Asymptotics of Discrete -Freud orthogonal polynomials from the -Riemann Hilbert Problem
Abstract.
We investigate a Riemann-Hilbert problem (RHP), whose solution corresponds to a group of -orthogonal polynomials studied earlier by Ismail et al. Using RHP theory we determine new asymptotic results in the limit as the degree of the polynomials approach infinity. The RHP formulation also enables us to obtain further properties. In particular, we consider how the class of polynomials and their asymptotic behaviours change under translations of the -discrete lattice and determine the asymptotics of related -Painlevé equations.
Key words and phrases:
Orthogonal polynomials, Riemann Hilbert Problem, -difference calculusKey words and phrases:
Riemann-Hilbert Problem, -orthogonal polynomials and -difference calculus. MSC classification: 33C45, 35Q15, 39A13.1. Introduction
Orthogonal polynomials are a key component of a wide array of mathematical problems. They provide the basis of solutions of Sturm-Liouville problems [7], their zeros are related to the eigenvalue distribution of random matrices [14], they describe transition probabilities in birth-death models [13] and are used in numerical spectral approximation methods, just to name a few examples. Their importance in describing physical phenomena was recognised over two hundred years ago (Legendre, Laplace) and they continue to be pivotal in describing mathematical and physical problems.
In this paper we study a class of -orthogonal polynomials and deduce new results concerning their asymptotic behaviour as the degree tends to infinity. The orthogonality measure of these polynomials is supported on the discrete lattice, , for , where . We also determine some properties of -Painlevé equations associated with these -orthogonal polynomials.
1.1. Notation
For completeness, we recall some well known definitions and notations from the calculus of -differences. These definitions can be found in [4]. Throughout the paper we will assume and .
Definition 1.1.
We define the Pochhammer symbol, -derivative and Jackson integral as follows:
-
(1)
The Pochhammer symbol is
We denote product of Pochhammer symbols in multiple variables by ,
-
(2)
The -derivative is defined by
(1.1) Note that
where
-
(3)
The Jackson integral from to for some integers is given by
The Jackson integral from to for some integers is given by
Definition 1.2.
In this paper will denote the set of natural numbers including zero (i.e. 0,1,2,3,..), unless otherwise stated.
We recall the definition of an appropriate Jordan curve and admissible weight function given in [10, Definition 1.2] (with slight modification).
Definition 1.3.
A positively oriented Jordan curve in with interior and exterior is called appropriate if
and,
Definition 1.4 ([10]).
Define by
(1.2) |
Note that satisfies the -difference equation
(1.3) |
In Appendix A, we show that has certain unique properties.
1.2. Background
Let be a class of monic polynomials which satisfy the orthogonality relation
(1.4) |
for some weight function . Equation (1.4) gives rise to the three term recurrence relation
(1.5) |
where the recurrence coefficients are given by
(1.6) |
For even weight functions, , we find for all .
Following the pioneering work of Freud and others, questions arising about the asymptotic locations of the zeros of and behaviour of as have led to many developments in orthogonal polynomials and approximation theory. Motivated by the work of Deift et al. [2] we use the setting of the Riemann Hilbert Problem (RHP) to answer such questions (in Theorems 1.5, 1.6 and 1.7) for a class of -orthogonal polynomials. In particular, we study polynomials which satisfy the orthogonality condition
Throughout this paper we will label these polynomials as -Freud II polynomials to be consistent with the nomenclature of the DLMF [3, Chapter 18]. -Freud II polynomials were studied earlier by Ismail et al. [8], however Ismail et al. considered orthogonality on a continuous measure over . We will show in Section 1.3 that this continuous class of polynomials can readily be extended to those with a discrete measure, which will be the focus of this paper.
A significant development in the theory of orthogonal polynomials is the observation that the recurrence coefficients often give rise to discrete Painlevé equations [5]. For example, the recurrence coefficients of orthogonal polynomials with weight function satisfy the discrete equation [6]
which is a case of the first discrete Painlevé equation, or [9].
The recurrence coefficients of -Freud II polynomials also satisfy a discrete Painlevé equation, where the non-autonomous term in the equation is now iterated on a multiplicative lattice. (For the terminology distinguishing types of discrete Painlevé equations, we refer to Sakai [12].) As detailed by Ismail et al. [8] the recurrence coefficients of -Freud II polynomials satisfy
(1.7) |
In Sections 1.3 and 8 we extend this result initially determined by Ismail et al. to a larger class of -Freud II polynomials, using techniques similar to those found in [1].
Through the connection to RHP theory developed in this paper we obtain new insights into the solutions of Equation (1.7). For example, Theorem 1.7 shows that there exists more than one real positive solution of Equation (1.7), and we notice that the asymptotic behaviour as varies between solutions.
There is an important feature of -difference equations that affects our discussion of -Freud II polynomials. The associated weight function satisfies a -difference equation, which gives rise to a family of weights involving a free -periodic function , where . If a weight in this family with were to be chosen, the resulting family of orthogonal polynomials may have properties that differ from the class we consider. We expand on this point below.
