• 検索結果がありません。

Representation theory of the higher-order peak algebras

N/A
N/A
Protected

Academic year: 2022

シェア "Representation theory of the higher-order peak algebras"

Copied!
31
0
0

読み込み中.... (全文を見る)

全文

(1)

DOI 10.1007/s10801-010-0223-y

Representation theory of the higher-order peak algebras

Jean-Christophe Novelli·Franco Saliola· Jean-Yves Thibon

Received: 2 July 2009 / Accepted: 11 March 2010 / Published online: 7 April 2010

© Springer Science+Business Media, LLC 2010

Abstract The representation theory (idempotents, quivers, Cartan invariants, and Loewy series) of the higher-order unital peak algebras is investigated. On the way, we obtain new interpretations and generating functions for the idempotents of de- scent algebras introduced in Saliola (J. Algebra 320:3866,2008).

Keywords Noncommutative symmetric functions·Peak algebras·Finite dimensional algebras·Descent algebras

1 Introduction

A descent of a permutationσ∈Snis an indexisuch thatσ (i) > σ (i+1). A descent is a peak if moreoveri >1 andσ (i) > σ (i−1). The sums of permutations with a given descent set span a subalgebra of the group algebra, the descent algebraΣn. The peak algebra ˚Pn ofSnis a subalgebra of its descent algebra, spanned by sums of permutations having the same peak set. This algebra has no unit.

The direct sum of the peak algebras is a Hopf subalgebra of the direct sum of all descent algebras, which can itself be identified with Sym, the Hopf algebra of noncommutative symmetric functions [11]. Actually, in [7], it was shown that most of the results on the peak algebras can be deduced from the caseq= −1 of aq-identity

J.-C. Novelli (

)·F. Saliola·J.-Y. Thibon

Institut Gaspard Monge, Université Paris-Est Marne-la-Vallée, 5 Boulevard Descartes, Champs-sur-Marne, 77454 Marne-la-Vallée cedex 2, France

e-mail:novelli@univ-mlv.fr F. Saliola

e-mail:saliola@gmail.com J.-Y. Thibon

e-mail:jyt@univ-mlv.fr

(2)

of [14]. Specializingq to other roots of unity, Krob and the third author introduced and studied higher-order peak algebras in [13]. Again, these are nonunital.

In [1], it has been shown that the peak algebra ofSn can be naturally extended to a unital algebra which is obtained as a homomorphic image of the descent algebra of the hyperoctahedral groupBn. This construction has been extended in [3]. It is shown there that unital versions of the higher-order peak algebras can be obtained as homomorphic images of the Mantaci–Reutenauer algebras of typeB.

Our purpose here is to investigate the representation theory of the unital higher- order peak algebras. The classical case has been worked out in [2]. In this reference, idempotents for the peak algebras were obtained from those of the descent algebras of typeBconstructed in [6].

To deal with the general case, we need a different construction of idempotents. It turns out that the recursive algorithm introduced in [17] for idempotents of descent algebras can be adapted to higher-order peak algebras.

In order to achieve this, we need a better understanding of the idempotents gener- ated by the algorithm of [17]. Interpreting them as noncommutative symmetric func- tions, we find that in typeA, these idempotents are associated with a known family of Lie idempotents, the so-called Zassenhaus idempotents, by the construction of [14].

We then show that similar Lie idempotents can be defined in typeB as well, which yields a simple generating function in terms of noncommutative symmetric functions of typeB.

This being understood, we obtain complete families of orthogonal idempotents for the higher-order peak algebras, which can be described either by recurrence relations as in [17] or by generating series of noncommutative symmetric functions.

Finally, we make use of these idempotents to study the quivers, Cartan invariants, and the Loewy series of the unital higher-order peak algebras.

2 Notations and background

2.1 Noncommutative symmetric functions

We will assume familiarity with the standard notations of the theory of noncommu- tative symmetric functions [11] and with the main results of [3,13]. We recall here only a few essential definitions.

