This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

Asymptotics of Discrete qq-Freud II\mathrm{II} orthogonal polynomials from the qq-Riemann Hilbert Problem

Nalini Joshi1 [email protected] 1School of Mathematics and Statistics F07, University of Sydney, Sydney NSW 2006, Australia, ORCID ID: 0000-0001-7504-4444  and  Tomas Lasic Latimer2 [email protected] 2School of Mathematics and Statistics F07, University of Sydney, Sydney NSW 2006, Australia, ORCID ID: 0000-0001-6859-7788
Abstract.

We investigate a Riemann-Hilbert problem (RHP), whose solution corresponds to a group of qq-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 qq-discrete lattice and determine the asymptotics of related qq-Painlevé equations.

Key words and phrases:
Orthogonal polynomials, Riemann Hilbert Problem, qq-difference calculus
Key words and phrases:
Riemann-Hilbert Problem, qq-orthogonal polynomials and qq-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 qq-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, qkq^{k}, for kk\in\mathbb{Z}, where 0<q<10<q<1. We also determine some properties of qq-Painlevé equations associated with these qq-orthogonal polynomials.

1.1. Notation

For completeness, we recall some well known definitions and notations from the calculus of qq-differences. These definitions can be found in [4]. Throughout the paper we will assume qq\in\mathbb{R} and 0<q<10<q<1.

Definition 1.1.

We define the Pochhammer symbol, qq-derivative and Jackson integral as follows:

  1. (1)

    The Pochhammer symbol (x;q)(x;q)_{\infty} is

    (x;q)=j=0(1xqj).(x;q)_{\infty}=\prod_{j=0}^{\infty}(1-xq^{j})\,.

    We denote product of Pochhammer symbols in multiple variables by (x1,x2;q)(x_{1},x_{2};q)_{\infty},

    (x1,x2;q)=j=0(1x1qj)(1x2qj).(x_{1},x_{2};q)_{\infty}=\prod_{j=0}^{\infty}(1-x_{1}q^{j})(1-x_{2}q^{j})\,.
  2. (2)

    The qq-derivative is defined by

    Dqf(x)=f(qx)f(x)x(q1).D_{q}f(x)=\frac{f(qx)-f(x)}{x(q-1)}. (1.1)

    Note that

    Dqxn=[n]qxn1,D_{q}x^{n}=[n]_{q}x^{n-1},

    where

    [n]q=qn1q1.[n]_{q}=\frac{q^{n}-1}{q-1}.
  3. (3)

    The Jackson integral from qjq^{j} to qiq^{i} for some integers i<ji<j is given by

    qjqif(x)dqx=k=ijf(qk)qk.\int_{q^{j}}^{q^{i}}f(x)d_{q}x=\sum_{k=i}^{j}f(q^{k})q^{k}.

    The Jackson integral from qj-q^{j} to qiq^{i} for some integers i,ji,j is given by

    qjqif(x)dqx=k=jf(qk)qk+k=if(qk)qk.\int_{-q^{j}}^{q^{i}}f(x)d_{q}x=\sum_{k=j}^{\infty}f(-q^{k})q^{k}+\sum_{k=i}^{\infty}f(q^{k})q^{k}.
Definition 1.2.

In this paper \mathbb{N} 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 Γ\Gamma in \mathbb{C} with interior 𝒟\mathcal{D}_{-}\subset\mathbb{C} and exterior 𝒟+\mathcal{D}_{+}\subset\mathbb{C} is called appropriate if

±qk{𝒟ifk0,(k),𝒟+ifk<0,(k),\pm q^{k}\in\begin{cases}&\mathcal{D}_{-}\quad{\mathrm{i}f}\,k\geq 0,\;(k\in\mathbb{Z}),\\ &\mathcal{D}_{+}\quad{\mathrm{i}f}\,k<0,\;(k\in\mathbb{Z}),\end{cases}

and,

ei(π+2nπ)4qk𝒟+,(n0,1,2,3andk).e^{\frac{i(\pi+2n\pi)}{4}}q^{-k}\in\mathcal{D}_{+},\;(n\in 0,1,2,3\;\mathrm{and}\,k\in\mathbb{N}).
Definition 1.4 ([10]).

Define hq:({0}{±qk}k=)h_{q}:\mathbb{C}\setminus(\{0\}\cup\{\pm q^{k}\}_{k=-\infty}^{\infty})\to\mathbb{C} by

hq(z)=k=2zqkz2q2k=k=(qkzqk+qkz+qk),h_{q}(z)=\sum_{k=-\infty}^{\infty}\frac{2zq^{k}}{z^{2}-q^{2k}}=\sum_{k=-\infty}^{\infty}\left(\frac{q^{k}}{z-q^{k}}+\frac{q^{k}}{z+q^{k}}\right), (1.2)

Note that hq(z)h_{q}(z) satisfies the qq-difference equation

hq(qz)=hq(z).h_{q}(qz)=h_{q}(z). (1.3)

In Appendix A, we show that hq(z)h_{q}(z) has certain unique properties.

1.2. Background

Let {Pn(x)}n=0\{P_{n}(x)\}_{n=0}^{\infty} be a class of monic polynomials which satisfy the orthogonality relation

Pn(x)Pm(x)w(x)𝑑x=γnδn,m,\int_{\mathbb{R}}P_{n}(x)P_{m}(x)w(x)dx=\gamma_{n}\delta_{n,m}, (1.4)

for some weight function w(x)w(x). Equation (1.4) gives rise to the three term recurrence relation

xPn(x)=Pn+1(x)+βnPn(x)+αnPn1(x),xP_{n}(x)=P_{n+1}(x)+\beta_{n}P_{n}(x)+\alpha_{n}P_{n-1}(x), (1.5)

where the recurrence coefficients are given by

αn=γnγn1,βn=xPn(x)2𝑑xγn.\alpha_{n}=\frac{\gamma_{n}}{\gamma_{n-1}}\,,\,\beta_{n}=\frac{\int_{\mathbb{R}}xP_{n}(x)^{2}dx}{\gamma_{n}}. (1.6)

For even weight functions, w(x)=w(x)w(-x)=w(x), we find βn=0\beta_{n}=0 for all nn.

Following the pioneering work of Freud and others, questions arising about the asymptotic locations of the zeros of Pn(x)P_{n}(x) and behaviour of αn\alpha_{n} as nn\to\infty 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 qq-orthogonal polynomials. In particular, we study polynomials which satisfy the orthogonality condition

Pn(x)Pm(x)(x4;q4)dqx=γnδn,m.\int_{-\infty}^{\infty}P_{n}(x)P_{m}(x)(-x^{4};q^{4})_{\infty}d_{q}x=\gamma_{n}\delta_{n,m}.

Throughout this paper we will label these polynomials as qq-Freud II polynomials to be consistent with the nomenclature of the DLMF [3, Chapter 18]. qq-Freud II polynomials were studied earlier by Ismail et al. [8], however Ismail et al. considered orthogonality on a continuous measure over \mathbb{R}. 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 αn\alpha_{n} often give rise to discrete Painlevé equations [5]. For example, the recurrence coefficients of orthogonal polynomials with weight function ex4/4dxe^{-x^{4}/4}dx satisfy the discrete equation [6]

αn(αn+1+αn+αn1)=n,\alpha_{n}(\alpha_{n+1}+\alpha_{n}+\alpha_{n-1})=n,

which is a case of the first discrete Painlevé equation, or dPI\mathrm{dP_{I}} [9].

The recurrence coefficients of qq-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 qq-Freud II polynomials satisfy

αn(αn+1+qn1αn+q2αn1q2n3αn+1αnαn1)=(qn1)q1n.\alpha_{n}(\alpha_{n+1}+q^{n-1}\alpha_{n}+q^{-2}\alpha_{n-1}-q^{2n-3}\alpha_{n+1}\alpha_{n}\alpha_{n-1})=(q^{-n}-1)q^{1-n}. (1.7)

In Sections 1.3 and 8 we extend this result initially determined by Ismail et al. to a larger class of qq-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 nn\to\infty varies between solutions.

There is an important feature of qq-difference equations that affects our discussion of qq-Freud II polynomials. The associated weight function satisfies a qq-difference equation, which gives rise to a family of weights involving a free qq-periodic function C(x)C(x), where C(qx)=C(x)C(qx)=C(x). If a weight in this family with C1C\not\equiv 1 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 qq-Freud II polynomials

We define the family of weight functions um:{eiπ(1+2n)2mqk}k=0u_{m}:\mathbb{C}\setminus\{e^{\frac{i\pi(1+2n)}{2m}}q^{-k}\}_{k=0}^{\infty}\to\mathbb{C}, where n=0,1,,2m1n=0,1,...,2m-1, as

um(x)=1(x2m,q2m),u_{m}(x)=\frac{1}{(-x^{2m},q^{2m})_{\infty}},

where mm\in\mathbb{N} [8]. They satisfy the qq-difference equation

Dqum(x)=x2m11qum(x).D_{q}u_{m}(x)=\frac{-x^{2m-1}}{1-q}u_{m}(x). (1.8)

This is analogous to classical Freudian weights vm(x)=ex2mv_{m}(x)=e^{-x^{2m}}, which satisfy the differential equation

ddxvm(x)=2mx2m1vm(x).\frac{d}{dx}v_{m}(x)=-2mx^{2m-1}v_{m}(x). (1.9)

However, a key difference between these two relations is that Equation (1.8) is a discrete relation. In particular if um(x)u_{m}(x) satisfies Equation (1.8) then um(x)C(x)u_{m}(x)C(x) also does, for any function C(x)C(x) satisfying C(qx)=C(x)C(qx)=C(x). Consider the case

u2(x)=1(x2,q2).u_{2}(x)=\frac{1}{(-x^{2},q^{2})_{\infty}}. (1.10)

This weight gives rise to discrete qq-Hermite II\mathrm{II} polynomials [3, Chapter 18.27]. The sequence of discrete qq-Hermite II\mathrm{II} polynomials, {Hn(x)}n=0\{H_{n}(x)\}_{n=0}^{\infty}, are orthogonal with respect to any measure u2(x)C(x)u_{2}(x)C(x). In particular they satisfy the continuous orthogonality condition

Hn(x)Hm(x)u2(x)𝑑x=γnδn,m,\int_{-\infty}^{\infty}H_{n}(x)H_{m}(x)u_{2}(x)dx=\gamma_{n}\delta_{n,m},

on the real line, and also satisfy the discrete orthogonality condition

Hn(cx)Hm(cx)u2(cx)dqx=γn(c)δn,m,\int_{-\infty}^{\infty}H_{n}(cx)H_{m}(cx)u_{2}(cx)d_{q}x=\gamma^{(c)}_{n}\delta_{n,m},

for any constant cc. In contrast, as we will show in Section 8 this is not true for the weight,

w(x)=u4(x)=1(x4,q4),w(x)=u_{4}(x)=\frac{1}{(-x^{4},q^{4})_{\infty}}, (1.11)

which is the focus of this paper. Thus, when describing qq-Freud II\mathrm{II} orthogonal polynomials, one also has to specify their orthogonality weight.