1.3. Defining -Freud II polynomials
We define the family of weight functions , where , as
where [8]. They satisfy the -difference equation
(1.8) |
This is analogous to classical Freudian weights , which satisfy the differential equation
(1.9) |
However, a key difference between these two relations is that Equation (1.8) is a discrete relation. In particular if satisfies Equation (1.8) then also does, for any function satisfying . Consider the case
(1.10) |
This weight gives rise to discrete -Hermite polynomials [3, Chapter 18.27]. The sequence of discrete -Hermite polynomials, , are orthogonal with respect to any measure . In particular they satisfy the continuous orthogonality condition
on the real line, and also satisfy the discrete orthogonality condition
for any constant . In contrast, as we will show in Section 8 this is not true for the weight,
(1.11) |
which is the focus of this paper. Thus, when describing -Freud orthogonal polynomials, one also has to specify their orthogonality weight.
For the remainder of this paper, we focus on the sequence of polynomials , , that satisfy
We will call these polynomials polynomials. In Section 8, we discuss the implications of our results to polynomials orthogonal with respect to the weights of the form
(1.12) |
for any constant . We will call these polynomials polynomials.
1.4. Main results
We are now in a position to state the main results of this paper, which are listed as Theorems 1.5, 1.6 and 1.7 below. The first main result concerns the asymptotic behaviour of orthogonal polynomials as their degree approaches infinity.
Theorem 1.5.
Suppose that is a family of monic polynomials, orthogonal with respect to the weight . Define . Then, as , for even :
where is a solution of Equation (4.2), defined in Lemma 4.2, and, is a solution of Equation (5.2), defined in Lemma 5.2 (they are both independent of ). Similarly has the asymptotic behaviour
where is also a solution of Equation (4.2), defined in Lemma 4.2, and, is a solution of Equation (5.2), defined in Lemma 5.2.
Note that in the statement of Theorem 1.5, , , , and are constants determined in Sections 4 and 5 which do not depend on or . Our second main result concerns the asymptotic behaviour of recurrence coefficients and norm of as approaches infinity.
Theorem 1.6.
Our third main theorem answers the question of uniqueness posed by Ismail et al. in [8, Remark 6.4].
Theorem 1.7.
There exists infinitely many real positive solutions of the discrete equation
(1.13) |
Furthermore, in general
(1.14) |
where is the sequence of recurrence coefficients corresponding to polynomials with orthogonality condition given by Equation (1.12).
1.5. Outline
This paper is structured as follows. In Section 2 we state and solve a RHP (Definition 2.1) whose solution is given in terms of polynomials. We then make a series of transformations to this RHP in Section 3. By taking the limit this motivates the form of a near-field and far-field RHP, whose solutions are determined in Sections 4 and 5 respectively. In Section 6 we glue together the near and far-field solutions to approximate our initial RHP in the limit . Consequently we prove Theorems 1.5 and 1.6 in Section 7. In Section 8 we discuss the implications of our results to the recurrence coefficients of polynomials and prove Theorem 1.7. In Appendix A we prove some important properties of the function which are used in solving the near and far-field RHPs. Finally, for completeness we prove a well-known result concerning RHPs whose jump approaches the identity in Appendix B.
2. Statement of RHP
We begin the main arguments of this paper by introducing and solving a RHP (Definition 2.1) whose solution is given in terms of orthogonal polynomials.
Definition 2.1 ( RHP).
Let be an appropriate curve (see Definition 1.3) with interior and exterior . A complex matrix function , , is a solution of the RHP if it satisfies the following conditions:
-
-
(i)
is meromorphic in , with simple poles at for .
-
(ii)
has continuous boundary values and as approaches from and respectively, where
(2.1a) and is defined in Equation (1.11).
-
(iii)
The residue at each pole for is given by
(2.1b) -
(iv)
satisfies
(2.1c) for such that , for all , for fixed .
Remark 2.2.
Note that the matrix has poles in its second column for . Thus, the asymptotic decay does not hold near these poles. This is why, following Equation (2.1c), we require the added condition must be such that , for all , for fixed .
We now determine the solution of the -RHP.
Lemma 2.3.
The unique solution of the -RHP given by Definition 2.1 is given by
(2.2) |
where satisfies the orthogonality condition
Proof.
The proof follows along similar lines to [10, Section 2(a)], with adjustments needed for the current case where the orthogonality weight is not contained in a compact set in .
We show that the second row of must be given by Equation (2.2). A similar argument can be carried out for the first row. To declutter notation we will label as for the rest of this proof.
It follows from the asymptotic condition, Equation (2.1c), that the entry of must have leading order as . As is analytic and its jump condition, Equation (2.1a), is given by the identity we immediately conclude that is a monic polynomial of degree . Similarly, it follows that is a polynomial of degree at most . We denote by .
