Classification_of_classical_orthogonal_polynomials
Classification_of_classical_orthogonal_polynomials
net/publication/238203450
CITATIONS READS
65 2,044
2 authors, including:
Lance Littlejohn
Baylor University
143 PUBLICATIONS 2,390 CITATIONS
SEE PROFILE
All content following this page was uploaded by Lance Littlejohn on 24 June 2014.
CLASSIFICATION OF CLASSICAL
ORTHOGONAL POLYNOMIALS
We first obtain new (algebraic) necessary and sufficient conditions on the coef-
ficients `1 (x) and `2 (x) for the above differential equation to have orthogonal
polynomial solutions. Using this result, we then obtain a complete classifi-
cation of all classical orthogonal polynomials : up to a real linear change of
variable, there are the six distinct orthogonal polynomial sets of Jacobi, Bessel,
Laguerre, Hermite, twisted Hermite, and twisted Jacobi.
1. Introduction
All polynomials in this work are assumed to be real polynomials in the
real variable x and we let P be the space of all these real polynomials.
We denote the degree of a polynomial π(x) by deg(π ) with the convention
that deg(0) = −1. By a polynomial system (PS), we mean a sequence of
polynomials {φn (x)}∞
n=0 with deg(φn ) = n, n ≥ 0. Note that a PS forms a
basis for P .
A PS {φn (x)}∞
n=0 is called orthogonal if there is a function µ : R → R of
bounded variation on the real line R such that
Z
(1.1) x n dµ(x)
R
where K n are non-zero real constants and δmn is the Kronecker delta function.
Furthermore, we shall say that {φn (x)}∞ n=0 is classi cal if each φn (x) (n ≥ 0)
satisfies a fixed second-order differential equation of the form
polynomials was handled by many authors thereafter : see, for example, [1],
[5], [6], [11], [27], and [31]. The problem was settled by Lesky [27] in 1962 at
least for classical orthogonal polynomials satisfying the orthogonality relation
(1.2) in which the function µ(x) is non-decreasing. Lesky [27] showed that
the only such orthogonal polynomials are Jacobi polynomials with α and
β > −1, Laguerre polynomials with α > −1, and Hermite polynomials.
It is easy to see that the PS {x n }∞
n=0 in case (d) above cannot be orthogonal.
The orthogonality of the Bessel polynomials was first observed by H.L. Krall
[18] and later investigated in depth by Krall and Frink [19]. Bochner [3]
observed the relation between the PS in case (e) above and the half-integer
Bessel functions and it is this relation which motivates the name Bessel
polynomials in [19]. The orthogonality of the Jacobi polynomials for α or
β < −1 and Laguerre polynomials for α < −1 was recently treated by
Morton and Krall [32].
A natural question arises : are these four PS’s of Jacobi, Laguerre, Hermite,
and Bessel the only classical orthogonal polynomials? Of course, if we allow
for a complex linear change of variable, as Bochner does in [3], the answer is
yes. However, if we restrict our attention to a real linear change of variable, as
we shall do in this paper, are there any more classical orthogonal polynomials?
As far as the authors know, no previous work on this classification problem
really exhausts all possibilities.
After obtaining necessary and sufficient conditions (see Theorem 2.9) for
the differential equation (1.3) to have orthogonal polynomials of solutions
in section two, we give a complete classification of classical orthogonal
polynomials in section three. Finally, in section four, we will discuss the
integral or distributional representation of orthogonality for each classical
orthogonal polynomial system found in section three.
or
Z
hσ, π i = π(x)φ(x) d x (π ∈ P ),
R
for every n ≥ 0.
Any PS {Pn (x)}∞ n=0 determines a moment functional σ (uniquely up to
a non-zero constant multiple), called a canonical moment functional for
{Pn (x)}∞
n=0 , by the conditions
If we further have
It is immediate from the orthogonality (2.3) that for any WOPS {Pn (x)}∞
n=0 ,
its orthogonalizing moment functional must be a canonical moment functional
for {Pn (x)}∞n=0 so that it is unique up to a non-zero constant multiple.
Classical orthogonal polynomials 977
It is well known (for example, see [4, Chapter 1]) that a moment functional
σ is quasi-definite (respectively, positive-definite) if and only if there is an
OPS (respectively, a positive-definite OPS) relative to σ . It is clear that
if {Pn (x)}∞ ∞
n=0 is an OPS relative to σ , then so is {C n Pn (x)}n=0 for every
sequence of non-zero constants Cn . Conversely if σ is any quasi-definite
moment functional and {Pn (x)}∞ n=0 is an OPS relative to σ , then each Pn (x)
is uniquely determined up to an arbitrary non-zero factor. In particular, for
any quasi-definite moment functional σ , there is a unique monic OPS relative
to σ given by
σ0 σ1 ... σn
σ1 σ2 ... σn+1
1 .. .. .. ..
(2.5) Pn (x) = det
. . . .
