

C

1

· · · C

k

otherwise where (τ, α ∈ A τ ) is the label of the root r ∈ L and where C

i

∈ R[(X

τ

)

τ∈T

, (Y

α

)

α∈A

] is associated to the i − th labelled principal sub- tree L(s

i

) of L as follows: C

i

= X

τi

if L(s

i

) is trivial of type τ

i

and

C

i

= w

ǫi

(α, α

i

)f (L

i

)

where ǫ

i

is the label of the edge joining the root r to its i − th son s

i

which is labelled (τ

i

, α

i

∈ A

τi

).

Define complementary weight-functions by ˜ w

ǫ

(α, β) = 1 − w

ǫ

(α, β) and compute the complementary energy f ˜ (L) of L as above using the comple- mentary weight-functions.

Call two labelled trees L

1

, L

2

edge-equivalent if they differ only on their edge-labels (but share the same underlying tree-structure and vertex-labels).

Let L be a subset of the set of all labelled (finite rooted planar) trees.

We denote by L

τ

⊂ L the subset of all labelled trees of type τ in L and assume that the set L satisfies the following three conditions:

(i) If L ∈ L and v is a leaf of L , then L(v) ∈ L where L(v) denotes the labelled subtree defined in the obvious way by considering the maximal subtree of L rooted at v.

(ii) If L

∈ L

τ

and v is a leaf of type τ in L ∈ L then the labelled tree obtained in the obvious way by gluing L

onto the leaf v ∈ L is again in L . (iii) All equivalence classes of the edge-equivalence relation are finite.

Otherwise stated, all maximal labelled subtrees of an element in L are

in L by (i) and L is “closed under composition” by (ii). Condition (iii) is a

finiteness condition (which can perhaps be slightly weakened or replaced by

(17)

a similar statement) ensuring the existence of the generating series Z

τ

and Z ˜

τ

defined below.

For such a set L we define Z

τ

= − X

L∈L

( − 1)

d(L)

f (L) and

Z ˜

τ

= X

L∈L

( − 1)

d(L)

f ˜ (L)

where d

(L) denotes the number of internal vertices of a labelled tree L and where d(L) denotes the number of leaves in L.

Theorem 6.2 We have

Z

τ

XL

( ˜ Z

σ

)

σ∈T

= ˜ Z

τ

XL

(Z

σ

)

σ∈T

= X

τ

where the notation Z

τ

XL

( ˜ Z

σ

)

σ∈T

means that every occurence of X

τ

in Z

τ

is replaced by the generating series Z ˜

τ

.

This result can be used for the formal inversion of power-series in several variables.

Sketch of Proof. Define grafted labelled trees in the obvious way and check that Proposition 5.2 remains valid in the present context.

7 Morphisms of rooted trees into posets

This section is a digression discussing a generalization of grafted trees.

The vertex set V of a rooted (not necessarily finite or planar) tree T with oriented edges E can be considered as a poset (partially ordered set) by considering the order relation induced by ω(e) > α(e) for e ∈ E.

Remark 7.1 Considering a rooted tree as a poset, one might wonder how many unordered pairs of comparable vertices are contained in (k − regular) rooted planar trees having n vertices. Let a

n

(respectively a

n,k

) denote this number. We have then

X

n=1

a

n

t

n

= ∂y

∂u (t, 1)

where y(t, u) = t+( terms of higher order in t ) satisfies the functional equa- tion

y(t, u) = t + t y(tu, u) 1 − y(tu, u) . The corresponding sequence a

2

, a

3

, a

4

, . . . starts as

1, 5, 22, 93, 386, 1586, 6476, 26333, 106762, 431910, . . . cf. A346 of [2].

(18)

Similarly, for the numbers a

n,k

associated to k − regular trees, we have X

n=1

a

n,k

t

n

= ∂y

∂u (t, 1)

with y(t, u) = t + ( terms of higher order in t ) satisfying y(t, u) = t + t (y(tu, u))

k

.