Consider the bottom right entry of Equation (2.2). By the jump condition, Equation (2.1a), we have
(2.3) |
If there was no residue condition, Equation (2.1b), then this scalar equation would be solved by the Cauchy transform
(2.4) |
which is analytic in and satisfies Equation (2.3). The residue condition can be readily resolved by letting
(2.5) |
which satisfies both Equations (2.1a) and (2.1b). The only step remaining is to prove the asymptotic condition, Equation (2.1c), for . Substituting our expression for into Equation (2.5), we find
which by Cauchy’s integral formula, for , becomes
where the sum to infinity is well defined on , as converges as , and the Jackson integral of an analytic function is well defined. Using the geometric series with remainder
(2.6) |
we find
Note that the asymptotic condition, Equation (2.1c), holds when the last term on the RHS is zero for . This is true iff
which is satisfied when is an orthogonal polynomial of degree on the -lattice with respect to the weight . This is the class of polynomials. We conclude that the solution of is given by
After appropriate scaling, and repeating the same arguments for the first row, it follows that Equation (2.2) is a solution of the -RHP given by Definition 2.1.
Uniqueness of this solution follows from consideration of the determinant. Observe that the jump matrix satisfies . It immediately follows that on . Furthermore, by the residue condition, Equation (2.1b), has no poles. Thus, is an entire function. By the asymptotic condition, Equation (2.1c), , and so by Liouville’s theorem, it follows that . This implies exists and is meromorphic in .
Now suppose that there exists a second solution of the -RHP and denote this solution by . If we define , it follows that the jump conditions (and residue conditions) effectively cancel and . Thus, is entire and as . Hence, by Liouville’s theorem . We conclude that and, therefore, there is a single unique solution of the -RHP. ∎
Remark 2.4.
Remark 2.5.
One can repeat the arguments above to show that there is unique solution of Definition 2.1 with a modified function
for some real constant , and that it is given by
where satisfies the orthogonality condition
3. Transformations of RHP
In this section we will transform the RHP given in Definition 2.1 to more easily determine the asymptotics of as . Throughout this section we will assume that is even. First, we introduce some new functions which will be used when transforming the RHP. We show that these functions satisfy certain difference equations.
Definition 3.1.
Define as
(3.1) |
Lemma 3.2.
satisfies the difference equation
(3.2) |
Proof.
Remark 3.3.
By induction using Equation (3.2) we find that for even
(3.3) |
Definition 3.4.
Define , where , as
(3.4) |
Lemma 3.5.
satisfies the difference equation
(3.5) |
Proof.
The proof follows from definition of (by applying the same arguments as in Lemma 3.2). ∎
Lemma 3.6.
The function defined in Equation (1.11) satisfies the difference equation
(3.6) |
Proof.
The proof follows from definition of (by applying the same arguments as in Lemma 3.2). ∎
Remark 3.7.
3.1. RHP transformations
Before proceeding we introduce some notation. Consider the contour , scaled such that the modulus of points on it are multiplied by . We denote this new contour by . (If were the unit circle, would be a circle with radius .) Consider the following transformation to the RHP given by Definition 2.1:
(3.12) |
This gives a new RHP for with two jumps (and also some more poles which will be discussed shortly). At we apply Equation (3.3) to determine
(3.13) |
Motivated by Equation (3.13) we make the transformation
(3.16) |
where
This new matrix satisfies the following RHP:
Definition 3.8 (-RHP).
Let be an appropriate curve (see Definition 1.3) with interior and exterior . A complex matrix function , , is a solution of the RHP if it satisfies the following conditions:
-
-
(i)
is meromorphic in , with simple poles at for .
-
(ii)
has continuous boundary values and as approaches from and respectively, where
(3.17a) -
(iii)
has continuous boundary values and as approaches from and respectively, where
(3.17b) -
(iv)
satisfies
(3.17c) Note that has poles in the second column for . Thus, the decay condition does not hold near these poles. For example: the decay condition holds for such that , for all integer , for fixed .
-
(v)
The residue at the poles for is given by
(3.17d) -
(vi)
The residue at the poles for is given by
(3.17e)
4. Near-field RHP
We will show that the solution of the RHP given in Definition 3.8 approaches a limiting solution . To do this, we are going to solve two separate RHPs, which we will call the near-field RHP and the far-field RHP. These RHPs will be chosen to mimic the two jump conditions satisfied by at and respectively. This section is devoted to the solution of the near-field RHP.
Motivated by the form of Equation (3.17a), we first introduce the following RHP.
Definition 4.1 (-RHP).
Let be an appropriate curve (see Definition 1.3) with interior and exterior . A complex matrix function , , is a solution of the -RHP if it satisfies the following conditions:
-
-
(i)
is meromorphic in , with simple poles in the first column at for .
- (ii)
-
(iii)
satisfies
(4.1b) where is a non-zero constant. Due to the simple poles in the first column of , the asymptotic decay condition only holds for for any .
-
(iv)
The residue at the poles for is given by
(4.1c)
To solve this RHP, a series of Lemmas are required.
Lemma 4.2.
