• Nebyly nalezeny žádné výsledky

Connection Formula for the Jackson Integral of Type A

N/A
N/A
Protected

Academic year: 2022

Podíl "Connection Formula for the Jackson Integral of Type A"

Copied!
42
0
0

Načítání.... (zobrazit plný text nyní)

Fulltext

(1)

Connection Formula for the Jackson Integral of Type A

n

and Elliptic Lagrange Interpolation

Masahiko ITO and Masatoshi NOUMI

Department of Mathematical Sciences, University of the Ryukyus, Okinawa 903-0213, Japan E-mail: mito@sci.u-ryukyu.ac.jp

Department of Mathematics, Kobe University, Rokko, Kobe 657-8501, Japan E-mail: noumi@math.kobe-u.ac.jp

Received January 23, 2018, in final form July 07, 2018; Published online July 24, 2018 https://doi.org/10.3842/SIGMA.2018.077

Abstract. We investigate the connection problem for the Jackson integral of typeAn. Our connection formula implies a Slater type expansion of a bilateral multiple basic hypergeo- metric series as a linear combination of several specific multiple series. Introducing certain elliptic Lagrange interpolation functions, we determine the explicit form of the connection coefficients. We also use basic properties of the interpolation functions to establish an ex- plicit determinant formula for a fundamental solution matrix of the associated system of q-difference equations.

Key words: Jackson integral of type An; q-difference equations; Selberg integral; Slater’s transformation formulas; elliptic Lagrange interpolation

2010 Mathematics Subject Classification: 33D52; 39A13

1 Introduction

Throughout this paper we fix the baseq ∈Cwith 0<|q|<1, and use the notation ofq-shifted factorials

(u)=

Y

l=0

1−uql

, (a1, . . . , ar)= (a1)· · ·(ar), (u)ν = (u)/ uqν

, (a1, . . . , ar)ν = (a1)ν· · ·(ar)ν.

For generic complex parameters a1, . . . , ar and b1, . . . , br, the bilateral basic hypergeometric series rψr [9, equation (5.1.1)] is defined by

rψr

a1, . . . , ar

b1, . . . , br;q, x

=

X

ν=−∞

(a1, . . . , ar)ν

(b1, . . . , br)ν xν, where |b1· · ·br/a1· · ·ar|<|x|<1.

An important summation for 1ψ1 series is Ramanujan’s theorem [9, equation (5.2.1)],

1ψ1 a

b;q, x

= (ax, q, b/a, q/ax)

(x, b, q/a, b/ax)

. (1.1)

There is also a summation formula for a very-well-poised 6ψ6 series due to Bailey [9, equa- tion (5.3.1)],

6ψ6

q√ a,−q√

a, b, c, d, e

√a, −√

a,aqb ,aqc,aqd,aqe ;q, a2q bcde

= aq,aqbc,aqbd,aqbe,aqcd,aqce,aqde, q,qa

aq

b ,aqc,aqd,aqe,qb,qc,qd,qe,bcdea2q

, (1.2)

This paper is a contribution to the Special Issue on Elliptic Hypergeometric Functions and Their Applications.

The full collection is available athttps://www.emis.de/journals/SIGMA/EHF2017.html

(2)

where|a2q/bcde|<1. Furthermore there are a lot of transformation formulas for therψr series.

In the paper [30] Slater proved the transformation formula

rψr

a1, . . . , ar b1, . . . , br;q, z

=

c1

a1, . . . ,ca1

r, c2, . . . , cr,cq

2, . . . ,cq

r,bc1q

1 , . . . ,bcrq

1 , Ac1z,Acq

1z

q

a1, . . . ,aq

r,cc1

2, . . . ,cc1

r,cc2q

1 , . . . ,ccrq

1 , b1, . . . , br, Azq,Az1

× rψr

"a1q

c1 , . . . ,acrq

1

b1q

c1 , . . . ,bcrq

1

;q, z

#

+ idem(c1;c2, . . . , cr), (1.3) whereA=a1· · ·ar/c1· · ·crand|b1· · ·br/a1· · ·ar|<|z|<1, which is called theSlater’s general transformation for a rψr series [9, equation (5.4.3)]. Here the symbol “idem(c1;c2, . . . , cr)”

stands for the sum of the r −1 expressions obtained from the preceding one by interchan- ging c1 with eachck, k= 2, . . . , r. Slater also proved a transformation formula for a very-well- poised 2rψ2r series, which is expressed as

