1. Introduction
In a recent paper, Littlejohn and Quintero [
1] studied the
Krein–Sobolev polynomials , which are orthogonal in the classical Sobolev space:
which is endowed with the inner product
Throughout this paper, unless otherwise specified, we assume
c is a fixed, positive constant. This is not immediately obvious, especially since the discrete term in (
2) has a minus sign, that
defines a (positive-definite) inner product. These Krein–Sobolev polynomials naturally arise in the spectral analysis of the shifted one-dimensional Krein Laplacian self-adjoint operator
defined by
here,
is the maximal domain associated with the expression
in
given by
We note that, when
the bilinear form
is only a pseudo inner product so, unless otherwise indicated, we assume
. We note that the subscript 1 in
refers to the fact that
is called the
first left-definite space (see
Section 2 and
Section 3) associated with
The Krein–Sobolev polynomials
are, in a sense, a generalization of
Althammer’s polynomials , first studied by Althammer [
2] in 1962, and later by Schäfke [
3] in 1972. Althammer showed the sequence
is orthogonal with respect to the Sobolev inner product
(where
is a fixed, positive parameter) in
. Notice that when
and
is an even function, then
(later, we see that both
and
are even; in fact, we show that
for all
).
In the literature, the sequence
is known as
Althammer’s polynomials or
Sobolev–Legendre polynomials, which has the latter name because the initial construction of
involves the difference of two Legendre polynomials. In [
2] and in a later 1972 contribution by Schäfke [
3], several explicit properties of these polynomials were established, including an exact formula for
Althammer’s work is considered to be one of the first publications on the subject of Sobolev orthogonal polynomials, which is an area that has seen massive growth since the 1990s. For informative and detailed accounts of the history of the study of Sobolev orthogonal polynomials, we recommend the sources [
4,
5]. Sobolev orthogonal polynomials have applications in various areas, particularly in numerical analysis and solving boundary value problems for differential equations, where their ability to provide smooth approximations with good convergence properties is valuable.
Littlejohn and Quintero [
1] found several properties of the Krein–Sobolev polynomials
. However, they were not able to find an explicit formula for
for all
It is the purpose of this note to give an explicit representation of these polynomials.
The contents of this paper are as follows. In
Section 2, we recall properties of the classical self-adjoint Krein Laplacian operator
and the shifted Krein Laplacian operator
when
The operator
is a positive, self-adjoint operator in
Consequently, by applying a general left-definite theory of positive self-adjoint operators developed by Littlejohn and Wellman in [
6,
7], there is a continuum of Hilbert spaces
the so-called
left-definite spaces, generated by positive powers
of
. In general, we call
the first left-definite space. In the case of the shifted Krein Laplacian
it is the case that the first left-definite space is
Section 3 motivates left-definite theory and discusses the main results of abstract left-definite theory needed for this paper. In
Section 4, we review properties of the Krein–Sobolev polynomials in the first left-definite space
for the shifted Krein Laplacian operator
in
The construction of
is intricate and similar to the construction of the Althammer polynomials developed in [
2,
3]. In
Section 5, we discuss these Althammer polynomials and present our new result, namely, an explicit formula of each Krein–Sobolev polynomial
Lastly, in
Section 6, we summarize our main result and suggest an open problem on density of polynomials in certain constrained Hilbert spaces.
2. The Krein Laplacian Self-Adjoint Operator
In
with standard inner product
there are uncountably many unbounded, self-adjoint operators
T generated by the second-order differential expression
In this continuum, one such self-adjoint operator is the Krein Laplacian operator
defined in (
3) and (
4) when
Complete details on an analytic study of this operator in
can be found in [
8] in Section 10. In
the Krein Laplacian operator
has a discrete spectrum given by
and (complete) eigenfunctions
where
is a solution of the transcendental equation
More specifically, the eigenspace of
, equivalently the null space of
is two-dimensional and spanned by
. The eigenspace for the eigenvalue
is one-dimensional and spanned by the eigenfunction
, while the eigenspace for each
is one-dimensional and spanned by
An application of the general left-definite theory developed by Littlejohn and Wellman (see [
6,
7]) to the Krein Laplacian
produces a first left-definite space that is only a
pseudo inner product space, since the two-dimensional eigenspace
acts like the
zero element for this inner product. To avoid the awkwardness of considering equivalence classes of functions, we shift the spectrum (
7) by
. This allows for a left-definite analysis in a Hilbert function space with a positive-definite inner product. Consequently, from hereon, we study the
shifted Krein Laplacian
in
when
is fixed. In this case,
is self-adjoint and bounded below by
in
that is,
moreover, the spectrum of
is given by
where
is given in (
7). The eigenfunctions of
are, of course, the same as for
The first left-definite space was computed in [
1] to be the Sobolev space
.