The Hopf algebra of noncommutative symmetric functions is denoted by Sym, or by Sym(A)if we consider the realization in terms of an auxiliary alphabetA. Linear bases of Symnare labeled by compositionsI =(i1, . . . , ir)ofn(we writeIn). The noncommutative complete and elementary functions are denoted bySn andΛn, and SI=Si1· · ·Sir. The ribbon basis is denoted byRI. The descent set ofI is Des(I )= {i1, i1+i2, . . . , i1+ · · · +ir−1}. The descent composition of a permutationσ∈Snis the compositionIofnwhose descent set is the descent setD(σ )ofσ.

2.2 The Mantaci–Reutenauer algebra of typeB

We denote by MR the free product SymSym of two copies of the Hopf algebra of noncommutative symmetric functions [15]. That is, MR is the free associative al- gebra on two sequences(Sn)and (Sn¯)(n≥1). We regard the two copies of Sym

(3)

as noncommutative symmetric functions on two auxiliary alphabets: Sn =Sn(A) andSn¯=Sn(A). We denote by¯ F → ¯F the involutive antiautomorphism which ex- changesSn andSn¯. The bialgebra structure is defined by the requirement that the series

σ1=

n0

Sn and σ¯1=

n0

Sn¯ (1)

are grouplike. The internal product of MR can be computed from the splitting for- mula

(f1· · ·fr)g=μr·(f1⊗ · · · ⊗fr)rΔrg, (2) whereμr isr-fold multiplication, andΔr the iterated coproduct with values in the rth tensor power and the conditions:σ1is neutral,σ¯1is central, andσ¯1∗ ¯σ1=σ1. 2.3 Noncommutative symmetric functions of typeB

Noncommutative symmetric functions of typeB were introduced in [9] as the right Sym-module BSym freely generated by another sequence(S˜n)(n≥0, S˜0=1) of homogeneous elements, withσ˜1grouplike. This is a coalgebra, but not an algebra. It is endowed with an internal product for which each homogeneous component BSymn is antiisomorphic to the descent algebra ofBn.

It should be noted that with this definition, the restriction of the internal product of BSym to Sym is not the internal product of Sym. To remedy this inconvenience, we use a different realization of BSym. We embed BSym as a subcoalgebra and sub-Sym-module of MR as follows. Define, forFSym(A),

F=F (A| ¯A)=F (AqA)¯ |q=−1 (3) called the supersymmetric version, or superization, ofF [16]. It is also equal to

F=Fσ1. (4)

Indeed,σ1is grouplike, and forF=SI, the splitting formula gives (Si1· · ·Sir)σ1=μr

(Si1⊗ · · · ⊗Sir)

σ1⊗ · · · ⊗σ1

=SI . (5) We have

σ1= ¯λ1σ1=

Λ¯iSj. (6)

The elementσ¯1is central for the internal product, and

¯

σ1F (A,A)¯ =F (A, A)¯ =F∗ ¯σ1. (7) The basis elementS˜I of BSym, whereI=(i0, i1, . . . , ir)is aB-composition (that is,i0may be 0), can be embedded as

S˜I=Si0(A)Si1i2···ir(A| ¯A). (8) We will identify BSym with its image under this embedding.

(4)

2.4 Other notation

For a partition λ, we denote by mi(λ) the multiplicity of i in λ and set mλ :=

i1mi(λ)!.

The reverse refinement order on compositions is denoted by. The nonincreasing rearrangement of a composition is denoted byI↓. The refinement order on partitions is denoted by≺p:λpμifλis finer thanμ, that is, each part ofμis a sum of parts ofλ.

3 Descent algebras of type A

3.1 Principal idempotents

In [17], a recursive construction of complete sets of orthogonal idempotents of de- scent algebras has been described. In [14], one finds a general method for construct- ing such families from an arbitrary sequence of Lie idempotents, as well as many remarkable families of Lie idempotents. It is therefore natural to investigate whether the resulting idempotents can be derived from a (possibly known) sequence of Lie idempotents. We shall show that it is indeed the case.