2rψ2r q√

a,−q√

a, b3, . . . , b2r

√a,−√ a,aqb

3, . . . ,baq

2r

;q,ar−1qr−2 b3. . . b2r

= a4, . . . , ar,aq

4, . . . ,aq

r,aa4, . . . ,aar,aqa

4, . . . ,aqa

r,ab3q

3 , . . . ,ab3q

2r,aaq

3b3, . . . ,aaq

3b2r

q

b3, . . . ,bq

2r,aqb

3, . . . ,baq

2r,aa4

3, . . . ,aar

3,aa3q

4 , . . . ,aa3q

r ,a3aa4, . . . ,a3aar,aaq

3a4, . . . ,aaq

3ar

× aq,qa

a23q

a ,aqa2 3

2rψ2r

"qa

3

a,−qaa3,a3ab3, . . . ,a3ab2r

a3

a,−a3

a,ab3q

3 , . . . ,ab3q

2r

;q,ar−1qr−2 b3. . . b2r

#

+ idem(a3;a4, . . . , ar), (1.4) where |ar−1qr−2/b3· · ·b2r| < 1, which is called Slater’s transformation for a very-well-poised balanced 2rψ2r series[9, equation (5.5.2)].

The aim of this paper is to give an explanation and a generalization of Ramanujan’s 1ψ1

summation (1.1) and Slater’s rψr transformation (1.3) in relation to the Selberg integral [29]

(see also the recent sources [7,8]), i.e., 1

n!

Z 1 0

· · · Z 1

0 n

Y

i=1

ziα−1(1−zi)β−1 Y

1≤j<k≤n

|zj−zk|dz1dz2· · ·dzn

=

n

Y

j=1

Γ(α+ (j−1)τ)Γ(β+ (j−1)τ)Γ(jτ)

Γ(α+β+ (n+j−2)τ)Γ(τ) , (1.5)

where Re(α) > 0, Re(β) > 0 and Re(γ) > −min{1/n,Re(α)/(n−1),Re(β)/(n−1)}, which is a multi-dimensional generalization of the evaluation of the Euler beta integral in terms of products of gamma functions. (This is not the first such paper using Selberg integrals to obtain multiple analogues of some of Slater’s transformation formulas. For example, the paper [14]

about a generalization for Bailey’s very-well-poised6ψ6 summation (1.2) and Slater’s very-well- poised2rψ2r transformation (1.4) withBCnsymmetry, is very similar in spirit and methodology to the current paper.) For this purpose we define some terminology about a q-analog of the Selberg type integral. For z= (z1, . . . , zn)∈(C)n, let Φ(z) be specified by

Φ(z) = Φs(z;α, a, b;q, t)

= (z1z2· · ·zn)α

n

Y

i=1 s

Y

m=1

qa−1m zi

(bmzi)

Y

1≤j<k≤n

zj2τ−1 qt−1zk/zj

(tzk/zj)

, (1.6)

withα∈C,a= (a1, . . . , as),b= (b1, . . . , bs)∈(C)s andt∈C, whereτ ∈Cis given byt=qτ. For z= (z1, . . . , zn)∈(C)n, let ∆(z) be the Vandermonde product

∆(z) = Y

1≤j<k≤n

(zj−zk). (1.7)

(3)

For a functionϕ=ϕ(z) of z= (z1, . . . , zn)∈(C)n, we denote by hϕ, zi=

Z z∞

0

ϕ(w)Φ(w)∆(w)dqw1

w1 ∧ · · · ∧ dqwn

wn

= (1−q)n X

ν∈Zn

ϕ(zqν)Φ(zqν)∆(zqν), (1.8)

the Jackson integral associated with the multiplicative lattice zqν = (z1qν1, . . . , znqνn)∈(C)n, ν = (ν1, . . . , νn) ∈ Zn. Note that the sum hϕ, zi is invariant under the shifts z → zqν for all ν ∈Znwhen it converges. This sum is called theJackson integral of A-typein the context of [1].

For the general setting of Jackson integrals see [3].

We remark that for s= 1, ϕ≡ 1 andz = a1, a1t, . . . , a1tn−1

∈(C)n, the Jackson integ- ral (1.8) is expressed as

1, a1, a1t, . . . , a1tn−1

= (1−q)n

n

Y

j=1

a1tj−1α+2(n−j)τ (q)(t) qαa1b1tn+j−2

(tj) qαtj−1

a1b1tj−1

, (1.9)

whose limiting case where q → 1 is equivalent to the Selberg integral (1.5). In this sense the Jackson integral (1.8) includes the Selberg integral as a special case. The formula (1.9) was discovered by Askey [5] and proved by Habsieger [11], Kadell [16], Evans [6] and Kaneko [19] in the case whereτ is a positive integer, while (1.9) for general complexτ was given by Aomoto [1].

See [13] for further details. Other closely related and relevant works are [17,18,33].

We denote bySnthe symmetric group of degreen; this group acts on the field of meromorphic functions on (C)n through the permutations of variables z1, . . . , zn. Setting

hhϕ, zii= hϕ, zi

Θ(z), Θ(z) =

n

Y

i=1

ziα

s

Q

m=1

θ(bmzi) Y

1≤j<k≤n

zjθ(zj/zk)

θ(tzj/zk) , (1.10)

where θ(u) = (u) qu−1

, we have the following.