(n ≥ 0),
1n−1 (σ )
σn−1 σn ... σ2n−1
1 x ... xn
X
i
`i (x) = `i j x j (i = 1, 2),
(2.6) j =0
λn = n(n − 1)`22 + n`11 (n ≥ 0),
For a new and somewhat simpler proof of Theorem 2.1, see [23] ; for
another proof of the general Krall characterization theorem, see [20] and
[25].
We call the recurrence relation (2.7) the moment equati on for the dif-
ferential equation (1.3). We may use Theorem 2.1 to classify all possible
classical OPS’s. However it is very difficult, in general, to solve the moment
equation (2.7) and to see whether the corresponding moment functional is
quasi-definite or not. The disadvantage of the conditions in Theorem 2.1 is
that the equation (2.7) contains not only the coefficients of (1.3) but also the
moments of a canonical moment functional of a classical OPS of which the
existence is not known apriori.
Below, we shall first obtain a necessary condition (see Theorem 2.5) and
then necessary and sufficient conditions (see Theorem 2.9) for the differential
equation (1.3) to have an OPS of solutions. Unlike those in Theorem 2.1,
these conditions involve only the coefficients of the differential equation (1.3).
We begin with introducing some formal calculus on moment functionals.
For a moment functional σ and π ∈ P , we let σ 0 , the derivative of σ and π σ ,
multiplication of σ by a polynomial, be those moment functionals defined by
(2.8) hσ 0 , pi = −hσ, p 0 i ( p ∈ P)
and
(2.9) hπ σ, pi = hσ, π pi ( p ∈ P ).
It is then easy to obtain the following Leibnitz rule for any moment functional
σ and polynomial π(x) :
(2.10) (π σ )0 = π 0 σ + π σ 0 .
X
K
0 = hπ σ, PN i = Ck hσ, Pk PN i = C N hσ, PN2 i
k=0
Note that the zero in the right hand side of the equation (2.11) means the
zero moment functional. In other words, the equation (2.11) means
h(`2 σ )0 − `1 σ, x n i = 0 (n ≥ 0),
X
n
Pn (x) = Ckn x k (Cnn = 1)
k=0
for any constant k. Hence L[y] = λ N y has infinitely many monic polynomial
solutions of degree N , which contradicts our assumption.
σ = cδ(`11 x + `10 ),
where c is an arbitrary constant and δ(`11 x + `10 ) is the Dirac delta moment
functional defined by
Theorem 2.5 was first proved by Lesky [27] only for positive-definite
classical OPS’s. However his method of proof cannot be extended to general
classical OPS’s since he used the following fact which holds only for positive-
definite OPS’s : for any positive-definite OPS {Pn (x)}∞
n=0 , the zeros of Pn (x),
n ≥ 1, are real and distinct and no two polynomials from {Pn (x)}∞ n=0 can
have common zeros (see Chihara [4]).
Classical orthogonal polynomials 983
The converse of Theorem 2.5 does not hold in general. For example, the
PS {x n }∞
n=0 satisfies the admissible differential equation
but {x n }∞
n=0 is not an OPS. However, we have the following partial converse
of Theorem 2.5.
THEOREM 2.6. If the differential operator L[·] in (1.3) is admissible, then
any PS {Pn (x)}∞
n=0 of solutions to the differential equation (1.3) is a WOPS.
Proof. By Lemma 2.4, we may assume that {Pn (x)}∞ n=0 is the unique
monic PS of solutions to (1.3). Let σ be a canonical moment functional of
{Pn (x)}∞
n=0 . Then σ 6= 0 by definition and, by Lemma 2.3, σ satisfies the
weight equation (2.11). Then we have for m and n ≥ 0
REMARK 2.4. In fact we can prove, by the same reasoning as in the proof
of Theorem 2.6, something more than Theorem 2.6. If L[ p] = λp and
L[q] = µq for some polynomials p(x) and q(x) and λ 6= µ, then hσ, pqi = 0
for any moment functional solution σ of the weight equation (2.11). Here we
do not need to assume L[·] is admissible.
We now seek a criterion for when a WOPS {Pn (x)}∞ n=0 is an OPS, which
does not involve a canonical moment functional of {Pn (x)}∞
n=0 .
For any monic PS {Pn (x)}n=0 , there are constants {αn }∞
∞ ∞
n=1 and {βn }n=1
such that
Pn
is a polynomial of degree ≤ n − 2. In fact if Pn (x) = k=0 Ckn x k (Cnn = 1 ;
n ≥ 1), then
(2.16)
αn = Cn−1
n
− Cnn+1
and
(2.17)
βn = Cn−2
n
− (Cn−1
n
− Cnn+1 )Cn−1
n
− Cn−1
n+1
(C−1
1
= 0).