Consider the difference equation
(4.2) |
There exists two entire solutions of Equation (4.2), one even and one odd.
Proof.
Definition 4.3.
Motivated by the entry of the right side of Equation (4.1a), we consider the properties of the product .
Lemma 4.4.
Define
then is a solution of the difference equation
(4.3) |
Furthermore, there exists two solutions to Equation (4.3) analytic in which can be represented by an even and odd power series at infinity.
Proof.
From the definition of and Lemmas 3.2 and 3.6 we find that
Substituting the above into Equation (4.2) we determine that satisfies the difference equation
Let , such a power series is a solution of Equation (4.3) iff
Lemma 4.4 follows immediately (note that the power series for converges everywhere). ∎
Definition 4.5.
Motivated by the entry of the right side of Equation (4.1a) for , we consider the properties of the product .
Lemma 4.6.
Define
then is a solution of the difference equation
(4.4) |
Furthermore, there exists two solutions to Equation (4.4) holomorphic for which can be represented by an even and odd power series at infinity.
Proof.
From the definition of and Equation (3.2) we find that
Substituting the above into Equation (4.2) we determine that satisfies the difference equation
which one can readily show is equivalent to Equation (4.4). Let , such a power series is a solution of Equation (4.4) iff
(4.5) |
For large index , we can deduce from Equation (4.5) that
Taking a telescopic product we conclude that the sum converges if . Lemma 4.4 follows immediately. ∎
Definition 4.7.
Lemma 4.8.
The function , where is defined in Lemma 4.2, can be written as
where are constants. Similarly can be written as
Proof.
From Lemma 4.6, we conclude that satisfies the same difference equation as and . It follows that
for some functions which satisfy . As is analytic everywhere and and are holomorphic for we conclude that is constant or has simple poles at the zeros of , which occur at for . Applying Corollary A.2 to and , and comparing even and odd terms we conclude that
and the first part of Lemma 4.8 follows immediately. The equation for also follows using similar arguments. ∎
Remark 4.9.
Lemma 4.10.
Proof.
From Lemma 4.4 we conclude that satisfies the same difference equation (Equation (4.3)) as and . It follows that
for some functions that satisfy . As and are analytic everywhere and is holomorphic in we conclude that is constant or has simple poles at the zeros of , which occur at for . Applying Corollary A.2 to and , and comparing even and odd terms we conclude that
(4.6) |
and the first part of Lemma 4.8 follows immediately. The equation for also follows using similar arguments. ∎
Remark 4.11.
We note that has poles at for , and . As is analytic at these locations we conclude
Thus,
We are now in a position to solve the RHP given in Definition 4.1.
Lemma 4.12.
Proof.
First we show condition (i) (meromorphicity) is satisfied. With the choice of given by Equations (4.10) and (4.11) it is clear from Remark 4.11 that has analytic entries in the second column for . In particular the second column has entries equal to and for . Meromorphicity of the LHS column follows immediately from the definition of and .
Conditions (ii) and (iv) follow immediately from the definition of . It is left to show condition (iii) is satisfied. Taking the limit of Equation (4.6) and applying Lemma 4.8 we find
Thus,
Let
(4.12) |
Hence, applying Lemma 4.8 again we find that as
We conclude that as the matrix behaves like
(4.13) | |||||
∎
5. Far-field RHP
In this section, we solve the far-field RHP, which we denote by -RHP (see Definition 5.1). The independent variable in the far-field and near-field RHPs are related through a scaling transformation. To distinguish the two, we use instead of to denote a complex variable in this section. Motivated by the form of Equation (3.17b) we introduce the following RHP.
Definition 5.1 (-RHP).
Let be an appropriate curve (see Definition 1.3) with interior and exterior . A complex matrix function , , is a solution of the RHP if it satisfies the following conditions:
-
-
(i)
is meromorphic in , with simple poles at for .
-
(ii)
has continuous boundary values and as approaches from and respectively, where
(5.1a) -
(iii)
satisfies
(5.1b) Due to the simple poles in the second column of , the asymptotic decay condition only holds for for any .
- (iv)
-
(v)
The residue at the poles for is given by
(5.1d)
We will explicitly solve this RHP, using a similar approach to Section 4. To do so, we prove a sequence of Lemmas.
Lemma 5.2.
Consider the difference equation
(5.2) |
There exists a solution of Equation (5.2), analytic in , which can be represented by the even power series
where we take (w.l.o.g.).
Similarly, there exists a solution analytic in , of the difference equation
(5.3) |
which can be represented by the odd power series
without loss of generality let .
Proof.
Lemma 5.3.
Define
where satisfies the difference equation given by Equation (5.2) and is defined in Equation (3.4). Then is a solution of the difference equation
(5.4) |
Furthermore, there exists two entire solutions of Equation (5.4) which can be represented by an even and odd power series.
Similarly, define
where satisfies the difference equation given by Equation (5.3), then satisfies the difference equation
(5.5) |
Furthermore, there exists two entire solutions which can be represented by an even and odd power series.
Proof.