Lemma 1.1. Let ϕ(z) be an Sn-invariant holomorphic function on (C)n, and suppose that the Jackson integral hϕ, zi of (1.8) converges as a meromorphic function on (C)n. Then the functionf(z) =hhϕ, ziidefined by (1.10)is anSn-invariant holomorphic function on(C)n. Fur- thermore it satisfies the quasi-periodicityTq,zif(z) =f(z)/(−zi)sqαtn−1b1· · ·bs for i= 1, . . . , n, where Tq,zi stands for the q-shift operator inzi.

For the proof of this lemma, see Lemma3.3.

We call hhϕ, zii the regularized Jackson integral of A-type, which is the main object of this paper. We remark that, when ϕ(z) is a symmetric polynomial such that degziϕ(z) ≤ s−1, i= 1, . . . , n, it satisfies the condition of Lemma 1.1under the assumption (3.2) below.

In order to state our first main theorem we define some terminology. We set Z =Zs,n =

µ= (µ1, µ2, . . . , µs)∈Ns12+· · ·+µs =n , (1.11) where N ={0,1,2, . . .}, so that |Zs,n|= s+n−1n

. We use the symbol for the lexicographic order onZs,n. Namely, forµ, ν ∈Zs,n, we denoteµ≺ν if there existsk∈ {1,2, . . . , n}such that µ11, µ22, . . . , µk−1k−1 and µk < νk. For an arbitrary x= (x1, x2, . . . , xs) ∈ (C)s, we consider the pointsxµ (µ∈Zs,n) in (C)n specified by

xµ= x1, x1t, . . . , x1tµ1−1

| {z }

µ1

, x2, x2t, . . . , x2tµ2−1

| {z }

µ2

, . . . , xs, xst, . . . , xstµs−1

| {z }

µs

∈(C)n. (1.12)

(4)

Theorem 1.2(connection formula). Suppose thatϕ(z)is anSn-invariant holomorphic function satisfying the condition of Lemma 1.1. Then, for genericx∈(C)s we have

hhϕ, zii= X

λ∈Zs,n

cλhhϕ, xλii, (1.13)

where the connection coefficients cλ are explicitly written as

cλ = X

K1t···tKs

={1,2,...,n}

s

Y

i=1

Y

k∈Ki

"

θ

qαb1· · ·bstn−1zk Q

1≤l≤s l6=i

xltλ(k−1)l

θ

qαb1· · ·bstn−1

s

Q

l=1

xltλ(k−1)l

× Y

1≤j≤s j6=i

θ zkx−1j t−λ

(k−1)

j

θ xitλ(k−1)i x−1j t−λ

(k−1)

j

#

, (1.14)

where λ(k)i =|Ki∩ {1,2, . . . , k}|, and the summation is taken over all partitionsK1t · · · tKs= {1,2, . . . , n} such that |Ki|=λi, i= 1,2, . . . , s.

We call the formula (1.13) the generalized Slater transformation. In fact, the connection formula (1.13) forn= 1 coincides with the transformation formula (1.3) ifϕ(z)≡1, as explained below.

Example 1.3. The case n = 1. Setting i = (0, . . . ,0,

i

^

1,0, . . . ,0) for i = 1,2, . . . , s, we have Zs,1={1, 2, . . . , s} and xi =xi ∈C, i.e., we obtain

hhϕ, zii=

s

X

i=1

cihhϕ, xiii, where ci = θ(qαb1· · ·bsx1· · ·xsz/xi) θ(qαb1· · ·bsx1· · ·xs)

Y

1≤j≤s j6=i

θ(z/xj)

θ(xi/xj).(1.15) The formula (1.15) for ϕ ≡ 1 (which was first stated by Mimachi [27, theorem in Section 4]) coincides with Slater’s transformation in (1.3). For z = ai, i = 1, . . . , s, and xj = b−1j , j = 1, . . . , s, it was given by Aomoto and Kato [4, equation (2.12) in Corollary 2.1]. For further details about the formula (1.15), see also [15, p. 255, Theorem 5.1 in the Appendix].

Example 1.4. The case s= 1. We have Z1,n ={(n)}, which has only one element, and have x(n)= (x, xt, . . . , xtn−1) for x∈C. Then

hhϕ, zii=c(n)hhϕ, x(n)ii, where c(n) =

n

Y

i=1

θ qαb1tn−1zi

θ qαb1tn−1xti−1. (1.16) The formula (1.16) for ϕ≡1 was given by Aomoto in [1]. See also [13, Lemma 3.6].

As we will see below the connection coefficients cµ in (1.14) are characterized, as functions of z∈(C)n, by an interpolation property.