At this moment, we need to recall Favard’s theorem (see [10]) which asserts
that a monic PS {Pn (x)}∞ n=0 is an OPS (respectively, a positive-definite OPS)
∞
if and only if {Pn (x)}n=0 satisfies a three term recurrence relation
PROPOSITION 2.7 (Krall and Sheffer [21]). A monic WOPS {Pn (x)}∞
n=0 is an
OPS (respectively, a positive-definite OPS) if and only if
∞
X
COROLLARY 2.8. Let {Pn (x)}∞
n=0 be a WOPS and Pn (x) = Ckn x k
k=0
(Cnn = 1) for n ≥ 0. Then {Pn (x)}∞ n=0 is an OPS (respectively, a positive-
definite OPS) if and only if
(2.20)
βn = Cn−2
n
− (Cn−1
n
− Cnn+1 )Cn−1
n
− Cn−1
n+1
6= 0 (respectively, βn > 0)
(n ≥ 1 ; C−1
1
= 0).
Proof. By Lemma 2.4, the above condition (i) is just the admissibility of
L[·] which is also equivalent to the fact that the differential equation (1.3) has a
Xn
∞
unique monic PS {Pn (x)}n=0 of solutions. If we set Pn (x) = Ckn x k (Cnn =
k=0
1 ; n ≥ 0), then Cn−1
n n
and Cn−2 are given by (2.21) and (2.22), respectively, by
solving the equation (2.14) for k = n − 1 and k = n − 2. Hence, Theorem 2.9
follows from Theorem 2.5, Theorem 2.6, and Corollary 2.8.
REMARK 2.5. If we assume that the differential equation (1.3) has a monic
PS {Pn (x)}∞ ∞
n=0 of solutions, then {Pn (x)}n=0 is an OPS if and only if the
condition (2.20) holds. For a proof of this statement, see [24, Proposition 3.7].
Note here that apriori we do not assume that {Pn (x)}∞ n=0 is a WOPS (as in
Proposition 2.7) or L[·] is admissible (as in Theorem 2.9). Furthermore, only
condition (2.2) must be checked but not with the conditions given in (2.21)
and (2.22). In general, these latter two equations are not well defined unless
the expression L[·] is admissible.
3. Classification
We say that any two OPS’s are equivalent if either one differs from the
other by non-zero constant factors or one is obtained from the other by a real
linear change of variable.
In this section, we will classify all classical OPS’s up to equivalence classes
using Theorem 2.9.
In the following, we let N be the set of all positive integers and use the
notation
α α α(α − 1) · · · (α − k + 1)
= 1 and =
0 k k!
for any complex number α and any integer k in N. As with Bochner, we
divide the cases according to the roots of the leading coefficient `2 (x) of the
differential expression L[·] in (1.3).
Cases 1: Jacobi polynomials
We assume `22 6= 0 and `221 − 4`22 `20 > 0. Then, by a real linear change
of variable, the equation (1.3) can be transformed into
L[y](x) = (1 − x 2 )y 00 (x) + [(β − α) − (α + β + 2)x]y 0 (x)
(3.1)
= − n(n + α + β + 1)y(x).
We assume −(α + β + 1) ∈ / N so that L[·] in (3.1) is admissible. Then the
(α,β)
equation (3.1) has a unique monic PS {Pn (x)}∞ n=0 , called the Jacobi PS, of
solutions :
(3.2)
n
(α,β) 2n + α + β −1 X n + α n + β
Pn (x) = (x − 1)n−k (x + 1)k
n k=0
k n−k
(n ≥ 0).
Classical orthogonal polynomials 987
(α,β)
PROPOSITION 3.1. The Jacobi PS {Pn (x)}∞ n=0 is
(i) a WOPS if −(α + β + 1) ∈ /N;
(ii) an OPS if and only if −α , −β , and −(α + β + 1) ∈/ N:
(iii) a positive-definite OPS if and only if α and β > −1.
Proof. The proof of (i) follows from Theorem 2.6. Now we assume
−(α + β + 1) ∈/ N. We then have, from (2.20), (2.21), (2.22), and (3.1),
(3.3)
4n(α + β + n)(α + n)(β + n)
βn = (n ≥ 1).
(α + β + 2n − 1)(α + β + 2n)2 (α + β + 2n + 1)
Hence, βn 6= 0 for n ≥ 1 if and only if α + n 6= 0 and β + n 6= 0 for n ≥ 1 so
that (ii) follows from Theorem 2.9. To prove (iii), it suffices to show βn > 0
for n ≥ 1 if and only if α and β > −1. If α and β > −1, then every factor
in (3.3) is positive so that βn > 0 for n ≥ 1. Conversely, assume βn > 0
for n ≥ 1 but α < −1 (when β < −1, the proof is essentially the same).
Then, from β1 > 0, we have (β + 1)(α + β + 3) < 0. If β + 1 < 0 and
α + β + 3 > 0, then α + β + 2 < 0 and 0 < α + 2, β + 2 < 1 so that β2 < 0,
which is a contradiction. If β + 1 > 0 and α + β + 3 < 0, then α < −2.
Then, from β2 > 0, we have α + β + 5 < 0 and so α < −4. Continuing
the same process, we have that α < −2k for any integer k ≥ 1, which is
impossible.