As mentioned in the introduction, it is not immediately clear that
is a positive-definite inner product on
Even though it necessarily is an inner product from left-definite theory, we give an elementary proof. For
we see from the Cauchy–Schwarz inequality that
It follows that
and, consequently,
Furthermore, if
we see from (
10) that
implying that
almost everywhere for
However, since
we must have
on
It now follows that
is a positive-definite inner product on
3. A Glimpse of Left-Definite Operator
Theory
Left-definite spectral theory has its origins in the seminal work of H. Weyl [
9] in his analytical study of second-order Sturm–Liouville differential equations. Indeed, consider the differential expression
where, for simplicity sake, we assume
are both positive functions; here,
There are two ‘natural’ Hilbert space settings to study symmetric and self-adjoint operators generated by
Indeed, consider the spectral equation
where
is positive almost everywhere on
. The first setting is the Hilbert space
since the weight function
w appears on the right-hand side of (
11), we call
the
right-definite setting. The
left-definite setting is a certain Sobolev space
S with inner product
The inner product in (
12) is formally obtained from the classic Dirichlet identity
where
is the Dirichlet form associated with
The term ‘
left-definite’ arises from the fact that the left-hand side of (
11) generates the inner product in (
12). The mathematical literature, especially during the period 1970–2005, contains numerous articles on the left-definite study of second-order Sturm–Liouville equations; for example, see [
10,
11,
12,
13]. A further discussion of left-definite theory applied to second-order Sturm–Liouville equations can be found in the texts [
14], Chapter 5, and [
15], Chapters 5 and 12.
In [
6,
7] (see also [
16]), the authors generalize the notion of left-definite theory to
arbitrary self-adjoint operators
A in a Hilbert space
, which are bounded below by a positive constant
k in
that is,
A satisfies the inequality
for some
Indeed, as shown in [
6,
7], Littlejohn and Wellman show that such an operator
A produces a continuum of left-definite Hilbert spaces
. The space
is called the
left-definite space associated with
. Left-definite spaces are special cases of Hilbert scales, as described in [
17,
18].
We briefly describe the salient results of this left-definite operator theory for the purposes of this paper. Definition 1 below, taken from Definitions 2.1 and 3.1 in [
6], is motivated by five common features observed in examples listed in the above-mentioned papers.
Suppose
V is a real or complex vector space and
is an inner product on
such that
is a separable Hilbert space. In addition, suppose
is self-adjoint and bounded below by
for some positive constant
that is, the inequality in (
14) is satisfied. From the Hilbert space spectral theorem, for each
,
is self-adjoint and bounded below by
Definition 1. Let ; suppose is a subspace of V, and is an inner product on Let We say that is an left-definite space associated with if each of the following properties are satisfied:
- (1)
is a Hilbert space;
- (2)
is a subspace of
- (3)
is dense in
- (4)
for
- (5)
for and
The following theorem clarifies the existence, and uniqueness, of each
The proof of this theorem can be found in [
6] in Theorems 3.1, 3.4, and Theorem 3.7.
Theorem 1. Suppose A is a self-adjoint operator in the separable Hilbert space , which is bounded below by for some Let , and defineThen, is the unique left-definite space associated with Moreover, we have the following: - (i)
If A is bounded, then and the inner products and are equivalent for all
- (ii)
If is unbounded, then is a proper subspace of V, and for , is a proper subspace of moreover, none of the inner products or are equivalent;
- (iii)
If is a (complete) set of orthogonal eigenfunctions of A in then they are also a (complete) orthogonal set in each left-definite space
From Theorem 1, we see that is the form domain of that is to say,
4. The Sobolev Space and the Krein–Sobolev Orthogonal
Polynomials
From Theorem 1(iii), the eigenfunctions of the shifted Krein Laplacian operator
given in (
8), form a complete orthogonal set in
The main purpose of the Littlejohn–Quintero paper [
1] was to construct a
different complete orthogonal set in
namely, the Krein–Sobolev polynomials
We remark, however, that the polynomial
is
not an eigenfunction of
for any
when
The construction of the Krein–Sobolev polynomials
is similar to Althammer’s construction of the polynomials
, which we discuss in
Section 5. Both polynomial sets
and
are even or odd polynomials, and both are normalized so that
Furthermore, as we will see in
Section 5, when
n is even and
(see (
5)),
. However, the Althammer and Krein–Sobolev polynomials of odd degrees
are different.
