Asymptotic behaviours of -orthogonal polynomials from a -Riemann Hilbert Problem
Abstract.
We describe a Riemann-Hilbert problem for a family of -orthogonal polynomials, , and use it to deduce their asymptotic behaviours in the limit as the degree, , approaches infinity. We find that the -orthogonal polynomials studied in this paper share certain universal behaviours in the limit . In particular, we observe that the asymptotic behaviour near the location of their smallest zeros, , and norm, , are independent of the weight function as .
Key words and phrases:
Riemann-Hilbert Problem, -orthogonal polynomials and -difference calculus. MSC classification: 33C45, 35Q15, 39A13.1. Introduction
The theory of orthogonal polynomials is a source of major developments in modern mathematical physics. But the spectacular outcomes of the classical theory of orthogonal polynomials with continuous measure are not yet matched by -orthogonal polynomials, which are orthogonal with respect to a discrete measure supported on a lattice , for some . In this paper, we focus on a class of such polynomials and deduce their asymptotic behaviours as their degree grows by expressing them in terms of a Riemann-Hilbert Problem (RHP).
We denote monic -orthogonal polynomials by . They satisfy the orthogonality relation
(1.1) |
where refers to the (discrete) Jackson integral (see Equation (1.4)). A fundamental consequence of Equation (1.1) is the 3-term recurrence relation
(1.2) |
where the recurrence coefficients are given by
A natural question, which motivated many studies through the past century [32, 9, 16], is to ask what the behaviours of and are as Classical results focused on polynomials which satisfy the orthogonality relation
where is a continuous measure on the real line and is a weight function whose rate of change satisfies certain conditions. Measures on the unit circle in the complex plane were also a focus of interest [32, 31]. Extensions to a wider class of weight functions, leading to so-called semi-classical orthogonal polynomials [30, 15], have attracted more attention in recent times, due to their appearance in Random Matrix Theory [24, 7] and relationship to the Painlevé equations [14].
More recently, discrete orthogonal polynomials have been of growing interest. For those on a multiplicative -lattice, particular attention has been paid to cases such as little -Jacobi and discrete -Hermite I [10, Chapter 18.27] polynomials. However, very little appears to be known about the asymptotic behaviour of -orthogonal polynomials outside these specific cases. More recent discoveries of -orthogonal polynomials related to multiplicative discrete Painlevé equations \citesfilipuk2018discrete,Boelen have reignited a need for further mathematical tools to answer questions about their asymptotic behaviours.
The aim of this paper is to consider such questions for a general class of -orthogonal polynomials which includes, but is not limited to, a large subset of the -Hahn class [10, Chapter 18.27]. Our main results are Theorems 1.5 and 1.6, where, under certain mild assumptions on the weight in Equation (1.1), we deduce the asymptotic behaviour of , and in the limit and show that the error term is of size .
1.1. Background
In the past two decades, -orthogonal polynomials have appeared in many areas of applied mathematics and physics \citesatakishiyev2008discrete,jafarov2010quantum,sasaki2009exactly, knizel2016moduli, particularly in quantum physics. However, little is known about the behaviour of -orthogonal polynomials. Earlier work in the field focused on specific examples. In 2003, Postelmans and Van Assche introduced two kinds of multiple little -Jacobi polynomials and described some asymptotic properties [25]. In 2005, Ismail described the asymptotic behaviour of discrete -Hermite II polynomials using -Airy functions. He then extended these results to -orthogonal polynomials satisfying a certain -difference equation [16]. In 2013, Driver and Jordaan studied the asymptotic behaviour of extreme zeros of -orthogonal polynomials [11]. In 2017, Chen and Filipuk studied generalised -Laguerre polynomials and determined resulting asymptotics for their recurrence coefficients [6].
Recently, there has been improved understanding of the asymptotics of larger families of -orthogonal polynomials. In 2020, Van Assche et al. [33] showed that the leading order of is for -orthogonal polynomials with weights satisfying
where . They also describe the location of (where are the zeros of ) in the limit . However, it remained an open question to obtain a more precise asymptotic description of large classes of -orthogonal polynomials.