(α,β)
The explicit orthogonality of the Jacobi PS {Pn (x)}∞
n=0 for α or β < −1
(but −α and −β ∈ / N) has been treated by Morton and Krall [32].
Case 2: Bessel polynomials
We assume `22 6= 0 and `221 − 4`22 `20 = 0. Then, by a real linear change
of variable, the equation (1.3) can be transformed into
(3.4) L[y](x) = x 2 y 00 (x) + (αx + β)y 0 (x) = n(n + α − 1)y(x).
We assume −(α −1) ∈ / N so that L[·] in (3.4) is admissible. Then the equation
(α,β)
(3.4) has a unique monic PS {Bn (x)}∞ n=0 of solutions :
n
x if β = 0
(3.5) Bn(α,β) (x) = 1 X
n
n! 0(α + n + k − 1) x k
β n 0(α + 2n − 1) if β 6= 0
(n − k)! k! β
k=0
(α,β)
(n ≥ 0). When β 6= 0, we call {Bn (x)}∞ n ∞
n=0 the Bessel PS. The PS {x }n=0
is a WOPS by Theorem 2.6 but it cannot be an OPS.
988 Kil H. Kwon and Lance L. Littlejohn
(α,β)
PROPOSITION 3.2. The Bessel PS {Bn (x)}∞ n=0 is an OPS (but not a
positive-definite OPS) if and only if −(α − 1) ∈
/ N and β 6= 0.
−nβ 2 (α + n − 2)
(3.6) βn = (n ≥ 1).
(α + 2n − 3)(α + 2n − 2)2 (α + 2n − 1)
The Bessel PS, as an OPS, was first observed by H.L. Krall [18]. Earlier
these polynomials were discussed by Romanovski [33] and Bochner [3]. In
[19], Krall and Frink studied the Bessel polynomials in detail and found,
explicitly, their complex orthogonality.
n
X
n + α (−x)k
(3.8) L (α)
n (x) = (−1) n!
n
(n ≥ 0).
k=0
n−k k!
∞
PROPOSITION 3.3. The Laguerre PS {L (α)
n (x)}n=0 is
(i) a WOPS for every α ;
(ii) an OPS if and only if −α ∈ /N;
(iii) a positive-definite OPS if and only if α > −1.
Classical orthogonal polynomials 989
Proof. (i) follows from Theorem 2.6. We have from (2.20), (2.21), (2.22),
and (3.7)
The case α = 0 is the one originally studied by Laguerre [26]. The case
α > −1 is due to Sonine [34] and the generalized Laguerre PS for α < −1
and −α ∈/ N has been recently studied by Morton and Krall [32].
Case 4: Hermite polynomials
We assume `22 = `21 = 0, `20 6= 0, and `11 < 0. Then, by a real linear
change of variable, the equation (1.3) can be transformed into
X
[n/2]
(−1)k x n−2k
(3.11) Hn (x) = n! (n ≥ 0),
k=0
k! (n − 2k)! 4k
X
[n/2]
1 x n−2k
(3.14) Ȟn (x) = i Hn (−i x) = n!
n
(n ≥ 0).
k=0
k! (n − 2k)! 4k
−n
(3.15) βn = (n ≥ 1).
2
Hence, the proposition follows from Theorem 2.9.
ie = β − α and d = α + β + 2.
Classical orthogonal polynomials 991
(α,β)
(n ≥ 0), where P̌n (x; d, e) = P̌n (x; α + β + 2, i (α − β)) = P̌n (x). Note
(α,β) (α,β)
that even though the expression for P̌n (x) in (3.17) involves i , P̌n (x)
is a real polynomial of degree n since β = ᾱ.
(α,β)
PROPOSITION 3.6. The twisted Jacobi PS { P̌n (x)}∞ n=0 is an OPS (but
not a positive-definite OPS) if and only if −(α + β + 1) ∈
/ N.
Proof. We have, from (2.20), (2.21), (2.22), and (3.16),
(3.18)
−4n(α + β + n)(α + n)(β + n)
βn = (n ≥ 1).
(α + β + 2n − 1)(α + β + 2n)2 (α + β + 2n + 1)
For each classical OPS, we can compute the moments {σn }∞ n=0 of its canoni-
cal moment functional σ by solving the corresponding moment equation (2.7)
successively starting from any non-zero value for σ0 . However, the moment
equation is, in general, a three term recurrence relation, which is not easy to
solve. Morton and Krall [32] introduced an idea by which we can always
reduce a three term recurrence relation to a two term recurrence relation. Let
t = x − x0 , where x0 is a constant, possibly complex, that will be chosen
later. Then, in terms of the new variable t, the differential equation (1.3) and
the corresponding moment equation (2.7) become
[`22 t 2 + (2`22 x0 + `21 )t + `22 x02 + `21 x0 + `20 ]y 00 (t)
(3.19)
+ [`11 t + (`11 x0 + `10 )]y 0 (t) = λn y(t),
and
(`11 + n`22 )σn+1 (x0 ) + [`11 x0 + `10 + n(2`22 x0 + `21 )]σn (x0 )
(3.20)
+ n(`22 x02 + `21 x0 + `20 )σn−1 (x0 ) = 0 (n ≥ 0),
where σn (x0 ) = hσ, (x − x0 )n i is the nth moment of σ about x0 . If we choose
x0 so that `22 x02 + `21 x0 + `20 = 0, then the equation (3.20) becomes a two
term recurrence relation and we have
(3.21)
Xn
n n−k
σn = hσ, x i = hσ, [(x − x0 ) + x0 ] i =
n n
x σk (x0 ) (n ≥ 0).