We next apply the connection formula (1.13) in Theorem 1.2 to establish a determinant formula associated with the Jackson integral (1.8). LetBs,n be the set of partitions defined by

B =Bs,n=

λ= (λ1, λ2, . . . , λn)∈Zn|s−1≥λ1 ≥λ2 ≥ · · · ≥λn≥0 , (1.17) so that|Bs,n|= s+n−1n

. We also use the symbolfor the lexicographic order ofBs,n. For each λ∈Bs,n, we denote by sλ(z) theSchur function

sλ(z) = det ziλj+n−j

1≤i,j≤n

det zin−j

1≤i,j≤n

= det ziλj+n−j

1≤i,j≤n

∆(z) ,

which is a Sn-invariant polynomial. Our second main theorem is the following.

(5)

Theorem 1.5 (determinant formula). Suppose that x∈(C)s is generic. Then we have det hhsλ, xµii

λ∈B µ∈Z =C

n

Y

k=1

θ qαx1· · ·xsb1· · ·bstn+k−2(s+k−2k−1 )

×

n

Y

k=1

n−k

Y

r=0

Y

1≤i<j≤s

xjtrθ xix−1j tn−k−2r

 (s+k−3k−1 )

, (1.18)

where the rows and the columns of the matrix are arranged by the lexicographic orders ofλ∈Bs,n

and µ∈Zs,n, respectively. Here C is a constant independent of x, which is explicitly written as

C =

n

Y

k=1

(1−q)s(q)s qt−(n−k+1)s

s

Q

i,j=1

qa−1i b−1j t−(n−k)

qt−1s

qαtn−k

q1−αt−(n+k−2)

s

Q

i=1

a−1i b−1i

 (s+k−2k−1 )

.

Remark. When s= 1, the determinant formula (1.18), combined with the connection formula hh1, zii=c(n)hh1, x(n)ii of (1.16), implies the summation formula

hh1, zii= (1−q)n

n

Y

j=1

(q) qt−j

qa−11 b−11 t−(j−1)

qt−1

qαtj−1

q1−αa−11 b−11 t−(n+j−2)

θ qαb1tn−1zj ,

for the Jackson integral hh1, zii for an arbitrary z ∈ (C)n. This formula for n = 1 coincides with Ramanujan’s formula (1.1). This is another multi-dimensional bilateral extension of Ra- manujan’s 1ψ1 summation theorem, which is different from the Milne–Gustafson summation theorem [10, 25]. Another class of extension relates to the theory of Macdonald polynomials;

see for example [19,20,26,33] as cited in [34].

In this paper, we prove Theorems 1.2 and 1.5 from the viewpoint of the elliptic Lagrange interpolation functions of typeAn. LetO((C)n) be theC-vector space of holomorphic functions on (C)n. In view of Lemma 1.1, fixing a constant ζ ∈ C we consider the C-linear subspace Hs,n,ζ ⊂ O((C)nconsisting of allSn-invariant holomorphic functionsf(z) such thatTq,zif(z) = f(z)/(−zi)sζ fori= 1, . . . , n, whereTq,zi stands for theq-shift operator in zi:

Hs,n,ζ =

f(z)∈ O((C)n)Sn|Tq,zif(z) =f(z)/(−zi)sζ fori= 1,2, . . . , n . (1.19) The dimension ofHs,n,ζ as aC-vector space will be shown to be n+s−1n

in Section2. Moreover, Theorem 1.6. For generic x∈(C)s there exists a unique C-basis {Eµ(x;z)|µ∈Zs,n} of the space Hs,n,ζ such that

Eµ(x;xν) =δµν for µ, ν ∈Zs,n, (1.20)

where xν ∈(C)n, ν∈Zs,n, are the reference points specified by (1.12) andδλµ is the Kronecker delta.

This theorem will be proved in the end of Section2.2. We callEµ(x;z) the elliptic Lagrange interpolation functions of type Anassociated with the set of reference points xν,ν∈Zs,n. Note that an arbitrary function f(z) ∈ Hs,n,ζ can be written as a linear combination of Eµ(x;z), µ∈Zs,n, with coefficientsf(xµ):

f(z) = X

µ∈Zs,n

f(xµ)Eµ(x;z). (1.21)

(6)

Theorem 1.7. The functionsEλ(x;z) are expressed as

Eλ(x;z) = X

K1t···tKs

={1,2,...,n}

s

Y

i=1

Y

k∈Ki

 θ

zkζ Q

1≤l≤s l6=i

xltλ(k−1)l

θ ζ

s

Q

l=1

xltλ(k−1)l Y

1≤j≤s j6=i

θ zkx−1j t−λ

(k−1)

j

θ xitλ(k−1)i x−1j t−λ

(k−1)

j

, (1.22)

where λ(k)i =|Ki∩ {1,2, . . . , k}|, and the summation is taken over all partitionsK1t · · · tKs= {1,2, . . . , n} such that |Ki|=λi, i= 1,2, . . . , s.

This theorem will be proved in Section2 as Theorem2.4.