RHPs have been extensively used to study the asymptotics of orthogonal polynomials \citeskuijlaars2003riemann,etna_vol25_pp369-392 since the asymptotic behaviour of semi-classical Freudian polynomials was derived by Deift et al. using a model RHP [9]. Their work was based on earlier advancements by Deift and Zhou on the steepest descent method for oscillatory RHPs [8]. Expanding on the approach of Deift et al., Baik et al. [3] deduced the asymptotics of orthogonal polynomials on a discrete lattice using what they call an interpolation problem, which can be seen as the discrete analogue to a RHP. Although this work yielded interesting results for general discrete weights, we find that it misses some key details of the behaviour of -orthogonal polynomials. In particular, the results do not accurately describe the behaviour of , and as . We note that in terms of Baik et al.’s notation, is an accumulation point of the -lattice.
Extending on our earlier theory [19], in this paper we will use a RHP to obtain detailed asymptotic results for a large class of -orthogonal polynomials. We will observe an interesting intersection with -RHP theory and provide explicit examples highlighting aspects of -RHP theory discussed in the literature [26, 29, 1]. In particular, we will solve the model -RHP by deducing an equivalent connection matrix between two solutions of a -difference equation represented by a power series about and .
The asymptotic results obtained in this paper also pertain to multiplicative discrete Painlevé equations. It has been shown that the recurrence coefficients of -orthogonal polynomials can satisfy multiplicative-type discrete Painlevé equations [4], for example Equation (1.3)
(1.3) |
where the non-autonomous term in the equation is iterated on multiplicative lattices. (For the terminology distinguishing types of discrete Painlevé equations, we refer to Sakai [27].)
Very little is known about the asymptotic behaviour of the solution to this equation. The results in this paper provide detailed asymptotics for the real positive solution to Equation (1.3).
1.2. Notation and previous results in the literature
For completeness, we recall some well known definitions and notations from the calculus of -differences. These definitions can be found in [12]. Throughout the paper we will assume and .
Definition 1.1.
We define the Pochhammer symbol , and Jackson integrals as follows.
-
(1)
The Pochhammer symbol is defined as
Furthermore, we define as
-
(2)
The unnormalised Jackson integral of from -1 to 1 is defined as
(1.4)
We remark on an equivalence between two types of -orthogonal polynomials seen in the literature.
Remark 1.2.
In general -orthogonal polynomials can be orthogonal with respect to a weight supported on the Jackson integral from or from [10, Chapter 18.27], where the latter is given by
Let be the normalised polynomials orthogonal with respect to the one-sided Jackson integral
Let , hence
(1.5) |
Define , it follows that is an even function. This implies that the corresponding set of orthogonal polynomials are even/odd for even/odd [17]. Thus, for positive integers , the orthogonality condition for is given by
(1.6) |
which is equivalent to Equation (1.5) up to scaling of the normalisation factor by . Hence, we proved that the class of -orthogonal polynomials with one-sided Jackson integrals are contained in the class of -orthogonal polynomials with two-sided Jackson integrals. It follows that it is sufficient to study -orthogonal polynomials with two-sided Jackson integrals.
We recall the definition of an appropriate Jordan curve and admissible weight function given in [19, Definition 1.2] (with slight modification).
Definition 1.3.
A positively oriented Jordan curve in with interior and exterior is called appropriate if
A weight function, , is called admissible if there exists an appropriate Jordan curve, , such that is bounded on , is analytic in , and, there exists constants and such that for
Furthermore, we require that
We define the function
(1.7) |
which satisfies the -difference equation
(1.8) |
Note that is equivalent to the definition of given in [19, Definition 1.3]. Consequently, we define the -RHP:
Definition 1.4 (-RHP).
Let be an appropriate curve (see Definition 1.3) with interior and exterior , and be a corresponding admissible weight. A complex matrix function , , is a solution to the -RHP if it satisfies the following conditions:
-
(i)
is analytic on .
-
(ii)
has continuous boundary values and as approaches from and respectively, where
(1.9a) -
(iii)
satisfies
(1.9b)
1.3. Main results
We are now in a position to state the main results of this paper, which are collected as Theorems 1.5 and 1.6. 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 -orthogonal polynomials, orthogonal with respect to the weight , where and is an admissible weight function. Define . Then for any given positive integer there exists an and such that for even , we have
and
where is a function of , independent of , and and , given in Definition 2.3, are entire functions independent of .