From the definition of , Equations (3.2) and (3.5) we find that
Substituting the above into Equation (5.2) we determine that satisfies the difference equation
Let , such a power series is a solution of Equation (5.4) iff
The first part of Lemma 5.3 follows immediately. The second part follows using similar arguments. ∎
Definition 5.4.
Lemma 5.5.
Similarly, is a solution of Equation (5.5).
Proof.
Lemma 5.6.
The function , where is defined in Lemma 5.2, can be written as
where are a constants. Similarly can be written as
(5.7) |
Proof.
From Lemma 5.5 we conclude that satisfies the same difference equation (Equation (5.2)) as and . It follows that
for some functions which satisfy . As is analytic in and and are entire we conclude that is constant or has simple poles at the zeros of , which occur at for . Applying Corollary A.2 and comparing even and odd terms we conclude that
and the first part of Lemma 5.6 follows immediately. The equation for also follows using similar arguments. ∎
Lemma 5.7.
Let , and be defined as in Lemma 5.6. Furthermore, define the constant as
(5.8) |
Then,
(5.9) |
as . It is also true that
(5.10) |
for .
Proof.
Remark 5.8.
Repeating the arguments of Lemma 5.7 one can readily show
(5.11) |
We require one last Lemma before determining the solution of the far-field RHP.
Lemma 5.9.
Let be defined as in Lemma 5.3, then
where is a non-zero constant and this limit clearly does not hold near the poles of , but holds for satisfying , for some fixed and all .
Proof.
We first show that the residue of the poles of are vanishing faster than as .
Consider the case (), from Equation (3.3) we find that
By Lemma 5.3 we know that as . Thus, we conclude
Note that the above statement is true for , with .
Consider the case . From Lemma 5.7 it is clear that a bound on as is equivalent to a bound on . By definition in Lemma 5.2 we determine that . Furthermore, using induction on Equation (3.5) we find
It follows
Let be the residue of at . Define the function
where this sum is well defined for all because we have just shown as . It follows is holomorphic in and can be represented by a Laurent series which converges everywhere. We will show that (i.e. there are no positive powers of ). Applying Equation (5.5) and Equation (3.2) we find that satisfies the difference equation
(5.12) |
Writing the above in matrix form we have
(5.13) |
Observe that the eigenvalues of the LHS matrix in the above equation are 1 and 0. Hence, repeatedly applying Equation (5.13) to determine the behaviour of as is essentially a Pochhammer symbol with matrix entries. Thus, , and consequently are bounded by a constant as . We now show that this constant is non-zero. From Lemma 5.3 we know that is an entire function (which is not the constant function), hence must grow in some direction. satisfies Equation (5.5),
It follows that as becomes large there must exist a ray where
Thus, along this ray
and applying Equation (3.2) we conclude approaches a constant along this ray. Thus, and . Lemma 5.9 follows immediately. ∎
Remark 5.10.
Repeating the same arguments as in Lemma 5.9 we can conclude
(5.14) |
where again this limit clearly does not hold near the poles of , but holds for satisfying , for some fixed and all .
Now we are in a position to solve the far-field RHP given in Definition 5.1. Let be given by
(5.15) |
Consider the conditions for this function to solve the far-field RHP. First, we note that condition (i), i.e. meromorphicity, is satisfied as by definition , , and are all analytic in (see Lemmas 5.2 and 5.3). Second, note that condition (ii), the jump condition, holds by direct calculation. Third, to show condition (iii), i.e. asymptotic decay, observe that as by the definition of , similarly as by the definition of . From Lemma 5.9 we conclude that as . Similarly from Remark 5.10 we conclude as . The remaining conditions (iv) and (v), the residue conditions, follow from Lemma 5.7 and Remark 5.8. Furthermore, from Lemmas 5.6, 5.7 and 5.9 we find
(5.16) |
6. Gluing together near- and far-field RHPs
We will now glue together the near- and far-field RHPs to approximate the RHP for as . The near- and far-field variables in Sections 4 and 5 are related by the linear transformation .
We first make a linear transformation to the near-field RHP solution. Let
(6.1) |
From Equations (1.11) and (3.4) one can readily determine that as . Thus, comparing Equations (4.12) and (5.8) we find that . Hence, applying Equations (4.13) and (5.16) we find
We next make a slight modification to given by Equation (5.15). Note that the residue condition for given in Equation (3.17e) is different to that for given in Equation (5.1d). To resolve this issue we define the new function
(6.2) |
Substituting in Equation (3.8) we find that
(6.3) |
Thus, the difference between and is bounded by for ().
Define
(6.6) |
and, furthermore
(6.7) |
Then, satisfies the following RHP.
Definition 6.1 ( RHP).
A complex matrix function , , is a solution of the RHP if it satisfies the following conditions:
-
-
(i)
is analytic in .
-
(ii)
has continuous boundary values and as approaches from and respectively, where
(6.8a) -
(iii)
satisfies
(6.8b)
7. Proofs of main theorems
Having proved Equation (6.9), we are now in a position to prove the first two main theorems of this paper.
