A Linear System of Differential Equations Related to Vector-Valued Jack Polynomials on the Torus
For each irreducible module of the symmetric group SN there is a set of parametrized nonsymmetric Jack polynomials in N variables taking values in the module. These polynomials are simultaneous eigenfunctions of a commutative set of operators, self-adjoint with respect to two Hermitian forms, one ca...
Збережено в:
Дата: | 2017 |
---|---|
Автор: | |
Формат: | Стаття |
Мова: | English |
Опубліковано: |
Інститут математики НАН України
2017
|
Назва видання: | Symmetry, Integrability and Geometry: Methods and Applications |
Онлайн доступ: | http://dspace.nbuv.gov.ua/handle/123456789/148638 |
Теги: |
Додати тег
Немає тегів, Будьте першим, хто поставить тег для цього запису!
|
Назва журналу: | Digital Library of Periodicals of National Academy of Sciences of Ukraine |
Цитувати: | A Linear System of Differential Equations Related to Vector-Valued Jack Polynomials on the Torus / C.F. Dunkl // Symmetry, Integrability and Geometry: Methods and Applications. — 2017. — Т. 13. — Бібліогр.: 11 назв. — англ. |
Репозитарії
Digital Library of Periodicals of National Academy of Sciences of Ukraineid |
irk-123456789-148638 |
---|---|
record_format |
dspace |
spelling |
irk-123456789-1486382019-02-19T01:31:47Z A Linear System of Differential Equations Related to Vector-Valued Jack Polynomials on the Torus Dunkl, C.F. For each irreducible module of the symmetric group SN there is a set of parametrized nonsymmetric Jack polynomials in N variables taking values in the module. These polynomials are simultaneous eigenfunctions of a commutative set of operators, self-adjoint with respect to two Hermitian forms, one called the contravariant form and the other is with respect to a matrix-valued measure on the N-torus. The latter is valid for the parameter lying in an interval about zero which depends on the module. The author in a previous paper [SIGMA 12 (2016), 033, 27 pages] proved the existence of the measure and that its absolutely continuous part satisfies a system of linear differential equations. In this paper the system is analyzed in detail. The N-torus is divided into (N−1)! connected components by the hyperplanes xi=xj, i<j, which are the singularities of the system. The main result is that the orthogonality measure has no singular part with respect to Haar measure, and thus is given by a matrix function times Haar measure. This function is analytic on each of the connected components. 2017 Article A Linear System of Differential Equations Related to Vector-Valued Jack Polynomials on the Torus / C.F. Dunkl // Symmetry, Integrability and Geometry: Methods and Applications. — 2017. — Т. 13. — Бібліогр.: 11 назв. — англ. 1815-0659 2010 Mathematics Subject Classification: 33C52; 32W50; 35F35; 20C30; 42B05 DOI:10.3842/SIGMA.2017.040 http://dspace.nbuv.gov.ua/handle/123456789/148638 en Symmetry, Integrability and Geometry: Methods and Applications Інститут математики НАН України |
institution |
Digital Library of Periodicals of National Academy of Sciences of Ukraine |
collection |
DSpace DC |
language |
English |
description |
For each irreducible module of the symmetric group SN there is a set of parametrized nonsymmetric Jack polynomials in N variables taking values in the module. These polynomials are simultaneous eigenfunctions of a commutative set of operators, self-adjoint with respect to two Hermitian forms, one called the contravariant form and the other is with respect to a matrix-valued measure on the N-torus. The latter is valid for the parameter lying in an interval about zero which depends on the module. The author in a previous paper [SIGMA 12 (2016), 033, 27 pages] proved the existence of the measure and that its absolutely continuous part satisfies a system of linear differential equations. In this paper the system is analyzed in detail. The N-torus is divided into (N−1)! connected components by the hyperplanes xi=xj, i<j, which are the singularities of the system. The main result is that the orthogonality measure has no singular part with respect to Haar measure, and thus is given by a matrix function times Haar measure. This function is analytic on each of the connected components. |
format |
Article |
author |
Dunkl, C.F. |
spellingShingle |
Dunkl, C.F. A Linear System of Differential Equations Related to Vector-Valued Jack Polynomials on the Torus Symmetry, Integrability and Geometry: Methods and Applications |
author_facet |
Dunkl, C.F. |
author_sort |
Dunkl, C.F. |
title |
A Linear System of Differential Equations Related to Vector-Valued Jack Polynomials on the Torus |
title_short |
A Linear System of Differential Equations Related to Vector-Valued Jack Polynomials on the Torus |
title_full |
A Linear System of Differential Equations Related to Vector-Valued Jack Polynomials on the Torus |
title_fullStr |
A Linear System of Differential Equations Related to Vector-Valued Jack Polynomials on the Torus |
title_full_unstemmed |
A Linear System of Differential Equations Related to Vector-Valued Jack Polynomials on the Torus |
title_sort |
linear system of differential equations related to vector-valued jack polynomials on the torus |
publisher |
Інститут математики НАН України |
publishDate |
2017 |
url |
http://dspace.nbuv.gov.ua/handle/123456789/148638 |
citation_txt |
A Linear System of Differential Equations Related to Vector-Valued Jack Polynomials on the Torus / C.F. Dunkl // Symmetry, Integrability and Geometry: Methods and Applications. — 2017. — Т. 13. — Бібліогр.: 11 назв. — англ. |
series |
Symmetry, Integrability and Geometry: Methods and Applications |
work_keys_str_mv |
AT dunklcf alinearsystemofdifferentialequationsrelatedtovectorvaluedjackpolynomialsonthetorus AT dunklcf linearsystemofdifferentialequationsrelatedtovectorvaluedjackpolynomialsonthetorus |
first_indexed |
2025-07-12T19:50:57Z |
last_indexed |
2025-07-12T19:50:57Z |
_version_ |
1837472014665777152 |
fulltext |
Symmetry, Integrability and Geometry: Methods and Applications SIGMA 13 (2017), 040, 41 pages
A Linear System of Differential Equations Related
to Vector-Valued Jack Polynomials on the Torus
Charles F. DUNKL
Department of Mathematics, University of Virginia,
PO Box 400137, Charlottesville VA 22904-4137, USA
E-mail: cfd5z@virginia.edu
URL: http://people.virginia.edu/~cfd5z/
Received December 11, 2016, in final form June 02, 2017; Published online June 08, 2017
https://doi.org/10.3842/SIGMA.2017.040
Abstract. For each irreducible module of the symmetric group SN there is a set of para-
metrized nonsymmetric Jack polynomials in N variables taking values in the module. These
polynomials are simultaneous eigenfunctions of a commutative set of operators, self-adjoint
with respect to two Hermitian forms, one called the contravariant form and the other is with
respect to a matrix-valued measure on the N -torus. The latter is valid for the parameter
lying in an interval about zero which depends on the module. The author in a previous
paper [SIGMA 12 (2016), 033, 27 pages] proved the existence of the measure and that its
absolutely continuous part satisfies a system of linear differential equations. In this paper
the system is analyzed in detail. The N -torus is divided into (N−1)! connected components
by the hyperplanes xi = xj , i < j, which are the singularities of the system. The main result
is that the orthogonality measure has no singular part with respect to Haar measure, and
thus is given by a matrix function times Haar measure. This function is analytic on each of
the connected components.
Key words: nonsymmetric Jack polynomials; matrix-valued weight function; symmetric
group modules
2010 Mathematics Subject Classification: 33C52; 32W50; 35F35; 20C30; 42B05
Contents
1 Introduction 2
2 Modules of the symmetric group 4
2.1 Jack polynomials . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5
3 The differential system 7
3.1 The adjoint operation on Laurent polynomials and L(x) . . . . . . . . . . . . . . . . . . . 10
4 Integration by parts 11
5 Local power series near the singular set 12
5.1 Behavior on boundary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 16
6 Bounds 18
7 Sufficient condition for the inner product property 22
8 The orthogonality measure on the torus 27
8.1 Maximal singular support . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 32
8.2 Boundary values for the measure . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 35
9 Analytic matrix arguments 38
References 40
mailto:cfd5z@virginia.edu
http://people.virginia.edu/~cfd5z/
https://doi.org/10.3842/SIGMA.2017.040
2 C.F. Dunkl
1 Introduction
The Jack polynomials form a parametrized basis of symmetric polynomials. A special case of
these consists of the Schur polynomials, important in the character theory of the symmetric
groups. By means of a commutative algebra of differential-difference operators the theory was
extended to nonsymmetric Jack polynomials, again a parametrized basis but now for all poly-
nomials in N variables. These polynomials are orthogonal for several different inner products,
and in each case they are simultaneous eigenfunctions of a commutative set of self-adjoint op-
erators. These inner products are invariant under permutations of the coordinates, that is, the
symmetric group. One of these inner products is that of L2
(
TN ,Kκ(x)dm(x)
)
, where
TN :=
{
x ∈ CN : |xj | = 1, 1 ≤ j ≤ N
}
,
dm(x) = (2π)−Ndθ1 · · · dθN , xj = exp(iθj), −π < θj ≤ π, 1 ≤ j ≤ N,
Kκ(x) =
∏
1≤i<j≤N
|xi − xj |2κ, κ > − 1
N
;
defining the N -torus, the Haar measure on the torus, and the weight function respectively.
Beerends and Opdam [1] discovered this orthogonality property of symmetric Jack polyno-
mials. Opdam [9] established orthogonality structures on the torus for trigonometric polynomials
associated with Weyl groups; the nonsymmetric Jack polynomials form a special case.
Griffeth [7] constructed vector-valued Jack polynomials for the family G(n, p,N) of complex
reflection groups. These are the groups of permutation matrices (exactly one nonzero entry
in each row and each column) whose nonzero entries are nth roots of unity and the product
of these entries is a (n/p)th root of unity. The symmetric groups and the hyperoctahedral
groups are the special cases G(1, 1, N) and G(2, 1, N) respectively. The term “vector-valued”
means that the polynomials take values in irreducible modules of the underlying group, and the
action of the group is on the range as well as the domain of the polynomials. The author [2]
together with Luque [5] investigated the symmetric group case more intensively. The basic setup
is an irreducible representation of the symmetric group, specified by a partition τ of N , and
a parameter κ restricted to an interval determined by the partition, namely −1/hτ < κ < 1/hτ
where hτ is the maximum hook-length of the partition τ . More recently [3] we showed that there
does exist a positive matrix measure on the torus for which the nonsymmetric vector-valued Jack
polynomials (henceforth NSJP’s) form an orthogonal set. The proof depends on a matrix-version
of Bochner’s theorem about the relation between positive measures on a compact abelian group
and positive-definite functions on the dual group, which is a discrete abelian group. In the
present situation the torus is the compact (multiplicative) group and the dual is ZN . By using
known properties of the NSJP’s we produced a positive-definite matrix function on ZN and this
implied the existence of the desired orthogonality measure. Additionally we showed that the
part of the measure supported by TNreg := TN\
⋃
i<j{x : xi = xj} is absolutely continuous with
respect to the Haar measure dm and satisfies a first-order differential system. In this paper we
complete the description of the measure by proving there is no singular part. The idea is to
use the functional equations satisfied by the inner product to establish a correspondence to the
differential system. The main reason for the argument being so complicated is that the “obvious”
integration-by-parts argument which works smoothly for the scalar case with κ > 1 has great
difficulty with the singularities of the measure of the form |xi − xj |−2|κ|. We use a Cauchy
principal-value argument based on a weak continuity condition across the faces {x : xi = xj}
(as an over-simplified one-dimensional example consider the integral
∫ 1
−1
d
dxf(x)dx with f(x) =
|2x + x2|−1/4: the integral is divergent but the principal value lim
ε→0+
{ ∫ −ε
−1 +
∫ 1
ε
}
f ′(x)dx =
f(1)− f(−1) + lim
ε→0+
{f(−ε)− f((ε))} and f(−ε)− f(ε) = O
(
ε3/4
)
hence the limit exists).
A System of Differential Equations on the Torus 3
The differential system is a two-sided version of a Knizhnik–Zamolodchikov equation (see [6])
modified to have solutions homogeneous of degree zero, that is, constant on circles {(ux1, . . .,
uxN ) : |u| = 1}. The purpose of the latter condition is to allow solutions analytic on connected
components of TNreg. Denote the degree of τ by nτ . The solutions of the differential system are
locally analytic nτ ×nτ matrix functions with initial condition given by a constant matrix. That
is, the solution space is of dimension n2
τ but only one solution can provide the desired weight
function. Part of the analysis deals with conditions specifying this solution – they turn out to be
commutation relations involving certain group elements. In the subsequent discussion it is shown
that the weight function property holds for a very small interval of κ values if these relations are
satisfied. This is combined with the existence theorem of the positive-definite matrix measure
to finally demonstrate that the measure has no singular part for any κ in −1/hτ < κ < 1/hτ .
In a subsequent development [4] it is shown that the square root of the matrix weight function
multiplied by vector-valued symmetric Jack polynomials provides novel wavefunctions of the
Calogero–Sutherland quantum mechanical model of identical particles on a circle with 1/r2
interactions.
Here is an outline of the contents of the individual sections:
• Section 2: a short description of the representation of the symmetric group associated
to a partition; the definition of Dunkl operators for vector-valued polynomials and the
definition of nonsymmetric Jack polynomials (NSJP’s) as simultaneous eigenvectors of
a commutative set of operators; and the Hermitian form given by an integral over the
torus, for which the NSJP’s form an orthogonal basis.
• Section 3: the definition of the linear system of differential equations which will be demon-
strated to have a unique matrix solution L(x) such that L(x)∗L(x)dm(x) is the weight
function for the Hermitian form; the proof that the system is Frobenius integrable and the
analyticity and monodromy properties of the solutions on the torus.
• Section 4: the use of the differential equation to relate the Hermitian form to L(x)∗L(x)
by means of integration by parts; the result of this is to isolate the role of the singularities
in the process of proving the orthogonality of the NSJP’s with respect to L∗Ldm.
• Section 5: deriving power series expansions of L(x) near the singular set
⋃
i<j
{
x ∈ TN :
xi = xj
}
, in particular near the set {x : xN−1 = xN}; description of commutation proper-
ties of the coefficients with respect to the reflection (N − 1, N); the behavior of L across
the mirror {x : xN−1 = xN}.
• Section 6: the derivation of global bounds on L(x) and local bounds on the coefficients of
the power series, needed to analyze convergence properties of the integration by parts.
• Section 7: the proof of a sufficient condition for the validity of the Hermitian form; the
condition is partly that κ lies in a small interval around 0 and that the boundary value
of L(x) satisfies a commutativity condition; the proof involves very detailed analysis of
bounds on L, since the local bounds have to be integrated over the entire torus.
• Section 8: further analysis of the orthogonality measure constructed in [3], in particular
the proof of the formal differential system satisfied by the Fourier–Stieltjes (Laurent) series
of the measure; this is used to show that the measure has no singular part on the open
faces, such as{(
eiθ1 , eiθ2 , . . . , eiθN−1 , eiθN−1
)
: θ1 < θ2 < · · · < θN−2 < θN−1 < θ1 + 2π
}
;
in turn this property is shown to imply the validity of the sufficient condition set up in
Section 7.
4 C.F. Dunkl
• Section 9: analyticity properties of the solutions of matrix equations with analytic coeffi-
cients; the results are used to extend the validity of the Hermitian form to the desired
interval −1/hτ < κ < 1/hτ from the smaller interval found in Section 7.
2 Modules of the symmetric group
The symmetric group SN , the set of permutations of {1, 2, . . . , N}, acts on CN by permutation
of coordinates. For α ∈ ZN the norm is |α| :=
N∑
i=1
|αi| and the monomial is xα :=
N∏
i=1
xαii .
Denote N0 := {0, 1, 2, . . .}. The space of polynomials P := spanC
{
xα : α ∈ NN0
}
. Elements
of spanC
{
xα : α ∈ ZN
}
are called Laurent polynomials. The action of SN is extended to
polynomials by wp(x) = p(xw) where (xw)i = xw(i) (consider x as a row vector and w as
a permutation matrix, [w]ij = δi,w(j), then xw = x[w]). This is a representation of SN , that is,
w1(w2p)(x) = (w2p)(xw1) = p(xw1w2) = (w1w2)p(x) for all w1, w2 ∈ SN .
Furthermore SN is generated by reflections in the mirrors {x : xi = xj} for 1 ≤ i < j ≤ N .
These are transpositions, denoted by (i, j), so that x(i, j) denotes the result of interchanging xi
and xj . Define the SN -action on α ∈ ZN so that (xw)α = xwα
(xw)α =
N∏
i=1
xαiw(i) =
N∏
j=1
x
αw−1(j)
j ,
that is (wα)i = αw−1(i) (take α as a column vector, then wα = [w]α).
The simple reflections si := (i, i+ 1), 1 ≤ i ≤ N − 1, generate SN . They are the key devices
for applying inductive methods, and satisfy the braid relations:
sisj = sjsi, |i− j| ≥ 2;
sisi+1si = si+1sisi+1.
We consider the situation where the group SN acts on the range as well as on the domain
of the polynomials. We use vector spaces, called SN -modules, on which SN has an irreducible
unitary (orthogonal) representation: τ : SN → Om(R)
(
τ(w)−1 = τ
(
w−1
)
= τ(w)T
)
. See James
and Kerber [8] for representation theory, including a modern discussion of Young’s methods.
Denote the set of partitions
NN,+0 :=
{
λ ∈ NN0 : λ1 ≥ λ2 ≥ · · · ≥ λN
}
.
We identify τ with a partition of N given the same label, that is τ ∈ NN,+0 and |τ | = N .
The length of τ is `(τ) := max{i : τi > 0}. There is a Ferrers diagram of shape τ (also given
the same label), with boxes at points (i, j) with 1 ≤ i ≤ `(τ) and 1 ≤ j ≤ τi. A tableau of
shape τ is a filling of the boxes with numbers, and a reverse standard Young tableau (RSYT)
is a filling with the numbers {1, 2, . . . , N} so that the entries decrease in each row and each
column. We exclude the one-dimensional representations corresponding to one-row (N) or one-
column (1, 1, . . . , 1) partitions (the trivial and determinant representations, respectively). We
need the important quantity hτ := τ1 + `(τ)− 1, the maximum hook-length of the diagram (the
hook-length of the node (i, j) ∈ τ is defined to be τi − j + #{k : i < k ≤ `(τ)&j ≤ τk} + 1).