Our second main result concerns the asymptotic behaviour of recurrence coefficients and norm of as approaches infinity.
Theorem 1.6.
Under the same hypotheses as Theorem 1.5 we have, for even , as :
Remark 1.7.
Theorem 1.6 gives information about as , but the methodology we present in this paper does not provide a similar level of information about the asymptotic behaviour of as . The reason lies in the fact that the model RHP (see Section 3) has a solution with an expansion of the form , as , where has zero main diagonal. This diagonal is where would typically appear and so our approach is only able to show that as , without further information about the rate at which vanishes.
Remark 1.9.
Remark 1.10.
There is a number of generalisations one can make to the results in this paper that require a slight change in methodology and lead to slightly different final results.
-
(i)
In Definition 1.3, the condition on admissible weights that there exists constants and c such that
for , can be relaxed to . However, this could result in a change of the asymptotic error terms in Theorems 1.5 and 1.6 (see the proof of Lemma 2.7). In general, is the smallest upper bound achievable on the error.
- (ii)
-
(iii)
The methodology presented in this paper can readily be extended to the Al-Salam Carlitz class of polynomials described in [17, Chapter 18]. In this case the discrete measure is supported on for , where is a constant. To enable such an extension, we need new functions
that replace and respectively. Furthermore, the -difference equation
should be used instead of Equation (3.1a). Repeating the methodology presented in this paper with these substitutions leads to similar asymptotic estimates.
1.4. Outline
The paper is structured as follows. In Section 2.1 we make a series of transformations to the the -RHP given in Definition 1.4. This motivates the form of a model RHP by taking the limit of the RHP defined by Equation (2.11). Consequently in Section 2.2, we prove that the solution of the -RHP approaches the solution to the model RHP and use this to prove Theorems 1.5 and 1.6. In Section 3, using -difference calculus we show that there exists a unique solution to the model RHP and determine its form. In Appendix A we prove important properties about the solution to the model RHP. In Appendix B we motivate some of the arguments presented in this paper using discrete -Hermite I polynomials as an example. In Appendix C we prove certain properties of required in Section 3.
2. Proofs of main results
In this section, we provide the proofs of Theorems 1.5 and 1.6. To carry out the proofs, we rely on a sequence of transformations to the RHP described in Definition 1.4. This sequence ends with a limiting RHP, referred to as a model RHP, which is studied further in Section 2.2 to deduce our main results.
2.1. Deriving the model RHP as , for even
We make a series of transformations to the RHP given by Definition 1.4. Recall (see Equation (1.10)) is the solution of this RHP. Inspired by a similar approach first described by Deift et al., we make the series of transformations
which will enable us to deduce a model RHP governing , such that as .
We will use the functions and in the sequence of transformations, where
(2.1) |
and
(2.2) |
It can verified by direct calculation that satisfies the -difference equation:
(2.3) |
and satisfies the -difference equation:
(2.4) |
By induction, using Equation (2.3), we find for even
To be concise, let us define as
(2.5) |
Thus,
(2.6) |
Furthermore,
it follows that for a fixed , as .
The transformations consist of four steps.
-
I.
We first define:
Note that the zeros of cancel with the simple poles of , at for . This allows us to deform the contour, , so that the poles of at for can lie in ext() without affecting analyticity of the solution (see [19, Section 2(a)] for a description of the holomorphicity of ). We observe that does not change the asymptotic condition; see Equation (1.9b).
-
II.
We now scale the contour so that the modulus of points on it are multiplied by . (If were the unit circle, it would now be a circle with radius .) Denote the new contour by .
-
III.
The next transformation is
Note we have now introduced simple poles at for .
-
IV.
Our final transformation is
(2.9) where is a constant, to be defined shortly. We observe that after these transformations has the asymptotic condition
Motivated by the form of Equation (2.6) we set
After these transformations we are left with the following transformed RHP for which the complex matrix function , defined in Equation (2.9), is the solution:
-
-
(i)
is meromorphic in with simple poles at for .