k=0
k 0
We illustrate the above procedure for the twisted Jacobi polynomials; see
Morton and Krall [32] for a similar discussion of the moments for the other
classical OPS’s, except the twisted Hermite PS. The moment equation for the
twisted Hermite PS is a two-term recurrence relation, which can be solved
easily.
Let σ̌ = σ̌ (α,β) be the canonical moment functional of the twisted Jacobi PS
(α,β)
{ P̌n (x)}∞ n=0 with σ̌0 = hσ̌ , 1i = 1. The corresponding moment equation
is
(3.22) (α + β + n + 2)σ̌n+1 + i (α − β)σ̌n + n σ̌n−1 = 0 (n ≥ 0),
which is a three-term recurrence relation unless α = β. If we choose x0 to be
i and let {σ̌n (i )}∞ ∞
n=0 be the moments of σ̌ about i , then {σ̌n (i )}n=0 satisfies a
two-term recurrence relation
(α + β + n + 2)σ̌n+1 (i ) + 2i (α + n + 1)σ̌n (i ) = 0 (n ≥ 0),
Classical orthogonal polynomials 993
Note that all σ̌n are real since the complex conjugate of σ̌n (recall β = ᾱ) in
(3.23) is exactly σ̌n in (3.24).
and found, via the Fourier transform, orthogonalizing weights for the Jacobi,
Laguerre, and Hermite PS’s. This formal δ-series expansion was also used in
Kim and Kwon [14] to produce an orthogonalizing hyperfunctional weight
for the Bessel PS {Bn(2,2)(x)}∞
n=0 .
In case of a classical OPS {Pn (x)}∞ n=0 satisfying the differential equation
(1.3), we may use the corresponding weight equation (2.11) to find an or-
thogonalizing weight for {Pn (x)}∞ n=0 . To do this, however, we must interpret
(2.11) as a classical differential equation with the right-hand side of (2.11)
replaced by a function (not necessarily identically zero) having zero moments.
To be precise we have the following Theorem, which is a special case
of Theorem 2.3 in [22] for second-order differential equations (see also [28,
Theorem 5.6]).
THEOREM 4.1. Let {Pn (x)}∞ n=0 be a classical OPS satisfying the differ-
ential equation (1.3). If w(x) is an orthogonalizing weight distribution for
{Pn (x)}∞
n=0 , then w(x) satisfies the distributional differential equation
and
(iii) w(x) is non-trivial as a moment functional,
then w(x) is an orthogonalizing weight distribution for {Pn (x)}∞
n=0 .
Condition (iii) in the above Theorem 4.1 means that hw, x n i 6= 0 for some
n ≥ 0. For any classical OPS, there always exists a distributional orthogo-
nalizing weight w(x) satisfying the conditions (i), (ii), (iii) in Theorem 4.1.
In fact, it is enough to take w(x) to be φ(x) in (4.2) where {σn }∞ n=0 are the
moments of any canonical moment functional σ of the given classical OPS.
We call the equation (4.4) the non-homogeneous weight equation for the
differential equation (1.3). When g(x) ≡ 0, the homogeneous weight equa-
tion
can be computed most easily from the three-term recurrence relation (2.18).
In fact, we have (see [4, Theorem 4.2 in Chap. 1])
Y
n
(4.8) hw(x), Pn2 i = βj ,
j =0
996 Kil H. Kwon and Lance L. Littlejohn
which is equivalent to
β
(4.11) w(α,β) (x) = (1 − x)α+ (1 + x)+ ,
2α+β+1 0(α+1)0(β+1)
for n ≥ 0, since hw(α,β) (x), 1i = 0(α+β+2)
.
Classical orthogonal polynomials 997
of which the only one linearly independent distributional solution with support
in [0, ∞) is
0 if x ≤ 0
(4.15) w0 (x) =
x α exp(−2/x) if x > 0.
Romanovski [33] used w0 (x) as an orthogonalizing weight for Bessel PS, but
w0 (x) cannot be an orthogonalizing weight since it does not decay rapidly at
infinity. In fact, we have
lim x n w0 (x) = ∞
x→∞
where g(x) is a function with zero moments. For x 6= 0, the general solution
of (4.16) is
c1 (−x)α e−2/x if x < 0
w(x) = Rx
x α e−2/x 0 e2/t t −2−α g(t) dt + c2 x α e−2/x if x > 0,
If we further take g(x) to be the function given in (4.7), then w(α) in (4.17) is a
continuous function on R satisfying the conditions (i) and (ii) in Theorem 4.1
(see [9], [22], and [30]). Hence, w(α) (x) in (4.17) (with g(x) in (4.7)) is an
orthogonalizing weight for Bessel PS {Bn(α)(x)}∞ n=0 if and only if
Z ∞ Z ∞
(α) α −2/x 2/t −2−α
(4.18) hw (x), 1i = − x e e t g(t) dt d x 6= 0.