Denote the set of RSYT’s of shape τ by Y(τ) and let
Vτ = span{T : T ∈ Y(τ)}
(the field is C(κ)) with orthogonal basis Y(τ). For 1 ≤ i ≤ N and T ∈ Y(τ) the entry i
is at coordinates (rw(i, T ), cm(i, T )) and the content is c(i, T ) := cm(i, T ) − rw(i, T ). Each
A System of Differential Equations on the Torus 5
T ∈ Y(τ) is uniquely determined by its content vector [c(i, T )]Ni=1. Let S1(τ) =
N∑
i=1
c(i, T ) (this
sum depends only on τ) and γ := S1(τ)/N . The SN -invariant inner product on Vτ is defined by
〈T, T ′〉0 := δT,T ′ ×
∏
1≤i<j≤N,
c(i,T )≤c(j,T )−2
(
1− 1
(c(i, T )− c(j, T ))2
)
, T, T ′ ∈ Y(τ).
It is unique up to multiplication by a constant.
The Jucys–Murphy elements
N∑
j=i+1
(i, j) satisfy
N∑
j=i+1
τ((i, j))T = c(i, T )T and thus the central
element
∑
1≤i<j≤N
(i, j) satisfies
∑
1≤i<j≤N
τ((i, j))T = S1(τ)T for each T ∈ Y(τ). The basis is
ordered such that the vectors T with c(N − 1, T ) = −1 appear first (that is, cm(N − 1, T ) = 1,
rw(N − 1, T ) = 2). This results in the matrix representation of τ((N − 1, N)) being[
−Imτ O
O Inτ−mτ
]
,
where nτ := dimVτ = #Y(τ) and mτ is given by tr(τ((N − 1, N))) = nτ − 2mτ . From the sum∑
i<j
τ((i, j)) = S1(τ)I it follows that
(
N
2
)
tr(τ((N −1, N))) = S1(τ)nτ and mτ = nτ
(
1
2 −
S1(τ)
N(N−1)
)
.
(The transpositions are conjugate to each other implying the traces are equal.)
2.1 Jack polynomials
The main concerns of this paper are measures and matrix functions on the torus associated to
Pτ := P ⊗ Vτ , the space of Vτ -valued polynomials, which is equipped with the SN action:
w (xα ⊗ T ) = (xw)α ⊗ τ(w)T, α ∈ NN0 , T ∈ Y(τ),
wp(x) = τ(w)p(xw), p ∈ Pτ ,
extended by linearity. There is a parameter κ which may be generic/transcendental or complex.
Definition 2.1. The Dunkl and Cherednik–Dunkl operators are (1 ≤ i ≤ N , p ∈ Pτ )
Dip(x) :=
∂
∂xi
p(x) + κ
∑
j 6=i
τ((i, j))
p(x)− p(x(i, j))
xi − xj
,
Uip(x) := Di(xip(x))− κ
i−1∑
j=1
τ((i, j))p(x(i, j)).
The commutation relations analogous to the scalar case hold:
DiDj = DjDi, UiUj = UjUi, 1 ≤ i, j ≤ N ;
wDi = Dw(i)w, ∀w ∈ SN ; sjUi = Uisj , j 6= i− 1, i;
siUisi = Ui+1 + κsi, Uisi = siUi+1 + κ, Ui+1si = siUi − κ.
The simultaneous eigenfunctions of {Ui} are called (vector-valued) nonsymmetric Jack poly-
nomials (NSJP). For generic κ these eigenfunctions form a basis of Pτ (this property fails for
certain rational numbers outside the interval −1/hτ < κ < 1/hτ ). There is a partial order on
NN0 × Y(τ) for which the NSJP’s have a triangular expression with leading term indexed by
6 C.F. Dunkl
(α, T ) ∈ NN0 ×Y(τ). The polynomial with this label is denoted by ζα,T , homogeneous of degree
N∑
i=1
αi and satisfies
Uiζα,T = (αi + 1 + κc (rα(i), T )) ζα,T , 1 ≤ i ≤ N,
rα(i) := #{j : αj > αi}+ #{j : 1 ≤ j ≤ i, αj = αi};
the rank function rα ∈ SN and rα = I if and only if α is a partition. The vector
[αi + 1 + κc(rα(i), T )]Ni=1
is called the spectral vector for (α, T ). The NSJP structure can be extended to Laurent polyno-
mials. Let eN :=
N∏
i=1
xi and 1 := (1, 1, . . . , 1) ∈ NN0 , then rα+m1 = rα for any α ∈ NN0 and m ∈ Z.
The commutation Ui(emNp) = emN (m + Ui)p for 1 ≤ i ≤ N and p ∈ Pτ imply that emNζα,T and
ζα+m1,T have the same spectral vector for any m ∈ N0. They also have the same leading term
(see [3, Section 2.2]) and hence emNζα,T = ζα+m1,T for α ∈ NN0 . This fact allows the definition
of ζα,T for any α ∈ ZN : let m = −mini αi then α+m1 ∈ NN0 and set ζα,T := e−mN ζα+m1,T .
For a complex vector space V a Hermitian form is a mapping 〈·, ·〉 : V ⊗ V → C such that
〈u, cv〉 = c〈u, v〉, 〈u, v1 + v2〉 = 〈u, v1〉 + 〈u, v2〉 and 〈u, v〉 = 〈v, u〉 for u, v1, v2 ∈ V , c ∈ C.
The form is positive semidefinite if 〈u, u〉 ≥ 0 for all u ∈ V . The concern of this paper is with
a particular Hermitian form on Pτ which has the properties (for all f, g ∈ Pτ , T, T ′ ∈ Y(τ),
w ∈ SN , 1 ≤ i ≤ N):
〈1⊗ T, 1⊗ T ′〉 = 〈T, T ′〉0, (2.1)
〈wf,wg〉 = 〈f, g〉,
〈xiDif, g〉 = 〈f, xiDig〉,
〈xif, xig〉 = 〈f, g〉.
The commutation Ui = Dixi − κ
∑
j<i
(i, j) = xiDi + 1 + κ
∑
j>i
(i, j) together with 〈(i, j)f, g〉 =
〈f, (i, j)g〉 show that 〈Uif, g〉 = 〈f,Uig〉 for all i. Thus uniqueness of the spectral vectors (for all
but a certain set of rational κ values) implies that 〈ζα,T , ζβ,T ′〉 = 0 whenever (α, T ) 6= (β, T ′). In
particular polynomials homogeneous of different degrees are mutually orthogonal, by the basis
property of {ζα,T }. For this particular Hermitian form, multiplication by any xi is an isometry
for all 1 ≤ i ≤ N . This involves an integral over the torus. The equations (2.1) determine the
form uniquely (up to a multiplicative constant if the first condition is removed).
Denote C× := C\{0} and CNreg := CN×\
⋃
i<j
{x : xi = xj}. The torus is a compact multiplicative
abelian group. The notations for the torus and its Haar measure in terms of polar coordinates
are
TN :=
{
x ∈ CN : |xj | = 1, 1 ≤ j ≤ N
}
,
dm(x) = (2π)−Ndθ1 · · · dθN , xj = exp(iθj), −π < θj ≤ π, 1 ≤ j ≤ N.
Let TNreg := TN ∩CNreg, then TNreg has (N − 1)! connected components and each component is
homotopic to a circle (if x is in some component then so is ux = (ux1, . . . , uxN ) for each u ∈ T).
Definition 2.2. Let ω := exp 2πi
N and x0 :=
(
1, ω, . . . , ωN−1
)
. Denote the connected component
of TNreg containing x0 by C0.
A System of Differential Equations on the Torus 7
Thus C0 is the set consisting of
(
eiθ1 , . . . , eiθN
)
with θ1 < θ2 < · · · < θN < θ1 + 2π.
In [3] we showed that if −1/hτ < κ < 1/hτ then there exists a positive matrix-valued
measure dµ on TN such that for f, g ∈ C(1)
(
TN ;Vτ
)
, w ∈ SN , 1 ≤ i ≤ N ,∫
TN
f(x)∗dµ(x)g(x) =
∫
TN
f(xw)∗τ(w)−1dµ(x)τ(w)g(xw),∫
TN
(xiDif(x))∗dµ(x)g(x) =
∫
TN
f(x)∗dµ(x)xiDig(x),∫
TN
(1⊗ T )∗dµ(x)(1⊗ T ) = 〈T, T 〉0, T ∈ Y(τ).
We introduced the notation
f(x)∗dµ(x)g(x) :=
∑
T,T ′∈Y(τ)
f(x)T g(x)T ′dµT,T ′(x),
where f, g ∈ Pτ have the components (fT ), (gT ) with respect to the orthonormal basis{
〈T, T 〉−1/2
0 T : T ∈ Y(τ)
}
.
Thus the Hermitian form 〈f, g〉 =
∫
TN f(x)∗dµ(x)g(x) satisfies (2.1). Furthermore we showed
that
dµ(x) = dµs(x) + L(x)∗H(C)L(x)dm(x),
where the singular part µs is the restriction of µ to TN\TNreg, H(C) is constant and positive-
definite on each connected component C of TNreg and L(x) is a matrix function solving a system
of differential equations. That system is the subject of this paper. In a way the main problem
is to show that µ has no singular part.
3 The differential system
Consider the system (with ∂i := ∂
∂xi
, 1 ≤ i ≤ N) for nτ × nτ matrix functions L(x)
∂iL(x) = κL(x)
∑
j 6=i
1
xi − xj
τ((i, j))− γ
xi
I
, 1 ≤ i ≤ N, (3.1)
γ :=
S1(τ)
N
=
1
2N
`(τ)∑
i=1
τi(τi − 2i+ 1).
The effect of the term γ
xi
I is to make L(x) homogeneous of degree zero, that is,
N∑
i=1
xi∂iL(x) = 0.
The differential system is defined on CNreg, Frobenius integrable and analytic, thus any local
solution can be continued analytically to any point in CNreg. Different paths may produce different
values; if the analytic continuation is done along a closed path then the resultant solution is a
constant matrix multiple of the original solution, called the monodromy matrix, however if the
closed path is contained in a simply connected subset of CNreg then there is no change.
Integrability means that ∂i(κL(x)Aj(x)) = ∂j(κL(x)Ai(x)) for i 6= j, writing the system as
∂iL(x) = κL(x)Ai(x), 1 ≤ i ≤ N (where Ai(x) is defined by equation (3.1)). The condition
becomes
κ2L(x)Ai(x)Aj(x) + κL(x)∂iAj(x) = κ2L(x)Aj(x)Ai(x) + κL(x)∂jAi(x).
8 C.F. Dunkl
Since ∂iAj(x) = τ((i,j))
(xi−xj)2 = ∂jAi(x) it suffices to show that Ai(x)′ :=
∑
k 6=i
τ((i,k))
xi−xk and Aj(x)′ :=∑̀
6=j
τ((j,`))
xj−x` commute with each other. The product Ai(x)′Aj(x)′ is a sum of − 1
(xi−xj)2 I, terms of
the form τ((i,k)(j,`))
(xi−xk)(xj−x`) + τ((i,`)(j,k))
(xi−x`)(xj−xk) for {i, k} ∩ {j, `} = ∅, and terms involving the 3-cycles
(i, j, k) and (j, i, k) occurring as
τ((i, j)(j, k))
(xi − xj)(xj − xk)
+
τ((i, k)(j, i))
(xi − xk)(xj − xi)
+
τ((i, k)(j, k))
(xi − xk)(xj − xk)
=
τ((i, j, k))
(xi − xk)(xj − xk)
+
τ((j, i, k))
(xi − xk)(xj − xk)
,
(because (i, j)(j, k) = (i, k)(j, i) = (i, j, k) and (i, k)(j, k) = (j, i, k)) and the latter two terms
are symmetric in i, j.
We consider only fundamental solutions, that is, detL(x) 6= 0. Recall Jacobi’s identity:
∂
∂t
detF (t) = tr
(
adj(F (t))
∂
∂t
F (t)
)
,
where F (t) is a differentiable matrix function and adj(F (t))F (t) = detF (t)I, that is, adj(F (t)) =
{detF (t)}F (t)−1 when F (t) is invertible; thus
∂i detL(x) = tr
(
{detL(x)}L(x)−1∂iL(x)
)
= κdetL(x) trAi(x).
This can be solved: from
∑
i<j
τ((i, j)) = S1(τ)I it follows that tr(τ((i, j))) =
(
N
2
)−1
S1(τ)nτ =
2
N−1γnτ (and nτ = #Y(τ)). We obtain the system
∂i detL(x) = κγnτ
2
N − 1
∑
j 6=i
1
xi − xj
− 1
xi
detL(x), 1 ≤ i ≤ N.
By direct verification
detL(x) = c
∏
1≤i<j≤N
(
−(xi − xj)2
xixj
)κλ/2
, λ :=
γnτ
2(N − 1)
= tr(τ((1, 2))),
is a local solution for any c ∈ C×, if xk = eiθk , 1 ≤ k ≤ N then − (xi−xj)2
xixj
= 4 sin2 θi−θj
2 (with
the principal branch of the power function, positive on positive reals). This implies detL(x) 6= 0
for x ∈ CNreg (and of course detL(x) is homogeneous of degree zero).
Proposition 3.1. If L(x) is a solution of (3.1) in some connected open subset U of CNreg then
L(xw)τ(w)−1 is a solution in Uw−1.
Proof. First let w = (j, k) for some fixed j, k. If i 6= j, k then replace x by x(j, k) in ∂iL to
obtain
∂iL(x(j, k))τ((j, k)) = κL(x(j, k))
∑
`6=i,j,k
τ((i, `))
xi − x`
+
τ((i, j))
xi − xk
+
τ((i, k))
xi − xj
− γ
xi
I
τ(j, k)
= κL(x(j, k))τ(j, k)
∑
`6=i,j,k
τ((i, `))
xi − x`
+
τ((i, k))
xi − xk
+
τ((i, j))
xi − xj
− γ
xi
I
,
A System of Differential Equations on the Torus 9
because (i, j)(j, k) = (j, k)(i, k). Next let w = (i, j), then ∂i[L(x(i, j))] = (∂jL)(x(i, j)) and
∂i[L(x(i, j))τ((i, j))] = κL(x(i, j))
∑
`6=i,j
τ((j, `))
xi − x`
+
τ((i, j))
xi − xj
− γ
xi
I
τ((i, j))
= κL(x(i, j))τ((i, j))
∑
`6=i,j
τ((i, `))
xi − x`
+
τ((i, j))
xi − xj
− γ
xi
I
,
by use of (j, `)(i, j) = (i, j)(i, `). Arguing by induction suppose L(xw)τ(w)−1 is a solution then
so is L(x(j, k)w)τ(w)−1τ((j, k)) = L(x(j, k)w)τ((j, k)w)−1, for any(j, k), that is, the statement
holds for w′ = (j, k)w. �
Let w0 := (1, 2, 3, . . . , N) = (12)(23) · · · (N − 1, N) denote the N -cycle and let 〈w0〉 denote
the cyclic group generated by w0. There are two components of TNreg which are set-wise invariant
under 〈w0〉 namely C0 and the reverse {θN < θN−1 < · · · < θ1 < θN + 2π}. Indeed 〈w0〉 is the
stabilizer of C0 as a subgroup of SN .
Henceforth we use L(x) to denote the solution of (3.1) in C0 which satisfies L(x0) = I.
Proposition 3.2. Suppose x ∈ C0 and m ∈ Z then L(xwm0 ) = τ(w0)−mL(x)τ(w0)m.
Proof. Consider the solution L(xw0)τ(w0)−1 which agrees with ΞL(x) for all x ∈ C0 for some
fixed matrix Ξ. In particular for x = x0 where x0w0 =
(
ω, . . . , ωN−1, 1
)
= ωx0 (recall (xw)i =
xw(i)) we obtain ΞL(x0) = L(x0w0)τ(w0)−1 = L(ωx0)τ(w0)−1 = L(x0)τ(w0)−1; because L(x) is
homogeneous of degree zero. Thus Ξ = τ(w0)−1. Repeated use of the relation shows L(xwm0 ) =
τ(w0)−mL(x)τ(w0)m. �
Because of its frequent use denote υ := τ(w0) (the letter υ occurs in the Greek word cycle).
Definition 3.3. For w ∈ SN set ν(w) := υ1−w(1). For any x ∈ TNreg there is a unique wx such
that wx(1) = 1 and xw−1
x ∈ C0. Set M(w, x) := ν(wxw).
As a consequence ν(wm0 w) = υ−mν(w) for any w ∈ SN and m ∈ Z; since wm0 w(1) − 1 =
(w(1) + m − 1) modN . Also M(I, x) = I. There is a 1-1 correspondence w 7→ C0w between
{w ∈ SN : w(1) = 1} and the connected components of TNreg.
Proposition 3.4. For any w1, w2 ∈ SN and x ∈ TNreg
M(w1w2, x) = M(w2, xw1)M(w1, x).
Proof. By definition M(w1w2, x) = ν(wxw1w2) and M(w1, x) = ν(wxw1) = υ−m where
wxw1(1) = m+1. Let w3 = wxw1 , that is, w3(1) = 1 and xw1w
−1
3 ∈ C0. From
(
xw−1
x
)(
wxw1w
−1
3
)
∈ C0 it follows that wxw1w
−1
3 ∈ 〈w0〉, in particular wxw1w
−1
3 = wm0 because wxw1w
−1
3 (1) =
wxw1(1) = m + 1 = wm0 (1). Thus M(w2, xw1) = ν(w3w2) = ν
(
w−m0 wxw1w2
)
= υmν(wxw1w2),
and υm = M(w1, x)−1. This completes the proof. �
Corollary 3.5. Suppose w ∈ SN and x ∈ TNreg then M
(
w−1, xw
)
= M(w, x)−1.
Proof. Indeed M
(
w−1, xw
)
M(w, x) = M
(
ww−1, x
)
= I. �
We can now extend L(x) to all of TNreg from its values on C0.
Definition 3.6. For x ∈ TNreg let
L(x) := L
(
xw−1
x
)
τ(wx).
10 C.F. Dunkl
Proposition 3.7. For any x ∈ TNreg and w ∈ SN
L(xw) = M(w, x)L(x)τ(w).
Proof. Let w1 = wxw, that is, w1(1) = 1 and xww−1
1 ∈ C0, then by definition L(xw) =
L
(
xww−1
1
)
τ(w1) and L
(
xw−1
x
)
= L(x)τ(wx)−1. Let m = wxw(1) − 1. Since wxww
−1
1 fixes C0
and wxww
−1
1 (1) = wxw(1) = m+ 1 it follows that wxww
−1
1 = wm0 . Thus w1 = w−m0 wxw,
L
(
xww−1
1
)
τ(w1) = L
(
xw−1
x wm0
)
τ
(
w−m0 wxw
)
= υ−mL
(
xw−1
x
)
υmτ
(
w−m0 wxw
)
= υ−mL
(
xw−1
x
)
τ(wxw) = υ−mL(x)τ(w)
and M(w, x) = ν(wxw) = υ−m. �
3.1 The adjoint operation on Laurent polynomials and L(x)
The purpose is to define an operation which agrees with taking complex conjugates of functions
and Hermitian adjoints of matrix functions when restricted to TN , and which preserves ana-
lyticity. The parameter κ is treated as real in this context even where it may be complex (to
preserve analyticity in κ). For x ∈ CN× define φx :=
(
x−1
1 , x−1
2 , . . . , x−1
N
)
, then φ(xw) = (φx)w
for all w ∈ SN .