For the remainder of this paper, we focus on the sequence of polynomials {Pn(x)}\{P_{n}(x)\}, 0n0\leq n\in\mathbb{N}, that satisfy

Pn(x)Pm(x)w(x)dqx=γnδn,m.\int_{-\infty}^{\infty}P_{n}(x)P_{m}(x)w(x)d_{q}x=\gamma_{n}\delta_{n,m}.

We will call these polynomials qFIIqF_{II} polynomials. In Section 8, we discuss the implications of our results to polynomials orthogonal with respect to the weights of the form

Pn(cx)Pm(cx)w(cx)dqx=γn(c)δn,m,\int_{-\infty}^{\infty}P_{n}(cx)P_{m}(cx)w(cx)d_{q}x=\gamma^{(c)}_{n}\delta_{n,m}, (1.12)

for any constant q<c1q<c\leq 1. We will call these polynomials qFII(c)qF_{II}^{(c)} 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 {Pn(z)}n=0\{P_{n}(z)\}_{n=0}^{\infty} is a family of monic polynomials, orthogonal with respect to the weight w(z)dqzw(z)d_{q}z. Define t=zqn/2t=zq^{n/2}. Then, as nn\to\infty, for even nn\in\mathbb{N}:

Pn(z)={(1)n/2qn2(n21)a(z)(μ2η2+O(qn/4))for |z|qn/4,zna(t)(1+O(qn/4))for |z|>qn/4,P_{n}(z)=\left\{\begin{array}[]{lr}(-1)^{n/2}q^{\frac{-n}{2}(\frac{n}{2}-1)}a(z)\left(\frac{\mu_{2}}{\eta_{2}}+O(q^{n/4})\right)&\text{for }|z|\leq q^{-n/4},\\ z^{n}a_{\infty}(t)\left(1+O(q^{n/4})\right)&\text{for }|z|>q^{-n/4},\end{array}\right.

where a(z)a(z) is a solution of Equation (4.2), defined in Lemma 4.2, and, a(t)a_{\infty}(t) is a solution of Equation (5.2), defined in Lemma 5.2 (they are both independent of nn). Similarly Pn1(z)P_{n-1}(z) has the asymptotic behaviour

Pn1(z)={(1)n/2qn2(n21)γn1b(z)(λ3cΨ+O(qn/4))for |z|qn/4,qn(n21)γn1μ4cΨznb(t)(1+O(qn/4))for |z|>qn/4,P_{n-1}(z)=\left\{\begin{array}[]{lr}(-1)^{n/2}q^{\frac{n}{2}(\frac{n}{2}-1)}\gamma_{n-1}b(z)\left(\frac{\lambda_{3}}{c_{\Psi}}+O(q^{n/4})\right)&\text{for }|z|\leq q^{-n/4},\\ \frac{q^{n(\frac{n}{2}-1)}\gamma_{n-1}}{\mu_{4}\mathcal{H}c_{\varPsi}}z^{n}b_{\infty}(t)\left(1+O(q^{n/4})\right)&\text{for }|z|>q^{-n/4},\end{array}\right.

where b(z)b(z) is also a solution of Equation (4.2), defined in Lemma 4.2, and, b(t)b_{\infty}(t) is a solution of Equation (5.2), defined in Lemma 5.2.

Note that in the statement of Theorem 1.5, μ2\mu_{2}, η2\eta_{2}, μ4\mu_{4}, \mathcal{H} and cΨc_{\varPsi} are constants determined in Sections 4 and 5 which do not depend on zz or nn. Our second main result concerns the asymptotic behaviour of recurrence coefficients and L2L_{2} norm of PnP_{n} as nn approaches infinity.

Theorem 1.6.

Under the same hypotheses as Theorem 1.5, we have for even nn\in\mathbb{N}:

γn\displaystyle\gamma_{n} =\displaystyle= qn2(1n)(A+O(qn/2)),\displaystyle q^{\frac{n}{2}(1-n)}\left(A+O(q^{n/2})\right),
γn1\displaystyle\gamma_{n-1} =\displaystyle= qn2(3n)(B1+O(qn/2)),\displaystyle q^{\frac{n}{2}(3-n)}\left(B^{-1}+O(q^{n/2})\right),
αn\displaystyle\alpha_{n} =\displaystyle= qn(q+O(qn/2)),\displaystyle q^{-n}(q+O(q^{n/2})),

for some constants AA and BB, where γn\gamma_{n} and αn\alpha_{n} are defined by Equations (1.4) and (1.6) respectively.

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

αn(αn+1+qn1αn+q2αn1q2n3αn+1αnαn1)=(qn1)q1n.\alpha_{n}(\alpha_{n+1}+q^{n-1}\alpha_{n}+q^{-2}\alpha_{n-1}-q^{2n-3}\alpha_{n+1}\alpha_{n}\alpha_{n-1})=(q^{-n}-1)q^{1-n}. (1.13)

Furthermore, in general

limnαn(c)αn(1)q1nαn(1)0,\lim_{n\to\infty}\frac{\alpha^{(c)}_{n}-\alpha^{(1)}_{n}}{q^{1-n}-\alpha^{(1)}_{n}}\neq 0, (1.14)

where {αn(c)}n=0\{\alpha^{(c)}_{n}\}_{n=0}^{\infty} is the sequence of recurrence coefficients corresponding to polynomials with orthogonality condition given by Equation (1.12).

Theorem 1.7 immediately follows from Theorem 8.3 proved in Section 8.

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 qFIIqF_{II} polynomials. We then make a series of transformations to this RHP in Section 3. By taking the limit nn\rightarrow\infty 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 nn\to\infty. 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 qFIIqF_{II} polynomials and prove Theorem 1.7. In Appendix A we prove some important properties of the function hq(z)h_{q}(z) 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 qFIIqF_{II} orthogonal polynomials.

Definition 2.1 (qFIIqF_{II} RHP).

Let Γ\Gamma be an appropriate curve (see Definition 1.3) with interior 𝒟\mathcal{D}_{-} and exterior 𝒟+\mathcal{D}_{+}. A 2×22\times 2 complex matrix function Yn(z)Y_{n}(z), zz\in\mathbb{C}, is a solution of the qFIIqF_{II} RHP if it satisfies the following conditions:

  1. (i)

    Yn(z)Y_{n}(z) is meromorphic in Γ\mathbb{C}\setminus\Gamma, with simple poles at z=±qkz=\pm q^{-k} for 1k1\leq k\in\mathbb{N}.

  2. (ii)

    Yn(z)Y_{n}(z) has continuous boundary values Yn(s)Y_{n}^{-}(s) and Yn+(s)Y_{n}^{+}(s) as zz approaches sΓs\in\Gamma from 𝒟\mathcal{D}_{-} and 𝒟+\mathcal{D}_{+} respectively, where

    Yn+(s)=Yn(s)[1hq(s)w(s)01],sΓ,\displaystyle Y_{n}^{+}(s)=Y_{n}^{-}(s)\begin{bmatrix}1&h_{q}(s)w(s)\\ 0&1\end{bmatrix},\;s\in\Gamma, (2.1a)

    and w(s)w(s) is defined in Equation (1.11).

  3. (iii)

    The residue at each pole z=±qkz=\pm q^{-k} for 1k1\leq k\in\mathbb{N} is given by

    Res(Yn(±qk))=limz±qkYn(z)[0(zqk)hq(z)w(z)00].\displaystyle\mathrm{Res}(Y_{n}(\pm q^{-k}))=\lim_{z\rightarrow\pm q^{-k}}Y_{n}(z)\begin{bmatrix}0&(z\mp q^{-k})h_{q}(z)w(z)\\ 0&0\end{bmatrix}. (2.1b)
  4. (iv)

    Yn(z)Y_{n}(z) satisfies

    Yn(z)[zn00zn]=I+O(1|z|), as |z|,Y_{n}(z)\begin{bmatrix}z^{-n}&0\\ 0&z^{n}\end{bmatrix}=I+O\left(\frac{1}{|z|}\right),\text{ as $|z|\rightarrow\infty$}, (2.1c)

    for zz such that |z±qk|>r|z\pm q^{-k}|>r, for all 1k1\leq k\in\mathbb{N}, for fixed r>0r>0.

Remark 2.2.

Note that the matrix Yn(±qk)Y_{n}(\pm q^{-k}) has poles in its second column for 1k1\leq k\in\mathbb{N}. Thus, the asymptotic decay does not hold near these poles. This is why, following Equation (2.1c), we require the added condition zz must be such that |z±qk|>r|z\pm q^{-k}|>r, for all 1k1\leq k\in\mathbb{N}, for fixed r>0r>0.

We now determine the solution of the qq-RHP.

Lemma 2.3.

The unique solution of the qq-RHP given by Definition 2.1 is given by

Yn(z)=[Pn(z)ΓPn(s)w(s)hq(s)2πi(zs)𝑑s+q1Pn(±s)w(±s)zsdqsγn11Pn1(z)ΓPn1(s)w(s)hq(s)2πi(zs)γn1𝑑s+q1Pn1(±s)w(±s)(zs)γn1dqs],\displaystyle Y_{n}(z)=\begin{bmatrix}P_{n}(z)&\oint_{\Gamma}\frac{P_{n}(s)w(s)h_{q}(s)}{2\pi i(z-s)}ds+\int_{q^{-1}}^{\infty}\frac{P_{n}(\pm s)w(\pm s)}{z\mp s}d_{q}s\\ \gamma_{n-1}^{-1}P_{n-1}(z)&\oint_{\Gamma}\frac{P_{n-1}(s)w(s)h_{q}(s)}{2\pi i(z-s)\gamma_{n-1}}ds+\int_{q^{-1}}^{\infty}\frac{P_{n-1}(\pm s)w(\pm s)}{(z\mp s)\gamma_{n-1}}d_{q}s\end{bmatrix}, (2.2)

where {Pn(z)}n=0\{P_{n}(z)\}_{n=0}^{\infty} satisfies the orthogonality condition

Pn(s)Pm(s)w(s)dqs=γnδn,m.\int_{-\infty}^{\infty}P_{n}(s)P_{m}(s)w(s)d_{q}s=\gamma_{n}\delta_{n,m}.
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 \mathbb{R}.

We show that the second row of Yn(z)Y_{n}(z) must be given by Equation (2.2). A similar argument can be carried out for the first row. To declutter notation we will label Yn(z)Y_{n}(z) as Y(z)Y(z) for the rest of this proof.

It follows from the asymptotic condition, Equation (2.1c), that the (1,1)(1,1) entry of YY must have leading order znz^{n} as zz\rightarrow\infty. As Y(1,1)(z)Y_{(1,1)}(z) is analytic and its jump condition, Equation (2.1a), is given by the identity we immediately conclude that Y(1,1)(z)Y_{(1,1)}(z) is a monic polynomial of degree nn. Similarly, it follows that Y(2,1)(z)Y_{(2,1)}(z) is a polynomial of degree at most n1n-1. We denote Y(2,1)(z)Y_{(2,1)}(z) by Qn1(z)Q_{n-1}(z).