For k = 2 we get the sequence

12

(a

3,2

, a

5,2

, a

7,2

, . . .) starting as

1, 6, 29, 130, 562, 2380, 9949, 41226, 169766, . . . , cf. A8549 of [2]

and for k = 3, the sequence

13

(a

4,3

, a

7,3

, a

10,3

, . . .) starting as

1, 9, 69, 502, 3564, 24960, 173325, 1196748, . . . , cf. A75045 of [2].

Given a second poset P , a morphism from T to P is an application µ : V −→ P such that µ(ω(e)) ≥ µ(α(e)) for every edge e of T .

Denote by { 1, 2 } (with 1 < 2) the obvious totally ordered poset. Call a morphism µ : T −→ { 1, 2 } restricted if µ

1

(2) contains all leaves of T . Proposition 7.2 If T is a rooted (planar) tree, then the set of grafted trees (A; B

1

, . . . , B

d(A)

) with skeleton T corresponds bijectively to the set of re- stricted morphisms µ : T −→ { 1, 2 } .

Proof. Given a restricted morphism µ : T −→ { 1, 2 } we set A = { µ

1

(1) ∪ sons of µ

1

(1) } where the root r ∈ T is by convention the son of the empty set. The rooted trees { B

1

, . . . , B

d

(A) } are the connected components of the maximal subforest with vertices µ

1

(2). It is obvious that (A; B

1

, . . . , B

d(A)

) is a grafted tree with skeleton T . Reciprocally, given a grafted tree (A; B

1

, . . . , B

d(A)

) with skeleton T , we get a morphism µ : T −→ { 1, 2 } by setting µ(v) = 2 if v ∈ B

j

for some j = 1, . . . , d(A) and µ(v) = 1 otherwise. Since the forest ∪

d(A)j=1

B

j

contains all leaves of T , the

morphism µ is restricted. 2

Proposition 7.2 suggests to generalize the notion of grafted trees by con- sidering sets of (suitable) morphisms from rooted trees into posets. We consider now a few special cases. Henceforth all rooted trees and posets will be finite. The number of morphisms of a rooted tree T having ♯(V ) vertices into a poset P with ♯(P ) elements is bounded by ♯(P )

♯(V)

.

7.1 Grafted trees

In this subsection we indicate how to count the number of grafted trees with

given skeleton T . A slight modification counts all morphisms from T into

(19)

{ 1, 2 } . Further generalizations consist in counting all (suitable) morphisms from T into the totally ordered set { 1, . . . , n } .

Given a (finite rooted planar) tree T we denote by γ (T ) the number of grafted trees with skeleton T . We denote by ˜ γ (T ) ≥ γ(T ) the number of all morphisms T −→ { 1, 2 } . We have γ( { r } ) = 1, γ ˜ ( { r } ) = 2 if T = { r } is the trivial tree reduced to its root. Remark that ˜ γ( ˜ T ) = γ (T ) where ˜ T is obtained by removing all leaves from the rooted tree γ (where ˜ γ( ∅ ) = 1 by convention). This remark generalizes easily to the analogous numbers enumerating (restricted) morphismes T −→ { 1, . . . , m } .

Proposition 7.3 For T a non-trivial (finite rooted planar) tree we have

γ(T ) = 1 +

k

Y

j=1

γ(T

i

) and

˜

γ(T ) = 1 +

k

Y

j=1

˜ γ(T

i

)

where T

1

, . . . , T

k

are the principal subtrees defined by the k sons s

1

, . . . , s

k

of the root r ∈ T .

Proof. We count all (restricted) morphisms µ from T into { 1, 2 } . Consider a morphism µ : V −→ { 1, 2 } . If µ(r) = 2 we have µ(v) = 2 for all v ∈ V and there is exactly one such morphism yielding a contribution of 1 to γ(T ) and ˜ γ (T ). If µ(r) = 1, the restriction µ

i

of µ to a principal subtrees T

i

is an arbitrary (restricted) morphism from T

i

into { 1, 2 } and the restrictions µ

1

, . . . , µ

k

can be arbitrary thus proving the formula. 2 Proposition 7.3 leads to a fast algorithm for computing γ (T ) illustrated by the following example.