LetPn be the sequence of partitions ofnordered in the following way: first, sort them by decreasing length, then, for each length, order them by reverse lexicographic order. We denote this order by≤. For example,

P5= [11111,2111,311,221,41,32,5]. (9) Now, start with

e1n:= 1

n!S1n (10)

and define by induction

eλ:= 1 mλSλ

Sn

μ<λ

eμ . (11)

Theorem 3.1 [17] The family(eλ)λnforms a complete system of orthogonal idem- potents for Symn.

Following [14], define the (left) Zassenhaus idempotentsζnby the generating se- ries

σ1=:

k1

eζk= · · ·eζ3eζ2eζ1. (12) For example,

S1=ζ1, S2=ζ2+1

2ζ12, S3=ζ3+ζ2ζ1+1

6ζ13, (13)

(5)

S4=ζ4+ζ3ζ1+1 2ζ22+1

2ζ2ζ12+ 1

24ζ14, (14)

S5=ζ5+ζ4ζ1+ζ3ζ2+1

2ζ3ζ12+1

2ζ22ζ1+1

6ζ2ζ13+ 1

120ζ15, (15) S6=ζ6+ζ5ζ1+ζ4ζ2+1

2ζ4ζ12+1

2ζ32+ζ3ζ2ζ1 +1

6ζ3ζ13+1 6ζ23+1

4ζ22ζ12+ 1

24ζ2ζ14+ 1

720ζ16, (16)

so that

ζ1=S1, ζ2=S2−1

2S11, ζ3=S3S21+1

3S111, (17)

ζ4=S4S31−1 2S22+3

4S211+1

4S112−1

4S1111, (18)

ζ5=S5S41S32+S311+S212−2

3S2111−1

3S1112+1

5S11111, (19)

ζ6=S6S51S42+S411−1 2S33+1

2S321+S312−5 6S3111 +1

3S222−1

6S2211+1

2S213−1

2S2121−2

3S2112+13 24S21111

−1

6S1122+ 1

12S11211−1

6S1113+1

6S11121+ 5

24S11112−1

6S111111. (20) Note that, in particular,

Sn=

λn

1

mλζλ1ζλ2· · ·ζλr. (21) For a compositionI=(i1, . . . , ir), define as usualζI:=ζi1· · ·ζir. SinceζnSn modulo smaller terms in the refinement order on compositions,ζISI modulo smaller terms. So theζI family is unitriangular on the basisSJ, so it is a basis of Sym.

In the sequel, we shall need a condition for a productSIζJ to be zero.

Lemma 3.2 LetIandJbe two compositions ofn. Then,

SIζJ=

⎧⎪

⎪⎩

0 ifJ↓ ≺pI,

mIζI ifJ↓ =I↓,

K↓=JcI JKζK otherwise,

(22)

wherecKI J is the number of ways of unshufflingJ intop=(I )subwords such that J(l)has sumil and whose concatenationJ(1)J(2)· · ·J(p)isK.

(6)

Proof Since the Zassenhaus idempotentsζm are primitive, we have, thanks to the splitting formula (2),

SIζJ=

J(1),...,J(p)

Si1ζJ(1)

· · ·

SipζJ(p)

, (23)

where the sum ranges over all possible ways of decomposingJ into p (possibly empty) subwords.

SinceSijζJ(j )=0 ifJ(j )is not a composition ofij, it follows thatSIζJ =0 ifJ↓ ≺pI↓. Moreover, ifJ↓ =I↓, then

SIζJ =

σSp

(Si1ζJσ (1))· · ·(SipζJσ (p))=mIζI. (24)

IfJ↓ ≺pI↓andJ↓ =I↓, then a term in the r.h.s. of (23) is nonzero iff allJ()are compositions ofi. In that case, we have

SIζJ=

J(1),...,J(p)

ζJ(1)· · ·ζJ(p)=

J(1),...,J(p)

ζJ(1)J(2)···J(p), (25)

whence the last case.

Theorem 3.3 For all partitionsλ=1, λ2, . . . , λk), eλ= 1

mλζλ1· · ·ζλk. (26) Proof Leteλbe the right-hand side of (26). We will show that these elements satisfy the same induction as theeλ(11).