Consider the bottom right entry of Equation (2.2). By the jump condition, Equation (2.1a), we have

Y(2,2)+(s)=Qn1(s)w(s)hq(s)+Y(2,2)(s).Y_{(2,2)}^{+}(s)=Q_{n-1}(s)w(s)h_{q}(s)+Y_{(2,2)}^{-}(s)\,. (2.3)

If there was no residue condition, Equation (2.1b), then this scalar equation would be solved by the Cauchy transform

Y(2,2)(z)=12πiΓQn1(s)w(s)hq(s)zs𝑑s,Y_{(2,2)}(z)=\frac{1}{2\pi i}\oint_{\Gamma}\frac{Q_{n-1}(s)w(s)h_{q}(s)}{z-s}ds\,, (2.4)

which is analytic in Γ\mathbb{C}\setminus\Gamma and satisfies Equation (2.3). The residue condition can be readily resolved by letting

Y(2,2)(z)=12πiΓQn1(s)w(s)hq(s)zs𝑑s+q1Qn1(±s)w(±s)(zs)dqs,Y_{(2,2)}(z)=\frac{1}{2\pi i}\oint_{\Gamma}\frac{Q_{n-1}(s)w(s)h_{q}(s)}{z-s}ds+\int_{q^{-1}}^{\infty}\frac{Q_{n-1}(\pm s)w(\pm s)}{(z\mp s)}d_{q}s\,, (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 Y(2,2)(z)Y_{(2,2)}(z). Substituting our expression for hq(s)h_{q}(s) into Equation (2.5), we find

Y(2,2)(z)\displaystyle Y_{(2,2)}(z) =12πiΓk=(qksqkQn1(s)w(s)zs+qks+qkQn1(s)w(s)zs)ds\displaystyle=\frac{1}{2\pi i}\oint_{\Gamma}\sum_{k=-\infty}^{\infty}\left(\frac{q^{k}}{s-q^{k}}\frac{Q_{n-1}(s)w(s)}{z-s}+\frac{q^{k}}{s+q^{k}}\frac{Q_{n-1}(s)w(s)}{z-s}\right)ds
+q1Qn1(±s)w(±s)(zs)dqs.\displaystyle\qquad+\int_{q^{-1}}^{\infty}\frac{Q_{n-1}(\pm s)w(\pm s)}{(z\mp s)}d_{q}s.

which by Cauchy’s integral formula, for zext(Γ)z\in\text{ext}(\Gamma), becomes

Y(2,2)(z)\displaystyle Y_{(2,2)}(z) =\displaystyle= k=0(qkQn1(qk)w(qk)zqk+qkQn1(qk)w(qk)z+qk)\displaystyle\sum_{k=0}^{\infty}\left(q^{k}\frac{Q_{n-1}(q^{k})w(q^{k})}{z-q^{k}}+q^{k}\frac{Q_{n-1}(-q^{k})w(-q^{k})}{z+q^{k}}\right)
+q1Qn1(±s)w(±s)(zs)dqs,\displaystyle\,+\int_{q^{-1}}^{\infty}\frac{Q_{n-1}(\pm s)w(\pm s)}{(z\mp s)}d_{q}s,
=\displaystyle= Qn1(x)w(x)zxdqx,\displaystyle\int^{\infty}_{-\infty}\frac{Q_{n-1}(x)w(x)}{z-x}d_{q}x\,,

where the sum to infinity is well defined on Γ\Gamma, as hq(s)h_{q}(s) converges as kk\rightarrow\infty, and the Jackson integral of an analytic function is well defined. Using the geometric series with remainder

1zx=k=0l(xkzk+1)+xl+1zl+1(zx),forxz,\frac{1}{z-x}=\sum^{l}_{k=0}\left(\frac{x^{k}}{z^{k+1}}\right)+\frac{x^{l+1}}{z^{l+1}(z-x)}\,,\quad\text{for}\,x\neq z\,, (2.6)

we find

Y(2,2)(z)\displaystyle Y_{(2,2)}(z) =\displaystyle= Qn1(x)w(x)xnzn(zx)dqx\displaystyle\int^{\infty}_{-\infty}\frac{Q_{n-1}(x)w(x)x^{n}}{z^{n}(z-x)}d_{q}x
+k=0n11zk+1Qn1(x)w(x)xkdqx.\displaystyle+\sum^{n-1}_{k=0}\frac{1}{z^{k+1}}\int^{\infty}_{-\infty}Q_{n-1}(x)w(x)x^{k}d_{q}x\,.

Note that the asymptotic condition, Equation (2.1c), holds when the last term on the RHS is zero for k=0,1,2,,n2k=0,1,2,...,n-2. This is true iff

Qn1(x)w(x)xkdqx=0,forkn2,\int^{\infty}_{-\infty}Q_{n-1}(x)w(x)x^{k}d_{q}x=0\,,\quad\text{for}\,k\leq n-2\,,

which is satisfied when Qn1Q_{n-1} is an orthogonal polynomial of degree n1n-1 on the qq-lattice with respect to the weight w(x)w(x). This is the class of qFIIqF_{II} polynomials. We conclude that the solution of Y(2,2)(z)Y_{(2,2)}(z) is given by

Y(2,2)(z)={1γn1Pn1(x)w(x)zxdqx1γn1Pn1(z)w(z)hq(z),for zint(Γ),1γn1Pn1(x)w(x)zxdqx,for zext(Γ).Y_{(2,2)}(z)=\left\{\begin{array}[]{lr}\frac{1}{\gamma_{n-1}}\int^{\infty}_{-\infty}\frac{P_{n-1}(x)w(x)}{z-x}d_{q}x-\frac{1}{\gamma_{n-1}}P_{n-1}(z)w(z)h_{q}(z),&\text{for }z\in\text{int}(\Gamma),\\ \frac{1}{\gamma_{n-1}}\int^{\infty}_{-\infty}\frac{P_{n-1}(x)w(x)}{z-x}d_{q}x,&\text{for }z\in\text{ext}(\Gamma).\end{array}\right.

After appropriate scaling, and repeating the same arguments for the first row, it follows that Equation (2.2) is a solution of the qq-RHP given by Definition 2.1.

Uniqueness of this solution follows from consideration of the determinant. Observe that the jump matrix J=Y1Y+J=Y_{-}^{-1}Y_{+} satisfies det(J)=1\text{det}(J)=1. It immediately follows that det(Y+)=det(Y)\text{det}(Y^{+})=\text{det}(Y^{-}) on Γ\Gamma. Furthermore, by the residue condition, Equation (2.1b), det(Y)\text{det}(Y) has no poles. Thus, det(Y)\text{det}(Y) is an entire function. By the asymptotic condition, Equation (2.1c), det(Y)1\text{det}(Y)\rightarrow 1, and so by Liouville’s theorem, it follows that det(Y)=1\text{det}(Y)=1. This implies Y1Y^{-1} exists and is meromorphic in /Γ\mathbb{C}/\Gamma.

Now suppose that there exists a second solution of the qq-RHP and denote this solution by Y^\widehat{Y}. If we define M=Y^Y1M=\widehat{Y}Y^{-1}, it follows that the jump conditions (and residue conditions) effectively cancel and M+=MM_{+}=M_{-}. Thus, MM is entire and MIM\rightarrow I as zz\rightarrow\infty. Hence, by Liouville’s theorem M=IM=I. We conclude that Y^=Y\widehat{Y}=Y and, therefore, there is a single unique solution of the qq-RHP. ∎

Remark 2.4.

It can be shown that solution of the RHP given by Definition 2.1 satisfies a Lax pair, namely a qq-difference equation in zz and a recurrence relation for the parameter nn. The proof follows using the same arguments as in [10].

Remark 2.5.

One can repeat the arguments above to show that there is unique solution of Definition 2.1 with a modified function

hq(z)c1hq(z/c),h_{q}(z)\to c^{-1}h_{q}(z/c),

for some real constant cc, and that it is given by

Yn(c)(z)=[Pn(c)(z)ΓPn(c)(s)w(s)hq(s/c)2cπi(zs)𝑑s+q1Pn(c)(±cs)w(±cs)zcsdqsγn11Pn1(c)(z)ΓPn1(c)(s)w(s)hq(s/c)2cπi(zs)γn1(c)𝑑s+q1Pn1(c)(±cs)w(±cs)(zcs)γn1(c)dqs],\displaystyle Y^{(c)}_{n}(z)=\begin{bmatrix}P^{(c)}_{n}(z)&\oint_{\Gamma}\frac{P^{(c)}_{n}(s)w(s)h_{q}(s/c)}{2c\pi i(z-s)}ds+\int_{q^{-1}}^{\infty}\frac{P^{(c)}_{n}(\pm cs)w(\pm cs)}{z\mp cs}d_{q}s\\ \gamma_{n-1}^{-1}P^{(c)}_{n-1}(z)&\oint_{\Gamma}\frac{P^{(c)}_{n-1}(s)w(s)h_{q}(s/c)}{2c\pi i(z-s)\gamma^{(c)}_{n-1}}ds+\int_{q^{-1}}^{\infty}\frac{P^{(c)}_{n-1}(\pm cs)w(\pm cs)}{(z\mp cs)\gamma^{(c)}_{n-1}}d_{q}s\end{bmatrix},

where {Pn(c)(z)}n=0\{P^{(c)}_{n}(z)\}_{n=0}^{\infty} satisfies the orthogonality condition

Pn(c)(cs)Pm(c)(cs)w(cs)dqs=γn(c)δn,m.\int_{-\infty}^{\infty}P^{(c)}_{n}(cs)P^{(c)}_{m}(cs)w(cs)d_{q}s=\gamma^{(c)}_{n}\delta_{n,m}.

3. Transformations of RHP

In this section we will transform the RHP given in Definition 2.1 to more easily determine the asymptotics of Yn(z)Y_{n}(z) as nn\to\infty. Throughout this section we will assume that nn 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 g:{0}g:\mathbb{C}\setminus\{0\}\to\mathbb{C} as

g(z)=(z2;q2)(q2z2;q2).g(z)=(z^{2};q^{2})_{\infty}(q^{2}z^{-2};q^{2})_{\infty}. (3.1)
Lemma 3.2.

g(z)g(z) satisfies the difference equation

g(qz)=z2g(z).g(qz)=-z^{-2}g(z). (3.2)
Proof.

Substituting qzqz into Equation (3.1) we find

g(qz)\displaystyle g(qz) =\displaystyle= (q2z2;q2)(z2;q2),\displaystyle(q^{2}z^{2};q^{2})_{\infty}(z^{-2};q^{2})_{\infty},
=\displaystyle= j=0(1q2z2q2j)j=0(1z2q2j),\displaystyle\prod_{j=0}^{\infty}(1-q^{2}z^{2}q^{2j})\prod_{j=0}^{\infty}(1-z^{-2}q^{2j}),
=\displaystyle= 1z21z2j=0(1z2q2j)j=0(1q2z2q2j),\displaystyle\frac{1-z^{-2}}{1-z^{2}}\prod_{j=0}^{\infty}(1-z^{2}q^{2j})\prod_{j=0}^{\infty}(1-q^{2}z^{-2}q^{2j}),
=\displaystyle= 1z21z2g(z),\displaystyle\frac{1-z^{-2}}{1-z^{2}}g(z),
=\displaystyle= z2g(z).\displaystyle-z^{-2}g(z).