Example. For the tree T of Figure 2 we have γ(v

1

) = γ(v

3

) = γ (v

8

) = 2

γ(v

4

) = 1 + γ(v

3

) · 1 = 3, γ(v

7

) = 1 + 1 · γ(v

8

) = 3 γ(v

2

) = 1 + γ(v

1

) · γ(v

4

) = 7, γ(v

6

) = 1 + 1 · γ(v

7

) = 4 γ(v

5

) = 1 + γ(v

2

) · γ(v

6

) = 29

and

˜

γ (v

1

) = ˜ γ (v

3

) = ˜ γ (v

8

) = 1 + 2 · 2 = 5

γ (v

4

) = 1 + γ(v

3

) · 2 = 11, γ(v

7

) = 1 + 2 · γ(v

8

) = 11 γ (v

2

) = 1 + γ(v

1

) · γ(v

4

) = 56, γ(v

6

) = 1 + 2 · γ(v

7

) = 23 γ (v

5

) = 1 + γ(v

2

) · γ(v

6

) = 1289

which shows that the tree T of Figure 2 is the skeleton of γ(T ) = γ (v

5

) = 29

different grafted trees (or equivalently, that T contains 29 different 2 − regular

(20)

subtrees sharing the root r with T ) and has 1289 distinct morphisms into the totally ordered poset { 1, 2 } .

Proposition 7.3 shows also that the generating function y(t) = X

T∈T2

γ (T ) t

d(T)

counting the number γ (k) = [t

k

]y of 2 − regular grafted trees whose skelettons have k leaves satisfies

y =

1214t

+ y

2

=

1

−1+2√ 1−4t

= t + 2t

2

+ 6t

3

+ 21t

4

+ 80t

2 5

+ 322t

6

+ 1348t

7

+ 5814t

8

+ . . . . The similar generating function ˜ y(t) = P

T∈T2

γ ˜ (T ) t

d(T)

(which counts all morphisms of binary rooted trees into { 1, 2 } ) is given by

˜

y = t +

1214t

+ y

2

=

1

−1−4t+2√ 1−4t

= 2t + 5t

2

+ 22t

3

+ 118t

4

+ 706t

2 5

+ 4530t

6

+ . . . .

More generally, consider a fixed natural integer k ≥ 2. We call a mor- phism µ : T −→ { 1, . . . , k } restricted if µ

1

contains all leaves of the rooted tree T . We denote by y

m

(t) the generating function counting the number of (non-isomorphic) restricted morphisms from a k − regular tree having n leaves into the completely ordered finite set { 1, . . . , m } . Similarly, we denote by ˜ y

m

(t) the analogous generating function counting all morphisms without restriction on the leaves.

Proposition 7.4 We have

y

1

(t) = ˜ y

1

(t) = t + (y

1

(t))

k

y

m

(t) = t +

m

X

h=1

(y

h

(t))

k

˜

y

m

(t) = mt +

m

X

h=1

(˜ y

h

(t))

k

Sketch of proof. The trivial tree T yields a contribution of t to y

k

and of mt to ˜ y(t). For a morphism µ from a nontrivial tree into { 1, . . . , m, }

we consider the image µ(r) of its root and the remaining possibilities of the

induced restrictions µ

1

, . . . , µ

k

on the principal subtrees T

1

, . . . , T

k

of T . 2

For arbitrary finite planar non-empty trees we consider the generating

function counting the number of different restricted morphisms µ from a

tree having n vertices into the completely ordered finite set { 1, . . . , m } . The

generating function ˜ y

m

(t) counts all such morphisms into { 1, . . . , m } . One

can then prove the following result.