Definition 3.8.
(1) If f(x) =
∑
α∈ZN
cαx
α is a Laurent polynomial then f∗(x) :=
∑
α∈ZN
cαx
−α.
(2) If f(x) =
∑
α∈ZN
Aαx
α is a Laurent polynomial with matrix coefficients then f∗(x) :=∑
α∈ZN
A∗αx
−α.
(3) if F (x) is a matrix-valued function analytic in an open subset U of CN× then F ∗(x) :=
(F (φx))
T
and F ∗ is analytic on φU , that is, if F (x)=[Fij(x)]Ni,j=1 then F ∗(x)=[Fji(φx)]Ni,j=1
(for example if F12(x) = c1κx1x
−1
3 + c2x
2
2x
−1
3 x−1
4 then F ∗21(x) = c1κx
−1
1 x3 + c2x
−2
2 x3x4).
Loosely speaking F ∗(x) is obtained by replacing x by φx, conjugating the complex constants
and transposing. The fundamental chamber C0 is mapped by φ onto
{(
eiθj
)N
j=1
: θ1 > θ2 > · · ·
> θN > θ1− 2π
}
, again set-wise invariant under w0. Using d
dt
{
f
(
1
t
)}
= − 1
t2
(
d
dtf
)(
1
t
)
we obtain
the system
∂iL(φx) = κL(φx)
∑
j 6=i
xj
xi
τ((i, j))
xi − xj
+
γ
xi
, 1 ≤ i ≤ N.
Transposing this system leads to (note τ(w)T = τ(w)∗ = τ
(
w−1
)
)
∂iL(φx)T = κ
∑
j 6=i
xj
xi
τ((i, j))
xi − xj
+
γ
xi
L(φx)T , 1 ≤ i ≤ N.
Now use part (3) of Definition 3.8 and set up the system whose solution of
∂iL
∗(x) = κ
∑
j 6=i
xj
xi
τ((i, j))
xi − xj
+
γ
xi
L∗(x), 1 ≤ i ≤ N. (3.2)
A System of Differential Equations on the Torus 11
satisfying L∗(x0) = I is denoted by L∗(x). The constants in the system are all real so replacing
complex constants by their complex conjugates preserves solutions of the system. The effect
is that L(x)∗ agrees with the Hermitian adjoint of L(x) for x ∈ C0 (for real κ). The goal here
is to establish conditions on a constant Hermitian matrix H so that K(x) := L∗(x)HL(x) has
desirable properties, such as K(xw) = τ(w)−1K(x)τ(w) and K(x) ≥ 0 (i.e., positive definite).
Similarly to the above τ((i, j))L∗(x(i, j)) is also a solution of (3.2), implying that τ(w)L∗(xw)
is a solution for any w ∈ SN , the inductive step is
τ((i, j))τ(w)L(x(i, j)w)∗ = τ((i, j)w)L(x(i, j)w)∗.
Also L∗(x0w0) = L∗
(
ω−1x0
)
= L∗(x0) = I (thus there is a matrix Ξ̃ such that τ(w0)L∗(xw0) =
L∗(φx)Ξ̃ for all x ∈ C0, and Ξ̃ = τ(w0) = υ. In analogy to L for x ∈ TNreg and the same
wx as above let L(φx0)T = I, L(φx)T := τ(wx)−1L
(
φxw−1
x
)T
(since φxw−1
x ∈ φC0). Then
L(φxw)T = τ(w)−1L(φx)TM(w, x)−1.
For any nonsingular constant matrix C the function CL(x) also satisfies (3.1) and the function
K(x) := L∗(x)C∗CL(x) satisfies the system
xi∂iK(x) = κ
∑
j 6=i
{
xj
xi − xj
τ((i, j))K(x) +K(x)τ((i, j))
xi
xi − xj
}
, 1 ≤ i ≤ N. (3.3)
This formulation can be slightly generalized by replacing C∗C by a Hermitian matrix H (not
necessarily positive-definite) without changing the equation.
For the purpose of realizing the form (2.1) we want K to satisfy K(xw) = τ(w)−1K(x)τ(w),
that is,
K(xw) = τ(w)−1L∗(x)M(w, x)−1HM(w, x)L(x)τ(w)
= τ(w)−1L∗(x)υmHυ−mL(x)τ(w)
(from Proposition 3.7), where m = wxw(1)− 1. The condition is equivalent to
υH = Hυ,
which is now added to the hypotheses, summarized here:
Condition 3.9. L(x) is the solution of (3.1) such that L(x0) = I and L(x) = L
(
xw−1
x
)
τ(wx) for
x ∈ TNreg where wx(1) = 1 and xw−1
x ∈ C0; L∗(x) is the solution of (3.2) satisfying L∗(x0) = I,
K(x) = L∗(x)HL(x) is a solution of (3.3) and H satisfies H∗ = H, υH = Hυ.
4 Integration by parts
In this section we establish the relation between the differential system and the abstract relation
〈xiDif, g〉 = 〈f, xiDig〉 holding for 1 ≤ i ≤ N and f, g ∈ C1
(
TN ;Vτ
)
. We demonstrate how
close L is to providing the desired inner product, by performing an integration-by-parts over an
SN -invariant closed set ⊂ TNreg. Here L(x) and H satisfy the hypotheses listed in Condition 3.9
above. We use the identity xi∂if
∗(x) = −(xi∂if)∗(x). For δ > 0 let
Ωδ :=
{
x ∈ TN : min
1≤i<j≤N
|xi − xj | ≥ δ
}
.
This set is invariant under SN and K(x) is bounded and smooth on it. Thus the following
integrals exist.
12 C.F. Dunkl
Proposition 4.1. Suppose H satisfies Condition 3.9 then for f, g ∈ C1
(
TN ;Vτ
)
and 1 ≤ i ≤ N∫
Ωδ
{
−(xiDif(x))∗K(x)g(x) + f(x)∗K(x)xiDig(x)
}
dm(x)
=
∫
Ωδ
xi∂i{f(x)∗K(x)g(x)}dm(x).
Proof. By definition
xiDig(x) = xi∂ig(x) + κ
∑
j 6=i
xi
xi − xj
τ((i, j))(g(x)− g(x(i, j))),
(xiDif(x))∗ = −xi∂if(x)∗ + κ
∑
j 6=i
xj
xj − xi
(f(x)∗ − f(x(i, j))∗)τ((i, j)).
Thus
−(xiDif(x))∗K(x)g(x) + f(x)∗K(x)xiDig(x)
= xi∂if(x)∗ + xi∂ig(x)
+ κf(x)∗
∑
j 6=i
{
xj
xi − xj
τ((i, j))K(x) +K(x)τ((i, j))
xi
xi − xj
}
g(x)
− κ
∑
j 6=i
1
xi − xj
{
xjf(x(i, j))∗τ((i, j))K(x)g(x) + xif(x)∗K(x)τ((i, j))g(x(i, j))
}
= xi∂i{f(x)∗K(x)g(x)} (4.1)
− κ
∑
j 6=i
1
xi − xj
{
xjf(x(i, j))∗τ((i, j))K(x)g(x) + xif(x)∗K(x)τ((i, j))g(x(i, j))
}
.
For each pair {i, j} the term inside {·} is invariant under x 7→ x(i, j), because K(x(i, j)) =
τ((i, j))K(x)τ((i, j)), and xi − xj changes sign under this transformation. Thus∫
Ωδ
xjf(x(i, j))∗τ((i, j))K(x)g(x) + xif(x)∗K(x)τ((i, j))g(x(i, j))
xi − xj
dm(x) = 0
for each j 6= i because Ωδ and dm are invariant under (i, j). �
Observe the value of κ is not involved in the proof. Since xj∂j = −i ∂
∂θj
when xj = eiθj and
dm(x) = (2π)−Ndθ1 · · · dθN one step of integration can be directly evaluated. Consider the case
i = N and for a fixed (N − 1)-tuple (θ1, . . . , θN−1) with θ1 < θ2 < · · · < θN−1 < θ1 + 2π such
that
∣∣eiθj − eiθi
∣∣ ≥ δ the integral over θN is over a union of closed intervals. These are the
complement of
⋃
1≤j≤N−1
{θ : θj − δ′ < θ < θj + δ′} in the circle, where sin δ′
2 = δ
2 . This results in
an alternating sum of values of f∗Kg at the end-points of the closed intervals. Analyzing the
resulting integral (over (θ1, . . . , θN−1) with respect to dθ1 · · · dθN−1) is one of the key steps in
showing that a given K provides the desired inner product. In other parts of this paper we find
that H must satisfy another commuting relation.
5 Local power series near the singular set
In this section assume κ /∈ Z+1
2 . We consider the system (3.1) in a neighborhood of the face
{x : xN−1 = xN} of C0. We use a coordinate system which treats the singularity in a simple way.
For a more concise notation define
x(u, z) = (x1, x2, . . . , xN−2, u− z, u+ z) ∈ CN×
A System of Differential Equations on the Torus 13
We consider the system in terms of the variable x(u, z) subject to the conditions that the
points x1, x2, . . . , xN−2, u are pairwise distinct and |z| < min
1≤j≤N−2
|xj−u|, also |z| < |u|, Im z
u > 0
(these conditions imply arg(u− z) < arg(u+ z)). This allows power series expansions in z.
For z1, z2 ∈ C× let
ρ(z1, z2) :=
[
z1Imτ O
O z2Inτ−mτ
]
.
Let σ := τ((N − 1, N)) = ρ(−1, 1). We analyze the local solution L(x(u − z, u + z)) with an
initial condition specified later. We obtain the differential system (using ∂z := ∂
∂z , ∂u := ∂
∂u)
∂zL(x) = ∂NL− ∂N−1L
= κL
N−2∑
j=1
(
τ((j,N))
u− xj + z
− τ((j,N − 1))
u− xj − z
)
+
τ(N − 1, N)
z
− γ
u+ z
I +
γ
u− z
I
,
∂uL(x) = ∂NL+ ∂N−1L
= κL
N−2∑
j=1
(
τ((j,N))
u− xj + z
+
τ((j,N − 1))
u− xj − z
)
− γ
u+ z
I − γ
u− z
I
,
∂jL(x) = κL(x)
N−2∑
i=1,i 6=j
τ((i, j))
xj − xi
− γ
xj
I +
τ((j,N − 1))
xj − u+ z
+
τ((j,N))
xj − u− z
,
1 ≤ j ≤ N − 2.
Using the expansion 1
t−z =
∞∑
n=0
zn
tn+1 for |z| < |t| we let
βn(x(u, 0)) :=
N−2∑
j=1
τ((j,N))
(u− xj)n+1
for n = 0, 1, 2, . . .and express the equations as (since στ((j,N))σ = τ((j,N − 1)))
∂zL(x) = κL(x)
{ ∞∑
n=0
{
(−1)nβn(x(u, 0))− σβn(x(u, 0))σ
}
zn +
σ
z
− γ
u+ z
I +
γ
u− z
I
}
,
∂uL(x) = κL(x)
{ ∞∑
n=0
{
(−1)nβn(x(u, 0)) + σβn(x(u, 0))σ
}
zn − γ
u+ z
I − γ
u− z
I
}
,
∂jL(x) = κL(x)
N−2∑
i=1,i 6=j
τ((i, j))
xj − xi
− γ
xj
I −
∞∑
n=0
τ((j,N − 1)) + (−1)nτ((j,N))
(u− xj)n+1
zn
,
1 ≤ j ≤ N − 2.
Set
Bn(x(u, 0)) = (−1)nβn(x(u, 0))− σβn(x(u, 0))σ, n = 0, 1, 2, . . . .
Note σBnx(u, 0)σ = (−1)n+1Bn(x(u, 0)). Suggested by the relation
∂
∂z
ρ
(
z−κ, zκ
)
=
κ
z
ρ
(
− z−κ, zκ
)
=
κ
z
ρ
(
z−κ, zκ
)
σ
14 C.F. Dunkl
we look for a solution of the form
L(x) =
(u2 − z2
)N−2∏
j=1
xj
−γκ ρ(z−κ, zκ) ∞∑
n=0
αn(x(u, 0))zn, (5.1)
where each an(x(u, 0)) is matrix-valued and analytic in x(u, 0), and the initial condition is
α0
(
x(0)
)
= I, where x(0) is a base point, chosen as
(
1, ω, ω2, . . . , ωN−3, ω−3/2, ω−3/2
)
(that is,
u = ω−3/2, z = 0), where ω := e2πi/N . Implicitly restrict (x1, . . . , xN−1, u) to a simply connected
open subset of CN−1
reg containing
(
1, ω, ω2, . . . , ωN−3, ω−3/2
)
. Substitute (5.1) in the ∂z equation
(suppressing the x(u, 0) argument in the αn’s)
∂zL = κγ
(
1
u+ z
− 1
u− z
)(u2 − z2
)N−2∏
j=1
xj
−γκ ρ(z−κ, zκ) ∞∑
n=0
αnz
n
+
(u2 − z2
)N−2∏
j=1
xj
−γκ κ
z
ρ
(
z−κ, zκ
)
σ
∞∑
n=0
αnz
n
+
(u2 − z2
)N−2∏
j=1
xj
−γκ ρ(z−κ, zκ) ∞∑
n=1
nαnz
n−1
= κ
(u2 − z2
)N−2∏
j=1
xj
−γκ ρ(z−κ, zκ)
×
∞∑
n=0
αnz
n
{ ∞∑
m=0
Bm(u)zm +
σ
z
− γ
(
1
u+ z
− 1
u− z
)}
,
which simplifies to
κ
z
∞∑
n=0
(σαn − αnσ)zn +
∞∑
n=1
nαnz
n−1 = κ
∞∑
n=0
αnz
n
∞∑
m=0
Bm(x(u, 0))zm. (5.2)
The equations for ∂u and ∂j simplify to(u2 − z2
)N−2∏
j=1
xj
−γκ{ ∞∑
n=0
∂uαnz
n − κγ
(
1
u+ v
+
1
u− v
) ∞∑
n=0
αnz
n
}
= κ
(u2 − z2
)N−2∏
j=1
xj
−γκ ∞∑
n=0
αnz
n
×
{ ∞∑
m=0
{
(−1)mβm(x(u, 0)) + σβm(x(u, 0))σ
}
zm − γ
u+ v
I − γ
u− v
I
}
,
leading to (with 1 ≤ j ≤ N − 2)
∞∑
n=0
∂uαn(x(u, 0))zn = κ
∞∑
n=0
αn(x(u, 0))zn
∞∑
m=0
{
(−1)mβm(x(u, 0)) + σβm(x(u, 0))σ
}
zm,
∞∑
n=0
∂jαn(x(u, 0))zn = κ
∞∑
n=0
αn(x(u, 0))zn
A System of Differential Equations on the Torus 15
×
N−2∑
i=1,i/∈j
τ((i, j))
xj − xi
−
∞∑
m=0
τ((j,N − 1)) + (−1)mτ((j,N))
(u− xj)m+1
zm
.
We only need the equations for α0(x(u, 0)) (that is, the coefficient of z0) to initialize the ∂z
equation (this is valid because the system is Frobenius integrable):
∂uα0(x(u, 0)) = κα0(x(u, 0))
{
β0(x(u, 0)) + σβ0(x(u, 0))σ
}
, (5.3)
∂jα0(x(u, 0)) = κα0(x(u, 0))
N−2∑
i=1,i/∈j
τ((i, j))
xj − xi
− τ((j,N − 1)) + τ((j,N))
(u− xj)
,
2 ≤ j ≤ N − 2.
Lemma 5.1. σα0(x(u, 0))σ = α0(x(u, 0)) and α0(x(u, 0)) is invertible.
Proof. By hypothesis α0
(
x(0)
)
= I. The right hand sides of the system are invariant un-
der the transformation Q 7→ σQσ thus α0(x(u, 0)) and σα0(x(u, 0))σ satisfy the same system.
They agree at the base-point x(0), hence everywhere in the domain. By Jacobi’s identity the
determinant satisfies (where λ := tr(σ) = nτ − 2mτ )
∂u detα0(x(u, 0)) = κdetα0(x(u, 0)) tr{β0(x(u, 0)) + σβ0(x(u, 0))σ}
= 2κdetα0(x(u, 0))λ
N−2∑
j=1
1
(u− xj)
,
∂j detα0(x(u, 0)) = κλdetα0(x(u, 0))
N−2∑
i=1,i/∈j
1
xj − xi
− 2
u− xj
, 1 ≤ j ≤ N − 2,
detα0(x(u, 0)) =
∏
1≤i<j≤N−2
(
xi − xj
x
(0)
i − x
(0)
j
)λκ N−2∏
i=1
(
xi − u
x
(0)
i − x
(0)
N−1
)2λκ
,
the multiplicative constant follows from α0
(
x(0)
)
= I. Thus α0(x(u, 0)) is nonsingular in its
domain. �
We turn to the inductive definition of {αn(x(u, 0)) : n ≥ 1}.
In terms of the block decomposition (mτ + (nτ −mτ ))× (mτ + (nτ −mτ )) (henceforth called
the σ-block decomposition) of a matrix
α =
[
α11 α12
α21 α22
]
σασ = α if and only if α12 = O = α21 and σασ = −α if and only if α11 = O = α22. Write the
σ-block decomposition of αn(u) as
αn =
[
αn,11 αn,12
αn,21 αn,22
]
then the coefficient of zn−1 on the left side of equation (5.2) is
κ(σαn − αnσ) + nαn =
[
nαn,11 (n− 2κ)αn,12
(n+ 2κ)αn,21 nαn,22
]
,
16 C.F. Dunkl
and on the right side it is
κSn(x(u, 0)) := κ
n−1∑
i=0
αn−1−iBi(x(u, 0)),
for n ≥ 1. Arguing inductively suppose σαmσ = (−1)mαm for 0 ≤ m ≤ n, then σSnσ =
n−1∑
i=0
(σαn−1−iσ)(σBiσ) =
n−1∑
i=0
(−1)n−1−i+i−1αn−1−iBi and thus σSn(u)σ = (−1)nSn(u). In terms
of the σ-block decomposition
[
Sn,11 Sn,12
Sn,21 Sn,22
]
of Sn(x(u, 0)) this condition implies Sn,12 = O =
Sn,21 when n is even, and Sn,11 = O = Sn,22 when n is odd. This implies (for n = 1, 2, 3, . . .)