From Lemma3.2we have

Sλeλ=mλeλ. (27)

Now, using (21), we get mλeλ=Sλ

Sn

μ=λ

eμ =Sλ

Sn

μ<λ

eμ , (28)

where the last equality follows again from Lemma3.2. Hence,eλ=eλ. Note that, thanks to Lemma3.2, the induction formula foreλsimplifies to

eλ= 1 mλSλ

Sn

μpλ

eμ . (29)

(7)

3.2 A basis of idempotents

As with any sequence of Lie idempotents, we can construct an idempotent basis of Symn from the ζn. Here, the principal idempotentseλ are members of the basis, which leads to a simpler derivation of the representation theory.

We start with a basic lemma, easily derived from the splitting formula (compare [14, Lemma 3.10]). Recall that the radical of(Symn,)isRn=RSymn, where Ris the kernel of the commutative image SymSym.

Lemma 3.4 Denote byS(J )the set of distinct rearrangements of a compositionJ. LetI=(i1, . . . , ir)andJ=(j1, . . . , js)be two compositions ofn. Then,

(i) If(J ) < (I ), thenζIζJ=0.

(ii) If(J ) > (I ), thenζIζJ∈VectζK:K∈S(J ) ∩R. More precisely, ζIζJ=

J1,...,Jr

|Jk|=ik

J, J1 · · · JrΓJ1· · ·ΓJr, (30)

where for a compositionKofk,ΓK:=ζkζK.

(iii) If(J )=(I ), thenζIζJ =0 only for J∈S(I ), in which caseζIζJ = mIζI.

Note that the ΓK are in the primitive Lie algebra. This follows from the

∗-multiplicativity of the coproduct: Δ(fg)=Δ(f )Δ(g), see [11, Proposi- tion 5.5].

Corollary 3.5 The elements

eI= 1

mIζI, In, (31)

are all idempotents and form a basis of Symn. This basis contains in particular the principal idempotentseλ.

3.3 Cartan invariants

By (iii) of Lemma 3.4, the indecomposable projective module Pλ=Symneλ

contains the eI for I ∈S(λ). For I ∈S(λ), (i) and (ii) imply that eIeλ is in VectζK:K∈S(λ). Hence, this space coincides withPλ. So, we get immediately the explicit decomposition

Symn=

λn

Pλ, Pλ=

IS(λ)

CeI. (32)

The Cartan invariants

cλμ=dim(eμSymneλ) (33)

(8)

are also easily obtained. The above space is spanned by the

eμeIeλ=eμeI, I∈S(λ). (34) By (ii) of Lemma3.4, this is the dimension of the space[Sμ(L)]λspanned by all sym- metrized products of Lie polynomials of degreesμ1, μ2, . . .formed fromζi1, ζi2, . . ., hence giving back the classical result of Garsia–Reutenauer [10].

3.4 Quiver andq-Cartan invariants (Loewy series)

Still relying upon point (ii) of Lemma3.4, we see thatcλμ=0 ifλis not finer than (or equal to)μ. Further, ifμis obtained fromλby adding up two partsλi, λj, then eμeI=0 ifλi=λj and is a nonzero element of the radical otherwise.

In [8], it is shown that the powers of the radical for the internal product coincide with the lower central series of Sym for the external product:

Rj=γj(Sym), (35)

where γj(Sym) is the ideal generated by the commutators [Sym, γj−1(Sym)].

Hence, forλ finer thanμ, eμeI is nonzero modulo R2 iff μis obtained from λ by summing two distinct parts. More generally, eμeI is in Rk and nonzero moduloRk+1iff(λ)(μ)=k.

Summarizing, we have the following:

Theorem 3.6 [8,17]

(i) In the quiver of Symn, there is an arrowλμiffμis obtained fromλby adding two distinct parts.

(ii) Theq-Cartan invariants are given by

cλμ(q)=q(λ)(μ) (36)

ifλis finer than (or equal to)μandcλμ(q)=0 otherwise.