- (ii)
-
(iii)
satisfies
(2.10b) -
(iv)
The residue at each pole for , is given by
(2.10c)
We now make the change in variables . Furthermore, we scale the orthogonal weight, , by . It follows from the RHP above for that solves the following RHP:
-
-
(i)
is meromorphic in , with simple poles at for .
-
(ii)
has continuous boundary values and as approaches from and respectively, where
(2.11a) -
(iii)
satisfies
(2.11b) -
(iv)
The residue at each pole for , is given by
(2.11c)
As seen in the statement of Theorems 1.5 and 1.6, we are interested in orthogonal weights which satisfy as . Taking the limit of the RHP for , motivates the following model RHP.
Definition 2.1 (Model RHP).
Assume that the contour and regions satisfy the conditions of Definition 1.3.
-
-
(i)
is meromorphic in , with simple poles at for .
-
(ii)
has continuous boundary values and as approaches from and respectively, where
(2.12a) -
(iii)
satisfies
(2.12b) However we also have that has poles in the LHS column for . Thus, the decay condition does not hold near these poles. For example the decay condition holds for such that , for all , for fixed .
-
(iv)
The residue at the poles for is given by
(2.12c)
In Section 3 we prove that there exists a unique solution to the model RHP. We now show that in the limit .
Remark 2.2.
In Section 3 we show restricted to can be analytically extended for to a matrix with entire entries. Let us denote this function as (note that = for ). We also show that , restricted to , can be meromorphically extended for to a matrix with simple poles for entries in the LHS column at , for . Let us denote this function as (note that = for ). For all the identity
holds. An analogous statement holds for such that
Definition 2.3.
We define , , and as the , , and entries of respectively.
In Section 3 we show that these four functions are entire and can be explicitly written in terms of a power series about 0.
2.2. Proofs of Theorems 1.5 and 1.6
We prove Theorem 1.5 and Theorem 1.6. First, Theorem 1.5 is proved by showing that as . To do this we will construct a RHP, given by Equation (2.14), that has the unique solution such that for , where is defined shortly. We will show that as , Theorem 1.5 and Theorem 1.6 then follow immediately.
Before stating the RHP for we define a number of identities.
Definition 2.4.
Define the piece-wise Jordan curve as , where:
-
•
, where is a free parameter, which will later be restricted in Theorem 2.9
-
•
, where there is a large degree of freedom in choosing . It is sufficient to choose such that the contours do not intersect and the orthogonality weight is analytic in .
-
•
, where there is a large degree of freedom in choosing . It is sufficient to choose such that the contours do not intersect and the orthogonality weight is analytic in .
See Figure 2.1 for an illustration of . Note that is composed of an infinite union of circles whilst is composed of a finite union .
Definition 2.5.
Define the three matrix functions:
(2.13a) | ||||
(2.13b) | ||||
(2.13c) |
In general these matrices have meromorphic entries with simple poles at , for .
We now prove the following lemma.
Lemma 2.6.
The unique solution to the RHP:
-
-
(i)
is analytic in , where is described above and illustrated in Figure 2.1,
-
(ii)
satisfies
(2.14a) where , and ,
-
(iii)
(2.14b)
is given by,
(2.15) |
Proof.
Existence. By definition is analytic in ext() (as and are analytic in ext()). Thus, , as defined in Equation (2.15), is analytic in ext(). We are left to show that (given by Equation (2.15)) is analytic in int().
First we look at the region . The matrix is defined in Equation (2.13a) as
By definition, which is analytic in int().
Next we look at . By definition
From the residue condition for , given by Equation (2.11c), we conclude that is analytic in int().
Finally we consider . By definition
From the residue condition for , given by Equation (2.11c), we know that is analytic for and thus is analytic in int().
Uniqueness. We note that . It follows that , and applying the same arguments as in Section 3.2 we conclude that if a solution exists to the RHP given by Equation (2.14), then it is unique.
∎
We now prove that under certain conditions the solution, , to the RHP given by Equation (2.14) approaches the identity. We first prove a lemma about the jump matrix .