Once Theorems 1.6 and 1.7 have been established, the connection formula (1.13) in Theo- rem 1.2 is immediately obtained. In the setting of Theorem 1.2, the regularization hhϕ, zii = hϕ, zi/Θ(z) belongs toHs,n,ζ withζ =qαtn−1b1· · ·bsby Lemma1.1. Hence, by (1.21) we obtain

hhϕ, zii= X

µ∈Zs,n

hhϕ, xµiiEµ(x;z). (1.23)

This means that the coefficients in (1.13) are given bycµ=Eµ(x;z). The explicit formula (1.14) follows from Theorem 1.7.

For the evaluation of the determinant of Theorem1.5, we make use of the following determi- nant formula for the elliptic Lagrange interpolation functions Eµ(x;z) withζ =qαtn−1b1· · ·bs. Theorem 1.8. Suppose that x, y∈(C)s are generic. Then,

det Eµ(x;yν)

µ,ν∈Zs,n =

n

Y

k=1

"

θ qαy1· · ·ysb1· · ·bstn+k−2 θ qαx1· · ·xsb1· · ·bstn+k−2

#(s+k−2k−1 )

×

n

Y

k=1

n−k

Y

r=0

Y

1≤i<j≤s

yitrθ t(n−k)−2ryjyi−1 xitrθ t(n−k)−2rxjx−1i

 (s+k−3k−1 )

,

where yν are specified as in (1.12).

This theorem will be proved as Theorem2.8 in Section2.

Applying the connection formula (1.13) in Theorem 1.2, we see that Theorem 1.5is reduced to the special case where x=a= (a1, . . . , as). Since

hhϕ, zii= X

µ∈Zs,n

hhϕ, aµiiEµ(a;z)

by (1.23), setting ϕ(z) =sλ(z) (λ∈Bs,n) andz=xν (ν ∈Zs,n), we have hhsλ, xνii= X

µ∈Zs,n

hhsλ, aµiiEµ(a;xν) for λ∈Bs,n, ν∈Zs,n, so that

det hhsλ, xνii

λ∈B ν∈Z

= det hhsλ, aµii

λ∈B µ∈Z

det Eµ(a;xν)

µ∈Z ν∈Z

.

Since we already know the explicit form of det Eµ(a;xν)

µ∈Z,ν∈Zby Theorem1.8, the evaluation of det hhsλ, xνii

λ∈B,ν∈Z reduces to that of det hhsλ, aµii

λ∈B,µ∈Z.

(7)

Lemma 1.9. We have

det hhsλ, aµii

λ∈B

µ∈Z =

n

Y

k=1

"

(1−q)s(q)s qt−(n−k+1)s

qt−1s

×

qαtn+k−2

s

Q

i=1

aibi

s

Q

i=1 s

Q

j=1

qa−1i b−1j t−(n−k)

qαtn−k

#(s+k−2k−1 )

×

n

Y

k=1

n−k

Y

r=0

Y

1≤i<j≤s

ajtrθ aia−1j tn−k−2r

 (s+k−3k−1 )

. (1.24)

Lemma1.9will be proved in Section 6.

Remark. Tarasov and Varchenko [32] have already obtained a determinant formula for a mul- tiple contour integrals of hypergeometric type, which is similar to (1.24). See [32, Theorem 5.9]

for the details.

In the succeeding sections, we give proofs for Theorems 1.6, 1.7, 1.8 and Lemmas 1.1, 1.9, which we use for proving our main theorems.

This paper is organized as follows. In the first part of Section 2 we prove Theorems 1.6 and 1.7 on the basis of an explicit construction of the elliptic Lagrange interpolation functions by means of a kernel function as in [22]. In the second part of Section2we investigate the tran- sition coefficients between two sets of elliptic interpolation functions. In particular we provide a proof of Theorem1.8 for the determinant of the transition matrix. A proof of Lemma1.1for the regularized Jackson integrals will be given in Section 3. In Section 4 we introduce certain interpolation polynomials which are a limiting case of the elliptic Lagrange interpolation func- tions. These polynomials are used in Section 5 for the construction of the q-difference system satisfied by the Jackson integrals. In particular we establish two-term difference equations with respect to α → α + 1 for the determinant of the Jackson integrals. Section 6 is devoted to analyzing the boundary condition for difference equations through asymptotic analysis of the Jackson integrals as α → +∞, which completes the proof of Lemma 1.9. In Appendix A, we provide a detailed proof of Lemma 5.4 which is omitted in Section 5. In Appendix B we give proofs for some propositions in Section 4 by using the kernel function in the similar way as in Section 2.

2 The elliptic Lagrange interpolation functions of type A

In this section we give proofs of Theorems1.6,1.7 and 1.8.

LetP+ be the set of partitions of length at mostn specified by P+=

λ= (λ1, λ2, . . . , λn)∈Zn1≥λ2 ≥ · · · ≥λn≥0 .