Proof of Theorem 1.5.
Proof of Theorem 1.6.
Using the transformations detailed in Section 3.1 we find that
Let,
Note that there is no difference in the term between and . Using the definition of given in Equation (6.7) and Equation (6.9) we find
Similarly in the bottom left term we find in the limit :
Taking the ratio of and we find that
However, we can determine that by considering the arguments presented in Theorem 8.3. ∎
8. Recurrence coefficients and -discrete Painlevé
As discussed in Remark 2.5, the class of monic polynomials satisfying the orthogonality condition
(8.1) |
where , satisfy a corresponding RHP. In this section we discuss the connection between the RHP, the asymptotic behaviour of and uniqueness results concerning their recurrence coefficients. First, we use the RHP to show that in general .
Lemma 8.1.
Let be the class of monic polynomials with orthogonality condition given by Equation (8.1). Then, the two classes of orthogonal polynomials corresponding to the cases and are the same. Furthermore,
Moreover, if then .
Proof.
Let
From the arguments in Section 2 it follows that is meromorphic with simple poles at location of the poles of . At these locations
(8.2) |
Consider the function
from Equation (8.2), Lemma A.3 and Remark A.6 we conclude that is meromorphic with simple poles at for . The residue of these poles is given by
Note that decays much faster than inverse polynomial decay and thus
Hence, the solution for the RHP corresponding to (see Remark 2.5) can also be written as
(8.6) |
Thus, we have shown that . One can readily deduce from Section 2 that
it follows from Remark 2.5,
Finally, we prove that in general , if . Assume to the contrary that . Let, be the solution of the corresponding RHP given in Remark 2.5. Note that the first column of is the same as that in .
By Remark 2.4, we know that the second column of must satisfy the same -difference equation as the second column of , where restricted to . If then by the analyticity of and comparing even and odd terms we conclude the second column of must satisfy
where is a constant. Denoting the -entry of , for , by , we conclude that .
By the analyticity of , we deduce that at the poles of , in particular at , we have
(8.7) |
Furthermore,
(8.8) |
Equations (8.7) and (8.8) can only be satisfied if the ratio is equal at the four points . By Lemma A.3 and Remark A.6, it follows that the equality can only hold at these four points if , which is the desired result. ∎
Theorem 8.2.
Suppose that the sequence of monic polynomials satisfies the orthogonality condition
where is given by Equation (1.11) and . Then, the recurrence coefficients , which occur in the recurrence relation
(8.9) |
solve the equation:
(8.10) |
with initial conditions for . Furthermore,
(8.11) |
Proof.
In order to show Equation (8.11), we observe that
(8.12) |
where we have shifted the index of summation and used Equations (1.1) and (3.6) to obtain the result. Applying the orthogonality condition in Equation (8.12), we find
(8.13) |
where , are constants depending on . We can immediately find explicitly by using the identity
Since the sequence consists of monic polynomials, we obtain
In order to derive , we first note that:
(8.14) |
We now take the -derivative of both sides of Equation (8.9). After gathering linearly independent terms in Equation (8.14) we find
(8.15) |
which leads to two equations for and :
(8.16) | |||||
(8.17) |
Equation (8.17) implies , for some constant .
We now proceed to show that satisfies Equation (8.10). Substituting into Equation (8.16) we find
(8.18) |
We will rearrange Equation (8.18) with the goal of obtaining a telescoping sum. Multiplying Equation (8.18) by , for some constant , we have
(8.19) |
Therefore, we find
Letting , we are led to
Thus, if we choose and take the sum from to we find
It remains to determine and . Observe that
(8.20) |
We now use Equation (8.20) to determine and . Without loss of generality assume that . From Equation (8.20) we find that
(8.21) | |||||
(8.22) |
Using Equation (8.9) we determine
(8.23) | |||||
Furthermore, by the orthogonality of we find
(8.25) | |||||
where we have used the recurrence relation, Equation (8.9), to determine . Applying Equations (8.21) to (8.25) we deduce
These values give Equations (8.10) and (8.11) and Theorem 8.2 follows immediately. ∎
As a consequence of Theorem 8.2 the sequences all provide positive solutions of Equation (8.10). However, we will show that the limits of these sequences, as , are in general not the same. In particular, we show that the asymptotic limit as of does not equal the limit of if .
Theorem 8.3.
Suppose the sequence of orthogonal polynomials and recurrence coefficients are defined as in Theorem 8.2. Then, it follows that
(8.26) |
if .
Proof.
From Lemma 8.1, we have . It remains to study the case . Let and be two solutions of Equation (4.2). Then, is a solution of the matrix equation
(8.27) |
Using the definition of the -derivative, and Equation (8.9) to write in terms of and , Equation (8.11) can be re-written as
(8.28) |
Taking and again using Equation (8.9), Equation (8.28) also allows us to express in terms of and . This results in the matrix difference equation
where
and is given by
(8.29) |
For the case we know that as Equation (8) approaches Equation (8.27) (this follows from the the proof that as , see Section 7). Assume that has the same asymptotic behaviour as . Then Equation (8) must similarly approach Equation (8.27). In particular, as and, , approach non-zero constants. Note that the condition as follows from the first order asymptotic behaviour i.e. as . In order for and to approach the same non-zero constant as the case we require the asymptotic behaviour to match to second order.