α2n(x(u, 0)) =
κ
2n
S2n(x(u, 0)), (5.4)
α2n−1(x(u, 0)) = ρ
(
κ
2n− 1− 2κ
,
κ
2n− 1 + 2κ
)
S2n−1(x(u, 0)),
and thus σαn(x(u, 0))σ = (−1)nαn(x(u, 0)). In particular
α1(x(u, 0)) = ρ
(
κ
1− 2κ
,
κ
1 + 2κ
)
α0(x(u, 0))B0(x(u, 0)),
and all the coefficients are determined; by hypothesis κ /∈ Z+1
2 and the denominators are of the
form 2m+ 1± 2κ.
Henceforth denote the series (5.1), solving (3.1) with the normalization α0
(
x(0)
)
= I by L1(x).
It is defined for all x(u, z) ∈ C0 subject to |z| < min
1≤j≤N−2
|xj − u|, also |z| < |u|, Im z
u > 0. The
radius of convergence depends on x(u, 0). Return to using L(x) to denote the solution from
Definition 3.6 (on all of TNreg and L(x0) = I). In terms of x(u, z) the point x0 corresponds to
u = 1
2
(
ω−1 +ω−2
)
, z = 1
2
(
ω−1−ω−2
)
, x(u, z) =
(
1, ω, . . . , ωN−3, u−z, u+z
)
, then min
1≤j≤N−2
|u−
xj | = sin π
N
(
5+4 cos 2π
N
)
and |z| = sin π
N (also z
u = i tan π
N ) and x0 is in the domain of convergence
of the series L1(x). Thus the relation L1(x) = L1(x0)L(x) holds in the domain of L1 in C0. This
implies the important fact that L1(x0) is an analytic function of κ, to be exploited in Section 9.
5.1 Behavior on boundary
The term ρ(z−κ, zκ) implies that L1(x) is not continuous at z = 0, that is, on the boundary
{x : xN−1 = xN}. However there may be a weak type of continuity, specifically
lim
xN−1−xN→0
(K(x)−K(x(N − 1, N))) = 0.
With the aim of expressing the desired K(x) in the form L(x)∗C∗CL(x) (and C is unknown at
this stage) we consider CL(x) in series form, that is CL1(x0)−1L1(x) (recall detL(x) 6= 0 in C0).
We analyze the effect of C on the weak continuity condition. Denote C ′ := CL1(x0)−1.
From Proposition 3.7 L(x(N − 1, N)) = ν((N − 1, N))L(x)τ((N − 1, N)) = L(x)σ, because
w(1) = 1 for w = (N−1, N), [for the special case N = 3, τ = (2, 1), T3
reg has two components and
we define L(x) = L(x(2, 3))σ for the component 6= C0]. By use of x(u, z)(N − 1, N) = x(u,−z)
it follows that
CL(x(u, z)(N − 1, N)) = CL(x(u, z))σ = C ′ (xNxN−1)−γκ ρ
(
z−κ, zκ
) ∞∑
n=0
αn(u)znσ
A System of Differential Equations on the Torus 17
= C ′σ(xNxN−1)−γκρ
(
z−κ, zκ
) ∞∑
n=0
αn(u)(−1)nzn,
because σαn(u)σ = (−1)nαn(u) and σ = ρ(−1, 1).
Recall L∗(x) is defined as L(φx)T with complex constants replaced by their conjugates. Then
φx(u, z) =
(
x−1
1 , x−1
2 , . . . , x−1
N−2,
1
u−z ,
1
u+z
)
. To compute L1(φx(u, z)) replace u by u′ = u
(u+z)(u−z)
and replace z by z′ = − z
(u+z)(u−z) . When restricted to the torus u′ = 1
2
(
1
xN−1
+ 1
xN
)
= u and
z′ = 1
2
(
1
xN−1
− 1
xN
)
= z. The terms βn(u) :=
N−2∑
j=1
τ((j,N))
(u−xj)n+1 in the intermediate formulae for L1
are replaced by their complex conjugates when x(u, z) ∈ TN . Similarly β̃k :=
∞∑
m=0
N−2∑
j=1
τ((j,N))
(u0−xj)k+1
transforms to (β̃k) because the constant u0 is conjugated. Thus for x(u, z) ∈ TNreg
L1(x(u, z))∗ =
∞∑
m=0
αm(u)∗zmρ
(
z−κ, zκ
)
C
′∗(xNxN−1)−γκ;
αm(u)∗ denotes the adjoint of the matrix αm(u). Then
L1(x(u, z)(N − 1, N))∗ =
∞∑
m=0
(−1)mαm(u)∗zmρ(z−κ, zκ)σC
′∗(xNxN−1)−γκ.
Furthermore (recall K(x(N − 1, N)) = σK(x)σ by definition)
K(x(u, z)) =
∞∑
m,n=0
zmznαm(u)∗ρ
(
z−κ, zκ
)
C ′∗C ′ρ
(
z−κ, zκ
)
αn(u),
K(x(u,−z)) =
∞∑
m,n=0
zmzn(−1)m+nαm(u)∗ρ
(
z−κ, zκ
)
σC ′∗C ′σρ
(
z−κ, zκ
)
αn(u).
The term of lowest order in z in K(x(u, z))−K(x(u,−z)) is
α0(u)∗ρ
(
z−κ, zκ
){
C ′∗C ′ − σC ′∗C ′σ
}
ρ
(
z−κ, zκ
)
α0(u).
In terms of the σ-block decomposition, with
C ′∗C ′ =
[
c11 c12
c∗12 c22
]
, α0(u) =
[
a11(u) O
O a22(u)
]
the expression equals
2
O
(z
z
)κ
a11(u)∗c12a22(u)(
z
z
)κ
a22(u)∗c∗12a11(u) O
,
which tends to zero as z → 0 if and only if c12 = 0, that is, σC∗Cσ = C∗C.
Proposition 5.2. Suppose C ′∗C ′ commutes with σ then
K(x(u, z))−K(x(u, z)(N − 1, N)) = O
(
|z|1−2|κ|).
18 C.F. Dunkl
Proof. The hypothesis implies C ′∗C ′ commutes with ρ(z−κ, zκ), thus
K(x(u, z))−K(x(u, z)(N − 1, N))
=
∞∑
m,n=0
zmzn
(
1− (−1)m+n
)
αm(u)∗ρ
(
|z|−2κ, |z|2κ
)
C ′∗C ′αn(u)
= 2zα0(u)∗ρ
(
|z|−2κ, |z|2κ
)
C ′∗C ′α1(u) + 2zα1(u)∗ρ
(
|z|−2κ, |z|2κ
)
C
′∗C ′α0(u)
+
∞∑
m+n≥2
zmzn
(
1− (−1)m+n
)
αm(u)∗ρ
(
|z|−2κ, |z|2κ
)
C ′∗C ′αn(u).
The dominant terms come from m = 0, n = 1 and m = 1, n = 0; both of order O
(
|z|1−2|κ|). �
We will see later for purpose of integration by parts, that the change in K between the points(
x1, . . . , xN−2, e
iθ, ei(θ−ε)) and
(
x1, . . . , xN−2, e
iθ, ei(θ+ε)
)
is a key part of the analysis; this uses
the relation K
((
x1, . . . , xN−2, e
iθ, ei(θ−ε))) = σK
((
x1, . . . , xN−2, e
i(θ−ε), eiθ
))
σ.
6 Bounds
In this section we derive bounds on L(x) of global and local type. Throughout we adopt the
normalization L(x0) = I. The operator norm on nτ × nτ complex matrices is defined by
‖M‖ = sup{|Mv| : |v| = 1}.
Theorem 6.1. There is a constant c depending on κ such that ‖L(x)‖ ≤ c
∏
1≤i<j≤N
|xi− xj |−|κ|
for each x ∈ TNreg.
The proof is a series of steps starting with a general result which applies to matrix functions
satisfying a linear differential equation in one variable.
Lemma 6.2. Suppose M(0) = I, d
dtM(t) = M(t)F (t) and ‖F (t)‖ ≤ f(t) for 0 ≤ t ≤ 1 then
‖M(t)− I‖ ≤ exp
∫ t
0 f(s)ds− 1 and ‖M(1)‖ ≤ exp
∫ 1
0 f(s)ds.
Proof from [11, Theorem 7.1.11]. Let `(t) := ‖M(t) − I‖ then the equation M(t) − I =∫ t
0 M(s)F (s)ds and the inequalities ‖M(t)‖ ≤ ‖M(t) − I‖ + ‖I‖ (and ‖I‖ = 1) imply that
`(t) ≤
∫ t
0 (`(s) + 1)f(s)ds. Define differentiable functions b(t) and h(t) by
h(t) := exp
∫ t
0
f(s)ds,
b(t)h(t) =
∫ t
0
(`(s) + 1)f(s)ds+ 1.
Apply d
dtto the latter equation:
b′(t)h(t) + b(t)f(t)h(t) = (`(t) + 1)f(t),
b′(t)h(t) = f(t)
{
`(t) + 1−
∫ t
0
(`(s) + 1)f(s)ds− 1
}
= f(t)
{
`(t)−
∫ t
0
(`(s) + 1)f(s)ds
}
≤ 0.
Hence b′(t) ≤ 0 and b(t) ≤ b(0) = 1 which implies
`(t) ≤
∫ t
0
(`(s) + 1)f(s)ds = b(t)h(t)− 1 ≤ h(t)− 1.
Finally ‖M(1)‖ ≤ ‖M(1)− I‖+ 1 ≤ exp
∫ 1
0 f(s)ds. �
A System of Differential Equations on the Torus 19
Next we set up a differentiable path p(t) = (p1(t), . . . , pN (t)) in CNreg starting at x0 and obtain
the equation
d
dt
L(p(t)) = κL(p(t))
N∑
i=1
∑
j 6=i
p′i(t)
pi(t)− pj(t)
τ((i, j))− γ p
′
i(t)
pi(t)
I
= κL(p(t))
∑
1≤i<j≤N
p′j(t)− p′i(t)
pj(t)− pi(t)
τ((i, j))− γ
N∑
i=1
p′i(t)
pi(t)
I
.
Suppose x =
(
eiθ1 , . . . , eiθN
)
∈ C0 and θ1 < θ2 < · · · < θN < θ1 + 2π. Define the path
p(t) =
(
eig1(t), . . . , eigN (t)
)
where gj(t) = (1 − t)2(j−1)π
N + tθj for 1 ≤ j ≤ N . Then p(t) ∈ C0
for 0 ≤ t ≤ 1 because gi+1(t) − gi(t) = (1 − t)2π
N + t(θi+1 − θi) > 0 for 1 ≤ i < N and
2π + g1(t) − gN (t) = 2π + tθ1 − (1 − t)2(N−1)π
N − tθN = (1 − t)2π
N + t(2π + θ1 − θN ) > 0. The
factor of τ((i, j)) in the equation is
i
g′j(t)e
igj(t) − g′i(t)eigi(t)
eigj(t) − eigi(t)
=
1
2
{
(g′j(t)− g′i(t))
cos
(
1
2(gj(t)− gi(t))
)
sin
(
1
2(gj(t)− gi(t))
) + i(g′j(t) + g′i(t))
}
.
We will apply Lemma 6.2 to L̃(x) =
N∏
j=1
xγκj L(x); this only changes the phase and removes the
N∑
i=1
p′i(t)
pi(t)
term. In the notation of Lemma 6.2
f(t) = |κ|
∑
i<j
∣∣∣∣∣ig′j(t)eigj(t) − g′i(t)eigi(t)
eigj(t) − eigi(t)
∣∣∣∣∣
(since ‖τ((i, j))‖ = 1). To set up the integral
∫ 1
0 f(t)dt let
φij(t) =
1
2
(gj(t)− gi(t)) =
1
2
{
(1− t)2(j − i)π
N
+ t(θj − θi)
}
so that φ
′
ij(t) = 1
2
(
θj − θi + 2(j−i)π
N
)
and 0 < φij(t) < π for i < j and 0 ≤ t ≤ 1. The terms
|i(g′j(t) + g′i(t))| ≤ 4π provide a simple bound (no singularities off TNreg). The dominant terms
come from
∫ 1
0
|φ′ij cosφij(t)|
sinφij(t)
dt. There are two cases. Let φ0, φ1 satisfy 0 < φ0, φ1 < π and let
φ(t) = (1 − t)φ0 + tφ1. The antiderivative
∫ φ′ cosφ(t)
sinφ(t) dt = log sinφ(t). The first case applies
when either 0 < φ0, φ1 ≤ π
2 or π
2 ≤ φ0, φ1 < π (assign φ0 = π
2 = φ1 to the first interval); then
φ′ cosφ(t) ≥ 0 if 0 < φ0 ≤ φ1 ≤ π
2 or π
2 ≤ φ1 ≤ φ0 < π and φ′ cosφ(t) < 0 otherwise. These
imply∫ 1
0
|φ′ cosφ(t)|
sinφ(t)
dt =
∣∣∣∣log
sinφ1
sinφ0
∣∣∣∣ .
The second case applies when either 0 < φ0 <
π
2 < φ1 < π (thus φ′ > 0) or 0 < φ1 <
π
2 < φ0 < π.
Let φ(t0) = π
2 (that is, t0 = π/2−φ0
φ1−φ0 ). In the first situation∫ 1
0
|φ′ cosφ(t)|
sinφ(t)
dt =
∫ t0
0
φ′ cosφ(t)
sinφ(t)
dt−
∫ 1
t0
φ′ cosφ(t)
sinφ(t)
dt = − log sinφ0 − log sinφ1,
20 C.F. Dunkl
since log sin π
2 = 0; and the same value holds for the second situation. We obtain∫ 1
0
f(t)dt ≤ |κ|
∑
1≤i<j≤N
{
− log sin
θj − θi
2
− log sin
(j − i)π
N
+ 4π
}
.
Taking exponentials and using the lemma (recall
∣∣eiφ1 − eiφ2
∣∣ = 2 sin
∣∣φ1−φ2
2
∣∣) we obtain
‖L(x)‖ ≤ c
∏
1≤i<j≤N
|xi − xj |−|κ|.
This bound applies to all of TNreg when L(x0) commutes with υ and L is extended to TNreg as in
Definition 3.6. This completes the proof of Theorem 6.1.
Next we find bounds on the series expansion from (5.1)
L1(x) =
(u2 − z2
)N−2∏
j=1
xj
−γκ ρ(z−κ, zκ) ∞∑
n=0
αn(x(u, 0))zn,
where |z| < δ0 := min
1≤j≤N−2
|u− xj | and Im z
u > 0. Recall the recurrence (5.4)
Sn :=
n−1∑
i=0
αn−1−i
{
(−1)iβi − σβiσ
}
, βi :=
N−2∑
j=0
τ((j,N))
(u− xj)i+1
,
α2n(x(u, 0)) =
κ
2n
S2n,
α2n+1(x(u, 0)) = ρ
(
κ
2n+ 1− 2κ
,
κ
2n+ 1 + 2κ
)
S2n+1.
Proposition 6.3. Suppose |κ| ≤ κ0 <
1
2 and λ := (N − 2)κ0 then for n ≥ 0
‖α2n(x(u, 0))‖ ≤ ‖α0(x(u, 0))‖
(λ)n
(
λ+ 1
2 − κ0
)
n
n!
(
1
2 − κ0
)
n
δ−2n
0 , (6.1)
‖α2n+1(x(u, 0))‖ ≤ ‖α0(x(u, 0))‖
(λ)n+1
(
λ+ 1
2 − κ0
)
n
n!
(
1
2 − κ0
)
n+1
δ−2n−1
0 .
Proof. Suppose n ≥ 1 then ‖Sn‖ ≤
n−1−i∑
i=0
‖αn−1−i‖(2N − 4)δ−i−1
0 (since ‖τ((j,N − 1))‖ = 1).
Furthermore, since
∣∣ κ
n±2κ
∣∣ ≤ κ0
n−2κ0
for n ≥ 2, we find
‖α2n+1‖ ≤
2λ
2n+ 1− 2κ0
2n∑
i=0
‖α2n−i‖δ−i−1
0 ,
‖α2n‖ ≤
λ
n
2n−1∑
i=0
‖α2n−1−i‖δ−i−1
0 .
To set up an inductive argument let tn denote the hypothetical bound on ‖αn(x(u, 0))‖ and
set vn =
n−1∑
i=0
tn−1−iδ
−i−1
0 ; then vn = δ−1
0 (tn−1 + vn−1) for n ≥ 2. Setting t2n = λ
nv2n and
t2n+1 = 2λ
2n+1−2κ0
v2n+1 the recurrence relations become
t2n =
λ
n
(
t2n−1 +
2n− 1− 2κ0
2λ
t2n−1
)
δ−1
0 =
2λ+ 2n− 1− 2κ0
2n
t2n−1δ
−1
0 ,
A System of Differential Equations on the Torus 21
t2n+1 =
2λ
2n+ 1− 2κ0
(
t2n +
n
λ
t2n
)
δ−1
0 =
2λ+ 2n
2n+ 1− 2κ0
t2nδ
−1
0 .
Starting with ‖α1‖ ≤ 2λ
1−2κ0
‖α0‖δ−1
0 = t1 the stated bounds are proved inductively. �
By use of Stirling’s formula for Γ(n+a)
Γ(n+b) ∼ na−b we see that tn behaves like (a multiple of)
n2λ−1 for large n. Also there is a constant c′ depending on N and κ0 such that
∞∑
n=2
‖αn(x(u, 0))‖|z|n ≤ c′‖α0(x(u, 0))‖
(
|z|
δ0
)2(
1− |z|
δ0
)−2λ−2
. (6.2)
We also need to analyze the effect of small changes in u. Fix a point x(ũ, 0) and consider series
expansions of αn(x(ũ, 0)) in powers of (u− ũ). Let δ1 := min
1≤j≤N−2
|ũ−xj |. Recall equation (5.3)
∂uα0(x(u, 0)) = κα0(x(u, 0)){β0(x(u, 0)) + σβ0(x(u, 0))σ},
and solve this in the form
α0(x(u, 0)) =
∞∑
n=0
α0,n(x(ũ, 0))(u− ũ)n.