Remark 3.3.

By induction using Equation (3.2) we find that for even nn

g(qn/2z)=(1)n/2qn2/4n/2zng(z).g(q^{-n/2}z)=(-1)^{n/2}q^{-n^{2}/4-n/2}z^{n}g(z). (3.3)
Definition 3.4.

Define ω:({0}{eiπ(1+2n)4qk}k=)\omega:\mathbb{C}\setminus(\{0\}\cup\{e^{\frac{i\pi(1+2n)}{4}}q^{k}\}_{k=-\infty}^{\infty})\to\mathbb{C}, where n=0,1,2,3n=0,1,2,3, as

ω(z)=1/(z4,q4z4;q4).\omega(z)=1/(-z^{4},-q^{4}z^{-4};q^{4})_{\infty}. (3.4)
Lemma 3.5.

ω(z)\omega(z) satisfies the difference equation

ω(qz)=z4ω(z).\omega(qz)=z^{4}\omega(z). (3.5)
Proof.

The proof follows from definition of ω(z)\omega(z) (by applying the same arguments as in Lemma 3.2). ∎

Lemma 3.6.

The function w(z)w(z) defined in Equation (1.11) satisfies the difference equation

w(qz)=(1+z4)w(z).w(qz)=(1+z^{4})w(z). (3.6)
Proof.

The proof follows from definition of w(z)w(z) (by applying the same arguments as in Lemma 3.2). ∎

Remark 3.7.

By induction using Equation (3.6) we find that for even nn

w(qn/2z)=z2nqn(n2+1)(q2n+4z4;q4)ω(z).w(q^{-n/2}z)=z^{-2n}q^{n(\frac{n}{2}+1)}(-q^{2n+4}z^{-4};q^{4})_{\infty}\omega(z). (3.7)

Taking zzqn/2z\to zq^{n/2}, Equation (3.7) gives

w(z)=z2nqn(n2+1)(z4q4;q4)ω(zqn/2).w(z)=z^{-2n}q^{n(-\frac{n}{2}+1)}(-z^{-4}q^{4};q^{4})_{\infty}\omega(zq^{n/2}). (3.8)

3.1. RHP transformations

Before proceeding we introduce some notation. Consider the contour Γ\Gamma, scaled such that the modulus of points on it are multiplied by qn/2q^{-n/2}. We denote this new contour by Γqn/2\Gamma_{q^{-n/2}}. (If Γ\Gamma were the unit circle, Γqn/2\Gamma_{q^{-n/2}} would be a circle with radius qn/2q^{-n/2}.) Consider the following transformation to the RHP given by Definition 2.1:

Un(z)={Yn(z),forz𝒟,Yn(z)[g(z)100g(z)],forz𝒟,qn/2𝒟,Yn(z)[zn00zn],forz𝒟+,qn/2.\displaystyle U_{n}(z)=\left\{\begin{array}[]{lr}Y_{n}(z),&\text{for}\,z\in\mathcal{D}_{-},\\ Y_{n}(z)\begin{bmatrix}g(z)^{-1}&0\\ 0&g(z)\end{bmatrix},&\text{for}\,z\in\mathcal{D}_{-,q^{-n/2}}\setminus\mathcal{D}_{-},\\ Y_{n}(z)\begin{bmatrix}z^{-n}&0\\ 0&z^{n}\end{bmatrix},&\text{for}\,z\in\mathcal{D}_{+,q^{-n/2}}.\end{array}\right. (3.12)

This gives a new RHP for UnU_{n} with two jumps (and also some more poles which will be discussed shortly). At zΓqn/2z\in\Gamma_{q^{-n/2}} we apply Equation (3.3) to determine

g(qn/2eiθ)(qn/2eiθ)n=g(eiθ)(1)n/2qn2/4n/2.g(q^{-n/2}e^{i\theta})(q^{-n/2}e^{i\theta})^{-n}=g(e^{i\theta})(-1)^{n/2}q^{n^{2}/4-n/2}. (3.13)

Motivated by Equation (3.13) we make the transformation

Wn(z)={[cnn00cnn]Un(z)[cnn00cnn],forz𝒟+,qn/2,[cnn00cnn]Un(z),forz𝒟,qn/2,\displaystyle W_{n}(z)=\left\{\begin{array}[]{lr}\begin{bmatrix}c_{n}^{n}&0\\ 0&c_{n}^{-n}\end{bmatrix}U_{n}(z)\begin{bmatrix}c_{n}^{-n}&0\\ 0&c_{n}^{n}\end{bmatrix},&\text{for}\,z\in\mathcal{D}_{+,q^{-n/2}},\\ \begin{bmatrix}c_{n}^{n}&0\\ 0&c_{n}^{-n}\end{bmatrix}U_{n}(z),&\text{for}\,z\in\mathcal{D}_{-,q^{-n/2}},\end{array}\right. (3.16)

where

cn=(1)1/2qn/41/2.c_{n}=(-1)^{1/2}q^{n/4-1/2}.

This new matrix Wn(z)W_{n}(z) satisfies the following RHP:

Definition 3.8 (WnW_{n}-RHP).

Let Γ\Gamma be an appropriate curve (see Definition 1.3) with interior 𝒟\mathcal{D}_{-} and exterior 𝒟+\mathcal{D}_{+}. A 2×22\times 2 complex matrix function Wn(z)W_{n}(z), zz\in\mathbb{C}, is a solution of the WnW_{n} RHP if it satisfies the following conditions:

  1. (i)

    Wn(z)W_{n}(z) is meromorphic in (ΓΓqn/2)\mathbb{C}\setminus(\Gamma\bigcup\Gamma_{q^{-n/2}}), with simple poles at z=±qkz=\pm q^{-k} for 1k1\leq k\in\mathbb{N}.

  2. (ii)

    Wn(z)W_{n}(z) has continuous boundary values Wn(s)W_{n}^{-}(s) and Wn+(s)W_{n}^{+}(s) as zz approaches sΓs\in\Gamma from 𝒟\mathcal{D}_{-} and 𝒟+\mathcal{D}_{+} respectively, where

    Wn+(s)=Wn(s)[g(s)1g(s)w(s)hq(s)0g(s)],sΓ.\displaystyle W_{n}^{+}(s)=W_{n}^{-}(s)\begin{bmatrix}g(s)^{-1}&g(s)w(s)h_{q}(s)\\ 0&g(s)\end{bmatrix},\;s\in\Gamma. (3.17a)
  3. (iii)

    Wn(z)W_{n}(z) has continuous boundary values Wn(s)W_{n}^{-}(s) and Wn+(s)W_{n}^{+}(s) as zz approaches sΓqn/2s\in\Gamma_{q^{-n/2}} from 𝒟,qn/2\mathcal{D}_{-,q^{-n/2}} and 𝒟+,qn/2\mathcal{D}_{+,q^{-n/2}} respectively, where

    Wn+(s)=Wn(s)[g(sqn/2)00g(sqn/2)1],sΓqn/2.\displaystyle W_{n}^{+}(s)=W_{n}^{-}(s)\begin{bmatrix}g(sq^{n/2})&0\\ 0&g(sq^{n/2})^{-1}\end{bmatrix},\;s\in\Gamma_{q^{-n/2}}. (3.17b)
  4. (iv)

    Wn(z)W_{n}(z) satisfies

    Wn(z)=I+O(1|z|), as |z|.W_{n}(z)=I+O\left(\frac{1}{|z|}\right),\text{ as $|z|\rightarrow\infty$}. (3.17c)

    Note that Wn(±qk)W_{n}(\pm q^{-k}) has poles in the second column for 1k1\leq k\in\mathbb{N}. Thus, the decay condition does not hold near these poles. For example: the decay condition holds for zz such that |z±qk|>r|z\pm q^{-k}|>r, for all integer k>n/2k>n/2, for fixed r>0r>0.

  5. (v)

    The residue at the poles z=±qkz=\pm q^{-k} for 1kn/21\leq k\leq n/2 is given by

    Res(Wn(±qk))=limz±qkWn(z)[00(zqk)g(z)2hq(z)1w(z)10].\displaystyle\text{Res}(W_{n}(\pm q^{-k}))=\lim_{z\rightarrow\pm q^{-k}}W_{n}(z)\begin{bmatrix}0&0\\ (z\mp q^{-k})g(z)^{-2}h_{q}(z)^{-1}w(z)^{-1}&0\end{bmatrix}. (3.17d)
  6. (vi)

    The residue at the poles z=±qkz=\pm q^{-k} for k>n/2k>n/2 is given by

    Res(Wn(±qk))=limz±qkWn(z)[0(zqk)z2nhq(z)w(z)cn2n00].\displaystyle\text{Res}(W_{n}(\pm q^{-k}))=\lim_{z\rightarrow\pm q^{-k}}W_{n}(z)\begin{bmatrix}0&(z\mp q^{-k})z^{2n}h_{q}(z)w(z)c_{n}^{2n}\\ 0&0\end{bmatrix}. (3.17e)

4. Near-field RHP

We will show that the solution Wn(z)W_{n}(z) of the RHP given in Definition 3.8 approaches a limiting solution G(z)G(z). 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 Wn(z)W_{n}(z) at Γ\Gamma and Γqn/2\Gamma_{q^{-n/2}} 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 (𝔚\mathfrak{W}-RHP).

Let Γ\Gamma be an appropriate curve (see Definition 1.3) with interior 𝒟\mathcal{D}_{-} and exterior 𝒟+\mathcal{D}_{+}. A 2×22\times 2 complex matrix function 𝔚(z)\mathfrak{W}(z), zz\in\mathbb{C}, is a solution of the 𝔚\mathfrak{W}-RHP if it satisfies the following conditions:

  1. (i)

    𝔚(z)\mathfrak{W}(z) is meromorphic in Γ\mathbb{C}\setminus\Gamma, with simple poles in the first column at z=±qkz=\pm q^{-k} for 1k1\leq k\in\mathbb{N}.

  2. (ii)

    𝔚(z)\mathfrak{W}(z) has continuous boundary values 𝔚(s)\mathfrak{W}^{-}(s) and 𝔚+(s)\mathfrak{W}^{+}(s) as zz approaches sΓs\in\Gamma from 𝒟\mathcal{D}_{-} and 𝒟+\mathcal{D}_{+}, and

    𝔚+(s)=𝔚(s)[g(s)1g(s)w(s)hq(s)0g(s)],sΓ,\displaystyle\mathfrak{W}^{+}(s)=\mathfrak{W}^{-}(s)\begin{bmatrix}g(s)^{-1}&g(s)w(s)h_{q}(s)\\ 0&g(s)\end{bmatrix},\;s\in\Gamma, (4.1a)

    where w(s)w(s) and g(s)g(s) are defined in Equations (1.11) and (3.1) respectively.