Forµ= (µ1, . . . , µn)∈Zn, we denote by zµthe monomialzµ11· · ·znµn. For the partitionsλ∈P+

let mλ(z) be the monomial symmetric functions [23] defined by mλ(z) = X

µ∈Snλ

zµ,

where Snλ = {wλ|w ∈ Sn} is the Sn-orbit of λ. For a function f(z) = f(z1, z2, . . . , zn) on (C)n, we define the action of the symmetric group Sn on f(z) by

(σf)(z) =f σ−1(z)

=f zσ(1), zσ(2), . . . , zσ(n)

for σ ∈Sn.

(8)

We say that a function f(z) on (C)n is symmetric or skew-symmetric if σf(z) = f(z) or σf(z) = (sgnσ)f(z) for all σ ∈ Sn, respectively. We denote by Af(z) the alternating sum overSn defined by

(Af)(z) = X

σ∈Sn

(sgnσ)σf(z), (2.1)

which is skew-symmetric.

For an arbitrarily fixed ζ ∈ C, we consider the C-linear subspace Hs,n,ζ ⊂ O((C)n) con- sisting of all Sn-invariant holomorphic functions f(z) such that Tq,zif(z) = f(z)/(−zi)sζ, i= 1, . . . , n, as in (1.19). In this section we use the symbol

e(u, v) =uθ(v/u), u, v∈C.

Since θ(u) =θ(q/u) and θ(qu) =−u−1θ(u), this symbol satisfies e(u, v) =−e(v, u), e(qu, v) = (−v/u)e(u, v).

In particular we have e(u, v)→u−v in the limit q→0.

2.1 Construction of the interpolation functions

This subsection is devoted to providing a proof of Theorem 1.6. In the first half, we define a family of functionsEλ(n)(x;z) recursively with respect to the numbernof variablesz1, . . . , zn, and show that those Eλ(n)(x;z) are expressed explicitly as (1.22) in Theorem 1.7. In the second half, we prove that they are in fact the elliptic interpolation functions in the sense of Theorem1.6;

we show thatEλ(n)(x;z)∈ Hs,n,ζ and Eλ(n)(x;xµ) =δλµ using the dual Cauchy kernel as in [22].

First of all we show the following lemma.

Lemma 2.1. dimCHs,n,ζn+s−1n .

Proof . For an arbitrary f(z) ∈ Hs,n,ζ, since f(z) is a holomorphic on (C)n, f(z) may be expanded as Laurent series as f(z) = P

λ∈Zncλzλ. Since f(z) is symmetric, all coefficients off(z) are determined fromcλ corresponding toλ∈P+. On the other hand, sincef(z) satisfies Tq,zif(z) =f(z)/(−zi)sζ fori= 1, . . . , n, we have

X

λ∈Zn

qλicλzλ= X

λ∈Zn

cλzλ/(−zi)sζ.

Equating coefficients of zλ on both sides, all coefficients of f(z) are determined from cλ corre- sponding to λ satisfying s−1 ≥ λi ≥ 0, i = 1, . . . , n. Therefore, f(z) is determined by the coefficients cλ corresponding to λ∈ Bs,n defined by (1.17). Since |Bs,n|= n+s−1n

, we obtain dimCHs,n,ζn+s−1n

.

Before introducingEλ(n)(x;z) we prove Theorem 1.6forn= 1 by independent means. From the definition (1.11), we have Zs,1 ={1, 2, . . . , s}, where i is specified in Example 1.3, and have xi =xi ∈C.

Lemma 2.2. Forx= (x1, . . . , xs)∈(C)s and z∈C the functions

Ei(x;z) = e

s

Q

k=1

xk, xi

e

xiζ

s

Q

k=1

xk, xi Y

1≤j≤s j6=i

e(z, xj)

e(xi, xj), i= 1, . . . , s, (2.2) satisfy Ei(x;xj) =δij. The set {Ei(x;z)|i= 1, . . . , s} is a basis of the C-linear space Hs,1,ζ. In particular dimCHs,1,ζ =s.

(9)

Proof . It is directly confirmed that Ei(x;xj) =δij from (2.2). It is obvious that Ei(x;z) ∈ Hs,1,ζ and the linearly independence of {Ei(x;z)|i = 1, . . . , s} follows since Ei(x;xj) = δij. Since we have dimCHs,1,ζ ≤ s by Lemma 2.1, we see that {Ei(x;z)|i = 1, . . . , s} is a basis

of Hs,1,ζ, and we obtain dimCHs,1,ζ =s.