0 x
Condition (4.18) was first proved in [22] for α = 0 and, recently, Maroni [30]
proved (4.18) for all α ≥ 12( π2 )4 − 2.
If we let σ (α) be the canonical moment functional of {Bn(α)(x)}∞ n=0 with
(α)
σ0 = 1, then we have from (3.6) and (4.8)
REMARK 4.2. Krall and Frink [19] found the complex orthogonality (now
called the Bessel orthogonality) of the Bessel PS through the contour integral
along the unit circle in the complex plane. Although the homogeneous weight
equation (4.14) cannot yield a distributional orthogonalizing weight for the
Bessel PS, it has a non-trivial hyperfunctional solution with support at {0}
with respect to which the Bessel PS is orthogonal (see [9], [14], and [15]).
Later in this section, we will discuss again real orthogonalizing weights
for {Bn(α)(x)}∞
n=0 for any α with −(α + 1) ∈/ N.
Classical orthogonal polynomials 999
xv 0 (x) − αv(x) = 0,
where c1 and c2 are arbitrary constants and x−α is the distribution on R with
support in (−∞, 0] (defined similarly as x+α ; see Remark 4.1 and Hörmander
[13, Chap. 3.3.2]). Hence, the general solution of (4.20) is
Since w(α) (x) in (4.21) satisfies the conditions (i), (ii), (iii) in Theorem 4.1,
∞
w(α) (x) is an orthogonalizing weight for {L (α)
n (x)}n=0 . Since
Since w(x) in (4.24) satisfies the conditions (i), (ii), (iii) in Theorem 4.1,
w(x) is an orthogonalizing weight for {Hn (x)}∞ n=0 . We then have, from (3.12)
and (4.8),
(4.25) Z ∞
√
hw(x), Hn (x)i =
2
Hn2 (x) exp(−x 2 ) d x = πn! 2−n (n ≥ 0).
−∞
w0 (x) = exp(x 2 ),
Then we have
0 if x ≤ 0
(4.30) w(x) = x2
Rx −t 2
e 0 e g(t) dt if x > 0.
Note that w(x) in (4.30) is a classical solution to (4.26) on R. If we further
assume
and so w(x) satisfies the conditions (i), (ii) in Theorem 4.1. Consequently,
w(x) in (4.30) is a real orthogonalizing weight for { Ȟn (x)}∞
n=0 if and only if
Z ∞ Z x
x2 −t 2
(4.32) hw(x), 1i = e e g(t) dt d x 6= 0.
0 0
The existence of a weight w(x) for the twisted Hermite PS, of the form
given in (4.30) and satisfying (4.29), is discussed below in Remark 4.3.
∞
√ let σ be the orthogonalizing moment functional for { Ȟn (x)}n=0 with
If we
σ0 = π , then we have, from (3.15) and (4.8),
√
(4.33) hσ, [ Ȟn (x)]2 i = (−1)n π n!2−n (n ≥ 0).
REMARK 4.3. We can easily see that there is a non-trivial function g(x)
with zero moments, which also satisfies the condition (4.29). Choose any
two linearly independent continuous functions g1 (x) and g2 (x) with zero
moments and support in [0, ∞). Set
Z ∞
e−x gi (x) d x
2
Ai = (i = 1, 2).
0
1002 Kil H. Kwon and Lance L. Littlejohn
If Ai 6= 0 (i = 1, 2), then
lim x n f (x) = ∞
x→∞
for n + 2 − d > 0. However, from (3.17) and (4.12), we can obtain the
complex orthogonality
(4.35)
β
h(1 − x)α+ (1 + x)+ , P̌m(α,β) (i x) P̌n(α,β) (i x)i
(−1)n 22n+α+β+1 0(n + α + 1)0(n + β + 1)0(n + α + β + 1)n!
= δmn
0(2n + α + β + 1)0(2n + α + β + 2)
(m and n ≥ 0), where i e = β − α and d = α + β + 2.
Let us now consider the non-homogeneous weight equation
Then we have
0 if x ≤ 0
(4.38) w̌(x) = f (x)
Rx − f (t) 2 −1
e 0 e (1 + t ) g(t) dt if x > 0.
Note that w̌(x) in (4.38) is a classical solution to (4.36). If g(x) satisfies
the condition (4.31), then w̌(x) satisfies the conditions (i) and (ii) in Theo-
rem 4.1. Consequently, w̌(x) in (4.38) is a real orthogonalizing weight for
(α,β)
{ P̌n (x)}∞n=0 if and only if
Z ∞ Z x
f (x) − f (t) 2 −1
(4.39) hw̌(x), 1i = e e (1 + t ) g(t) dt d x 6= 0.