  3. (iii)

    𝔚(z)\mathfrak{W}(z) satisfies

    𝔚(z)=[10c0hq(z)1]+O(1z),\displaystyle\mathfrak{W}(z)=\begin{bmatrix}1&0\\ c_{0}h_{q}(z)&1\end{bmatrix}+O\left(\frac{1}{z}\right), (4.1b)

    where c0c_{0} is a non-zero constant. Due to the simple poles in the first column of 𝔚(z)\mathfrak{W}(z), the asymptotic decay condition only holds for |z±qk|>qk/R|z\pm q^{-k}|>q^{-k/R} for any R>0R>0.

  4. (iv)

    The residue at the poles z=±qkz=\pm q^{-k} for 1k1\leq k\in\mathbb{N} is given by

    Res(𝔚(±qk))=limz±qk𝔚(z)[00(zqk)g(z)2hq(z)1w(z)10].\displaystyle\text{Res}(\mathfrak{W}(\pm q^{-k}))=\lim_{z\rightarrow\pm q^{-k}}\mathfrak{W}(z)\begin{bmatrix}0&0\\ (z\mp q^{-k})g(z)^{-2}h_{q}(z)^{-1}w(z)^{-1}&0\end{bmatrix}. (4.1c)

To solve this RHP, a series of Lemmas are required.

Lemma 4.2.

Consider the difference equation

y(q2z)+(q3z2(1+q1)(1+q1))y(q1z)+q1(1+q4z4)y(z)=0.y(q^{-2}z)+\bigl{(}q^{-3}z^{2}(1+q^{-1})-(1+q^{-1})\bigr{)}y(q^{-1}z)+q^{-1}(1+q^{-4}z^{4})y(z)=0. (4.2)

There exists two entire solutions of Equation (4.2), one even and one odd.

Proof.

Let

y(z)=i=0yizi.y(z)=\sum_{i=0}^{\infty}y_{i}z^{i}.

Substituting this into Equation (4.2) and comparing coefficients of zz, we find that y(z)y(z) is a solution iff

yi=(1+q1)qiyi2+q4yi4q2i+1(1+q)qi+1.y_{i}=-\frac{(1+q^{-1})q^{-i}y_{i-2}+q^{-4}y_{i-4}}{q^{-2i+1}-(1+q)q^{-i}+1}.

Lemma 4.2 follows immediately. ∎

Definition 4.3.

Following from Lemma 4.2 we define a(z)a(z) as the even entire solution to Equation (4.2) normalised such that its first non-zero Taylor series coefficient (atO(1))(\mathrm{at}\,O(1)) is unity. Similarly, we define b(z)b(z) as the odd entire solution normalised such that its first non-zero Taylor series coefficient (atO(z))(\mathrm{at}\,O(z)) is unity.

Motivated by the (1,2)(1,2) entry of the right side of Equation (4.1a), we consider the properties of the product y(z)g(z)w(z)y(z)g(z)w(z).

Lemma 4.4.

Define

v(z)=y(z)g(z)w(z),v(z)=y(z)g(z)w(z),

then v(z)v(z) is a solution of the difference equation

(q6z4+q2)v(q2z)+(z2q(1+q)q1(1+q1))v(q1z)+q1v(z)=0.(q^{6}z^{-4}+q^{-2})v(q^{-2}z)+(z^{-2}q(1+q)-q^{-1}(1+q^{-1}))v(q^{-1}z)+q^{-1}v(z)=0. (4.3)

Furthermore, there exists two solutions to Equation (4.3) analytic in {0}\mathbb{C}\setminus\{0\} which can be represented by an even and odd power series at infinity.

Proof.

From the definition of v(z)v(z) and Lemmas 3.2 and 3.6 we find that

y(q1z)\displaystyle y(q^{-1}z) =\displaystyle= v(q1z)g(q1z)w(q1z),\displaystyle\frac{v(q^{-1}z)}{g(q^{-1}z)w(q^{-1}z)},
=\displaystyle= q2z2(1+q4z4)v(q1z)g(z)w(z).\displaystyle-\frac{q^{2}z^{-2}(1+q^{-4}z^{4})v(q^{-1}z)}{g(z)w(z)}.

Substituting the above into Equation (4.2) we determine that v(z)v(z) satisfies the difference equation

(q6z4+q2)v(q2z)(q1(1+q1)(q2+q)z2)v(q1z)+q1v(z)=0.(q^{6}z^{-4}+q^{-2})v(q^{-2}z)-(q^{-1}(1+q^{-1})-(q^{2}+q)z^{-2})v(q^{-1}z)+q^{-1}v(z)=0.

Let v(z)=i=0viziv(z)=\sum_{i=0}^{\infty}v_{i}z^{-i}, such a power series is a solution of Equation (4.3) iff

vi=q(1+q)qivi2+q2ivi4q(1+q)qi+q2i.v_{i}=-\frac{q(1+q)q^{i}v_{i-2}+q^{2i}v_{i-4}}{q-(1+q)q^{i}+q^{2i}}.

Lemma 4.4 follows immediately (note that the power series for v(z)v(z) converges everywhere). ∎

Definition 4.5.

Following from Lemma 4.4 we define ϕeven(z)\phi_{even}(z) for z{0}z\in\mathbb{C}\setminus\{0\}, as the even solution to Equation (4.3) normalised such that its first non-zero series coefficient (atO(1))(\mathrm{at}\,O(1)) is unity. Similarly, we define ϕodd(z)\phi_{odd}(z) as the odd solution normalised such that its first non-zero series coefficient (atO(z1))(\mathrm{at}\,O(z^{-1})) is unity.

Motivated by the (1,1)(1,1) entry of the right side of Equation (4.1a) for z𝒟+z\in\mathcal{D}_{+}, we consider the properties of the product y(z)g(z)1y(z)g(z)^{-1}.

Lemma 4.6.

Define

u(z)=y(z)g(z)1,u(z)=y(z)g(z)^{-1},

then u(z)u(z) is a solution of the difference equation

q5u(q2z)+q1(z2(1+q1)q3(1+q1))u(q1z)+(z4+q4)u(z)=0.q^{-5}u(q^{-2}z)+q^{-1}(z^{-2}(1+q^{-1})-q^{-3}(1+q^{-1}))u(q^{-1}z)+(z^{-4}+q^{-4})u(z)=0. (4.4)

Furthermore, there exists two solutions to Equation (4.4) holomorphic for |z|>q|z|>q which can be represented by an even and odd power series at infinity.

Proof.

From the definition of u(z)u(z) and Equation (3.2) we find that

y(q1z)\displaystyle y(q^{-1}z) =\displaystyle= u(q1z)g(q1z),\displaystyle u(q^{-1}z)g(q^{-1}z),
=\displaystyle= q2z2u(q1z)g(z).\displaystyle-q^{-2}z^{2}u(q^{-1}z)g(z).

Substituting the above into Equation (4.2) we determine that u(z)u(z) satisfies the difference equation

q6z4u(q2z)q2z2(q3z2(1+q1)(1+q1))u(q1z)+q1(1+q4z4)u(z)=0,q^{-6}z^{4}u(q^{-2}z)-q^{-2}z^{2}(q^{-3}z^{2}(1+q^{-1})-(1+q^{-1}))u(q^{-1}z)+q^{-1}(1+q^{-4}z^{4})u(z)=0,

which one can readily show is equivalent to Equation (4.4). Let u(z)=i=0uiziu(z)=\sum_{i=0}^{\infty}u_{i}z^{-i}, such a power series is a solution of Equation (4.4) iff

ui=(1+q)qi+3ui2+q5ui4q(1+q)qi+q2i.u_{i}=-\frac{(1+q)q^{i+3}u_{i-2}+q^{5}u_{i-4}}{q-(1+q)q^{i}+q^{2i}}. (4.5)

For large index ii, we can deduce from Equation (4.5) that

max(|ui|,|ui+2|)=(q4+O(qi))max(|ui4|,|ui2|).\mathrm{max}(|u_{i}|,|u_{i+2}|)=(q^{4}+O(q^{i}))\mathrm{max}(|u_{i-4}|,|u_{i-2}|).

Taking a telescopic product we conclude that the sum i=0uizi\sum_{i=0}^{\infty}u_{i}z^{-i} converges if |z|>q|z|>q. Lemma 4.4 follows immediately. ∎

Definition 4.7.

Following from Lemma 4.6 we define φeven(z)\varphi_{even}(z) as the even solution to Equation (4.4), analytic in |z|>q|z|>q, normalised such that its first non-zero series coefficient (atO(1))(\mathrm{at}\,O(1)) is unity. Similarly, we define φodd(z)\varphi_{odd}(z) as the odd solution, analytic in |z|>q|z|>q, normalised such that its first non-zero series coefficient (atO(z1))(\mathrm{at}\,O(z^{-1})) is unity.

Lemma 4.8.

The function a(z)g(z)1a(z)g(z)^{-1}, where a(z)a(z) is defined in Lemma 4.2, can be written as

a(z)g(z)1=η1hq(z)φodd(z)+η2φeven(z),a(z)g(z)^{-1}=\eta_{1}h_{q}(z)\varphi_{odd}(z)+\eta_{2}\varphi_{even}(z),

where η1,η2\eta_{1},\eta_{2} are constants. Similarly b(z)g(z)1b(z)g(z)^{-1} can be written as

b(z)g(z)1=η3φodd(z)+η4hq(z)φeven(z).b(z)g(z)^{-1}=\eta_{3}\varphi_{odd}(z)+\eta_{4}h_{q}(z)\varphi_{even}(z).
Proof.

From Lemma 4.6, we conclude that a(z)a(z) satisfies the same difference equation as g(z)φeven(z)g(z)\varphi_{even}(z) and g(z)φodd(z)g(z)\varphi_{odd}(z). It follows that

a(z)=g(z)(C1(z)φeven(z)+C2(z)φodd(z)),a(z)=g(z)(C_{1}(z)\varphi_{even}(z)+C_{2}(z)\varphi_{odd}(z)),

for some functions Ci(z)C_{i}(z) which satisfy Ci(qz)=Ci(z)C_{i}(qz)=C_{i}(z). As a(z)a(z) is analytic everywhere and φeven(z)\varphi_{even}(z) and φodd(z)\varphi_{odd}(z) are holomorphic for |z|>q|z|>q we conclude that Ci(z)C_{i}(z) is constant or has simple poles at the zeros of g(z)g(z), which occur at ±qk\pm q^{k} for kk\in\mathbb{Z}. Applying Corollary A.2 to C1(z)C_{1}(z) and C2(z)C_{2}(z), and comparing even and odd terms we conclude that

a(z)=g(z)(η1hq(z)φodd(z)+η2φeven(z)),a(z)=g(z)(\eta_{1}h_{q}(z)\varphi_{odd}(z)+\eta_{2}\varphi_{even}(z)),

and the first part of Lemma 4.8 follows immediately. The equation for b(z)/g(z)b(z)/g(z) also follows using similar arguments. ∎

Remark 4.9.

It follows from the above Lemma that η1\eta_{1}, η2\eta_{2}, η3\eta_{3} and η4\eta_{4} are all non-zero. We know that a(z)a(z) is a non-zero analytic function and thus a(qk)0a(q^{k})\neq 0 for large enough kk. However, as shown in Corollary A.4, g(qk)=hq(qk+1/2)=0g(q^{k})=h_{q}(q^{k+1/2})=0 for kk\in\mathbb{Z}. Thus, if η1=0\eta_{1}=0 or η2=0\eta_{2}=0 then we arrive at a contradiction. (Note that by Equation (4.4) φ(z)\varphi(z) cannot have poles along the real axis.)