Lemma 2.7.
Proof.
First we consider . By the asymptotic condition, Equation (2.12b), we know that
(2.19) |
where the radius can be ignored as we will be considering the case . Applying Equation (2.19) and Definition 2.4 we find
By definition , hence we conclude
Next we study . Applying Equation (2.19) we find for
Furthermore, we observe that the matrix
vanishes much faster than . To see this, recall from Equation (2.5) that
Thus, is vanishingly small for large . (To see this, expand out the first few terms of the product ). Hence we conclude that
It follows there exists an such that for .
Next we study . Define the function
(2.20) |
where and are given in Equations (2.5) and (2.2) respectively. For large , we can take a Taylor series expansion of to find
(2.21) | |||||
Furthermore, define
Then, expanding Equation (2.13a) we find
where are defined in Definition 2.3. We note that . Applying Equation (2.21) it is clear that for a fixed , there exists an such that for
where is a function of . Note that we choose because there exists poles of the jump function, , at for integer values of . ∎
Remark 2.8.
Lemma 2.7 holds for lying on . However, from Equations (2.13a), (2.13b) and (2.13c), the matrix functions are well defined for all . In general, the matrices do not approach the identity everywhere, but, they will have simple poles of order at , for . Furthermore,
This follows by direct computation, observing that and are vanishingly small for large , and applying Equation (A.1) which demonstrates that and also become vanishingly small for large positive integer values of .
Having proved Lemma 2.7 we are now in a position to show that the solution, , to the RHP given by Equation (2.14) approaches the identity.
Lemma 2.9.
Proof.
Define
It immediately follows that for in
By the asymptotic condition, Equation (2.14b), we conclude that
Let be defined as . As is analytic in it follows achieves its maximum on the boundary (i.e. on ). Therefore,
for some on . Furthermore, for , as the points lie in for . As , we can also determine using the Neumann series
Thus we find that,
(2.23) | |||||
It follows from Remark 2.8 the sum on the RHS of Equation (2.23) converges and
Thus, Equation (2.23) gives:
(2.24) |
It follows there exists an such that for , . Hence, we have just determined an upper bound for inside int(). Observing that , Equation (2.2) implies that for any fixed radius
(2.25) |
By definition for lying on ,
(2.26) |
Recall that achieves its maximum on . Applying Equations (2.25) and (2.26) and Lemma 2.7, where we showed , we conclude . ∎
Proof of Theorem 1.5.
For the result follows immediately from Lemma 2.9. Lemma 2.9 implies that
for large . Theorem 1.5 follows immediately after reversing the transformations .
For , we observe that Equation (2.25) implies that
for (for some fixed ). We also observe that is bounded for and , and goes to zero as . Furthermore, Equation (A.2) implies that the poles of and vanish for large , much faster than the function defined in Equation (2.1) grows. As,
this allows one to more accurately describe the behaviour of , , as . ∎
We now prove Theorem 1.6.
Proof of Theorem 1.6.
. Theorem 1.6 follows from Lemma 2.9. We note that in transforming from Equation (2.10) to Equation (2.11) the weight function was scaled by . Let . Substituting in
we can evaluate the expression
using the transformations detailed in Section 2.1. We note that the function , defined in Equation (2.3), is even and does not impact on the coefficient during the transformations. Let,
In Section 3 we show that this is a valid representation of the solution for large . Applying Lemma 2.9 we observe that . Comparing coefficients of the power in the top right term we find in the limit :
Similarly in the bottom left term we find in the limit :
The constants and can be evaluated by observing that generalised discrete -Hermite I polynomials must satisfy these asymptotic conditions. Thus, and . Theorem 1.6 follows.
∎
3. On the existence of a unique solution to the model RHP
In this section we prove that there exists a unique solution to the model RHP given by Equation (2.12).