4 Descent algebras of typeB

4.1 Reinterpretation of some results of Chow

We begin by showing that in our realization of BSym, Chow’s map Θ (see [9, Sect. 3.4]) corresponds to the left internal product by the reproducing kernelσ1of the superization map. Chow’s conditionΘ(S˜n)=Sntranslates into the obvious equality σ1σ1=σ1. The second conditionΘ(Sn(A))=Sn(2A)amounts to

σ1σ1= σ12

, (37)

which is an easy consequence of the splitting formula σ1σ1=(¯λ1σ1)σ1=λ¯1σ1

σ1= σ12

(38)

(9)

since

λ¯1σ1=1σ1)(λ1λ1)=σ1. (39) Here we used the fact that left∗-multiplication byλ¯1is an antiautomorphism. Denot- ing byμas in [9] the twisted product

μ(ABC)=1B)AC, (40)

we have the following:

Lemma 4.1 ForF, GBSym,

σ1(F G)=μ σ1F

Δ(G)

. (41)

Proof

σ1(F G)=μ

1σ1)(ΔF ΔG)

=

(F ),(G)

μ

1F1G1)F2G2

=

1G1)

1F1)F2

G2=μ σ1F

Δ(G)

. (42)

This is Chow’s third condition, which completes the characterization ofΘ.

4.2 Idempotents in BSym

Define the elementsζnBSym by the generating series σ1=:

eζ1eζ2eζ3· · ·

· · ·eζ3eζ2eζ1

. (43)

For example, collecting the terms of weights 1, 2, and 3, respectively, we have S1=2ζ1, S2=2ζ2+2ζ12, S3=2ζ3+2ζ2ζ1+2ζ1ζ2+4

3ζ13, (44) so that

ζ1=1

2S1, ζ2=1 2S2−1

4S11, ζ3=1 2S3−1

4S21−1

4S12+1 6S111.

(45) Note that the elementsζn are well defined and that they are primitive. We shall use the notation

eζ1eζ2· · ·

=:E(ζ ),

· · ·eζ2eζ1

=:E(ζ ). (46)

Next, define the elementsζ˜nBSym by the generating series

σ1=:

n0

ζ˜n · · ·eζ2eζ1

=: ˜ζE(ζ ). (47)

(10)

For example,

S1=ζ1+ ˜ζ1, S2=ζ2+1

2ζ12+ ˜ζ1ζ1+ ˜ζ2, (48) S3=1

6ζ13+ζ2ζ1+ζ3+ ˜ζ1ζ2+1

2ζ˜1ζ12+ ˜ζ2ζ1+ ˜ζ3, (49) so that

ζ˜1=S1−1

2S1, ζ˜2=S2−1 2S2−1

2S1S1+3

8S11, (50)

ζ˜3=S3−1

2S2S1−1

2S1S2+3

8S1S11−1 2S3+1

4S21+1

2S12− 5

16S111. (51) Sinceσ1 is grouplike and sinceeζn is grouplike for alln≥1, the series ζ˜ is also grouplike.

The next two lemmas describe some properties of the elementsζ˜nandζn. Lemma 4.2 The ordered exponentials are exchanged as follows:

λ1E(ζ )=E(ζ ). (52)

In particular,λ1ζi=ζi for alli≥0.

Proof λ1∗ ·is an antiautomorphism, so the left-hand side is λ1eζ1

λ1eζ2

· · ·. (53)

Taking into account (39) and recalling that (43) characterizes theζi, we see that if we set

ζi:=λ1ζi, (54)

then

σ1=

eζ1eζ2· · ·

· · ·eζ2eζ1

, (55)

so thatζi=ζi.

Lemma 4.3 For alln≥1,

σ1∗ ˜ζn=0. (56)

Proof By definition,

E(ζ )E(ζ )=σ1=σ1σ1=σ1ζ˜E(ζ )

=1σ1)ζ˜E(ζ )

=μ

1σ1)ζ˜E(ζ )⊗ ˜ζE(ζ )

=

λ1∗ ˜ζE(ζ )ζ˜E(ζ )

(11)

=

λ1E(ζ )

1∗ ˜ζ )ζ˜E(ζ )=E(ζ )(λ1∗ ˜ζ )(σ1∗ ˜ζ )E(ζ )

=E(ζ ) σ1∗ ˜ζ

E(ζ ), (57)

so thatσ1∗ ˜ζ =1.