We conclude using Remarks 2.4 and 2.5, and similar arguments to those of Lemma 4.8, that for , approaches the matrix
(8.30) |
for some real constants (recall and are defined in Lemma 4.2). By the meromorphicity of at the poles of we find
(8.31) |
Thus,
(8.32) |
Hence, is either real or imaginary. We conclude from Lemma A.5 and Remark A.6 that Equation (8.32) is satisfied iff . It follows that has the same asymptotic behaviour as (i.e. Equation (8.26) is zero) iff . ∎
9. Conclusion
In this paper, we described the asymptotic behaviour of a class of polynomials, defined in Section 1.3, by using the -RHP setting [10]. Our main results are Theorems 1.5, 1.6 and 1.7. In Theorems 1.5 and 1.6, we provided detailed asymptotic results for polynomials. In Theorem 1.7, we detailed the implications of our analysis for the -Painlevé equation satisfied by the recurrence coefficients, , of polynomials (see Equation (1.5)).
Perhaps the most unexpected results of this paper concern the effect of the properties of on the class of orthogonal polynomials that arise when the lattice was varied. The values of at the poles of the weight function play an important role in determining the behaviour of orthogonal polynomials supported on the shifted lattice , . This observation enabled us to determine whether the class of orthogonal polynomials were invariant as varies. Furthermore, we were able to compare the variation in the asymptotic behaviours of polynomials when the lattice was shifted.
This paper focused on the weight . But, the methodology can readily be extended to describe discrete -Hermite II polynomials [3, Chapter 18.27] with weight . However, generalising the results to higher order weights i.e. , for , remains an open problem. The key difficulty is describing and solving the near-field RHP for higher order weights. One important aspect of this problem is accounting for the increased number of poles when dealing with higher order weights.
Another possible direction of future research could be determining all of the positive solutions of Equation (1.7) and the asymptotic behaviour of different solutions as . In this paper we described the asymptotic behaviour of one particular solution, which satisfies the limit
(9.1) |
Numerical evidence suggests that the RHS of Equation (9.1) oscillates for other positive solutions of Equation (1.7). It would be interesting to see if the -RHP formalism is able to accurately capture this behaviour. One avenue to achieve this would be to extend the detailed asymptotic results obtained in this paper to polynomials, defined in Section 1.3.
Appendix A Properties of
In this section we prove some properties of the function defined in Equation (1.2), which are used in this paper. Before discussing we first prove a necessary Lemma.
Lemma A.1.
Let be a function defined on , which is analytic everywhere except for simple poles at for . Then, .
Proof.
We prove the result by contradiction. Assume . Define
By direct calculation one can show . Furthermore, by definition, is zero on the -lattice , . Let
then it follows is analytic in and satisfies the difference equation
(A.1) |
As is analytic in we can write as the Laurent series
Comparing the coefficients of in Equation (A.1), one can readily determine
(A.2) |
However, there is only one solution with coefficients given by Equation (A.2) (up to scaling by a constant) and it follows that . Thus, if , then , and has no poles. ∎
Corollary A.2.
Let be a function defined on , which is analytic everywhere except for simple poles at for . Furthermore, suppose satisfies . Then, , where and are constants and is as defined in Definition 1.4.
Proof.
As both and have simple poles at we conclude that there exists a such that
Furthermore, both and are invariant under the transformation , hence for all
Thus, the function
is meromorphic in , with possible simple poles at for , and satisfies . However, by Lemma A.1, can not have simple poles at for . Hence, is analytic in and it follows that can be written as a convergent Laurent series. Thus,
Substituting this into the -difference equation , we conclude and Corollary A.2 follows immediately. ∎
Lemma A.3.
The function has zero real part along the circles and .
Proof.
Let . Substituting this into Equation (1.2) and determining the real part we find
(A.3) |
First we consider the case . The RHS of Equation (A.3) becomes
Next we consider the case . The RHS of Equation (A.3) can be written as
∎
Corollary A.4.
Proof.
Lemma A.5.
Along the ray , the real part of is non-zero except at , for where is complete imaginary.
Proof.
From Equation (A.3) we determine that the real part of is given by
Define the function
where is a complex variable. We note that the sum is well defined and converges for all , where and . Furthermore, satisfies the -difference equation . We now multiply by a function with zeros at the poles of to give us a function analytic in . Define
Note that is an even function and satisfies . Thus, is analytic in , satisfies the -difference equation , and is an odd function. Let us represent with the convergent Laurent series
As satisfies the difference equation we find
Hence, we conclude that there are four linearly independent solutions, which can be chosen such that two are odd and two are even. As is odd we conclude it is the sum of two linearly independent odd solutions. Consider the two functions
We note that has zeros at , for and has zeros at , for . Furthermore, both and are odd and satisfy the difference equation . Thus,
for some constants and . By Lemma A.3 we conclude that and only has zeros at the zeros of which occur at , for . Hence, this is where the zeros of are and Lemma A.5 follows immediately. ∎
Appendix B RHP theory
For completeness we recall a well known result from RHP theory and prove it below. Let be a solution of the following RHP:
Definition B.1.