Definition 2.3. For x = (x1, . . . , xs) ∈ (C)s and z = (z1, . . . , zn) ∈ (C)n, let Eλ(n)(x;z), λ ∈ Zs,n, be functions defined inductively by E(1)k (x;z1) = Ek(x;z1), k = 1, . . . , s, and for n≥2 by

Eλ(n)(x;z1, . . . , zn) = X

1≤k≤s λk>0

E(n−1)λ−

k (x;z1, . . . , zn−1)E(1)k xtλ−k;zn

, (2.3)

where

xtµ= x1tµ1, x2tµ2, . . . , xstµs

∈(C)s. (2.4)

By the repeated use of (2.3) we have Eλ(n)(x;z) = X

(i1,...,in)

∈{1,...,s}n, i1+···+in

Ei1(x;z1)Ei2 xti1;z2

Ei3 xti1+i2;z3

· · ·Ein xti1+···+in−1;zn

= X

(i1,...,in)∈{1,...,s}n i1+···+in

n

Y

k=1

Eik xtλ(k−1);zk

, (2.5)

where λ(k) = i1 +· · ·+ik. Rewriting this formula as in [14, p. 373, Theorem 3.4] we obtain the following.

Theorem 2.4. The functionsE(n)λ (x;z) defined by (2.3) are expressed explicitly as Eλ(n)(x;z) = X

K1t···tKs

={1,2,...,n}

s

Y

i=1

Y

k∈Ki

Ei xtλ(k−1);zk

,

where λ(k)= λ(k)1 , . . . , λ(k)s

∈Zs,k, λ(k)i =|Ki∩ {1,2, . . . , k}| and the summation is taken over all index sets Ki, i= 1,2, . . . , s satisfying |Ki|=λi and K1t · · · tKs ={1,2, . . . , n}.

We remark that these functions forλ=ni ∈Zs,nhave simple factorized forms; this fact will be used in the next subsection.

Corollary 2.5. For ni ∈Zs,n, i= 1, . . . , s, one has

En(n)i(x;z) =

n

Q

k=1

e

zkζ

s

Q

m=1

xm, xi

e

xiζ

s

Q

m=1

xm, xi

n

Y

1≤j≤s j6=i

n

Q

k=1

e(zk, xj)

e(xi, xj)n , (2.6)

where

e(u, v)r=e(u, v)e(ut, v)· · ·e utr−1, v

for r= 0,1,2, . . . . (2.7)

(10)

Proof . If we put λ=ni in the formula (2.5) then the right-hand side reduces to a single term with (i1, i2, . . . , in) = (i, i, . . . , i). Therefore, using (2.2) we obtain

En(n)i(x;z) =Ei(x;z1)Ei xti;z2

Ei xt2i;z3

· · ·Ei xt(n−1)i;zn ,

which coincides with (2.6).

We simply write Eλ(x;z) = Eλ(n)(x;z) when there is no fear of misunderstanding. In the remaining part of this subsection we show that Eλ(x;z)∈ Hs,n,ζ and Eλ(x;xµ) =δλµ.

Forx= (x1, . . . , xs)∈(C)sandw= (w1, . . . , ws)∈(C)s, letFµ(x;w),µ∈Ns, be functions specified by

Fµ(x;w) =

s

Y

i=1 s

Y

j=1

e(xi, wj)µi, (2.8)

where e(u, v)r is given by (2.7). By definition the functionsFµ(x;w) satisfy

Fµ(x;w)Fν(xtµ;w) =Fµ+ν(x;w), (2.9)

where xtµ is given by (2.4). For z = (z1, . . . , zn) ∈ (C)n and w = (w1, . . . , ws) ∈ (C)s, let Ψ(z;w) be function specified by

Ψ(z;w) =

n

Y

i=1 s

Y

j=1

e(zi, wj),

which we call thedual Cauchy kernel. IfζQs

i=1wi= 1, then by definition Ψ(z;w) as a holomor- phic function of z ∈ (C)n satisfies Ψ(z;w) ∈ Hs,n,ζ. Note that Fµ(x;w) of (2.8) is expressed as

Fµ(x;w) = Ψ(xµ;w).

Lemma 2.6 (duality). Under the conditionζQs

i=1wi= 1, Ψ(z;w) expands as Ψ(z;w) = X

µ∈Zs,n

Eµ(x;z)Fµ(x;w). (2.10)

Proof . We proceed by induction onn, the cardinality of z. We consider the case n= 1 as the base case. Under the condition ζQs

i=1wi = 1, we have Ψ(z1;w) =Qs

j=1e(z1, wj)∈ Hs,1. Then, from Lemma2.2, Ψ(z1;w) is expanded as Ψ(z1;w) =Pn

i=1Ψ(xi;w)Ei(x;z), whose coefficient Ψ(xi;w) is written as Ψ(xi;w) = Ψ(xi;w) =Qs

j=1e(xi, wj) =Fi(x;w). This indicates (2.10) of the case n= 1.

Next we supposen≥2. We assume (2.10) holds for the number of variables forzless thann.