3.1. Existence
We first show that there exists a solution to the model RHP. This is achieved by determining the connection matrix [5] between three solutions of a -difference equation (Equation (3.1b)), , and (defined shortly). We prove that the connection matrix is equivalent to the jump condition, Equation (2.12a), of the model RHP. Consequently, we show that the model RHP is satisfied by , and , after appropriate transformations. To begin, let us consider the two -difference equations
(3.1a) | |||
(3.1b) |
where and are vectors with complex entries and is a real parameter. Note that: . We motivate the form of Equation (3.1a) in Appendix B. Writing the entries of as a power series in , we find by direct substitution into Equation (3.1a) that can be written in terms of the odd power series
where
Likewise, can be written as an even power series
where
From the recurrence relations we can deduce that both entries of are entire. To see this, observe that for
Similarly, can be written in terms of power series which converge everywhere. However, in this case is odd and is even.
Now consider the -difference equation
(3.1c) |
Note the similarity to Equation (3.1b). One can readily show that there exists a solution to Equation (3.1c), which can be represented by a power series at infinity
(3.2) |
which converges everywhere (except obviously at 0). Earlier, in Equation (2.2) we defined the even function , which satisfies
We also earlier defined the function in Equation (1.7), which satisfies the -difference equation
As both and satisfy the -difference equation (3.1b) we conclude that
(3.3) |
where and . This is the equivalent to a column of the connection matrix in [5]. As is a meromorphic function with simple poles at for we conclude that must be a constant and must be either be a constant or a meromorphic function with simple poles at . Thus, by Corollary C.2 we conclude
where and are constants, and is defined in Equation (1.7). Comparing odd and even terms in Equation (3.3) we conclude that and . Thus, after absorbing constants into the power series of and , we have shown,
(3.4) |
and
Furthermore, in Section A we show that satisfies the asymptotic condition,
where is a constant. Hence, in summary we have proved
(3.5) |
where and are analytic everywhere and is analytic everywhere except . Furthermore, we have proved the asymptotic behaviour
(3.6) |
Thus, after appropriate scaling we have found a solution which satisfies condition of the model RHP: by the holomorphicity of , and , condition : by Equation (3.5), condition : by Equation (3.6), and condition : by Equation (3.5).
3.2. Uniqueness
Uniqueness follows by considering the determinant of a solution, , to the model RHP, Equation (2.12). By the residue condition, Equation (2.12c), we can deduce that det() is analytic in . Furthermore, by the jump condition, Equation (2.12a), we deduce that det() is entire. Applying Louiville’s theorem we conclude that the asymptotic condition, Equation (2.12b), implies that everywhere.
Suppose that there exists another solution, to the model RHP. By the residue condition, Equation (2.12c), is analytic everywhere. Furthermore, the jump conditions cancel and we can conclude is entire. Applying Louiville’s theorem we conclude that the asymptotic condition, Equation (2.12b), implies that . Thus, .
4. Conclusion
In this paper, we determined the asymptotic behaviour of a general class of -orthogonal polynomials by using the -RHP setting [19]. The work is motivated by the methods developed by Deift et al. [9], which used the RHP setting to determine the asymptotic behaviour of semi-classical orthogonal polynomials. The main results are Theorems 1.5 and 1.6 which provide more detailed asymptotic results for a large class of -orthogonal polynomials than we could find in the literature.
There are a number of observations we can make from the results of this paper. In particular we proved that is only dependent on the , where is the orthogonality measure. Furthermore, we note that the results in this paper hold even if for some . That is we do not require positivity of the weight function. When determining a solution to the model RHP we observed some interesting examples of -RHP theory. For example we demonstrated how to explicitly determine a connection matrix between two solutions of a -difference equation and, in Appendix A, also saw the relationship between -Borel transforms and divergent power series arising in -difference equations.
An interesting avenue for future exploration would be to extend the results of the current paper to a larger class of -orthogonal polynomials. Another possible direction could be determining if the theory presented in this paper can be applied to other settings not just -discrete weights and orthogonal polynomials.
Appendix A Properties of the solution to the model RHP
In this section we prove some important properties of the solution to the RHP given by Equation (2.12). These results are used in Sections 2 and 3. The section concludes with a remark which highlights an interesting connection between the present work and -Stoke’s phenomena. Note this is a side observation and not necessary for the proofs of the main results of this paper. As shown in Section 3 studying the solution to the RHP given by Equation (2.12) is equivalent to studying the solutions , and of the -difference equations given in Section 3.