Lemma 4.4 For alln≥1,

σ1ζn=2ζn. (58)

Proof We have

σ1ζn=1σ1)ζn=μ

1σ1)n⊗1+1⊗ζn)

=2ζn. (59) Proposition 4.5 Let I = (i0, . . . , ip) be a B-composition of n, and let λ = 0, . . . , λk)be aB-partition ofn.

S˜I∗ ˜ζλ0ζλ1· · ·ζλk =

0 ifλI,

(2p

j1mj!)˜ζi0ζi1· · ·ζip ifλ=I, (60) wheremj is the multiplicity ofj in(i1, i2, . . . , ip)(not countingi0!).

Proof The splitting formula yields S˜I∗ ˜ζλ0ζλ1· · ·ζλk =μp

Si0Si

1⊗ · · · ⊗Si

p

ζ˜α0ζα(0)⊗ · · · ⊗ ˜ζαpζα(p)

=

Si0∗ ˜ζα

0ζα(0) Si

1∗ ˜ζα

1ζα(1)

· · · Si

p∗ ˜ζα

pζα(p)

. (61) By (4) and Lemma4.3, a summand is zero if anyαi>0 fori≥1, so that

S˜I∗ ˜ζλ0ζλ1· · ·ζλk=

Si0∗ ˜ζα0ζα(0) Si

1ζα(1)

· · · Si

rζα(p)

. (62) If a term is nonzero in this equation, thenλI↓. This proves the first case. By (4) and Lemma4.4,Siζi=2ζi, which proves the second case.

We are now in a position to give an explicit formula for the idempotents of [17].

Theorem 4.6 For allB-partitionsλ=0, λ1, . . . , λk)ofn, define the elementseλBSymnrecursively by the formula

eλ= 1

2k

jmj!S˜λ

Sn

μ<λ

eμ , (63)

wheremj is the multiplicity ofj in(λ1, . . . , λk)(not countingλ0!). Then eλ= 1

jmj!ζ˜λ0ζλ1· · ·ζλk. (64)

(12)

Proof Leteλbe the right-hand side of the above equation. By Proposition4.5, S˜λeλ = ˜Sλ∗ 1

jmj!ζ˜λ0ζλ1· · ·ζλk=

2k

j

mj! eλ. (65)

By (47),Sn=

eλ, where the sum ranges over allB-partitionsλofn. Together with the above and Proposition4.5, we have

2k

j

mj! eλ= ˜Sλeλ= ˜Sλ

Sn

μ=λ

eμ = ˜Sλ

Sn

μ<λ

eμ . (66)

Since theeλ satisfy the same induction as theeλ, they are equal.

Theorem 4.7 [17] The family(eλ)whereλruns overB-partitions ofnforms a com- plete system of orthogonal idempotents for BSymn.

5 Idempotents in the higher order peak algebras

Letq be a primitiverth root of unity. We denote byθq the endomorphism of Sym defined by

f˜=θq(f )=f

(1q)A

=f (A)σ1

(1q)A

. (67)

We denote by ˚P(r)the image ofθq and byP(r)the right ˚P(r)-module generated by Sn forn≥0. Note that ˚P(r)is by definition a left∗-ideal of Sym. Forr=2, ˚P(2) is the classical peak ideal, andP(2)is the unital peak algebra. For generalr, ˚P(r)is the higher-order peak algebra of [13], andP(r)is its unital extension defined in [3].

These objects depend only onrand not on the choice of the primitive root of unity.

Bases ofP(r)can be labeled byr-peak compositionsI=(i0;i1, . . . , ip)with at most one parti0divisible byr.

5.1 The radical

By definition,Pn(r)is a∗-subalgebra of Symn. The radical of Symnconsists of those elements whose commutative image is zero (see [14], Lemma 3.10). The radical of Pn(r)is therefore spanned by theSi0·θq(SISI)such thatIis a permutation ofI. Indeed, the quotient ofPn(r)by the span of those elements is a semisimple commuta- tive algebra, the∗-subalgebra of Symnspanned by thepλn) such that at most one part ofλis multiple ofr. This special part will be denoted byλ0.