Denoting zbn= (z1, . . . , zn−1)∈(C)n−1 forz= (z1, . . . , zn)∈(C)n, (2.3) is rewritten as Eλ(n)(x;z) =

s

X

i=1

Eλ−(n−1)

i (x;z

bn)E(1)i xtλ−i;zn

, (2.11)

where we regard Eλ−(n−1)

i (x;z

bn) = 0 if λ−i 6∈Zs,n−1. Then, we obtain Ψ(z;w) =

n

Y

i=1 s

Y

j=1

e(zi, wj) =

n−1

Y

i=1 s

Y

j=1

e(zi, wj

s

Y

j=1

e(zn, wj)

(11)

= X

µ∈Zs,n−1

Eµ(n−1)(x;z

nb)Fµ(x;w)

X

ν∈Zs,1

Eν(1)(xtµ;zn)Fν(xtµ;w)

(by the assumption of induction)

= X

µ∈Zs,n−1

ν∈Zs,1

Eµ(n−1)(x;znb)Eν(1)(xtµ;zn)Fµ(x;w)Fν(xtµ;w)

= X

µ∈Zs,n−1

ν∈Zs,1

Eµ(n−1)(x;z

nb)Eν(1)(xtµ;zn)Fµ+ν(x;w) (by (2.9))

= X

λ∈Zs,n

X

µ∈Zs,n−1, ν∈Zs,1

µ+ν=λ

Eµ(n−1)(x;zbn)Eν(1)(xtµ;zn)

Fλ(x;w)

= X

λ∈Zs,n

Eµ(n)(x;z)Fλ(x;w), (by (2.11))

as desired.

When we consider the family of functions Fµ(x;w), µ ∈ Zs,n, on the hypersurface w ∈ (C)s|ζQs

i=1wi = 1 , we regard Fµ(x;w) as a function of s−1 variables (w1, . . . , ws−1) ∈ (C)s−1 with the relation ws = ζQs−1

i=1wi−1

. Here, for x = (x1, . . . , xs) ∈ (C)s and ν = (ν1, . . . , νs)∈Zs,n, we define a special pointην(x) on the hypersurface as

ην(x) =

x1tν1, . . . , xs−1tνs−1, ζ

s−1

Y

i=1

xitνi

!−1

.

Lemma 2.7(triangularity). For eachµ, ν ∈Zs,n,Fµ(x;ην(x)) = 0unlessµi≤νi fori= 1, . . . , s−1. In particular, Fµ(x;ην(x)) = 0 for µν with respect to the lexicographic order of Zs,n. Moreover, if x∈(C)s is generic, then Fµ(x;ηµ(x))6= 0 for all µ∈Zs,n.

This lemma implies that the matrix F = Fµ(x;ην(x))

µ,ν∈Z is upper triangular, and also invertible if x∈Cs is generic.

Proof . If there existsj ∈ {1,2, . . . , s−1}such that νj < µj, thenFµ(x;ην(x)) = 0. In fact, in the expression

Fµ(x;ην(x)) =

s

Y

i=1

e

xi, ζ

s−1

Y

i=1

xitνi

!−1

µi

s−1

Y

j=1

e xi, xjtνj

µi

, the function e(xj, xjtνj)µj has the factorθ(tνj)θ tνj−1

· · ·θ tνj−(µj−1)

= 0 ifνj < µj. Ifν ≺µ, then νi < µi for some i∈ {1,2, . . . , s−1} by definition, and hence we obtain Fµ(x;ην(x)) = 0 ifν ≺µ. Forν =µ,

Fµ(x;ηµ(x)) =

s

Y

i=1

"

xµiit(µi2)

µi−1

Y

j=0

θ

x−1i t−j ζ

s−1

Y

k=1

xktµk

!−1

×

s−1

Y

j=1

θ x−1i xjtµj

θ x−1i xjtµj−1

· · ·θ x−1i xjtµj−µi+1

#

does not vanish if we impose an appropriate genericity condition on x∈(C)s.

Odkazy

Související dokumenty

In Section 3 we introduce in the complex plane C the series and integral representations of the general Wright function denoted by W λ,μ z and of the two related auxiliary functions F

In this section we construct an action of the elliptic dynamical quantum group associated with gl 2 on the extended equivariant elliptic cohomology ˆ E T (X n ) of the union

We have derived in Section I a variational formula for the Szeg5 kernel and ob- tained in Section 2 remarkable identities for the variational expressions which

Finally, in section V we make some applications of the theorems which precede to the theory of functions.. We have other such applications in

As already mentioned, in this section we introduce the arithmetic-geometric-harmonic operator mean which possesses many of the properties of the standard one. In what follows, we

While using the same technique as in [5], but replacing the Cauchy type kernel function by Mimachi’s dual-Cauchy type one (as to the kernel functions, see [9, 15]), we can study

• In Section 3, we introduce the Riemann–Hilbert problem for the orthogonal polynomials and transform this problem into one which can be controlled as n → ∞... • In Section 4,

In the case when the roots of Van Vleck and Stieltjes polynomials are real we can still rely on the result of Stieltjes mentioned above, which make ordering of Stieltjes