(21)

Proposition 7.5 We have

y

1

(t) = ˜ y

1

(t) = t + ty

1

1 − y

1

y

m

(t) = t + t

m

X

h=1

y

h

1 − y

h

˜

y

m

(t) = mt + t

m

X

h=1

˜ y

h

1 − y ˜

h

The first instances are y

1

= ˜ y

1

=

1214t

(defining the Catalan numbers, cf. A108 of [2])

y

2

= 3 − 2t − √

1 − 4t − q

2 − 16t + 4t

2

+ (2 + 4t) √ 1 − 4t

= t + 2t

2

+ 5t

3

+ 15t

4

+ 50t

5

4 + 178t

6

+ 663t

7

+ 2553t

8

+ . . . which is sequence A7853 of [2] and

˜

y

2

= 3 − √

1 − 4t − p

2 − 20t + 2 √ 1 − 4t

= 2t + 3t

2

+ 9t

3

+ 34t 4

4

+ 145t

5

+ 667t

6

+ 3231t

7

+ 16247t

8

+ . . . Finally, let us mention the well-known fact that the function p(m) count- ing all morphisms from a fixed finite poset E into the totally ordered poset { 1, . . . , m } is a polynomial of degree ♯(E). Indeed

p(m) = X

k=1

α

k

m

k

where α

k

denotes the number of surjective morphismes of E into { 1, . . . , k } . One can thus also consider generating functions associated to such polyno- mials. The number of restricted morphisms from a fixed rooted tree into { 1, . . . , m } is of course also a polynomial function. Its degree in m is the number of interior leaves in T .

7.2 Surjective morphisms

Given a finite rooted tree T having n vertices, the number σ(T ) of surjective morphisms from T into { 1, . . . , n } can be recursively computed by remarking that

σ(T ) = (n − 1)!

k

Y

j=1

σ(T

j

)

n

j

!

(22)

where T

1

, . . . , T

k

are the principal (rooted) subtrees of T having n

1

, . . . , n

k

vertices. Denoting by

nn1

1,...,nk

=

n(n1)!

1!···nk!

the obvious multinomial coefficient (where n − 1 = P

n

i

) we get for our favorite (binary) tree T of Figure 2 the following numbers: σ(v) = 1 if v is a leaf and

σ(v

1

) = σ(v

3

) = σ(v

8

) = 2

1, 1

= 2

σ(v

4

) = σ(v

7

) = 4

1, 3

2 = 8 σ(v

2

) =

8 3, 5

2 · 8 = 896, σ(v

6

) = 6

1, 5

1 · 8 = 48 σ(T ) = σ(v

5

) =

16 9, 7

896 · 48 = 492011520 The generating function

y(t) = X

k=1

α

k

t

k

encoding the number α

n

of surjective morphisms into { 1, . . . , n } from all rooted binary planar trees on n vertices satisfies

α

n

= (n − 1)!

n−2

X

k=1

α

k

α

n1k

k! (n − 1 − k)!

(cf. sequence A182 of [2]). Otherwise stated, the exponential generating function

z(t) = X

k=1

α

k

t

k

k!

satisfies z

= z

2

− 1 thus proving that z(t) = tanh(t).

Similarly, considering the exponential generating function z(t) =

X

k=1

β

k

t

k

k!

enumerating the number β

n

of all surjective homomorphisms from rooted planar trees having n vertices into { 1, . . . , n } we have

z

= z

1 − z + 1 = 1 1 − z

which implies z(t) =

1214t

(cf. sequence A108 of [2]) and shows that β

n

= n!

2(nn11)

/n =

(2n(n1)!2)!

.

(23)

8 Loday’s example (i)

The aim of this section is a partial analysis of example (i) in [3]. The framework is somewhat simpler as in the previous sections: We work over the commutative ground field C of complex numbers. The alphabet A consists of nine elements and equals

A = {◦ , N, N W , W, SW , S, SE, E, N E }

suggesting the graphical notations of [3]. We set X = t in order to stick to [3].