Lemma 4.10.

ϕeven(z)\phi_{even}(z) as defined in Lemma 4.4 can be written as

ϕeven(z)=g(z)w(z)(λ3hq(z)b(z)+λ4a(z)),\phi_{even}(z)=g(z)w(z)(\lambda_{3}h_{q}(z)b(z)+\lambda_{4}a(z)),

where λ3\lambda_{3}, λ4\lambda_{4} are constants. Similarly

ϕodd(z)=g(z)w(z)(λ1hq(z)a(z)+λ2b(z)),\phi_{odd}(z)=g(z)w(z)(\lambda_{1}h_{q}(z)a(z)+\lambda_{2}b(z)),
Proof.

From Lemma 4.4 we conclude that ϕeven(z)\phi_{even}(z) satisfies the same difference equation (Equation (4.3)) as a(z)g(z)w(z)a(z)g(z)w(z) and b(z)g(z)w(z)b(z)g(z)w(z). It follows that

ϕeven(z)=g(z)w(z)(C1(z)a(z)+C2(z)b(z)),\phi_{even}(z)=g(z)w(z)(C_{1}(z)a(z)+C_{2}(z)b(z)),

for some functions Ci(z)C_{i}(z) that satisfy Ci(qz)=Ci(z)C_{i}(qz)=C_{i}(z). As a(z)a(z) and b(z)b(z) are analytic everywhere and ϕeven(z)\phi_{even}(z) is holomorphic in {0}\mathbb{C}\setminus\{0\} we conclude that Ci(z)C_{i}(z) is constant or has simple poles at the zeros of g(z)g(z), which occur at ±qk\pm q^{k} for kk\in\mathbb{Z}. Applying Corollary A.2 to C1(z)C_{1}(z) and C2(z)C_{2}(z), and comparing even and odd terms we conclude that

ϕeven(z)=g(z)w(z)(λ4a(z)+λ3hq(z)b(z)),\phi_{even}(z)=g(z)w(z)(\lambda_{4}a(z)+\lambda_{3}h_{q}(z)b(z)), (4.6)

and the first part of Lemma 4.8 follows immediately. The equation for ϕodd(z)\phi_{odd}(z) also follows using similar arguments. ∎

Remark 4.11.

We note that w(z)w(z) has poles at ei(π+2nπ)4qke^{\frac{i(\pi+2n\pi)}{4}}q^{-k} for kk\in\mathbb{N}, and n0,1,2,3n\in 0,1,2,3. As ϕeven\phi_{even} is analytic at these locations we conclude

λ4a(ei(π+2nπ)4qk)+λ3hq(ei(π+2nπ)4)b(ei(π+2nπ)4qk)=0.\lambda_{4}a(e^{\frac{i(\pi+2n\pi)}{4}}q^{-k})+\lambda_{3}h_{q}(e^{\frac{i(\pi+2n\pi)}{4}})b(e^{\frac{i(\pi+2n\pi)}{4}}q^{-k})=0.

Thus,

λ4=λ3hq(eiπ4)b(eiπ4)a(eiπ4).\lambda_{4}=-\lambda_{3}\frac{h_{q}(e^{\frac{i\pi}{4}})b(e^{\frac{i\pi}{4}})}{a(e^{\frac{i\pi}{4}})}.

We are now in a position to solve the RHP given in Definition 4.1.

Lemma 4.12.

The solution of the 𝔚\mathfrak{W}-RHP (Definition 4.1) is given by:

𝔚(z)={[η21a(z)λ2η21λ11w(z)b(z)λ3b(z)λ4w(z)a(z)],forz𝒟,[η21a(z)λ2η21λ11w(z)b(z)λ3b(z)λ4w(z)a(z)][g(z)1w(z)g(z)hq(z)0g(z)],forz𝒟+.\displaystyle\mathfrak{W}(z)=\left\{\begin{array}[]{lr}\begin{bmatrix}\eta_{2}^{-1}a(z)&\lambda_{2}\eta_{2}^{-1}\lambda_{1}^{-1}w(z)b(z)\\ \lambda_{3}b(z)&\lambda_{4}w(z)a(z)\end{bmatrix},&\text{for}\,z\in\mathcal{D}_{-},\\ \begin{bmatrix}\eta_{2}^{-1}a(z)&\lambda_{2}\eta_{2}^{-1}\lambda_{1}^{-1}w(z)b(z)\\ \lambda_{3}b(z)&\lambda_{4}w(z)a(z)\end{bmatrix}\begin{bmatrix}g(z)^{-1}&w(z)g(z)h_{q}(z)\\ 0&g(z)\end{bmatrix},&\text{for}\,z\in\mathcal{D}_{+}.\end{array}\right. (4.9)

where,

λ2\displaystyle\lambda_{2} =\displaystyle= λ1hq(eiπ4)b(eiπ4)a(eiπ4),\displaystyle-\lambda_{1}\frac{h_{q}(e^{\frac{i\pi}{4}})b(e^{\frac{i\pi}{4}})}{a(e^{\frac{i\pi}{4}})}, (4.10)
λ4\displaystyle\lambda_{4} =\displaystyle= λ3hq(eiπ4)b(eiπ4)a(eiπ4).\displaystyle-\lambda_{3}\frac{h_{q}(e^{\frac{i\pi}{4}})b(e^{\frac{i\pi}{4}})}{a(e^{\frac{i\pi}{4}})}. (4.11)
Proof.

First we show condition (i) (meromorphicity) is satisfied. With the choice of λi\lambda_{i} given by Equations (4.10) and (4.11) it is clear from Remark 4.11 that 𝔚(z)\mathfrak{W}(z) has analytic entries in the second column for z𝒟+z\in\mathcal{D}_{+}. In particular the second column has entries equal to η21λ11ϕodd(z)\eta_{2}^{-1}\lambda_{1}^{-1}\phi_{odd}(z) and ϕeven(z)\phi_{even}(z) for z𝒟+z\in\mathcal{D}_{+}. Meromorphicity of the LHS column follows immediately from the definition of a(z)a(z) and b(z)b(z).

Conditions (ii) and (iv) follow immediately from the definition of 𝔚(z)\mathfrak{W}(z). It is left to show condition (iii) is satisfied. Taking the limit zz\to\infty of Equation (4.6) and applying Lemma 4.8 we find

λ3η4hq(z)2+λ4η21g(z)2w(z).\lambda_{3}\eta_{4}h_{q}(z)^{2}+\lambda_{4}\eta_{2}\sim\frac{1}{g(z)^{2}w(z)}.

Thus,

λ3=limk1η4g(qk)2w(qk)hq(qk)2.\lambda_{3}=\lim_{k\to\infty}\frac{1}{\eta_{4}g(q^{-k})^{2}w(q^{-k})h_{q}(q^{-k})^{2}}.

Let

c0=limk1g(qk)2w(qk)hq(qk)2.c_{0}=\lim_{k\to\infty}\frac{1}{g(q^{-k})^{2}w(q^{-k})h_{q}(q^{-k})^{2}}. (4.12)

Hence, applying Lemma 4.8 again we find that as zz\to\infty

λ3b(z)g(z)1hq(z)c0.\lambda_{3}b(z)g(z)^{-1}\sim h_{q}(z)c_{0}.

We conclude that as zz\to\infty the matrix 𝔚(z)\mathfrak{W}(z) behaves like

𝔚(z)\displaystyle\mathfrak{W}(z) =\displaystyle= [η21a(z)λ2η21λ11w(z)b(z)λ3b(z)λ4w(z)a(z)][g(z)1w(z)g(z)hq(z)0g(z)]\displaystyle\begin{bmatrix}\eta_{2}^{-1}a(z)&\lambda_{2}\eta_{2}^{-1}\lambda_{1}^{-1}w(z)b(z)\\ \lambda_{3}b(z)&\lambda_{4}w(z)a(z)\end{bmatrix}\begin{bmatrix}g(z)^{-1}&w(z)g(z)h_{q}(z)\\ 0&g(z)\end{bmatrix} (4.13)
=\displaystyle= [10hq(z)c01]+O(1/z)\displaystyle\begin{bmatrix}1&0\\ h_{q}(z)c_{0}&1\end{bmatrix}+O(1/z)

5. Far-field RHP

In this section, we solve the far-field RHP, which we denote by 𝒲\mathcal{W}-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 tt instead of zz to denote a complex variable in this section. Motivated by the form of Equation (3.17b) we introduce the following RHP.

Definition 5.1 (𝒲\mathcal{W}-RHP).

Let Γ\Gamma be an appropriate curve (see Definition 1.3) with interior 𝒟\mathcal{D}_{-} and exterior 𝒟+\mathcal{D}_{+}. A 2×22\times 2 complex matrix function 𝒲(t)\mathcal{W}(t), tt\in\mathbb{C}, is a solution of the RHP if it satisfies the following conditions:

  1. (i)

    𝒲(t)\mathcal{W}(t) is meromorphic in Γ\mathbb{C}\setminus\Gamma, with simple poles at t=±qkt=\pm q^{k} for kk\in\mathbb{Z}.

  2. (ii)

    𝒲(t)\mathcal{W}(t) has continuous boundary values 𝒲(s)\mathcal{W}^{-}(s) and 𝒲+(s)\mathcal{W}^{+}(s) as tt approaches sΓs\in\Gamma from 𝒟\mathcal{D}_{-} and 𝒟+\mathcal{D}_{+} respectively, where

    𝒲+(s)=𝒲(s)[g(s)00g(s)1],sΓ.\displaystyle\mathcal{W}^{+}(s)=\mathcal{W}^{-}(s)\begin{bmatrix}g(s)&0\\ 0&g(s)^{-1}\end{bmatrix},\;s\in\Gamma. (5.1a)
  3. (iii)

    𝒲(t)\mathcal{W}(t) satisfies

    𝒲(t)=I+O(1|t|), as |t|.\mathcal{W}(t)=I+O\left(\frac{1}{|t|}\right),\text{ as $|t|\rightarrow\infty$}. (5.1b)

    Due to the simple poles in the second column of 𝒲(t)\mathcal{W}(t), the asymptotic decay condition only holds for |z±qk|>R|z\pm q^{-k}|>R for any R>0R>0.

  4. (iv)

    The residue at the poles t=±qkt=\pm q^{k} for kk\in\mathbb{N} is given by

    Res(𝒲(±qk))=limt±qk𝒲(t)[00(tqk)g(t)2hq(t)1ω(t)10],\displaystyle\text{Res}(\mathcal{W}(\pm q^{-k}))=\lim_{t\rightarrow\pm q^{-k}}\mathcal{W}(t)\begin{bmatrix}0&0\\ (t\mp q^{-k})g(t)^{-2}h_{q}(t)^{-1}\omega(t)^{-1}&0\end{bmatrix}, (5.1c)

    where ω(t)\omega(t) is defined in Equation (3.4).