This leads to the recurrence (suppressing the arguments x(ũ, 0))
∞∑
n=1
nα0,n(u− ũ)n−1
= κ
∞∑
n=0
α0,n(u− ũ)n
∞∑
m=0
(−1)m(u− ũ)m
N−2∑
j=0
τ((j,N − 1)) + τ((j,N))
(ũ− xj)m+1
,
(n+ 1)α0,n+1 = κ
n∑
m=0
α0,n−mβ̃m(x(ũ, 0)),
β̃m(x(ũ, 0)) := (−1)m
N−2∑
j=0
τ((j,N − 1)) + τ((j,N))
(ũ− xj)m+1
.
Thus ‖β̃m‖ ≤ 2(N−2)
δm+1
1
and by a similar method as above we find
‖α0,n(x(ũ, 0))‖ ≤ (2λ)n
n!
δ−n1 ‖α0(x(ũ, 0))‖, (6.3)
where λ = (N − 2)κ0. From
α1(x(u, 0)) = ρ
(
κ
1− 2κ
,
κ
1 + 2κ
)
α0(x(u, 0))
N−2∑
j=0
τ((j,N − 1))− τ((j,N))
(u− xj)
= ρ
(
κ
1− 2κ
,
κ
1 + 2κ
) ∞∑
n=0
α0,n(u− ũ)n
×
∞∑
m=0
(−1)m(u− ũ)m
N−2∑
j=0
τ((j,N − 1))− τ((j,N))
(ũ− xj)m+1
,
22 C.F. Dunkl
we can derive a recurrence for the coefficients in α1(x(u, 0)) =
∞∑
n=0
α1,n(x(ũ, 0))(u− ũ)n. Also
‖α1,n(x(ũ, 0))‖ ≤ 2λ
1− 2κ0
‖α0(x(ũ, 0))‖
n∑
j=0
(2λ)n−j
(n− j)!
δj−n1 δ−j−1
1
=
2λ(2λ+ 1)n
(1− 2κ0)n!
δ−n−1
1 ‖α0(x(ũ, 0))‖;
note 2λ(2λ+ 1)n = (2λ)n+1.
Essentially we are setting up bounds on behavior of L(x(u, z)) for points near x(ũ, 0) in terms
of ‖α0(x(ũ, 0))‖ which is handled by the global bound.
In the series
∞∑
n=0
αn(x(u, 0))zn =
∞∑
m,n=0
αn,m(x(ũ, 0))zn(u− ũ)m
the first order terms are
α00(x(ũ, 0)) + α0,1(x(ũ, 0))(u− ũ) + α1,0(x(ũ, 0))z,
and the bounds (6.1) for the omitted terms
∞∑
n=2
‖αn(x(u, 0))‖|z|n ≤ c′‖α0(x(u, 0))‖
(
|z|
δ0
)2(
1− |z|
δ0
)−2λ−2
,
∞∑
n=2
‖α0,n(x(ũ, 0))‖|u− ũ|n ≤ (2λ)2
(
|u− ũ|
δ1
)2(
1− |u− ũ|
δ1
)−2λ−2
‖α0(x(ũ, 0))‖, (6.4)
|z|
∞∑
n=1
‖α1,n(x(ũ, 0))‖|u− ũ|n ≤ (2λ)2
1− 2κ0
(
|z(u− ũ)|
δ2
1
)(
1− |u− ũ|
δ1
)−2λ−2
‖α0(x(ũ, 0))‖.
Note there is a difference between δ0 and δ1: δ0, δ1 are the distances from the nearest xj
(1 ≤ j ≤ N − 2) to u, ũ respectively; thus the double series converges in |z| + |u − ũ| < δ1
because this implies |z| < δ1 − |u− ũ| ≤ δ0, by the triangle inequality: δ1 ≤ |u− ũ|+ δ0 .
7 Sufficient condition for the inner product property
In this section we will use the series
L1(y, u− z, u+ z) =
N∏
j=1
xj
−γκ ρ(z−κ, zκ) ∞∑
n=0
αn(x(u, 0))zn,
normalized by α0
(
x(0)
)
= I where x(0) =
(
1, ω, ω2, . . . , ωN−3, ω−3/2, ω−3/2
)
, ω = e2πi/N . The
hypothesis is that there exists a Hermitian matrix H such that υH = Hυ (recall υ := τ(w0))
and the matrix H1 defined by
L1(x)∗H1L1(x) = L(x)∗HL(x) (7.1)
commutes with σ = τ(N − 1, N) (recall L(x0) = I). Setting x = x0 we find that H =
L1(x0)∗H1L1(x0). The analogous condition has to hold for each face of C0 and any such face
can be obtained from {xN−1 = xN} by applying x 7→ xwm0 with suitable m. For notational
simplicity we will work out only the {xN−2 = xN−1} case. From the general relation w(i, j)w−1 =
A System of Differential Equations on the Torus 23
(w(i), w(j)) we obtain w−1
0 (N − 1, N)w0 = (N − 2, N − 1). A matrix M commutes with τ(N −
2, N − 1) if and only if υMυ−1 commutes with σ. Let
x′ =
(
x′1, . . . , x
′
N−3, u− z, u+ z, x′N
)
,
x′w−1
0 =
(
x′N , x
′
1, . . . , x
′
N−3, u− z, u+ z
)
= x,
L2(x′) := υ−1L1
(
x′w−1
0
)
υ =
N∏
j=1
xj
−γκ υ−1ρ
(
z−κ, zκ
) ∞∑
n=0
αn(y, u)υzn.
This is a solution of (3.1) by Proposition 3.1. This has the analogous behavior to L1; writing
υ−1ρ
(
z−κ, zκ
)
αn(y, u)υ =
{
υ−1ρ
(
z−κ, zκ
)
υ
}{
υ−1αn(y, u)υ
}
implies the relations
τ(N − 2, N − 1)
{
υ−1ρ
(
z−κ, zκ
)
υ
}
=
{
υ−1ρ
(
z−κ, zκ
)
υ
}
τ(N − 2, N − 1),
τ(N − 2, N − 1)
{
υ−1αn(y, u)υ
}
τ(N − 2, N − 1) = (−1)n
{
υ−1αn(y, u)υ
}
, n ≥ 0.
We claim that the Hermitian matrix H2 defined by
L2(x)∗H2L2(x) = L(x)∗HL(x) (7.2)
commutes with τ(N−2, N−1). There is a subtle change: the base point x(0) =
(
1, ω, . . . , ωN−2,
ω−3/2, ω−3/2
)
is replaced by
(
ω, . . . , ωN−2, ω−3/2, ω−3/2, 1
)
and now ωx0 =
(
ω, . . . , ωN−1, 1
)
is
in the domain of convergence of L2. Set x = ωx0 in (7.2) to obtain
L2(ωx0) = υ−1L1
(
ωx0w
−1
0
)
υ = υ−1L1(x0)υ,
H2 = (L2(ωx0)∗)−1HL2(ωx0)−1 = υ−1(L1(x0)∗)−1υHυ−1L1(x0)−1υ
= υ−1(L1(x0)∗)−1HL1(x0)−1υ = υ−1H1υ,
because H commutes with υ (and L(ωx0) = L(x0) = I by the homogeneity). Thus H2 commutes
with τ(N − 2, N − 1).
From Theorem 6.1 we have the bound
‖L(x)∗HL(x)‖ ≤ c
∏
i<j
|xi − xj |−2|κ|.
Denote K(x) = L(x)∗HL(x). We will show that there is an interval −κ1 < κ < κ1 where κ1
depends on N such that∫
TN
{
(xNDNf)∗(x)K(x)g(x)− f∗(xt)K(x)xNDNg(x)
}
dm(x) = 0,
for each f, g ∈ C1
(
TN ;Vτ
)
. Consider the Haar measure of
{
x : min
i<j
|xi − xj | < ε
}
; let sin ε′
2 = ε
2
and i < j then m
{
x : |xi − xj | ≤ ε
}
= 1
πε
′, thus m
{
x : min
i<j
|xi − xj | < ε
}
≤
(
N
2
)
ε′
π . The integral
is broken up into three pieces. The aim is to let δ → 0; where δ satisfies an upper bound
δ < min
((
2 sin π
N
)2
, 1
9
)
; the first term comes from the maximum spacing of N points on T and
the second is equivalent to 3δ < δ1/2. Also δ′ := 2 arcsin δ
2 .
1. min
i<j
|xi − xj | < δ, done with the integrability of
∏
i<j
|xi − xj |−2|κ| for |κ| < 1
N (from the
Selberg integral
∫
TN
∏
i<j
|xi−xj |−2|κ|dm(x) = Γ1−N |κ|)
Γ(1−|κ|)N ), and the measure of the set is O(δ).
The limit as δ → 0 is zero by the dominated convergence theorem.
24 C.F. Dunkl
2. δ ≤ min
1≤i<j≤N−1
|xi − xj | < δ1/2 and δ ≤ min
1≤i≤N−1
|xi − xN |; this case uses the same bound
on K and the TN−1-Haar measure of
{
x ∈ TN−1 : min
1≤i<j≤N−1
|xi − xj | < δ1/2
}
;
3. min
1≤i<j≤N−1
|xi−xj | ≥ δ1/2 and min
1≤i≤N−1
|xi−xN | ≥ δ. This is done with a detailed analysis
using the double series from (6.4).
The total of parts (2) and (3), that is, the integral over Ωδ, equals∫
Ωδ
xN∂N{f(x)∗K(x)g(x)}dm(x).
We use the coordinates xj = eiθj , 1 ≤ j ≤ N ; thus xN∂N = −i ∂
∂θN
. For fixed (θ1, . . . , θN−1)
the condition x ∈ Ωδ implies that the set of θN -values is a union of disjoint closed inter-
vals (it is possible there is only one, in the extreme case θj = jδ′ for 1 ≤ j ≤ N − 1
the interval is Nδ′ ≤ θN ≤ 2π). In case (2) the θN -integration results in a sum of terms
(f∗Kg)
(
eiθ1 , . . . , eiθN−1 , eiφ
)
with coefficients ±1 where min
1≤i≤N−1
|φ− θi| = δ. Each such sum is
bounded by 2(N − 1)c‖f‖∞‖g‖∞δ−N(N−1)|κ|, because
∏
i<j
|xi−xj | ≥ δN(N−1)/2 on Ωδ. Thus the
integral for part (2) is bounded by
2(N − 1)‖f‖∞‖g‖∞δ−N(N−1)|κ|
(
N − 1
2
)(
2 arcsin
δ1/2
2
)
≤ c′δ1/2−N(N−1)|κ|,
for some finite constant c′ (depending on f , g). This term tends to zero as δ → 0 if |κ| < 1
2N(N−1) .
In part (3) the intervals [θi − δ′, θi + δ′] are pairwise disjoint because |θi − θj | ≥ 3δ′ for i 6= j
(recall
√
δ > 3δ). To simplify the notation assume θ1 < θ2 < · · · (the other cases follow from
the group invariance of the setup). Then the θN -integration yields
(2π)1−N
N−1∑
j=1
∫
Rδ
{
(f∗Kg)
(
eiθ1 , . . . , eiθj , . . . , ei(θj−δ′)
)
−(f∗Kg)
(
eiθ1 , . . . , eiθj , . . . , ei(θj+δ
′)
)}dθ1 · · · dθN−1,
where Rδ :=
{
(θ1, . . . , θN−1) : θ1 < θ2 < · · · < θN−1 < θ1+2π,min
∣∣eiθj−eiθk
∣∣ ≥ √δ}. It suffices
to deal with the term with j = N − 1; this allows the use of the double series. It is fairly easy
to show that (f∗Kg)
(
eiθ1 , . . . , eiθN−1 , ei(θN−1−δ′)
)
− (f∗Kg)
(
eiθ1 , . . . , eiθN−1 , ei(θN−1+δ′)
)
tends to
zero with δ but this is not enough to control the integral. The idea is to show that∣∣(f∗Kg)
(
eiθ1 , . . . , eiθN−1 , ei(θN−1−δ′)
)
− (f∗Kg)
(
eiθ1 , . . . , eiθN−1 , ei(θN−1+δ′)
)∣∣
≤ c′′δ1/2−2|κ|∣∣(f∗Kg)
(
eiθ1 , . . . , eiθN−1 , ei(θN−1+δ′)
)∣∣
for some constant c′′. This can then be bounded using the ‖K‖ bound for sufficiently small |κ|.
Fix y =
(
eiθ1 , . . . , eiθN−2
)
and let x(y, u− v, u+ v) denote
(
eiθ1 , . . . , eiθN−2 , u− z, u+ z
)
. We will
use the form K = L∗1H1L1 from (7.1) with two pairs of values along with ũ = eiθN−1 , and set
ζ = eiδ′
1) η(1) = x(y, u1−z1, u1 +z1) = x
(
y, eiθN−1 , ei(θN−1+δ′)
)
, then u1 = 1
2 ũ(1+ ζ), z1 = 1
2 ũ(ζ−1),
u1 − ũ = z1, |z1| = δ,
2) η(2) = x(y, u2 − z2, u2 + z2) = x
(
y, ei(θN−1−δ′), eiθN−1
)
, then u2 = 1
2 ũ
(
1 + ζ−1
)
, z2 =
1
2 ũ
(
1− ζ−1
)
= ζ−1z1, u2 − ũ = −z2, |z2| = δ.
A System of Differential Equations on the Torus 25
Let η(3) =x
(
y, eiθN−1 , ei(θN−1−δ′)
)
=η(2)(N−1, N), then by construction K
(
η(3)
)
=σK
(
η(2)
)
σ.
We start by disposing of the f and g factors: by uniform continuous differentiability there is
a constant c′′′ such that
∥∥f(η(1)
)
− f
(
η(3)
)∥∥ ≤ c′′′δ′ and
∥∥g(η(1)
)
− g
(
η(3)
)∥∥ ≤ c′′′δ′ (same
constant for all of TN ). So the error made by assuming f and g are constant is bounded by
c′′′δ′
(∥∥K(η(1)
)∥∥ +
∥∥K(η(2)
)∥∥). The problem is reduced to bounding K
(
η(1)
)
− σK
(
η(2)
)
σ. To
add more detail about the effect of the ∗-operation on u and z we compute
z∗ =
1
2
(
1
u+ z
− 1
u− z
)
= − z
u2 − z2
, u∗ =
1
2
(
1
u+ z
+
1
u− z
)
=
u
u2 − z2
and if u− z = eiθN−1 , u+ z = eiθN then
z =
1
2
ei(θN−1+θN )/2
(
ei(θN−θN−1)/2 − ei(θN−1−θN )/2
)
= iei(θN−1+θN )/2 sin
θN − θN−1
2
,
z∗ = ie−i(θN−1+θN )/2 sin
θN−1 − θN
2
= z,
u =
1
2
ei(θN−1+θN )/2
(
ei(θN−θN−1)/2 + ei(θN−1−θN )/2
)
= ei(θN−1+θN )/2 cos
θN − θN−1
2
,
u∗ = e−i(θN−1+θN )/2 cos
θN−1 − θN
2
= u;
the ∗-operation agrees with complex conjugate on the torus and ρ(z−κ, zκ)∗ = ρ(z−κ, zκ). The
reason for this is to emphasize that L(x)∗ is an analytic function agreeing with the (Hermitian)
adjoint of L(x). Thus
K
(
η(1)
)
=
∞∑
n,m=0
αn(x(u1, 0))∗ρ
(
z−κ1 , zκ1
)∗
H1ρ
(
z−κ1 , zκ1
)
αm(x(u1, 0))(z∗1)mzn1
=
∞∑
n,m=0
αn(x(u1, 0))∗H1ρ
(
|z1|−2κ, |z1|2κ
)
αm(x(u1, 0))z1
mzn1 , (7.3)
because H1 commutes with σ and hence with ρ(z−κ1 , zκ1 ), and
K
(
η(3)
)
= σK
(
η(2)
)
σ
=
∞∑
n,m=0
(−1)m+nαn(x(u2, 0))∗H1ρ
(
|z2|−2κ, |z2|2κ
)
αm(x(u2, 0))z2
mzn2 , (7.4)
because σαn(x(u2, 0))σ = (−1)nαn(x(u2, 0)) for n ≥ 0. Now we use the expansion in powers of
(u− ũ)n to evaluate K
(
η(1)
)
−K
(
η(3)
)
. From the inequality (6.2)
∞∑
n=2
‖αn(x(u, 0))‖|z|n ≤ c′‖α0(x(u, 0))‖
(
|z|
δ0
)2(
1− |z|
δ0
)−2λ−2
= c′δ‖α0(x(u, 0))‖
(
1− δ1/2
)−2λ−2
(7.5)
with δ0 = min
1≤j≤N−2
|u − xj | = δ1/2 we can restrict the problem to 0 ≤ n,m ≤ 1. The omitted
terms in K
(
η(1)
)
−K
(
η(3)
)
are bounded by c′′δ1−2|κ|‖B1‖‖α0(x(u2, x))‖2, for some constant c′′.
Then
L1
(
η(1)
)
=
N∏
j=1
x
(1)
j
−γκ ρ(z−κ1 , zκ1
){α00(x(ũ, 0)) + α0,1(x(ũ, 0))(u1 − ũ)
+α1,0(x(ũ, 0))z1 +O(δ)
}
,
26 C.F. Dunkl
σL1
(
η(2)
)
σ =
N∏
j=1
x
(2)
j
−γκ ρ(z−κ2 , zκ2
){α00(x(ũ, 0)) + α0,1(x(ũ, 0))(u2 − ũ)
−α1,0(x(ũ, 0))z2 +O(δ)
}
,
because σα1(x(u, 0))σ = (−1)nα1(x(u, 0)). The terms O(δ) correspond to the bound in (7.5).
Drop the argument x(ũ, 0) for brevity. Combining these with (7.3) and (7.4) we obtain
K
(
η(1)
)
−K
(
η(3)
)
= {α0,1(u1 − ũ) + α1,0z1}∗H1ρ
(
|z1|−2κ, |z1|2κ
)
α00
+ α∗00H1ρ
(
|z1|−2κ, |z1|2κ
)
{α0,1(u1 − ũ) + α1,0z1}
+ {α0,1(u1 − ũ) + α1,0z1}∗H1ρ
(
|z1|−2κ, |z1|2κ
)
× {α0,1(u1 − ũ) + α1,0z1}
− {α0,1(u2 − ũ)− α1,0z2}∗H1ρ
(
|z2|−2κ, |z2|2κ
)
α00
− α∗00H1ρ
(
|z2|−2κ, |z2|2κ
)
{α0,1(u1 − ũ)− α1,0z1}
− {α0,1(u2 − ũ) + α1,0z2}∗H1ρ
(
|z2|−2κ, |z2|2κ
)
× {α0,1(u2 − ũ) + α1,0z2}+O(δ).
The key fact is that the α∗00H1ρ
(
|z1|−2κ, |z1|2κ
)
α00 terms cancel out (|z1| = |z2|).