We denote byPn(r)the subset of partitions ofnwith at most one part divisible byr.

The simplePn(r)-modules, and the principal idempotents, can therefore be labeled byPn(r).

(13)

5.2 An induction for the idempotents

Define a total order<onPn(r)as follows: sort the partitions by decreasing length and sort partitions of the same length by reverse lexicographic order. For example,

P5(2)= [11111,2111,311,41,23,5], (68) P7(2)= [1111111,211111,31111,4111,2311,511,331,61,43,25,7]. (69) Now, set

e1(r)n := 1

n!S1n (70)

and define by induction

eλ(r):= 1 mλ

Tλ

Sn

μ<λ

e(r)μ , (71)

whereTm=Rmifr|m,Tm=Rrijifrmandm=ir+jwith 0< j < r, andTλ= Tλ0Tλ1· · ·Tλp forλ=0;λ1, . . . , λp)Pn(r). It follows from [13, Corollary 3.17]

and from the definition ofP(r)thatTλP(r). We want to prove that(e(r)λ )

λPn(r)is a complete system of orthogonal idempotents forP(r).

To this aim, we introduce the sequence of (left) Zassenhaus idempotentsζn(r)of levelras the unique solution of the equation

σ1=

p0

ζpr(r) i1,ri

eζi(r). (72)

Note thatζn(r)=ζnforn <2r.

For example, forr=2,

ζ1(2)=S1; ζ2(2)=S2−1

2S11; ζ3(2)=S3S21+1

3S111, (73) ζ4(2)=S4S31+1

2S211−1

8S1111, (74)

ζ5(2)=S5S41+1

2S311S23+S221−1

2S2111+1 2S113

−1

2S1121+1

5S11111; (75)

and forr=3,

ζ1(3)=S1; ζ2(3)=S2−1

2S11; ζ3(3)=S3S21+1

3S111, (76)

(14)

ζ4(3)=S4S31−1 2S22+3

4S211+1

4S112−1

4S1111, (77)

ζ5(3)=S5S41S32+S311+S212−2

3S2111−1

3S1112+1

5S11111, (78)

ζ6(3)=S6S51S42+S411+S312−2

3S3111+1

3S222−1 6S2211

−2

3S2112+3

8S21111−1

6S1122+ 1

12S11211+ 5

24S11112−1

9S111111. (79) Define now forλ=0;λ1, . . . , λk)Pn(r),

e(r)λ := 1 mλζλ(r)

0 ζλ(r)

1 · · ·ζλ(r)

k . (80)

We will show thate(r)λ =e(r)λ for allλPn(r). We begin with two lemmas.

Lemma 5.1

Δ ζn(r)

=

1⊗ζn(r)+ζn(r)⊗1 ifrn, n/r

i=0ζir(r)ζn(r)ir ifr|n.

(81)

Proof This means that ζn(r) are primitive if r |n and that the generating series

p0ζrp(r)is grouplike. If we define new elementsYpby

σ1=:

p

eYrp

ri

eYi, (82)

the standard argument showing that the Zassenhaus elements are primitive shows as well that all theYi are primitive. Now identify the first product in the right-hand side with the generating series

p0ζrp(r). Thenζi(r)=Yi ifri. Since the exponential of a primitive element is grouplike and a product of grouplike series is group-like, both products in the right-hand side above are grouplike. By identification,

p0ζrp(r)is

grouplike.

Lemma 5.2 Letλ=0;λ1, . . . , λk)Pn(r) and I =(i0, i1, . . . , ip)be anr-peak composition ofn. Then,

TIζλ(r)

0 ζλ(r)

1 · · ·ζλ(r)

k =

0 ifI< λ,

mIζi(r)

0 ζi(r)

1 · · ·ζi(r)

k ifI↓ =λ. (83)

(15)

Proof To simplify the notation, we letζ(r)λ=ζλ(r)

0 ζλ(r)

1 · · ·ζλ(r)

k forλPn(r). IfFI=

Fi1· · ·Fip with eachFijSymij, the splitting formula and (81) yield

FIζ(r)λ=

λ(1)∨···∨λ(p)=1,...,λk) ra1+···+rap=λ0

p j=1

Fijζra(r)

jζ(r)λ

(j )

, (84)

whereλ0is the part (possibly 0) ofλthat is divisible byr, and whereαβdenotes the partition obtained by reordering the concatenation of the partitionsαandβ.