  5. (v)

    The residue at the poles t=±qkt=\pm q^{-k} for 1k1\leq k\in\mathbb{N} is given by

    Res(𝒲(±qk))=limt±qk𝒲(t)[0(tqk)hq(t)ω(t)00].\displaystyle\text{Res}(\mathcal{W}(\pm q^{-k}))=\lim_{t\rightarrow\pm q^{-k}}\mathcal{W}(t)\begin{bmatrix}0&(t\mp q^{-k})h_{q}(t)\omega(t)\\ 0&0\end{bmatrix}. (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

y1(q2t)q7t4(1q2(q+1)t2)y1(q1t)+y1(t)=0.y_{1}(q^{-2}t)q^{7}t^{-4}-(1-q^{2}(q+1)t^{-2})y_{1}(q^{-1}t)+y_{1}(t)=0. (5.2)

There exists a solution of Equation (5.2), analytic in {0}\mathbb{C}\setminus\{0\}, which can be represented by the even power series

a(t)=j=0a2jt2j,a_{\infty}(t)=\sum_{j=0}^{\infty}a_{2j}t^{-2j},

where we take a0=1a_{0}=1 (w.l.o.g.).

Similarly, there exists a solution analytic in {0}\mathbb{C}\setminus\{0\}, of the difference equation

y2(q2t)q7t4(q1q2(q+1)t2)y2(q1t)+y2(t)=0,y_{2}(q^{-2}t)q^{7}t^{-4}-(q^{-1}-q^{2}(q+1)t^{-2})y_{2}(q^{-1}t)+y_{2}(t)=0, (5.3)

which can be represented by the odd power series

b(t)=j=0b2j+1t2j1,b_{\infty}(t)=\sum_{j=0}^{\infty}b_{2j+1}t^{-2j-1},

without loss of generality let b1=1b_{1}=1.

Proof.

Let

y1(z)=i=0yiti.y_{1}(z)=\sum_{i=0}^{\infty}y_{i}t^{-i}.

Substituting this into Equation (5.2) and comparing coefficients of tt, we find that y(t)y(t) is a solution iff

yi=(1+q)qiyi2+q2i1yi4qi1.y_{i}=-\frac{(1+q)q^{i}y_{i-2}+q^{2i-1}y_{i-4}}{q^{i}-1}.

The first part of Lemma 5.2 follows immediately. The second part follows using similar arguments. ∎

Lemma 5.3.

Define

α(t)=y1(t)g(t)ω(t),\alpha(t)=y_{1}(t)g(t)\omega(t),

where y1(t)y_{1}(t) satisfies the difference equation given by Equation (5.2) and ω(t)\omega(t) is defined in Equation (3.4). Then α(t)\alpha(t) is a solution of the difference equation

qα(q2t)+(t2q2(1+q))α(q1t)+α(t)=0.q\alpha(q^{-2}t)+(t^{2}q^{-2}-(1+q))\alpha(q^{-1}t)+\alpha(t)=0. (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

β(t)=y2(t)g(t)ω(t),\beta(t)=y_{2}(t)g(t)\omega(t),

where y2(t)y_{2}(t) satisfies the difference equation given by Equation (5.3), then β(t)\beta(t) satisfies the difference equation

qβ(q2t)+(t2q3(1+q))β(q1t)+β(t)=0.q\beta(q^{-2}t)+(t^{2}q^{-3}-(1+q))\beta(q^{-1}t)+\beta(t)=0. (5.5)

Furthermore, there exists two entire solutions which can be represented by an even and odd power series.

Proof.

From the definition of α(t)\alpha(t), Equations (3.2) and (3.5) we find that

y1(q1t)\displaystyle y_{1}(q^{-1}t) =\displaystyle= α(q1t)g(q1t)ω(q1t),\displaystyle\frac{\alpha(q^{-1}t)}{g(q^{-1}t)\omega(q^{-1}t)},
=\displaystyle= q2t2α(q1t)g(t)w(t).\displaystyle-\frac{q^{-2}t^{2}\alpha(q^{-1}t)}{g(t)w(t)}.

Substituting the above into Equation (5.2) we determine that α(t)\alpha(t) satisfies the difference equation

α(q2t)q+(q2t2(q+1))α(q1t)+α(z)=0.\alpha(q^{-2}t)q+(q^{-2}t^{2}-(q+1))\alpha(q^{-1}t)+\alpha(z)=0.

Let α(t)=i=0αiti\alpha(t)=\sum_{i=0}^{\infty}\alpha_{i}t^{i}, such a power series is a solution of Equation (5.4) iff

αi=q1iαk21(1+q)qi+q12i.\alpha_{i}=-\frac{q^{-1-i}\alpha_{k-2}}{1-(1+q)q^{-i}+q^{1-2i}}.

The first part of Lemma 5.3 follows immediately. The second part follows using similar arguments. ∎

Definition 5.4.

Following from Lemma 5.3 we define Ψeven(t)\Psi_{even}(t) as the even entire solution to Equation (5.4), normalised such that its first non-zero Taylor series coefficient (t0)(t^{0}) is unity. Similarly we define Ψodd(t)\Psi_{odd}(t) as the odd entire solution to Equation (5.4), normalised such that its first non-zero Taylor series coefficient (t1)(t^{1}) is unity.

Furthermore, we define Ψeven(t)\varPsi_{even}(t) as the even entire solution to Equation (5.5), normalised such that its first non-zero Taylor series coefficient (t0)(t^{0}) is unity. Similarly we define Ψodd(t)\varPsi_{odd}(t) as the odd entire solution to Equation (5.5), normalised such that its first non-zero Taylor series coefficient (t1)(t^{1}) is unity.

Lemma 5.5.

Define

Θ1(t)=y1(t)g(t)1,\Theta_{1}(t)=y_{1}(t)g(t)^{-1},

where y1(t)y_{1}(t) is a solution of Equation (5.2). Then Θ1(t)\Theta_{1}(t) is a solution of Equation (5.4).

Similarly, Θ2(t)=y2(t)g(t)1\Theta_{2}(t)=y_{2}(t)g(t)^{-1} is a solution of Equation (5.5).

Proof.

From the definition of Θ1(t)\Theta_{1}(t) we find that

Θ1(q1t)\displaystyle\Theta_{1}(q^{-1}t) =\displaystyle= y1(q1t)g(q1t),\displaystyle\frac{y_{1}(q^{-1}t)}{g(q^{-1}t)}, (5.6)
=\displaystyle= α(q1t)g(q1t)2w(q1t),\displaystyle\frac{\alpha(q^{-1}t)}{g(q^{-1}t)^{2}w(q^{-1}t)},
=\displaystyle= α(q1t)g(t)2w(t),\displaystyle\frac{\alpha(q^{-1}t)}{g(t)^{2}w(t)},

where we have used Equation (3.2) and Equation (3.5) to arrive at the final line. The first part Lemma 5.5 follows immediately from Equation (5.6). The proof for Θ2(t)\Theta_{2}(t) follows from the same arguments. ∎

Lemma 5.6.

The function a(t)g(t)1a_{\infty}(t)g(t)^{-1}, where a(t)a_{\infty}(t) is defined in Lemma 5.2, can be written as

a(t)g(t)1=μ1hq(t)Ψodd(t)+μ2Ψeven(t),a_{\infty}(t)g(t)^{-1}=\mu_{1}h_{q}(t)\Psi_{odd}(t)+\mu_{2}\Psi_{even}(t),

where μ1,μ2\mu_{1},\mu_{2} are a constants. Similarly aodd(t)g(t)1a_{odd}(t)g(t)^{-1} can be written as

b(t)g(t)1=μ3Ψodd(t)+μ4hq(t)Ψeven(t).b_{\infty}(t)g(t)^{-1}=\mu_{3}\varPsi_{odd}(t)+\mu_{4}h_{q}(t)\varPsi_{even}(t). (5.7)
Proof.

From Lemma 5.5 we conclude that a(t)a_{\infty}(t) satisfies the same difference equation (Equation (5.2)) as g(t)Ψeven(t)g(t)\Psi_{even}(t) and g(t)Ψodd(t)g(t)\Psi_{odd}(t). It follows that

a(t)=g(t)(C1(t)Ψeven(t)+C2(t)Ψodd(t)),a_{\infty}(t)=g(t)(C_{1}(t)\Psi_{even}(t)+C_{2}(t)\Psi_{odd}(t)),

for some functions Ci(t)C_{i}(t) which satisfy Ci(qt)=Ci(t)C_{i}(qt)=C_{i}(t). As a(t)a_{\infty}(t) is analytic in {0}\mathbb{C}\setminus\{0\} and Ψeven(t)\Psi_{even}(t) and Ψodd(t)\Psi_{odd}(t) are entire we conclude that Ci(t)C_{i}(t) is constant or has simple poles at the zeros of g(t)g(t), which occur at ±qk\pm q^{k} for kk\in\mathbb{Z}. Applying Corollary A.2 and comparing even and odd terms we conclude that

a(t)=g(t)(μ1hq(t)Ψodd(t)+μ2Ψeven(t)),a_{\infty}(t)=g(t)(\mu_{1}h_{q}(t)\Psi_{odd}(t)+\mu_{2}\Psi_{even}(t)),

and the first part of Lemma 5.6 follows immediately. The equation for b(t)g(t)1b_{\infty}(t)g(t)^{-1} also follows using similar arguments. ∎

Lemma 5.7.

Let b(t)b_{\infty}(t), μ4\mu_{4} and Ψeven(t)\varPsi_{even}(t) be defined as in Lemma 5.6. Furthermore, define the constant \mathcal{H} as

=ω(q)g(q)2hq(q)2.\mathcal{H}=\omega(q)g(q)^{2}h_{q}(q)^{2}. (5.8)

Then,

1μ4b(t)g(t)1hq(t),\frac{1}{\mu_{4}\mathcal{H}}b_{\infty}(t)g(t)^{-1}\sim\frac{h_{q}(t)}{\mathcal{H}}, (5.9)

as t0t\to 0. It is also true that

Res(1μ4b(qk)ω(qk)hq(qk))=Res(Ψeven(qk)g(qk)1),\mathrm{Res}\left(\frac{1}{\mu_{4}\mathcal{H}}b_{\infty}(q^{k})\omega(q^{k})h_{q}(q^{k})\right)=\mathrm{Res}(\varPsi_{even}(q^{k})g(q^{k})^{-1}), (5.10)

for kk\in\mathbb{Z}.

Proof.

Equation (5.9) follows immediately from taking the limit t0t\to 0 in Equation (5.7). Multiplying Equation (5.7) by ω(t)g(t)hq(t)\omega(t)g(t)h_{q}(t) we find

b(t)ω(t)hq(t)=ω(t)g(t)hq(t)μ3Ψodd(t)+ω(t)g(t)2hq(t)2μ4(Ψeven(t)g(t)1).b_{\infty}(t)\omega(t)h_{q}(t)=\omega(t)g(t)h_{q}(t)\mu_{3}\varPsi_{odd}(t)+\omega(t)g(t)^{2}h_{q}(t)^{2}\mu_{4}(\varPsi_{even}(t)g(t)^{-1}).