From ‖α1,0(x(ũ, 0))‖ ≤ (2λ)2
(1−2κ0)δ
−1
1 ‖α0(x(ũ, 0))‖ and ‖α0,1(x(ũ, 0))‖ ≤ 2λδ−1
1 ‖α0(x(ũ, 0))‖
(from (6.3)) δ1 = δ0 − |u| = δ1/2 − δ = δ1/2(1 − δ1/2). Thus the sum of the first order terms
in K
(
η(1)
)
−K
(
η(3)
)
is bounded by c′′′‖α0(x(ũ, 0))‖2δ1/2−2|κ|(1− δ1/2
)−1‖H1‖, where the con-
stant c′′′ is independent of x(ũ, 0) (but is dependent on κ0 and N). Note |u1 − ũ| = |u2 − ũ| =
|z1| = |z2| = δ. The second last step is to relate ‖α0(x(ũ, 0))‖ to ‖L1
(
η(1)
)
‖; indeed
L1
(
η(1)
)
=
N∏
j=1
x
(1)
j
−γκ ρ(z−κ1 , zκ1
){
α0(x(ũ, 0)) +
∞∑
n=1
αn(x(ũ, 0))zn1
}
.
Similarly to (7.5)
∞∑
n=1
‖αn(x(ũ, 0))‖|z|n ≤ c′‖α0(x(ũ, 0))‖
(
|z|
δ0
)(
1− |z|
δ0
)−2λ−1
= c′‖α0(x(ũ, 0))‖δ1/2
(
1− δ1/2
)−2λ−1 ≤ c′′‖α0(x(ũ, 0))‖δ1/2,
(if δ < 1
9 then 1− δ1/2 > 2
3); thus
‖α0(x(ũ, 0))‖
(
1− c′′δ1/2
)
≤ δ−|κ|
∥∥L1
(
η(1)
)∥∥.
By Theorem 6.1
∥∥L(η(1)
)∥∥ ≤ c ∏
1≤i<j≤N−1
|xi − xj |−|κ|
N−2∏
j=1
∣∣eiθj − ei(θN−1+δ′)
∣∣−|κ|∣∣eiθN−1 − ei(θN−1+δ′)
∣∣−|κ|
≤ cδ−|κ|{(N+1)(N−2)/2+1},
because the first two groups of terms satisfy the bound |xi − xj | ≥ δ1/2. Combining everything
we obtain the bound∥∥K(η(1)
)
−K
(
η(3)
)∥∥ ≤ c′′′‖α0(x(ũ, 0))‖2δ1/2−2|κ|(1− δ1/2
)−1‖B1‖
≤ c′′‖H1‖δ1/2−2|κ|−|κ|{(N+1)(N−2)+2}.
A System of Differential Equations on the Torus 27
The constant is independent of η(1) and the exponent on δ is 1
2 − |κ|
(
N2 − N + 2
)
. Thus the
integral of part (3) goes to zero as δ → 0 if |κ| <
(
2
(
N2 − N + 2
))−1
. This is a crude bound,
considering that we know everything works for −1/hτ < κ < 1/hτ , but as we will see, an open
interval of κ values suffices.
Theorem 7.1. If there exists a Hermitian matrix H such that
υH = Hυ and (L1(x0)∗)−1HL1(x0)−1
commutes with σ, and −
(
2
(
N2 −N + 2
))−1
< κ <
(
2
(
N2 −N + 2
))−1
then∫
TN
{(xiDif(x))∗L(x)∗HL(x)g(x)− f(x)∗L(x)∗HL(x)xiDig(x)}dm(x) = 0
for f, g ∈ C(1)
(
TN ;Vτ
)
and 1 ≤ i ≤ N .
It is important that we can derive uniqueness of H from the relation, because the conditions
〈wf,wg〉 = 〈f, g〉, 〈xif, xig〉 = 〈f, g〉, and 〈xiDif, g〉 = 〈f, xiDig〉 for w ∈ SN and 1 ≤ i ≤ N
determine the Hermitian form uniquely up to multiplication by a constant. Thus the measure
K(x)dm(x) is similarly determined, by the density of Laurent polynomials.
8 The orthogonality measure on the torus
At this point there are two logical threads in the development. On the one hand there is
a sufficient condition implying the desired orthogonality measure is of the form L∗HLdm,
specifically if H commutes with υ, (L1(x0)∗)−1HL1(x0)−1 commutes with σ, and |κ| < (2(N2−
N + 2))−1. However we have not yet proven that H exists. On the other hand in [3] we
showed that there does exist an orthogonality measure of the form dµ = dµS +L∗HLdm where
sptµS ⊂ TN\TNreg, H commutes with υ, and −1/hτ < κ < 1/hτ (the support of a Baire
measure ν, denoted by spt ν, is the smallest compact set whose complement has ν-measure
zero). In the next sections we will show that (L1(x0)∗)−1HL1(x0)−1 commutes with σ and
that H is an analytic function of κ in a complex neighborhood of this interval. Combined with
the above sufficient condition this is enough to show that there is no singular part, that is,
µS = 0. The proof involves the formal differential equation satisfied by the Fourier–Stieltjes
series of µ, which is used to show µS = 0 on
{
x ∈ TN : #{xj}Nj=1 = N − 1
}
(that is, x has at
least N − 1 distinct components). In turn this implies (L∗1(x0))−1HL1(x0)−1 commutes with σ.
The proofs unfortunately are not short. In the sequel H refers to the Hermitian matrix in
the formula for dµ and K denotes L∗HL. Also H is positive-definite since the measure µ
is positive (else there exists a vector v with Hv = 0 and then the C(1)
(
TNreg;Vτ
)
function
given by f(x) := L(x)−1vg(x) where g is a smooth scalar nonnegative function with support
in a sufficiently small neighborhood of x0, has norm 〈f, f〉 = 0, a contradiction). Thus H
has a positive-definite square root C which commutes with υ. Now extend CL(x) from C0 to
all of TNreg by Definition 3.6 and so K(x) = L∗(x)C∗CL(x) for all x ∈ TNreg (this follows from
K(xw) = τ(w)−1K(x)τ(w)).
Furthermore
∫
TN ‖K(x)‖dm(x) < ∞ because Kdm is the absolutely continuous part of the
finite Baire measure µ.
We will show that (L∗1(x0))−1C∗CL1(x0)−1 commutes with σ. The proof begins by establish-
ing a recurrence relation for the Fourier coefficients of K(x), which comes from equation (3.3).
For F (x) integrable on TN , possibly matrix-valued, and α ∈ ZN let F̂α =
∫
TN F (x)x−αdm(x).
Clearly
∫
TN x
βF (x)x−αdm(x) = F̂α−β; and if ∂iF (x) is also integrable then (integration-by-
parts)∫
TN
xi∂iF (x)x−αdm(x) = αi
∫
TN
F (x)x−αdm(x). (8.1)
28 C.F. Dunkl
For a subset J ⊂ {1, 2, . . . , N} let εJ ∈ NN0 be defined by (εJ)i = 1 if i ∈ J and = 0 otherwise;
also εi := ε{i}. For 1 ≤ i ≤ N let
Ei := {1, 2, . . . , N}\{i}, Eij := Ei\{j},
pi(x) :=
∏
j 6=i
(xi − xj) =
N−1∑
`=0
(−1)`xN−1−`
i
∑
J⊂Ei,#J=`
xεJ .
Equation (3.3) can be rewritten as
pi(x)xi∂iK(x) = κ
∑
j 6=i
∏
` 6=i,j
(xi − x`){xjτ((i, j))K(x) +K(x)τ((i, j))xi}; (8.2)
this is a polynomial relation which shows that pi(x)xi∂iK(x) is integrable and which has impli-
cations for the Fourier coefficients of K.
Proposition 8.1. For 1 ≤ i ≤ N and α ∈ ZN the Fourier coefficients K̂ satisfy
N−1∑
`=0
(−1)`(αi + `)
∑
J⊂Ei,#J=`
K̂α+`εi−εJ
= κ
∑
j 6=i
N−2∑
`=0
(−1)`
∑
J⊂Eij ,#J=`
{
τ((i, j))K̂α+`εi−εj−εJ + K̂α+(l+1)εi−εJ τ((i, j))
}
. (8.3)
Proof. Multiply both sides of (8.2) by x1−N
i ; this makes the terms homogeneous of degree zero.
Suppose j 6= i then
x1−N
i
∏
6̀=i,j
(xi − x`) =
∏
` 6=i,j
(
1− x`
xi
)
=
N−2∑
`=0
(−1)`x−`i
∑
J⊂Eij ,#J=`
xεJ .
Multiply the right side by x−αdm(x) and integrate over TN to obtain
κ
∑
j 6=i
N−2∑
`=0
(−1)`
∑
J⊂Eij ,#J=`
{
τ((i, j))K̂α+`εi−εj−εJ + K̂α+(l+1)εi−εJ τ((i, j))
}
.
The sum is zero unless α ∈ ZN where ZN :=
{
α ∈ ZN :
N∑
j=1
αj = 0
}
, by the homogeneity. For
the left side start with (8.1) applied to x1−N
i pi(x)xi∂iK(x)
(αi +N − 1)
∫
TN
pi(x)K(x)x1−N
i x−αdm(x)
=
∫
TN
{
(xi∂ipi(x))K(x) + pi(x)(xi∂iK(x))
}
x1−N
i x−αdm(x),∫
TN
pi(x)(xi∂iK(x))x1−N
i x−αdm(x)
=
∫
TN
((αi +N − 1)pi(x)− xi∂ipi(x))K(x)x1−N
i x−αdm(x)
=
∫
TN
N−1∑
`=0
(−1)`(αi + `)x−`i
∑
J⊂Ei,#J=`
xεJK(x)x−αdm(x)
A System of Differential Equations on the Torus 29
=
N−1∑
`=0
(−1)`(αi + `)
∑
J⊂Ei,#J=`
K̂α+`εi−εJ .
Combining the two sides finishes the proof. If α /∈ ZN then both sides are trivially zero. �
This system of recurrences has the easy (and quite undesirable) solution K̂α = I for all
α ∈ ZN and 0 otherwise. The right side becomes 2κ
∑
j 6=i
τ((i, j))
N−2∑̀
=0
(−1)`
(
N−2
`
)
= 0 (for N ≥ 3,
an underlying assumption), and the left side is
N−1∑̀
=0
(−1)`(αi+`)
(
N−1
`
)
I = 0. This K̂ corresponds
to the measure 1
2πdθ on the circle
{
eiθ(1, . . . , 1) : − π < θ ≤ π
}
. Next we show that µ̂α :=∫
TN x
−αdµ(x) satisfies the same recurrences. Proposition 5.2 of [3] asserts that if α, β ∈ NN0 and
N∑
j=1
(αj − βj) = 0 then
(αi − βi)µ̂α−β = κ
∑
αj>αi
αj−αi∑
`=1
τ((i, j))µ̂α+`(εi−εj)−β
− κ
∑
αi>αj
αi−αj−1∑
`=0
τ((i, j))µ̂α+`(εj−εi)−β − κ
∑
βj>βi
βj−βi∑
`=1
µ̂α−`(εi−εj)−βτ((i, j))
+ κ
∑
βi>βj
βi−βj−1∑
`=0
µ̂α−`(εj−εi)−βτ((i, j)). (8.4)
The relation τ(w)∗µ̂wατ(w) = µ̂α is shown in [3, Theorem 4.4]. Introduce Laurent series∑
α∈ZN
B
(i,j)
α xα (i 6= j) satisfying
B(i,j)
α −B(i,j)
α+εi−εj = µ̂α, B
(i,j)
α−αj(εj−εi) = 0,
note
α− αj(εj − εi) =
(
. . . ,
i
αi + αj , . . . ,
j
0, . . . ,
`
α`, . . .
)
, ` 6= i, j.
The purpose of the definition is to produce a formal Laurent series satisfying(
1− xj
xi
)∑
α
B(i,j)
α xα =
∑
α
µ̂αx
α.
The ambiguity in the solution is removed by the second condition (note that
∑
α
(
B
(i,j)
α − cI
)
xα
also solves the first equation for any constant c).
Proposition 8.2. Suppose i 6= j and α ∈ ZN then B
(i,j)
α τ((i, j)) = τ((i, j))B
(j,i)
(i,j)α.
Proof. Start with µ̂ατ((i, j)) = τ((i, j))µ̂(i,j)α and the defining relations
B(i,j)
α τ((i, j))−B(i,j)
α+εi−εjτ((i, j)) = µ̂ατ((i, j)),
τ((i, j))B
(j,i)
(i,j)α − τ((i, j))B
(j,i)
(i,j)α+εj−εi = τ((i, j))µ̂(i,j)α;
30 C.F. Dunkl
subtract the second equation from the first:
B(i,j)
α τ((i, j))− τ((i, j))B
(j,i)
(i,j)α = B
(i,j)
α+εi−εjτ((i, j))− τ((i, j))B
(j,i)
(i,j)α+εj−εi .
By two-sided induction
B(i,j)
α τ((i, j))− τ((i, j))B
(j,i)
(i,j)α = B
(i,j)
α+s(εi−εj)τ((i, j))− τ((i, j))B
(j,i)
(i,j)α+s(εj−εi)
for all s ∈ Z, in particular for s = αj where the right hand side vanishes by definition. �
Theorem 8.3. For γ ∈ ZN and 1 ≤ i ≤ N
γiµ̂γ = κ
∑
j 6=i
{
−τ((i, j))B(j,i)
γ +B(i,j)
γ τ((i, j))
}
. (8.5)
Proof. The proof involves a number of cases (for each (i, j) whether γi ≥ 0 or γi < 0, γj ≥ 0
or γj < 0). Consider equation (8.4), in the terms on the first line (with τ((i, j)) acting on the
left) use the substitution µ̂δ = B
(j,i)
δ − B(j,i)
δ−εi+εj , and for the terms on the second line (with
τ((i, j)) acting on the right) use the substitution µ̂δ = B
(i,j)
δ − B(i,j)
δ+εi−εj . Set α` = max(γ`, 0)
and β` = max(0,−γ`) for 1 ≤ ` ≤ N , thus γ = α−β. The left hand side is (αi−βi)µ̂α−β = γiµ̂γ .
We consider two possibilities separately: (i) αi ≥ 0, βi = 0; (ii) αi = 0, βi > 0; and describe
the typical τ((i, j)) terms. The sums over ` telescope. In the following any term of the form
τ((i, j))µ̂· or µ̂·τ((i, j)) not mentioned explicitly is zero. Proposition 8.2 is used in each case.
For case (i) and αj > αi
τ((i, j))
αj−αi∑
`=1
µ̂α+`(εi−εj)−β = τ((i, j))
αj−αi∑
`=1
(
B
(j,i)
γ+`(εi−εj) −B
(j,i)
γ+`(εi−εj)−εi+εj
)
= τ((i, j))
αj−αi∑
`=1
(
B
(j,i)
γ+`(εi−εj) −B
(j,i)
γ+(`−1)(εi−εj)
)
= τ((i, j))
(
B
(j,i)
γ+(αj−αi)(εi−εj) −B
(j,i)
γ
)
= τ((i, j))
(
B
(j,i)
(i,j)γ −B
(j,i)
γ
)
= −τ((i, j))B(j,i)
γ +B(i,j)
γ τ((i, j))
For case (i) and αi > αj ≥ 0 = βj
−τ((i, j))
αi−αj−1∑
`=0
µ̂α+`(εj−εi)−β = −τ((i, j))
αi−αj−1∑
`=0
(
B
(j,i)
γ+`(εj−εi) −B
(j,i)
γ+(`+1)(εi−εj)
)
= −τ((i, j))
(
B(j,i)
γ −B(j,i)
(i,j)γ
)
= −τ((i, j))B(j,i)
γ +B(i,j)
γ τ((i, j)),
note γ + (αi − αj)(εj − εi) = (i, j)γ. For case (i) and αi > αj = 0 > −βj
−τ((i, j))
αi−1∑
`=0
µ̂α+`(εj−εi)−β = −τ((i, j))
αi−1∑
`=0
(
B
(j,i)
γ+`(εj−εi) −B
(j,i)
γ+(`+1)(εj−εi)
)
= −τ((i, j))
(
B(j,i)
γ −B(j,i)
γ+γi(εj−εi)
)
,
−
βj∑
`=1
µ̂α−`(εi−εj)−βτ((i, j)) = −
βj∑
`=1
(
B
(i,j)
γ−`(εi−εj) −B
(i,j)
γ−`(εi−εj)+εi−εj
)
τ((i, j))
A System of Differential Equations on the Torus 31
= −
βj∑
`=1
(
B
(i,j)
γ−`(εi−εj) −B
(i,j)
γ−(`−1)(εi−εj)
)
τ((i, j))
=
(
B(i,j)
γ −B(i,j)
γ+γj(εi−εj)
)
τ((i, j))
let δ = γ+γi(εj−εi) then δk = γk for k 6= i, j, δi = 0, and δj = γi+γj ; also (i, j)δ = γ+γj(εi−εj).
Thus the sum of the terms for this case is
−τ((i, j))B(j,i)
γ +B(i,j)
γ τ((i, j)) + τ((i, j))B
(j,i)
δ +B
(i,j)
(i,j)δτ((i, j))
= −τ((i, j))B(j,i)
γ +B(i,j)
γ τ((i, j)).
For case (ii) and βj = −γj > βi = −γi > 0
−
βj−βi∑
`=1
µ̂α−`(εi−εj)−βτ((i, j)) = −
βj−βi∑
`=1
(
B
(i,j)
γ−`(εi−εj) −B
(i,j)
γ−(`−1)(εi−εj)
)
τ((i, j))
=
(
−B(i,j)
γ−(γi−γj)(εi−εj) +B(i,j)
γ
)
τ((i, j))
=
(
−B(i,j)
(i,j)γ +B(i,j)
γ
)
τ((i, j))
= −τ((i, j))B(j,i)
γ +B(i,j)
γ τ((i, j)).
For case (ii) and βi > βj = −γj ≥ 0 (and αj = 0)
βi−βj−1∑
`=0
µ̂α−`(εj−εi)−βτ((i, j)) =
βi−βj−1∑
`=0
(
B
(i,j)
γ−`(εj−εi) −B
(i,j)
γ−(`+1)(εj−εi)
)
τ((i, j))
=
(
B(i,j)
γ −B(i,j)
γ−(γj−γi)(εj−εi)
)
τ((i, j))
=
(
B(i,j)
γ −B(i,j)
(i,j)γ
)
τ((i, j))
= −τ((i, j))B(j,i)
γ +B(i,j)
γ τ((i, j)).