Let be an appropriate curve (see Definition 1.3) with interior and exterior . A complex matrix function , , is a solution of the RHP (B.1) if it satisfies the following conditions:
-
-
(i)
is analytic in .
-
(ii)
has continuous boundary values and as approaches from and respectively, where
(B.1a) for a matrix .
-
(iii)
satisfies
(B.1b)
Theorem B.2.
Suppose that can be analytically extended to a neighbourhood of . Furthermore, for a given suppose that
in this neighbourhood (where is the matrix norm). Then, the solution of the RHP given by Definition B.1 satisfies
for all .
Proof.
Multiple sources give a proof of Theorem B.2 with various conditions on the jump matrix \citesDeift1999strongkuijlaars2003riemann. For our setting the jump matrix is quite well behaved and satisfies all these constraints. We include a brief proof for completeness. Let
substituting into Equation (B.1a) gives
By the asymptotic condition, Equation (B.1b), we conclude that
(B.2) |
Let be defined as , and let the maximum be at . As is analytic in it follows achieves its maximum on the boundary (i.e. on ). By assumption and are also analytic for some fixed distance from , let us call this curve . Therefore,
(B.3) |
where can be determined using the Nuemann series
We conclude from Equation (B.3)
(B.4) |
and hence . Thus we find that,
(B.5) |
For some constant
which is independent of . ∎
Funding
Nalini Joshi’s research was supported by an Australian Research Council Discovery Projects #DP200100210 and #DP210100129. Tomas Lasic Latimer’s research was supported the Australian Government Research Training Program and by the University of Sydney Postgraduate Research Supplementary Scholarship in Integrable Systems.
References
- [1] L. Boelen, C. Smet and W. Van Assche “-Discrete Painlevé equations for recurrence coefficients of modified -Freud orthogonal polynomials” In Journal of Difference Equations and Applications 16.1, 2010, pp. pp. 37–53
- [2] P. Deift et al. “Strong asymptotics of orthogonal polynomials with respect to exponential weights” In Communications on Pure and Applied Mathematics 52.12 Wiley Online Library, 1999, pp. 1491–1552
- [3] “NIST Digital Library of Mathematical Functions” F. W. J. Olver, A. B. Olde Daalhuis, D. W. Lozier, B. I. Schneider, R. F. Boisvert, C. W. Clark, B. R. Miller, B. V. Saunders, H. S. Cohl, and M. A. McClain, eds., http://dlmf.nist.gov/, Release 1.1.0 of 2020-12-15 URL: http://dlmf.nist.gov/
- [4] T. Ernst “A comprehensive treatment of q-calculus” Springer Science & Business Media, 2012
- [5] A.S. Fokas, A.R. Its and A.V. Kitaev “Discrete Painlevé equations and their appearance in quantum gravity” In Communications in Mathematical Physics 142, 1991, pp. 313–344
- [6] G. Freud “On the Coefficients in the Recursion Formulae of Orthogonal Polynomials” In Proc. Roy. Irish Acade. sect. A 76 Royal Irish Academy, 1976, pp. 1–6
- [7] D. Gómez-Ullate, N. Kamran and R. Milson “An extended class of orthogonal polynomials defined by a Sturm–Liouville problem” In Journal of Mathematical Analysis and Applications 359.1, 2009, pp. 352–367
- [8] Mourad E.H. Ismail and Z.S.I. Mansour “-Analogues of Freud weights and nonlinear difference equations” In Advances in Applied Mathematics 45.4, 2010, pp. 518–547
- [9] N. Joshi “Discrete Painlevé Equations” American Mathematical Soc., 2019
- [10] N. Joshi and T. Lasic Latimer “On a class of -orthogonal polynomials and the -Riemann-Hilbert problem” In Proc. R. Soc. A. 477, 2021
- [11] A. Kuijlaars “Riemann-Hilbert analysis for orthogonal polynomials” In Orthogonal polynomials and special functions Springer, 2003, pp. 167–210
- [12] H. Sakai “Rational surfaces associated with affine root systems and geometry of the Painlevé equations” In Communications in Mathematical Physics 220, 2001, pp. 165–229
- [13] R. Sasaki “Exactly solvable birth and death processes” In Journal of Mathematical Physics 50.10 American Institute of Physics, 2009, pp. 103509
- [14] W. Van Assche “Orthogonal and multiple orthogonal polynomials, random matrices, and Painlevé equations” In AIMS-Volkswagen Stiftung Workshops Springer, 2018, pp. 629–683