Definition A.1.
For conciseness, we will adopt the notation throughout the appendix.
Let
where and satisfy Equations (3.1a) and (3.1b) respectively. From Equation (3.1a) it can be deduced that satisfies the second order -difference equation
Let us consider evaluated at , for large integer , note that these locations coincide with the poles of and the zeros of defined in Equations (1.7) and (2.2) respectively. If
then this implies that grows as . Now consider the jump condition given in Equation (3.4)
At the points of interest, , it can readily be verified by direct calculation from Equations (2.4) and (1.8) that satisfies the -difference equation:
Note that and has a simple pole at for . If grows as then grows as , but the first entry in , , decays as for large which is a contradiction (remembering that as at our points of interest). Thus, shrinks like . Hence,
(A.1) |
at . Note that this indicates that the residue of the poles of are rapidly shrinking, such that
(A.2) |
Furthermore, as is an entire function by Liouville’s theorem there must be a such that
Thus, it follows that along this ray (given by iterating ) in the complex plane,
which importantly means that
(A.3) |
Note that is a meromorphic function, of the form
with vanishingly small poles (Equation (A.2)). Hence, Equation (A.3) implies that is non-zero and approaches a non-zero constant for large . A similar argument for the second entry of shows that
for , (with fixed ).
Remark A.2.
By solving the -discrete equation satisfied by one can determine a divergent power series representation for at infinity. Taking a -Borel transformation [26] we expect this series to represent the presence of a theta function ‘switching’, analogous to Stokes phenomena. This is reflected in the vanishing poles found in Equation (A.2).
Appendix B -Hermite -difference equation
In this section we motivate the form of Equation (3.1a) by studying the -difference equation satisfied by discrete -Hermite I polynomials. Discrete -Hermite I polynomials satisfy the -difference equation:
where
After making the transformation we find
After taking the linear transformation
We find that satisfies the -difference equation
(B.1) |
Taking the limit the -difference equation for becomes,
(B.2) |
We would expect that the solution to this difference equation solves the model RHP for the case , and indeed that is what we find.
Appendix C Functions invariant under
In this section we prove some properties about meromorphic functions with simple poles which are invariant under the transformation , i.e. .
Lemma C.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
(C.1) |
As is analytic in we can write as the Laurent series
Comparing the coefficients of in Equation (C.1), one can readily determine
(C.2) |
However, there is only one solution with coefficients given by Equation (C.2) (up to scaling by a constant) and it follows that . Thus, if , then , and has no poles. ∎
Corollary C.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 Equation 1.7.
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 C.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 C.2 follows immediately. ∎
Acknowledgment
The authors would like to thank Dr. Pieter Roffelsen for helpful discussions during the inception of the paper.
Funding
Nalini Joshi’s research was supported by 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] C.R. Adams “On the linear ordinary q-difference equation” In Annals of mathematics JSTOR, 1928, pp. 195–205
- [2] N.. Atakishiyev, A.. Klimyk and K.. Wolf “A discrete quantum model of the harmonic oscillator” In Journal of Physics A: Mathematical and Theoretical 41.8 IOP Publishing, 2008, pp. 085201
- [3] J. Baik, T. Kriecherbauer, K. McLaughlin and P.D. Miller “Discrete Orthogonal Polynomials: Asymptotics and Applications” Princeton University Press, 2007
- [4] 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
- [5] R.D. Carmichael “The General Theory of Linear q-Difference Equations” In American Journal of Mathematics 34, 1912, pp. 147–168