0 0
(α,β)
If we let σ̌ be the orthogonalizing moment functional of { P̌n (x)}∞
n=0
with
2α+β+1 0(α + 1)0(β + 1)
σ̌0 = ,
0(α + β + 2)
then we have, from (3.18) and (4.8),
(4.40)
hσ̌ ,[ P̌n(α,β) (x)]2 i
(−1)n 22n+α+β+1 0(n + α + 1)0(n + β + 1)0(n + α + β + 1)n!
= .
0(2n + α + β + 1)0(2n + α + β + 2)
(n ≥ 0)
REMARK 4.4. In the formula (4.12), the parameters α and β are real
numbers with −(α + β + 1), −α, and −β ∈ / N . However, by analytic
continuation, the same formula holds for complex parameters α and β as long
as − Re(α + β + 1), − Re α, and − Re β ∈ / −N. This fact is used in (4.35),
where β = ᾱ, Re α = Re β = 2 , and −(d − 1) = −(α + β + 1) ∈
d−2
/ N.
Constructing explicit real orthogonalizing weights for classical OPS’s by
solving the non-homogeneous weight equation (4.4) has been successful ex-
cept, at the moment, for the Bessel PS {Bn(α) (x)}∞
n=0 when 0 6= α < 12( π ) −2,
2 4
1004 Kil H. Kwon and Lance L. Littlejohn
(α,β)
the twisted Hermite PS { Ȟn (x)}∞
n=0 , and the twisted Jacobi PS { P̌n (x)}∞
n=0 .
For any OPS (classical or not), its real orthogonalizing weight can be explic-
itly constructed by the following remarkable result on the general moment
problem.
THEOREM 4.2 (Duran [8]). For any sequence of real or complex numbers
{σn }∞n=0 , define a function w(x) by
0 if x ≤ 0
(4.41) w(x) = 1 R ∞ P∞ √
2 0
( n=0 σn cn t n h(λn t))J0 ( xt) dt if x > 0,
(−1)n P
where cn = 22n+1 (n!)2
, λn = n + nk=0 cn , J0 (x) is the Bessel function of the
first kind, and h(x) is a C ∞ -function on R with compact support satisfying
h(0) = 1 and h (n) (0) = 0 (n ≥ 1). Then, w(x) is a function in the Schwartz
space S and satisfies
Z ∞ Z ∞
x w(x) d x =
n
x n w(x) d x = σn (n ≥ 0).
−∞ 0
In particular, if we take {σn }∞ in Theorem 4.2 to be the moments of a
n=0
canonical moment functional of any classical OPS, then the function w(x)
in (4.41) is a real orthogonalizing weight for the OPS. Moreover, by Theo-
rem 4.1, the function
g(x) = (`2 (x)w(x))0 − `1 (x)w(x)
is a function, in the Schwartz space S with zero moments and support in
[0, ∞), satisfying the condition (4.31). In the case of the twisted Hermite or
the twisted Jacobi polynomials, this g(x) also satisfies (4.29) and (4.32) or
(4.37) and (4.39) respectively.
REMARK 4.5. In the case of the Bessel, the twisted Hermite, and the twisted
Jacobi polynomials, their complex orthogonality seems more natural than
their real orthogonality. In fact, through the hyperfunctional representations
of orthogonalizing weights, we can see that any OPS has both real and
complex orthogonality : see, for example, [14].
ACKNOWLEDGEMENTS. The first author (KHK) thanks Professor L. Duane
Loveland of the Department of Mathematics and Statistics at Utah State
University for the opportunity to visit Utah State during the 1992-93 academic
year. In addition, he thanks the Korea Science and Engineering Foundation
(95-0701-02-01-3), GARC, and Korea Ministry of Education(BSRI 1420) for
their research support.
Classical orthogonal polynomials 1005
References
[1] J. Aczél, Eine Bemerkung über die Characterisierung der klassischen orthogonale Poly-
nome, Acta Math. Sci. Hung. 4 (1953), 315–321.
[2] R. P. Boas, The Stieltjes moment problem for functions of bounded variation, Bull. Amer.
Math. Soc. 45 (1939), 399–404.
[3] S. Bochner, Über Sturm-Liouvillesche Polynomsysteme, Math. Z. 29 (1929), 730–736.
[4] T. S. Chihara, An introduction to orthogonal polynomials, Gordon and Breach, New York,
1977.
[5] C. W. Cryer, Rodrigues’ formulas and the classical orthogonal polynomials, Boll. Unione
Mat. Ital. 25 (1970), 1–11.
[6] Á. Császár, Sur les polynômes orthogonaux classiques, Annales Univ. Sci. Budapest Sec.
Math. 1 (1958), 33–39.
[7] A. J. Duran, The Stieltjes moment problem for rapidly decreasing functions, Proc. Amer.