Loday’s example (i) corresponds to k = 2. The two 9 × 9 matrices M

1

, M

2

are given by

1 0 0 0 0 0 0 0 0 1 1 1 0 0 0 0 0 0 1 1 1 1 0 0 0 0 0 1 0 1 1 0 0 0 0 0 1 1 1 1 1 0 0 0 0 1 1 0 0 0 0 0 0 0 1 1 1 1 0 1 0 0 0 1 1 0 1 0 0 0 0 0 1 1 1 1 0 0 0 0 0

 ,

0 1 1 1 0 0 0 0 0 0 0 0 1 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 1 0 0 0 0 0 0 0 0 0 0 0 0 1 0 0 0 0 0 1 1 1 1 0 0 0 0 0 0 0 1 0 1 1 1 0 0 1 0 0 0 0 1 1 0 0 0 0 0 0 0 1 1

Writing V = (g

, g

N

, g

N W

, g

W

, g

SW

, g

S

, g

SE

, g

E

, g

N E

)

t

we get the equations g

= ( − t + g

)( − t + g

N

+ g

N W

+ g

W

)

g

N

= ( − t + g

+ g

N

+ g

N W

)( − t + g

W

) g

N W

= ( − t + g

+ g

N

+ g

N W

+ g

W

)( − t) g

W

= ( − t + g

+ g

N W

+ g

W

)( − t + g

N

)

g

SW

= ( − t + g

+ g

N

+ g

N W

+ g

W

+ g

SW

)( − t + g

S

) g

S

= ( − t + g

+ g

N

)( − t + g

N W

+ g

W

+ g

SW

+ g

S

) g

SE

= ( − t + g

+ g

N

+ g

N W

+ g

W

+ g

S

)

( − t + g

SW

+ g

SE

+ g

E

+ g

N E

)

g

E

= ( − t + g

+ g

N

+ g

W

)( − t + g

N W

+ g

E

+ g

N E

) g

N E

= ( − t + g

+ g

N

+ g

N W

+ g

W

)( − t + g

E

+ g

N E

) One sees easily that we have

g

o

= g

N

= g

W

. Eliminating g

N

, g

W

we get the simpler equations:

g

= ( − t + g

)( − t + 2g

+ g

N W

) g

N W

= ( − t + 3g

+ g

N W

)( − t)

g

SW

= ( − t + 3g

+ g

N W

+ g

SW

)( − t + g

S

) g

S

= ( − t + 2g

)( − t + g

+ g

N W

+ g

S

+ g

SW

) g

E

= ( − t + 3g

)( − t + g

N W

+ g

E

+ g

N E

) g

N E

= ( − t + 3g

+ g

N W

)( − t + g

E

+ g

N E

)

g

SE

= ( − t + 3g

+ g

N W

+ g

S

)( − t + g

SW

+ g

SE

+ g

E

+ g

N E

)

Références

Documents relatifs

Assume A is a family of connected planar graphs that satisfies the previous condition, and let X n denote the number of connected components that belong to A in a random planar

In this paper, we have introduced what we may call variational dilation, erosion, opening, and closing. We have seen that these operators can be applied in graph processing and

This complete redistribution of glutathione pools between the cytosol and mitochondria in MPTP-intoxicated mice, lead to a dynamic flux of reduced glutathione within the

Regarding the in-hospital care pathway, we collected the admission type (from home, from the ER, or trans- ferred from another hospital), the discharge type (to home, to another

trum using the Climax Neutron Monitor data in the frequency range from 10 − 9 Hz to 10 − 7 Hz (which corresponds to peri- odicities from 11 years to a few months) during the

å âmèoäUê åê åõ „ä9ðöK÷[ölåõ âmêtä†äU÷ mæ åê.ä o㠐øoëläú å„ð bäâCãGë ú Gÿ ®ò7ãõ .è5ãGå+ú Gÿ |òoäiãGõ .ègãGå7ú ÿ[ÿ âtèoöK÷

In the second section we review the Hopf algebras of renormalization for quantum electrodynamics in the context of planar binary trees, and their dual groups of tree-expanded

Many of these cases can be pruned from the exploration tree by calling gl0w(w,s,n) only when n is not smaller than the number of pending symbols, what we call here the control