For
define
where
is the sequence of classical Legendre polynomials normalized by
for each
For various properties of the Legendre polynomials, and orthogonal polynomials in general, see the classic text [
19], Chapter IV, of Szegö.
Calculations show that , etc.; moreover, for
In [
1], the following technical result was established for
.
Proposition 1. The following properties for the polynomials hold:
- (i)
For ,
- (ii)
- (iii)
- (iv)
- (v)
- (vi)
We are now in position to define the polynomials
from
as shown in [
1], the sequence
forms a complete orthogonal set (see Theorem 3) with respect to the inner product
Definition 2. where and Note that, by (
16), each
is well defined when
Moreover, calculations show that
, and
. We recall from
Section 2 that both
and
are eigenfunctions of the shifted Krein Laplacian operator
(corresponding to the eigenvalue
). However,
has no other polynomial eigenfunctions.
In [
1], the authors proved the following results.
Theorem 2. The Krein–Sobolev polynomials and the sequence have the following properties:
- (i)
- (ii)
- (iii)
The sequence satisfies and, for - (iv)
moreover, for each
- (v)
- (vi)
is a polynomial of degree exactly n, and .
Theorem 3. The Krein–Sobolev polynomials defined in (17), form a complete orthogonal set of polynomials in the first left-definite space defined in (1) and (2), associated with the shifted Krein Laplacian operator in Moreover,where is the Kronecker delta function. Theorem 4. The roots of the Krein–Sobolev orthogonal polynomials are real, simple, and lie in the interval
5. An Explicit Representation of the Krein–Sobolev
Polynomials
As shown by Schäfke [
3], the Althammer polynomials
satisfy the following properties:
- (i)
and, for
where
is defined in (
15), and the sequence
satisfies the recurrence relation
and
Furthermore, each
and is explicitly given by
- (ii)
- (iii)
For each
satisfy the orthogonality relationship
where
is defined in (
5).
From (
20), when
(see (
6)), we see that the Althammer coefficients
satisfy the
exact same recurrence relation as the Krein–Sobolev coefficients
in (
19). Moreover, since
it follows that
and, consequently, when
we see that
However, since
, the odd connection coefficients
for the Krein–Sobolev polynomials are different from the Althammer coefficients
given explicitly in (
21). In fact, the unique solution to (
19) when
n is odd, that is to say,
with the initial conditions
is given explicitly by
An algebraic proof (
22) that satisfies this initial value difference problem is tedious to calculate. Remarkably, we note that Wolfram’s Mathematica
instantly verifies the solution (
22) with the command FullSimplify.
We summarize our main result in the following theorem.
Theorem 5. With the polynomials defined in (15) we have the following formula for each Krein-Sobolev polynomial of degree whereandwhere As discussed in Theorem 3 and Theorem 4, these polynomials form a complete orthogonal set in the Sobolev space and, for has exactly n distinct real roots in .
We list the first few Krein–Sobolev polynomials
:
6. Conclusions
The first left-definite space
defined in (
1) and (
2), for the self-adjoint shifted Krein operator
contains polynomials as a dense subspace. Furthermore, the Krein–Sobolev polynomials
form a (complete) orthogonal set in
and, for
has
n simple roots in
These results were established by Littlejohn and Quintero in the earlier contribution [
1]. In this present paper, the main objective was to present an explicit form of these Krein–Sobolev polynomials; this explicit form is given in (
23) and (
24).
A natural question to ask regards the set of polynomials that are dense in the
secondleft-definite space
given explicitly by
and
We do not know the answer to this question. A straightforward calculation shows that a basis for polynomials in
is given by
for
Consequently, there is no polynomial of degree 2 or 3 belonging to
Even though this basis is not algebraically complete, it is not clear if they are analytically complete in
Note that the theory of exceptional orthogonal polynomials provides several examples of algebraically incomplete sequences of polynomials that are analytically complete. This problem suggests a larger open problem with regard to finding conditions for the density of polynomials in Hilbert spaces constrained by boundary conditions (as is the case in (
25)).