Math. Soc. 107(3) (1989), 731–741.
[8] , Functions with given moments and weight functions for orthogonal polynomials,
Rocky Mountain J. Math. 23(1) (1993), 87–104.
[9] W. D. Evans, W. N. Everitt, K. H. Kwon, and L. L. Littlejohn, Real orthogonalizing weights
for Bessel polynomials, J. Comp. Appl. Math. 49 (1993), 51-57.
[10] J. Favard, Sur les polynomes de Tchebicheff, C.R. Acad. Sci. Paris 200 (1935), 2052–2053.
[11] J. Feldmann, On a characterization of classical orthogonal polynomials, Acta. Sc. Math.
17 (1956), 129–133.
[12] G. H. Hardy, On Stieltjes’ “Probleme des Moments", Messenger of Math. 4 (1917), 175–182.
[13] L. Hörmander, The analysis of linear partial differential operators I, Springer-Verlag, New
York, 1983.
[14] S. S. Kim and K. H. Kwon, Hyperfunctional weights for orthogonal polynomials, Results
in Math. 18 (1990), 273–281.
[15] , Generalized weights for orthogonal polynomials, Diff. and Int. Equations 4 (1991),
601–608.
[16] A. M. Krall and L. L. Littlejohn, On the classification of differential equations having
orthogonal polynomial solutions II, Ann. Mat. Pura Appl. 4 (1987), 77–102.
[17] H. L. Krall, Certain differential equations for Tchebychev polynomials, Duke Math. J. 4
(1938), 705–719.
[18] , On derivatives of orthogonal polynomials II, Bull. Amer. Math. Soc. 47 (1941),
261–264.
[19] H. L. Krall and O. Frink, A new class of orthogonal polynomials : The Bessel polynomials,
Trans. Amer. Math. Soc. 65 (1949), 100–115.
[20] H. L. Krall and I. M. Sheffer, A characterization of orthogonal polynomials, J. Math. Anal.
Appl. 8 (1964), 232–244.
[21] , Differential equations of infinite order for orthogonal polynomials, Ann. Mat. Pura
Appl. 4 (1966), 135–172.
[22] K. H. Kwon, S. S. Kim, and S. S. Han, Orthogonalizing weights for Tchebychev sets of
polynomials, Bull. London Math. Soc. 24 (1992), 361–367.
[23] K. H. Kwon, J. K. Lee, and B. H. Yoo, Characterizations of classical orthogonal polynomi-
als, Results in Math. 24 (1993), 119-128.
1006 Kil H. Kwon and Lance L. Littlejohn
[24] K. H. Kwon and L. L. Littlejohn, Sobolev orthogonal polynomials and second-order differ-
ential equations, Rocky Mt. J. Math. (to appear).
[25] K. H. Kwon, L. L. Littlejohn, and B. H. Yoo, Characterizations of orthogonal polynomials
satisfying differential equations,R∞ SIAM J. Math. Anal. 25(3) (1994), 976–990.
[26] E. N. Laguerre, Sur l’integrale x x −1 e−x d x, Bull. de la Socíeté Math. de France 7 (1879),
72–81.
[27] P. Lesky, Die Charakterisierung der klassischen orthogonalen Polynome durch Sturm-
Liouvillesche Differentialgleichungen, Arch. Rat. Mech. Anal. 10 (1962), 341–352.
[28] L. L. Littlejohn, On the classification of differential equations having orthogonal polynomial
solutions, Ann. Mat. Pura Appl. 4 (1984), 35–53.
[29] L. L. Littlejohn and D. Race, Symmetric and symmetrisable ordinary differential expressions,
Proc. London Math. Soc. 1 (1990), 334–356.
[30] P. Maroni, An integral representation for the Bessel form, J. Comp. Appl. Math. 57 (1995),
251–260.
[31] M. Mikolás, Common characterization of the Jacobi, Laguerre and Hermite-like polynomi-
als (in Hungarian), Mate. Lapok 7 (1956), 238–248.
[32] R. D. Morton and A. M. Krall, Distributional weight functions for orthogonal polynomials,
SIAM J. Math. Anal. 9(4) (1978), 604–626.
[33] M. V. Romanovski, Sur quelques classes nouvelle de polynomes orthogonaux, C.R. Acad.
Sci. Paris 188 (1929), 1023–1025.
[34] N. J. Sonine, Recherches sur les fonctions cylindriques et le développment des fonctions
continues en series, Math. Ann. 16 (1880), 1–80.
[35] T. J. Stieltjes, Recherches sur les fractions continues, Ann. de la Faculté des Sci. de Toulouse
8 (1894), J1–122; 9 (1895), A1–47; Oeuvres 2, 398–566.
K. H. Kwon
Department of Mathematics
KAIST
Taejon 305-701, Korea
E-mail: khkwon@ jacobi.kaist.ac.kr
L. L. Littlejohn
Department of Mathematics and Statistics
Utah State University
Logan, Utah, 84322-3900
E-mail: lance@sunfs.math.usu.edu