For case (ii) and −βi = γi < 0 < γj = αj (and βj = 0)
τ((i, j))
αj∑
`=1
µ̂α+`(εi−εj)−β +
βi−1∑
`=0
µ̂α−`(εj−εi)−βτ((i, j))
= τ((i, j))
αj∑
`=1
(
B
(j,i)
γ+`(εi−εj) −B
(j,i)
γ+(`−1)(εi−εj)
)
+
βi−1∑
`=0
(
B
(i,j)
γ−`(εj−εi) −B
(i,j)
γ−(`+1)(εj−εi)
)
τ((i, j))
= τ((i, j))
(
B
(j,i)
γ+γj(εi−εj) −B
(j,i)
γ
)
+
(
B(i,j)
γ −B(i,j)
γ+γi(εj−εi)
)
τ((i, j))
= −τ((i, j))B(j,i)
γ +B(i,j)
γ τ((i, j)),
because (i, j)(γ+γj(εi−εj)) = γ+γi(εj−εi). In the trivial case γi = γj so that (i, j)γ = γ where
are no nonzero τ((i, j)) terms the equation −τ((i, j))B
(j,i)
γ − B(i,j)
γ τ((i, j)) = 0 applies. Thus
in each case and for each j 6= i the right hand side contains the expression −κ
(
τ((i, j))B
(j,i)
γ −
B
(i,j)
γ τ((i, j))
)
. �
In the following there is no implied claim about convergence, because any term xα appears
only a finite number of times in the equation.
32 C.F. Dunkl
Theorem 8.4. For 1 ≤ i ≤ N the formal Laurent series F (x) :=
∑
α∈ZN
µ̂αx
α satisfies the
equation
pi(x)xi∂iF (x) = κ
∑
j 6=i
∏
` 6=i,j
(xi − x`){xjτ((i, j))F (x) + F (x)τ((i, j))xi}. (8.6)
Proof. Start with multiplying equation (8.5) by x1−N
i pi(x)xγ and sum over γ ∈ ZN to obtain
N∏
j=1, j 6=i
(
1− xj
xi
) ∑
γ∈ZN
γiµ̂γx
γ = κ
∑
j 6=i
∏
k 6=i,j
(
1− xk
xi
)(
1− xj
xi
)
×
−τ((i, j))
∑
γ∈ZN
B(j,i)
γ xγ +
∑
γ∈ZN
B(i,j)
γ xγτ((i, j))
.
By construction(
1− xj
xi
) ∑
γ∈ZN
B(i,j)
γ xγ =
∑
γ∈ZN
(
B(i,j)
γ −B(i,j)
γ+εi−εj
)
xγ =
∑
γ∈ZN
µ̂γx
γ
and (
1− xj
xi
) ∑
γ∈ZN
B(j,i)
γ xγ = −xj
xi
(
1− xi
xj
) ∑
γ∈ZN
B(j,i)
γ xγ = −xj
xi
∑
γ∈ZN
µ̂γx
γ .
Thus the equation becomes
N∏
j=1, j 6=i
(
1− xj
xi
) ∑
γ∈ZN
γiµ̂γx
γ
= κ
∑
j 6=i
∏
k 6=i,j
(
1− xk
xi
)xjxi τ((i, j))
∑
γ∈ZN
µ̂γx
γ +
∑
γ∈ZN
µ̂γx
γτ((i, j))
.
This completes the proof. �
Corollary 8.5. The coefficients {µ̂α} satisfy the same recurrences as
{
K̂α
}
in (8.3).
8.1 Maximal singular support
Above we showed that µ and K satisfy the same Laurent series differential systems (8.2)
and (8.6), thus the singular part µS also satisfies this relation. The singular part µS is the
restriction of µ to
⋃
i<j
{
x ∈ TN : xi = xj
}
, a closed set. For each pair {k, `} let Ek` =
{
x ∈
TN : xk 6= x`
}
, an open subset of TN . For i 6= j let
Ti,j =
{
x ∈ TN : xi = xj
}
∩
⋂
{k,`}∩{i,j}=∅
{Ek` ∩ Eik ∩ Ejk};
this is an intersection of a closed set and an open set, hence Ti,j is a Baire set and the restric-
tion µi,j of µ to Ti,j is a Baire measure. Informally Ti,j =
{
x ∈ TN : xi = xj ,#{xk} = N − 1
}
.
We will prove that µi,j = 0 for all i 6= j. That is, µS is supported by
{
x ∈ TN : #{xk} ≤ N −2
}
A System of Differential Equations on the Torus 33
(the number of distinct coordinate values is ≤ N − 2). In [3, Corollary 4.15] there is an approx-
imate identity
σN−1
n (x) :=
n∑
k=0
(−n)k
(1− n−N)k
∑
α∈ZN , |α|=2k
xα,
which satisfies σN−1
n (x) ≥ 0 and σN−1
n ∗ν → ν as n→∞, in the weak-∗ sense for any finite Baire
measure ν on TN/D (referring to functions and measures on TN homogeneous of degree zero as
Laurent series). The set Ti,j is pointwise invariant under (i, j) thus dµi,j(x) = dµi,j(x(i, j)) =
τ((i, j))dµi,j(x)τ((i, j)).
Remark 8.6. The density of Laurent polynomials in C(1)
(
TN
)
can be shown by using an
approximate identity, for example: un(x) =
{
1
n+1
n∑
j=−n
(n− |j|+ 1)xj1
}
σN−1
n (x); for any α ∈ ZN
the coefficient of xα in un(x) tends to 1 as n→∞ (express α = (α1−m)ε1 + (−m,α2, . . . , αN )
where m =
N∑
j=2
αj). Then f ∗ un → f in the C(1)
(
TN
)
norm.
Let Ks
n = σN−1
n ∗ µS (convolution), a Laurent polynomial, fix ` in 1 ≤ ` ≤ N , and consider
the functionals F`,n, G`,n on scalar functions p ∈ C(1)
(
TN
)
F`,n(p) :=
∫
TN
p(x)
∏
j 6=`
(
1− xj
x`
)
x`∂`K
s
n(x)dm(x),
G`,n(p) := κ
∑
i 6=`
∫
TN
p(x)
∏
j 6=`,i
(
1− xj
x`
){
xi
x`
τ (`, i)Ks
n(x) +Ks
n(x)τ(`, i)
}
dm(x).
By construction the functionals annihilate xα for α /∈ ZN . For a fixed α ∈ ZN the value
F`,n(x−α)−G`,n(x−α) is
α`Aαbn(α) +
N−1∑
i=1
(−1)i
∑
J⊂E`,#J=i
(α` + i)Aα+iε`−εJ bn(α+ iε` − εJ)
−κ
N∑
j=1, j 6=`
N−2∑
i=0
(−1)`
∑
J⊂E`,j ,#J=i
{
τ(`, j)Aα+(i+1)ε`−εj−εJ bn(α+ (i+ 1)ε` − εj − εJ)
+Aα+iε`−εJ τ(`, j)bn(α+ iε` − εJ)
}
,
where bn(γ) :=
(−n)|γ|/2
(1−N−n)|γ|/2
(from the Laurent series of σN−1
n ), and Aγ :=
∫
TN x
−γdµS . Thus
for fixed α the coefficients bn(·) → 1 as n → ∞ and the expression tends to the differential
system 8.2 and
lim
n→∞
(
F`,n(x−α)−G`,n(x−α)
)
= 0.
This result extends to any Laurent polynomial by linearity. From the approximate identity
property
lim
n→∞
G`,n(p) = κ
∑
i 6=`
∫
TN
p(x)
∏
j 6=`,i
(
1− xj
x`
){
xi
x`
τ(`, i)dµS(x) + dµS(x)τ(`, i)
}
,
and
‖G`,n(p)‖ ≤M sup
x, i
∣∣∣∣∣∣p(x)
∏
j 6=`,i
(
1− xj
x`
)∣∣∣∣∣∣ ,
34 C.F. Dunkl
where M depends on µS . Also
F`,n(p) = −
∫
TN
x`∂`
p(x)
∏
j 6=`
(
1− xj
x`
)Kn(x)dm(x),
and
lim
n→∞
F`,n(p) = −
∫
TN
x`∂`
p(x)
∏
j 6=`
(
1− xj
x`
) dµS(x)
for Laurent polynomials p. By density of Laurent polynomials in C(1)
(
TN/D
)
(D={(u, u, . . . , u) :
|u| = 1} thus functions homogeneous of degree zero on TN can be considered as functions on
the quotient group TN/D) we obtain
−
∫
TN
x`∂`
p(x)
∏
j 6=`
(
1− xj
x`
) dµS(x)
= κ
∑
i 6=`
∫
TN
p(x)
∏
j 6=`,i
(
1− xj
x`
){
xi
x`
τ (`, i) dµS(x) + dµS(x)τ(`, i)
}
, (8.7)
for all p ∈ C(1)
(
TN/D
)
.
Theorem 8.7. For 1 ≤ i < j ≤ N the restriction µS |Ti,j = 0.
Proof. It suffices to take i = 1, j = 2. Let E be an open neighborhood of a point in T1,2
such that if x ∈ E (the closure) and xi = xj for some pair i < j then i = 1 and j = 2. Let
f(x) ∈ C(1)
(
TN/D
)
have support ⊂ E. Thus f(x) = 0 = ∂1f(x) at each point x such that
xi = xj for some pair {i, j} 6= {1, 2} (f = 0 on a neighborhood of
⋃
i<j
{x : xi = xj}\T1,2). Then
in formula (8.7) (with ` = 1) applied to f the measure µS can be replaced with µ1,2. Evaluate
the derivative
x1∂1
f(x)
∏
j 6=1
(
1− xj
x1
) =
(
1− x2
x1
)
f(x)x1∂1
∏
j>2
(
1− xj
x1
)
+ f(x)
x2
x1
∏
j>2
(
1− xj
x1
)
+ (x1∂1f(x))
(
1− x2
x1
)∏
j>2
(
1− xj
x1
)
.
Each term vanishes on
⋃
i<j
{x : xi = xj}\T1,2, and restricted to T1,2 the value is f(x)
∏
j>2
(
1− xj
x1
)
.
Thus
−
∫
TN
x1∂1
f(x)
∏
j 6=1
(
1− xj
x1
)dµS(x) = −
∫
TN
x1∂1
f(x)
∏
j 6=1
(
1− xj
x1
)dµ1,2(x)
= −
∫
TN
f(x)
∏
j>2
(
1− xj
x1
)
dµ1,2(x).
The right hand side of the formula reduces to
κ
∫
TN
f(x)
∏
j>2
(
1− xj
x1
){
x2
x1
τ(1, 2)dµ1,2(x) + dµ1,2(x)τ(1, 2)
}
A System of Differential Equations on the Torus 35
= 2κτ(1, 2)
∫
TN
f(x)
∏
j>2
(
1− xj
x1
)
dµ1,2(x),
since dµ1,2(x)τ(1, 2) = τ(1, 2)dµ1,2(x). Thus the integral is a matrix F (f) such that
(I + 2κτ(1, 2))F (f) = 0,
which implies F (f) = 0 provided κ 6= ±1
2 . Replacing f(x) by f(x)
∏
j>2
(
1 − xj
x1
)−1
shows that
µ1,2 = 0, since E was arbitrarily chosen. �
8.2 Boundary values for the measure
In this subsection we will show that K satisfies the weak continuity condition
lim
xN−1−xN→0
(K(x)−K(x(N − 1, N))) = 0
at the faces of C0 and then deduce that H1 commutes with σ (as described in Theorem 7.1).
The idea is to use the inner product property of µ on functions supported in a small enough
neighborhood of x(0) =
(
1, ω, . . . , ωN−3, ω−3/2, ω−3/2
)
where µS vanishes, so that only K is
involved, then argue that a failure of the continuity condition leads to a contradiction.
Let 0 < δ ≤ 2π
3N and define the boxes
Ωδ =
{
x ∈ TN :
∣∣xj − x(0)
j
∣∣ ≤ 2 sin δ
2 , 1 ≤ j ≤ N
}
,
Ω′δ =
{
x ∈ TN−1 :
∣∣xj − x(0)
j
∣∣ ≤ 2 sin δ
2 , 1 ≤ j ≤ N − 1
}
(so if xj = eiθj then
∣∣θj − 2π(j−1)
N
∣∣ ≤ δ, for 1 ≤ j ≤ N −2 and
∣∣θj − (2N−3)π
2
∣∣ for N −1 ≤ j ≤ N).
Then x ∈ Ωδ implies |xi − xj | ≥ 2 sin δ
2 for 1 ≤ i < j ≤ N except for i = N − 1, j = N
(that is, |θi − θj | ≥ δ). Further Ωδ is invariant under (N − 1, N), while Ωδ ∩ Ωδ(i,N) = ∅ for
1 ≤ i ≤ N − 2. For brevity set φ0 = (2N−3)π
2 , eiφ0 = ω−3/2. We consider the identity∫
TN
(xNDNf(x))∗dµ(x)g(x)−
∫
TN
f(x)∗dµ(x)xNDNg(x) = 0
for f, g ∈ C(1)
(
TN ;Vτ
)
whose support is contained in Ωδ. Then spt((xNDNf(x))∗g(x)) ⊂ Ωδ
and spt(f(x)∗xNDNg(x)) ⊂ Ωδ.
The support hypothesis and the construction of Ωδ imply that Ωδ∩
(
TN\TNreg
)
⊂ TN−1,N and
thus dµ can be replaced by K(x)dm(x) in the formula. Recall the general identity (4.1)
−(xNDNf(x))∗K(x)g(x) + f(x)∗K(x)xNDNg(x)
= xN∂N{f(x)∗K(x)g(x)} − κ
∑
1≤j≤N−1
1
xN − xj
{
xjf(x(j,N))∗τ((j,N))K(x)g(x)
+ xNf(x)∗K(x)τ((j,N))g(x(j,N))
}
.
Specialize to spt(f) ⊂ Ωδ and spt(g) ⊂ Ωδ and x ∈ Ωδ then only the j = N − 1 term in the
sum remains, and this term changes sign under x 7→ x(N − 1, N).
For ε > 0 let Ωδ,ε =
{
x ∈ Ωδ : |xN−1 − xN | ≥ 2 sin ε
2
}
, then∫
Ωδ,ε
{
xN∂N (f(x)∗K(x)g(x))
+ (xNDNf(x))∗K(x)g(x)− f(x)∗K(x)xNDNg(x)
}
dm(x) = 0,
36 C.F. Dunkl
because Ωδ,ε is (N − 1, N)-invariant (similar argument to Proposition 4.1). By integrability
lim
ε→0+
∫
Ωδ,ε
{
(xNDNf(x))∗K(x)g(x)− f(x)∗K(x)xNDNg(x)dm(x)
}
= 0,
hence
lim
ε→0+
∫
Ωδ,ε
xN∂N (f(x)∗K(x)g(x))dm(x) = 0.
Now we use iterated integration. For fixed θN−1 the ranges for θN are obtained by inserting
suitable gaps into the interval [φ0 − δ, φ0 + δ] (as usual, x =
(
eiθ1 , . . . , eiθN
)
):
1) φ0 − δ ≤ θN−1 ≤ φ0 − δ + ε : [θN−1 + ε, φ0 + δ],
2) φ0 − δ + ε ≤ θN−1 ≤ φ0 + δ − ε : [φ0 − δ, θN−1 − ε] ∪ [θN−1 + ε, φ0 + δ],
3) φ0 + δ − ε ≤ θN−1 ≤ φ0 + δ : [φ0 − δ, θN−1 − ε].
From xN∂N = −i ∂
∂θN
it follows that
1
2π
∫ b
a
xN∂N (f∗Kg)
((
eiθ1 , . . . , eiθN
))
dθN
=
1
2πi
{
(f∗Kg)
((
eiθ1 , . . . , eiθN−1 , eib
))
− (f∗Kg)
((
eiθ1 , . . . , eiθN−1 , eia
))}
.
Since f and g are at our disposal we can take their supports contained in Ωδ/2 then for 0 < ε ≤ δ
4
the dθN -integrals for (1) and (3) vanish and the integrals in (2) have the value
1
2πi
{
(f∗Kg)
((
eiθ1 , . . . , eiθN−1 , ei(θN−1−ε)
))
− (f∗Kg)
((
eiθ1 , . . . , eiθN−1 , ei(θN−1+ε)
))}
.
We use the power series (from (5.1))
L1(x(u, z)) =
(u2 − z2
)N−2∏
j=1
xj
−γκ ρ(z−κ, zκ) ∞∑
n=0
αn(x(u, 0))zn,
with the notation x(u, z) = (x1, . . . , xN−2, u− z, u+ z) for x ∈ Ωδ. Recall αn(x(u, 0)) is analytic
for a region including Ωδ and σαn(x(u, 0))σ = (−1)nαn(x(u, 0)). Also α0(x(u, 0)) is invertible.
As in Section 5 define C1 := CL1(x0)−1 so that L1(x)∗C∗1C1L1(x) = L(x)∗HL(x) on their
common domain, and set H1 = C∗1C1. It suffices to use the approximation
∞∑
n=0
αn(x(u, 0))zn =
α0(x(u, 0)) +O(|z|), uniformly in Ω2π/3N .
Let
η(1)(x, θ, ε) :=
(
x1, . . . , xN−2, e
iθ, ei(θ+ε)
)
,
η(2)(x, θ, ε) :=
(
x1, . . . , xN−2, e
i(θ−ε), eiθ
)
,
η(3)(x, θ, ε) :=
(
x1, . . . , xN−2, e
iθ, ei(θ−ε))
with η(1), η(2) ∈ Ωδ ∩ C0 and η(3) = η(2)(N − 1, N). Set ζ = eiε. Then
η(1) = x(u1 − z1, u1 + z1), u1 =
1
2
eiθ(1 + ζ), z1 =
1
2
eiθ(ζ − 1),
η(2) = x(u2 − z2, u2 + z2), u2 =
1
2
eiθ
(
1 + ζ−1
)
, z2 =
1
2
eiθ
(
1− ζ−1
)
= ζ−1z1.