Observe that since at most oneij is divisible byr, a product in the above summa- tion is 0 if at least two of the partitionsλ(1), . . . , λ(p)are empty. If(I ) > (λ), then kp−2, so this hypothesis is always satisfied. Thus,

FIζ(r)λ=0 if(I ) > (λ). (85)

SupposeI↓ ≤λ. Then, by the definition of the order,(I )(λ). Hence, forI such thatI↓ ≤λand(I ) > (λ), the result follows by takingFI=TI in (85). So suppose instead that(I )=(λ). By definition,TI=Ti1· · ·Tip, whereTm=Rmif r|mandTm=Rrijifm=ir+jwith 0< j < r. SinceRJcan be written as a linear combination ofSK for whichJ is a refinement ofK, it follows thatTI is equal to SIplus a linear combination ofSKwith(K) > (I ). By takingFI=SK in (85), it follows thatTIζ(r)λ=SIζ(r)λ.

It remains to show that, forI andλof the same length,SIζ(r)λ=mIζ(r)λ if I↓ =λand is 0 otherwise. Ifλcontains no part divisible byr, then it follows from (84) that ifSIζ(r)λ=0, we must haveI↓ =λ, in which caseSIζ(r)λ=mIζ(r)λ. Suppose instead thatλcontains a part that is divisible byr. Then each decomposi- tion(1), . . . , λ(p))in (84) contains at least oneλ(j )= ∅. Thus, ifI contains no part that is divisible byr, thenSIζ(r)λ=0. Otherwise, the part ofI that is divisible byr is bounded byλ0. This implies thatλI↓. Since we began by assuming that I↓ ≤λ, it follows thatI↓ =λ, and the result follows from (84) as before.

Theorem 5.3 For all partitionsλ=0;λ1, . . . , λk)Pn(r), e(r)λ =e(r)λ := 1

mλ

ζλ(r)

0 ζλ(r)

1 · · ·ζλ(r)

k . (86)

Proof From the definition of ζm(r) it follows that Sn =

λPn(r)e(r)λ . Hence, by Lemma5.2,

mλe(r)λ =Tλeλ=Tλ

Sn

μ=λ

e(r)μ =Tλ

Sn

μ<λ

e(r)μ .

Thus, the elementse(r)λ andeλ(r)satisfy the same induction equation (71).

参照

関連したドキュメント

However its power ∇ / 2 , though not conformally covariant, has positive definite leading symbol (in fact, leading symbol |ξ| 2 Id), and so satisfies our analytic and

Similarly, an important result of Garsia and Reutenauer characterizes which elements of the group algebra k S n belong to the descent algebra Sol( A n−1 ) in terms of their action

Abstract The classical abelian invariants of a knot are the Alexander module, which is the first homology group of the the unique infinite cyclic covering space of S 3 − K ,

We study the classical invariant theory of the B´ ezoutiant R(A, B) of a pair of binary forms A, B.. We also describe a ‘generic reduc- tion formula’ which recovers B from R(A, B)

• characters of all irreducible highest weight representations of principal W-algebras W k (g, f prin ) ([T.A. ’07]), which in particular proves the conjecture of

Giuseppe Rosolini, Universit` a di Genova: rosolini@disi.unige.it Alex Simpson, University of Edinburgh: Alex.Simpson@ed.ac.uk James Stasheff, University of North

Before we proceed to investigate locally finite simple Lie algebras which are of a more general form than those discussed in the previous sections, we need some results on certain

The explicit treatment of the metaplectic representa- tion requires various methods from analysis and geometry, in addition to the algebraic methods; and it is our aim in a series