- [6] H. Chen, G. Filipuk and Y. Chen “Nonlinear difference equations for the generalized little q-Laguerre polynomials” In Journal of Difference Equations and Applications 23.12 Taylor & Francis, 2017, pp. 1943–1973
- [7] P. Deift “Orthogonal Polynomials and Random Matrices: A Riemann-Hilbert Approach” 3, Courant Lecture Notes in Mathematics Amer. Math. Soc., Providence, RI, 1999
- [8] P. Deift and X. Zhou “A steepest descent method for oscillatory Riemann-Hilbert problems. Asymptotics for the MKdV equation” In Ann. of Math. 137.2 JSTOR, 1993, pp. 295–368
- [9] P. Deift et al. “Strong asymptotics of orthogonal polynomials with respect to exponential weights” In Communications on Pure and Applied Mathematics: A Journal Issued by the Courant Institute of Mathematical Sciences 52.12 Wiley Online Library, 1999, pp. 1491–1552
- [10] “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/
- [11] K. Driver and K. Jordaan “Inequalities for extreme zeros of some classical orthogonal and q-orthogonal polynomials” In Mathematical Modelling of Natural Phenomena 8.1 EDP Sciences, 2013, pp. 48–59
- [12] T. Ernst “A comprehensive treatment of q-calculus” Springer Science & Business Media, 2012
- [13] G. Filipuk and W. Van Assche “Discrete orthogonal polynomials with hypergeometric weights and Painlevé VI” In SIGMA. Symmetry, Integrability and Geometry: Methods and Applications 14 SIGMA. Symmetry, IntegrabilityGeometry: MethodsApplications, 2018, pp. 088
- [14] 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
- [15] G. Freud “On the coefficients in the recursion formulae of orthogonal polynomials” In Proc. Roy. Irish Acade. sect. A, 1976, pp. 1–6 JSTOR
- [16] M.E.H. Ismail “Asymptotics of q-orthogonal polynomials and a q-Airy function” In International Mathematics Research Notices 2005.18 Hindawi Publishing Corporation, 2005, pp. 1063–1088
- [17] M.E.H. Ismail and W. Van Assche “Classical and Quantum Orthogonal Polynomials in One Variable” Cambridge university press, 2005
- [18] E.I. Jafarov and J. Van der Jeugt “Quantum state transfer in spin chains with q-deformed interaction terms” In Journal of Physics A: Mathematical and Theoretical 43.40 IOP Publishing, 2010, pp. 405301
- [19] N. Joshi and T. Lasic Latimer “On a class of -orthogonal polynomials and the -Riemann- Hilbert problem” In Proc. R. Soc. A. 477, 2021 DOI: 10.1098/rspa.2021.0452
- [20] A. Knizel “Moduli spaces of q-connections and gap probabilities” In International Mathematics Research Notices 2016.22 Oxford University Press, 2016, pp. 6921–6954
- [21] R. Koekoek, P.. Lesky and R.. Swarttouw “Hypergeometric orthogonal polynomials and their q-analogues” Springer Science & Business Media, 2010
- [22] A. Kuijlaars “Riemann-Hilbert analysis for orthogonal polynomials” In Orthogonal polynomials and special functions Springer, 2003, pp. 167–210
- [23] A. Martínez-Finkelshtein “Szegő polynomials: a view from the Riemann-Hilbert window” In Electron. Trans. Numer. Anal. 25, 2006, pp. 369–392
- [24] M.. Mehta “Random matrices” Elsevier, Amsterdam, 2004
- [25] K. Postelmans and W. Van Assche “Multiple little q-Jacobi polynomials” In Journal of computational and applied mathematics 178.1-2 Elsevier, 2005, pp. 361–375
- [26] J. Ramis, J. Sauloy and C. Zhang “Local analytic classification of q-difference equations” Société mathématique de France, 2013
- [27] 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
- [28] R. Sasaki “Exactly solvable birth and death processes” In Journal of Mathematical Physics 50.10 American Institute of Physics, 2009, pp. 103509
- [29] J. Sauloy “Galois theory of Fuchsian -difference equations” In Annales Scientifiques de l’Ecole Normale Supérieure 36.6, 2003, pp. 925–968
- [30] J. Shohat “A differential equation for orthogonal polynomials” In Duke Mathematical Journal 5, 1939, pp. 401–417
- [31] B. Simon “Orthogonal polynomials on the unit circle: New results” In International Mathematics Research Notices 2004.53 OUP, 2004, pp. 2837–2880
- [32] G. Szegő “Orthogonal polynomials” American Mathematical Soc., 1939
- [33] W. Van Assche and Q. Van Baelen “Zero Distribution of Orthogonal Polynomials on a q-Lattice” In Constructive Approximation Springer, 2020, pp. 1–28