A System of Differential Equations on the Torus 37
The invariance properties of K imply K
(
η(3)
)
= σK
(
η(2)
)
σ. Then
K
(
η(1)
)
= α0(x(u1, 0))∗ρ
(
z−κ1 , zκ1
)∗
H1ρ
(
z−κ1 , zκ1
)
α0(x(u1, 0)) +O
(
|z1|1−2|κ|),
σK
(
η(2)
)
σ = α0(x(u2, 0))∗ρ
(
z−κ2 , zκ2
)∗
σH1σρ
(
z−κ2 , zκ2
)
α0(x(u2, 0)) +O
(
|z2|1−2|κ|),
because σα0(x(u, 0))σ = α0(x(u, 0)) and σ = ρ(−1, 1) commutes with ρ(z−κ2 , zκ2 ). To express
K
(
η(1)
)
− σK
(
η(2)
)
σ let
A1 = ρ
(
z−κ1 , zκ1
)∗
H1ρ
(
z−κ1 , zκ1
)
= O
(
|z1|−2|κ|),
A2 = ρ
(
z−κ2 , zκ2
)∗
σH1σρ
(
z−κ2 , zκ2
)
= O
(
|z2|−2|κ|),
then
K
(
η(1)
)
− σK
(
η(2)
)
σ
= α0(x(u1, 0))∗A1α0(x(u1, 0))− α0(x(u2, 0))∗A2α0(x(u2, 0)) +O
(
|z1|1−2|κ|).
Also u2−u1 = 1
2xN−1ξ1
(
ζ−1−ζ
)
= O(|z1|) (since |z1| = |1−ζ|) thus α0(x(u1, 0))−α0(x(u2, 0)) =
O(|z1|) and
K
(
η(1)
)
− σK
(
η(2)
)
σ = α0(x(u1, 0))∗(A1 −A2)α0(x(u1, 0)) +O
(
|z1|1−2|κ|).
Using the σ-decomposition write
H1 :=
[
H11 H12
H12
∗ H11
]
,
then
A1 −A2 =
H11|z1|−2κ H12
(
z1
z1
)κ
H12
∗
(
z1
z1
)−κ
H11|z1|2κ
−
H11|z2|−2κ −H12
(
z2
z2
)κ
−H12
∗
(
z2
z2
)−κ
H11|z2|2κ
=
O H12
{(
z1
z1
)κ
+
(
z2
z2
)κ}
H12
∗
{(
z1
z1
)−κ
+
(
z2
z2
)−κ}
O
,
because |z2| = |z1|. Next
z1
z1
= e2iθ e
iε − 1
e−iε − 1
= −ei(2θ+ε),
z2
z2
= e2iθ 1− e−iε
1− eiε
= −ei(2θ−ε),(
z1
z1
)κ
+
(
z2
z2
)κ
=
(
−e2iθ
)κ{
eiεκ + e−iεκ
}
= 2
(
−e2iθ
)κ
cos εκ,
where some branch of the power function is used (the interval where it is applied is a small arc
of the unit circle), and φ0 − δ1 < θ < φ0 − δ1.
We show that H12 = O, equivalently H1 commutes with σ. By way of contradiction suppose
some entry hij 6= 0 (1 ≤ i ≤ mτ < j ≤ nτ ). There exist r > 0, δ1 ≥ δ2 > 0 and c ∈ C with
|c| = 1 such that
Re
(
2c
(
−e2iθ
)κ
hij
)
> r
38 C.F. Dunkl
for φ0 − δ2 ≤ θ ≤ φ0 + δ2. Let p(x) ∈ C(1)
(
TN
)
such that spt(p) ⊂ Ωδ2/2, 0 ≤ p(x) ≤ 1 and
p(x) = 1 for x ∈ Ωδ2/4. Let f(x) = p(x)α0(x(u, 0))−1εi and g(x) = cp(x)α0(x(u, 0))−1εj (for
x ∈ Ωδ2/2). Also impose the bound 0 < ε < δ2
4 . Then
f
(
η(1)
)∗
K
(
η(1)
)
g
(
η(1)
)
= p
(
η(1)
)2
c
(
z1
z1
)κ
hij +O
(
|z1|1−2|κ|),
f
(
η(3)
)∗
σK
(
η(2)
)
σg
(
η(3)
)
= −p
(
η(3)
)2
c
(
z2
z2
)κ
hij +O
(
|z2|1−2|κ|),
Suppose x =
(
eiθ1 , . . . , eiθN
)
∈ Ωδ2/2 then p(x) = 1 for φN−1− δ2
4 ≤ θN−1, θN ≤ φN−1− δ2
4 , thus
p
(
η(1)(x, θ, ε)
)
= 1 for φ0− δ2
4 ≤ θ ≤ φ0− δ2
4 −ε and p
(
η(3)(x, θ, ε)
)
= 1 for φ0− δ2
4 +ε ≤ θ ≤ φ0− δ2
4 .
By the continuous differentiability it follows that for φ0− δ2
4 ≤ θ ≤ φ0− δ2
4 both p
(
η(1)
)
= 1+O(ε)
and p
(
η(3)
)
= 1 +O(ε). Thus
p
(
η(1)
)
f
(
η(1)
)∗
K
(
η(1)
)
g
(
η(1)
)
p
(
η(1)
)
− p
(
η(3)
)
f
(
η(3)
)∗
σK
(
η(2)
)
σg
(
η(3)
)
p
(
η(3)
)
= p
(
η(1)
)2
c
{(
z1
z1
)κ
+
(
z2
z2
)κ}
hij +O
(
|z1|1−2|κ|)+O(ε).
By construction
Re
(
c
{(
z1
z1
)κ
+
(
z2
z2
)κ}
hij
)
> r cos εκ,
multiply the inequality by p
((
eiθ1 , . . . , eiθN−1 , eiθN−1
))2
and integrate over the (N − 1)-box Ω′δ2
with respect to dmN−1 =
(
1
2π
)N−1
dθ1 · · · dθN−1. This integral dominates the integral over Ω′δ2/4,
thus
Re
∫
Ω′δ2
p
(
η(1)
)2
c
{(
z1
z1
)κ
+
(
z2
z2
)κ}
hijdmN−1 ≥ r cos εκ
(
δ2
2π
)N−1
.
This contradicts the limit of the integral being zero as ε→ 0. The ignored parts of the integral
are O
(
ε1−2|κ|) and |κ| < 1
2 . We have proven the following:
Theorem 8.8. For −1/hτ < κ < 1/hτ the matrix H1 = (L∗1(x0))−1HL1(x0)−1 commutes
with σ.
9 Analytic matrix arguments
In this section we set up some tools from linear algebra dealing with matrices whose entries are
analytic functions of one variable. The aim is to establish the existence of an analytic solution
for the matrices described in Theorem 8.8. The key fact is that the solution L1(x;κ) of (3.1) is
analytic in κ for |κ| < 1
2 , in fact for κ ∈ C\(Z+ 1
2); the series expansion in (5.1) does not apply to
κ ∈ Z+ 1
2 and a logarithmic term has to be included for this case. Set bN := (2(N2−N + 2))−1,
the bound from Section 7. One would like use analytic continuation to extend the inner product
property of L∗HL from the interval −bN < κ < bN to −1/hτ < κ < 1/hτ but the Bochner
theorem argument for the existence of µ does not allow κ to be a complex variable. The
following arguments work around this obstacle.
Theorem 9.1. Suppose M(κ) is an m×n complex matrix with m ≥ n− 1 such that the entries
are analytic in κ ∈ Dr := {κ ∈ C : |κ| < r}, some r > 0 and rank(M(κ)) = n − 1 for a real
interval −r1 < κ < r1 then rank(M(κ)) = n−1 for all κ ∈ Dr except possibly at isolated points λ
where rank(M(λ)) < n − 1, and there is a nonzero vector function v(κ), analytic on Dr such
that M(κ)v(κ) = 0 and v(κ) is unique up to multiplication by a scalar function.
A System of Differential Equations on the Torus 39
Proof. Let M ′(κ) be any n × n submatrix of M(κ) (when m ≥ n), that is, M ′ is composed
of n rows of M(κ), then detM ′(κ) is analytic for κ ∈ Dr and by hypothesis detM ′(κ) = 0 for
−r1 < κ < r1. This implies detM ′(κ) ≡ 0 for all κ, by analyticity. Thus rank(M(κ)) ≤ n − 1
for all κ ∈ Dr. For each subset J = {j1, . . . , jn−1} with 1 ≤ j1 < · · · < jn−1 ≤ m let MJ,k(κ)
be the (n − 1) × (n − 1) submatrix of M(κ) consisting of rows # j1, . . . , jn−1 and deleting
column #k, and XJ(κ) := [detMJ,1(κ), . . . ,detMJ,n(κ)], an n-vector of analytic functions.
There exists at least one set J for which XJ(0) 6= [0, . . . , 0] otherwise rank(M(0)) < n− 1. By
continuity there exists δ > 0 such that at least one detMJ,k(κ) 6= 0 for |κ| < δ and v(κ) =[
(−1)k−1 detMJ,k(κ)
]n
k=1
is a nonzero vector in the null-space of M(κ) (by Cramer’s rule and
the rank hypothesis). The analytic equation M(κ)v(κ) = 0 holds in a neighborhood of κ = 0
and thus for all of Dr. If v(κ) = 0 for isolated points κ1, . . . , κ` in |κ| ≤ r2 < r then v(κ) can
be multiplied by
∏̀
j=1
(
1 − κ
κj
)−aj for suitable positive integers a1, . . . , a` to produce a solution
never zero in |κ| ≤ r2 < r. (It may be possible that there are infinitely many zeros in the open
set Dr.) �
We include the parameter in the notations for L and L1. The ∗ operation replaces xj by x−1
j ,
the constants by their conjugates, and transposing, but κ is unchanged to preserve the analytic
dependence, see Definition 3.8. For x ∈ TNreg and real κ the Hermitian adjoint of L1(x0;κ))
agrees with L1(x0;κ)∗. The matrix M(κ) is implicitly defined by the linear system with the
unknown B1
B1 = σB1σ,
υL∗1(x0;κ)B1L1(x0;κ) = L∗1(x0;κ)B1L1(x0;κ)υ. (9.1)
(Recall υ = τ(w0).) The entries of M(κ) are analytic in |κ| < 1
2 . The equation B1 = σB1σ
implies that B1 has n := m2
τ+(nτ−mτ )2 possible nonzero entries, by the σ-block decomposition.
The number of equations m = n2
τ −dim{A : Aυ = υA}. Because w0 and (N −1, N) generate SN
and τ is irreducible Aσ = σA and Aυ = υA imply A = cI for c ∈ C by Schur’s lemma.
This implies n ≥ m − 1 (else there are two linearly independent solutions). By a result of
Stembridge [10, Section 3] n can be computed from the following: (recall ω := exp 2πi
N ) for
0 ≤ j ≤ N − 1 set ej equal to the multiplicity of ωj in the list of the nτ eigenvalues of υ and set
Fτ (q) := qe0 + qe1 + · · ·+ qeN−1 then
Fτ (q) =
qn(τ)
N∏
i=1
(
1− qi
) ∏
(i,j)∈τ
(
1− qh(i,j)
)−1
mod
(
1− qN
)
,
where n(τ) :=
`(τ)∑
i=1
(i − 1)τi and h(i, j) is the hook length at (i, j) in the diagram of τ (note
Fτ (1) = nτ ). Thus dim{A : Aυ = υA} =
N−1∑
j=0
e2
j . For example let τ = (4, 2) then nτ = 9, mτ = 3
and n = 45 while Fτ (q) = 2 + q + 2q2 + q3 + 2q4 + q5 and dim{A : Aυ = υA} = 15, m = 66.
Theorem 9.2. For −1/hτ < κ < 1/hτ there exists a unique Hermitian matrix H such that
dµ = L∗HLdm. Also (L1(x0)∗)−1HL1(x0)−1 commutes with σ.
Proof. For any Hermitian nτ × nτ matrix B define the Hermitian form
〈f, g〉B :=
∫
TN
f(x)∗L(x)∗BL(x)g(x)dm(x)
40 C.F. Dunkl
for f, g ∈ C(1)
(
TN ;Vτ
)
. If the form satisfies 〈wf,wg〉B = 〈f, g〉B for all w ∈ SN and 〈xiDif, g〉B
= 〈f, xiDig〉B for 1 ≤ i ≤ N then B is determined up to multiplication by a constant.
This follows from the density of the span of the nonsymmetric Jack (Laurent) polynomials
in C(1)
(
TN ;Vτ
)
. By Theorem 8.8 there exists a nontrivial solution of the system (9.1) for
−1/hτ < κ < 1/hτ . Thus rankM(κ) ≤ n − 1 in this interval. Now suppose that B1 is a non-
trivial solution of (9.1) for some κ such that −bN < κ < bN . Then both B(1) := B1 + B∗1 and
B(2) := i(B1 − B∗1) are also solutions (by the invariance of (9.1) under the adjoint operation).
Let H(i) := L∗1(x0;κ)B(i)L1(x0;κ) for i = 1, 2 then by Theorem 7.1 the forms 〈·, ·〉H(1) and
〈·, ·〉H(2) satisfy the above uniqueness condition. Hence either B1 is Hermitian or B(1) = rB(2)
for some r 6= 0 which implies B1 = 1
2(r − i)B(2), that is, B1 is a scalar multiple of a Hermitian
matrix. Thus there is a unique (up to scalar multiplication) solution of (9.1) which implies
rankM(κ) ≥ n− 1 in −bN < κ < bN .
Hence the hypotheses of Theorem 9.1 are satisfied, and there exists a nontrivial solution B1(κ)
which is analytic in |κ| < 1
2 . Since the Hermitian form is positive definite for −1/hτ < κ < 1/hτ
we can use the fact that B1(κ) is a multiple of a positive-definite matrix when κ is real (in fact,
of the matrix H1 arising from µ as in Section 8) and its trace is nonzero (at least on a complex
neighborhood of {κ : − 1/hτ < κ < 1/hτ} by continuity). Set B′1(κ) :=
(
nt/
nτ∑
i=1
B1(κ)ii
)
B1(κ),
analytic and tr(B′1(κ)) = 1 thus the normalization produces a unique analytic (and Hermitian
for real κ) matrix in the null-space of M(κ). Let H(κ) = L1(x0;κ)∗B1(κ)L1(x0;κ) then for
fixed f, g ∈ C(1)
(
TN ;Vτ
)
and 1 ≤ i ≤ N∫
TN
{
(xiDif(x))∗L∗(x;κ)H(κ)L(x;κ)g(x)
−f(x)∗L∗(x;κ)H(κ)L(x;κ)xiDig(x)
}
dm(x)
is an analytic function of κ which vanishes for −bN < κ < bN hence for all κ in −1/hτ < κ <
1/hτ ; this condition is required for integrability. This completes the proof. �
By very complicated means we have shown that the torus Hermitian form for the vector-
valued Jack polynomials is given by the measure L∗HLdm. The orthogonality measure we
constructed in [3] is absolutely continuous with respect to the Haar measure. We conjecture
that L∗(x;κ)H(κ)L(x;κ) is integrable for −1/τ1 < κ < 1/`(τ) but H(κ) is not positive outside
|κ| < 1/hτ (the length of τ is `(τ) := max{i : τi ≥ 1}). In as yet unpublished work we have found
explicit formulas for L∗HL for the two-dimensional representations (2, 1) and (2, 2) of S3 and S4
respectively, using hypergeometric functions. It would be interesting to find the normalization
constant, that is, determine the scalar multiple of H(κ) which results in 〈1 ⊗ T, 1 ⊗ T 〉H(κ) =
〈T, T 〉0 (see (2.1)) the “initial condition” for the form. In [3, Theorem 4.17(3)] there is an infinite
series for H(κ) but it involves all the Fourier coefficients of µ.
Acknowledgement
Some of these results were presented at the conference “Dunkl operators, special functions and
harmonic analysis” held at Universität Paderborn, Germany, August 8–12, 2016.
References
[1] Beerends R.J., Opdam E.M., Certain hypergeometric series related to the root system BC, Trans. Amer.
Math. Soc. 339 (1993), 581–609.
[2] Dunkl C.F., Symmetric and antisymmetric vector-valued Jack polynomials, Sém. Lothar. Combin. 64 (2010),
Art. B64a, 31 pages, arXiv:1001.4485.
[3] Dunkl C.F., Orthogonality measure on the torus for vector-valued Jack polynomials, SIGMA 12 (2016),
033, 27 pages, arXiv:1511.06721.
https://doi.org/10.2307/2154288
https://doi.org/10.2307/2154288
http://arxiv.org/abs/1001.4485
https://doi.org/10.3842/SIGMA.2016.033
http://arxiv.org/abs/1511.06721
A System of Differential Equations on the Torus 41
[4] Dunkl C.F., Vector-valued Jack polynomials and wavefunctions on the torus, J. Phys. A: Math. Theor. 50
(2017), 245201, 21 pages, arXiv:1702.02109.
[5] Dunkl C.F., Luque J.G., Vector-valued Jack polynomials from scratch, SIGMA 7 (2011), 026, 48 pages,
arXiv:1009.2366.
[6] Felder G., Veselov A.P., Polynomial solutions of the Knizhnik–Zamolodchikov equations and Schur–Weyl
duality, Int. Math. Res. Not. 2007 (2007), rnm046, 21 pages, math.RT/0610383.
[7] Griffeth S., Orthogonal functions generalizing Jack polynomials, Trans. Amer. Math. Soc. 362 (2010),
6131–6157, arXiv:0707.0251.
[8] James G., Kerber A., The representation theory of the symmetric group, Encyclopedia of Mathematics and
its Applications, Vol. 16, Addison-Wesley Publishing Co., Reading, Mass., 1981.
[9] Opdam E.M., Harmonic analysis for certain representations of graded Hecke algebras, Acta Math. 175
(1995), 75–121.
[10] Stembridge J.R., On the eigenvalues of representations of reflection groups and wreath products, Pacific J.
Math. 140 (1989), 353–396.
[11] Stoer J., Bulirsch R., Introduction to numerical analysis, Springer-Verlag, New York – Heidelberg, 1980.
https://doi.org/10.1088/1751-8121/aa6fb9
http://arxiv.org/abs/1702.02109
https://doi.org/10.3842/SIGMA.2011.026
http://arxiv.org/abs/1009.2366
https://doi.org/10.1093/imrn/rnm0046
http://arxiv.org/abs/math.RT/0610383
https://doi.org/10.1090/S0002-9947-2010-05156-6
http://arxiv.org/abs/0707.0251
https://doi.org/10.1007/BF02392487
https://doi.org/10.2140/pjm.1989.140.353
https://doi.org/10.2140/pjm.1989.140.353
https://doi.org/10.1007/978-0-387-21738-3
1 Introduction
2 Modules of the symmetric group
2.1 Jack polynomials
3 The differential system
3.1 The adjoint operation on Laurent polynomials and L(x)
4 Integration by parts
5 Local power series near the singular set
5.1 Behavior on boundary
6 Bounds
7 Sufficient condition for the inner product property
8 The orthogonality measure on the torus
8.1 Maximal singular support
8.2 Boundary values for the measure
9 Analytic matrix arguments
References
|