Studying the residue at t=qkt=q^{k} for kk\in\mathbb{Z} we find that

Res(b(qk)ω(qk)hq(qk))\displaystyle\mathrm{Res}(b_{\infty}(q^{k})\omega(q^{k})h_{q}(q^{k})) =\displaystyle= Res(ω(qk)g(qk)2hq(qk)2μ4(Ψeven(qk)g(qk)1)),\displaystyle\mathrm{Res}\left(\omega(q^{k})g(q^{k})^{2}h_{q}(q^{k})^{2}\mu_{4}(\varPsi_{even}(q^{k})g(q^{k})^{-1})\right),
=\displaystyle= Res(μ4Ψeven(qk)g(qk)1).\displaystyle\mathrm{Res}\left(\mathcal{H}\mu_{4}\varPsi_{even}(q^{k})g(q^{k})^{-1}\right).

Remark 5.8.

Repeating the arguments of Lemma 5.7 one can readily show

Res(a(qk)ω(qk)hq(qk))=Res(μ2Ψeven(qk)g(qk)1).\mathrm{Res}(a_{\infty}(q^{k})\omega(q^{k})h_{q}(q^{k}))=\mathrm{Res}\left(\mathcal{H}\mu_{2}\Psi_{even}(q^{k})g(q^{k})^{-1}\right). (5.11)

We require one last Lemma before determining the solution of the far-field RHP.

Lemma 5.9.

Let Ψeven(t)\varPsi_{even}(t) be defined as in Lemma 5.3, then

Ψeven(t)g(t)1=cΨ+O(t1),ast,\varPsi_{even}(t)g(t)^{-1}=c_{\varPsi}+O(t^{-1}),\;\mathrm{as}\;t\to\infty,

where cΨc_{\varPsi} is a non-zero constant and this limit clearly does not hold near the poles of Ψeven(t)g(t)1\varPsi_{even}(t)g(t)^{-1}, but holds for tt satisfying |tqk|>r|t-q^{k}|>r, for some fixed r>0r>0 and all kk\in\mathbb{Z}.

Proof.

We first show that the residue of the poles of Ψeven(t)g(t)1\varPsi_{even}(t)g(t)^{-1} are vanishing faster than qk2/2q^{k^{2}/2} as |k||k|\to\infty.

Consider the case k+k\to+\infty (t0t\to 0), from Equation (3.3) we find that

g(qn/2z)=(1)n/2qn2/4+n/2g(z).g(q^{n/2}z)=(-1)^{n/2}q^{-n^{2}/4+n/2}g(z).

By Lemma 5.3 we know that Ψeven(qk)1\varPsi_{even}(q^{k})\sim 1 as kk\to\infty. Thus, we conclude

Res(Ψeven(qk)g(qk)1)<O(qk2/2),ask+.\mathrm{Res}(\varPsi_{even}(q^{k})g(q^{k})^{-1})<O(q^{k^{2}/2}),\;\text{as}\;k\to+\infty.

Note that the above statement is true for O(qck2)O(q^{ck^{2}}), with c<1c<1.

Consider the case kk\to-\infty. From Lemma 5.7 it is clear that a bound on Res(b(qk)ω(qk)hq(qk))\mathrm{Res}(b_{\infty}(q^{k})\omega(q^{k})h_{q}(q^{k})) as kk\to-\infty is equivalent to a bound on Res(Ψeven(qk)g(qk)1)\mathrm{Res}(\varPsi_{even}(q^{k})g(q^{k})^{-1}). By definition in Lemma 5.2 we determine that b(qk)=O(qk)b_{\infty}(q^{k})=O(q^{-k}). Furthermore, using induction on Equation (3.5) we find

ω(qkt)=q2k(k1)t4kw(t).\omega(q^{k}t)=q^{2k(k-1)}t^{4k}w(t).

It follows

Res(b(qk)ω(qk)hq(qk))<O(qk2),ask.\mathrm{Res}(b_{\infty}(q^{k})\omega(q^{k})h_{q}(q^{k}))<O(q^{k^{2}}),\;\text{as}\;k\to-\infty.

Let RkR_{k} be the residue of Ψeven(t)g(t)1\varPsi_{even}(t)g(t)^{-1} at t=qkt=q^{k}. Define the function

F(t)=Ψeven(t)g(t)1k=Rktqk,F(t)=\varPsi_{even}(t)g(t)^{-1}-\sum_{k=-\infty}^{\infty}\frac{R_{k}}{t-q^{k}},

where this sum is well defined for all tt because we have just shown Rk<O(qk2/2)R_{k}<O(q^{k^{2}/2}) as |k||k|\to\infty. It follows F(t)F(t) is holomorphic in {0}\mathbb{C}\setminus\{0\} and can be represented by a Laurent series which converges everywhere. We will show that F(t)=j=0FjtjF(t)=\sum_{j=0}^{\infty}F_{j}t^{-j} (i.e. there are no positive powers of tt). Applying Equation (5.5) and Equation (3.2) we find that v(t)=Ψeven(t)g(t)1v(t)=\varPsi_{even}(t)g(t)^{-1} satisfies the difference equation

v(q2t)+((1+q)q3t21)v(q1t)+q5t4v(t)=0.v(q^{-2}t)+((1+q)q^{3}t^{-2}-1)v(q^{-1}t)+q^{5}t^{-4}v(t)=0. (5.12)

Writing the above in matrix form we have

[v(q2t)v(q1t)]=([1010]q3t2[(1+q)q2t400])[v(q1t)v(t)].\displaystyle\begin{bmatrix}v(q^{-2}t)\\ v(q^{-1}t)\end{bmatrix}=\left(\begin{bmatrix}1&0\\ 1&0\end{bmatrix}-q^{3}t^{-2}\begin{bmatrix}-(1+q)&-q^{2}t^{-4}\\ 0&0\end{bmatrix}\right)\begin{bmatrix}v(q^{-1}t)\\ v(t)\end{bmatrix}. (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 v(t)v(t) as tt\to\infty is essentially a Pochhammer symbol with matrix entries. Thus, v(t)v(t), and consequently F(t)F(t) are bounded by a constant as tt\to\infty. We now show that this constant is non-zero. From Lemma 5.3 we know that Ψeven(t)\varPsi_{even}(t) is an entire function (which is not the constant function), hence Ψeven(t)\varPsi_{even}(t) must grow in some direction. Ψeven(t)\varPsi_{even}(t) satisfies Equation (5.5),

qΨeven(q2t)+(t2q3(1+q))Ψeven(q1t)+Ψeven(t)=0.q\varPsi_{even}(q^{-2}t)+(t^{2}q^{-3}-(1+q))\varPsi_{even}(q^{-1}t)+\varPsi_{even}(t)=0.

It follows that as tt becomes large there must exist a ray where

t2q3Ψeven(q1t)Ψeven(t).t^{2}q^{-3}\varPsi_{even}(q^{-1}t)\gg\varPsi_{even}(t).

Thus, along this ray

Ψeven(q2t)=t2q3Ψeven(q1t)(1+O(t2)),\varPsi_{even}(q^{-2}t)=-t^{2}q^{-3}\varPsi_{even}(q^{-1}t)(1+O(t^{-2})),

and applying Equation (3.2) we conclude Ψeven(t)g(t)1\varPsi_{even}(t)g(t)^{-1} approaches a constant along this ray. Thus, F(t)=j=0FjtjF(t)=\sum_{j=0}^{\infty}F_{j}t^{-j} and F00F_{0}\neq 0. Lemma 5.9 follows immediately. ∎

Remark 5.10.

Repeating the same arguments as in Lemma 5.9 we can conclude

Ψodd(t)g(t)1O(t1),ast,\Psi_{odd}(t)g(t)^{-1}\to O(t^{-1}),\,\mathrm{as}\,t\to\infty, (5.14)

where again this limit clearly does not hold near the poles of Ψeven(t)g(t)1\varPsi_{even}(t)g(t)^{-1}, but holds for tt satisfying |tqk|>r|t-q^{k}|>r, for some fixed r>0r>0 and all kk\in\mathbb{Z}.

Now we are in a position to solve the far-field RHP given in Definition 5.1. Let 𝒲(t)\mathcal{W}(t) be given by

𝒲(t)={[a(t)g(t)1μ2Ψodd(t)b(t)g(t)1μ4cΨΨeven(t)cΨ1]forz𝒟,[a(t)μ2Ψodd(t)g(t)1b(t)μ4cΨΨeven(t)g(t)1cΨ1],forz𝒟+.\mathcal{W}(t)=\begin{cases}\begin{bmatrix}a_{\infty}(t)g(t)^{-1}&\mu_{2}\mathcal{H}\Psi_{odd}(t)\\ \dfrac{b_{\infty}(t)g(t)^{-1}}{\mu_{4}\mathcal{H}c_{\varPsi}}&\varPsi_{even}(t)c_{\varPsi}^{-1}\end{bmatrix}&\text{for}\,z\in\mathcal{D}_{-},\\ &\\ \begin{bmatrix}a_{\infty}(t)&\mu_{2}\mathcal{H}\Psi_{odd}(t)g(t)^{-1}\\ \dfrac{b_{\infty}(t)}{\mu_{4}\mathcal{H}c_{\varPsi}}&\varPsi_{even}(t)g(t)^{-1}c_{\varPsi}^{-1}\end{bmatrix},&\text{for}\,z\in\mathcal{D}_{+}.\end{cases} (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 aa_{\infty}, bb_{\infty}, Ψodd\Psi_{odd} and Ψeven\varPsi_{even} are all analytic in {0}\mathbb{C}\setminus\{0\} (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 𝒲(1,1)(t)=1+O(1/t2)\mathcal{W}_{(1,1)}(t)=1+O(1/t^{2}) as tt\to\infty by the definition of a(t)a_{\infty}(t), similarly 𝒲(1,2)(t)=O(1/t)\mathcal{W}_{(1,2)}(t)=O(1/t) as tt\to\infty by the definition of b(t)b_{\infty}(t). From Lemma 5.9 we conclude that 𝒲(2,2)(t)=1+O(1/t)\mathcal{W}_{(2,2)}(t)=1+O(1/t) as tt\to\infty. Similarly from Remark 5.10 we conclude 𝒲(2,1)(t)=O(1/t)\mathcal{W}_{(2,1)}(t)=O(1/t) as tt\to\infty. 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

𝒲(t)=[μ20hq(t)cΨcΨ1]+O(t),ast0.\displaystyle\mathcal{W}(t)=\begin{bmatrix}\mu_{2}&0\\ \dfrac{h_{q}(t)}{\mathcal{H}c_{\varPsi}}&c_{\varPsi}^{-1}\end{bmatrix}+O(t),\;\text{as}\,t\to 0. (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 Wn(z)W_{n}(z) as nn\to\infty. The near- and far-field variables in Sections 4 and 5 are related by the linear transformation t=zqn/2t=zq^{n/2}.

We first make a linear transformation to the near-field RHP solution. Let

𝔚~(z)=[μ200cΨ1]𝔚(z).\displaystyle\widetilde{\mathfrak{W}}(z)=\begin{bmatrix}\mu_{2}&0\\ 0&c_{\varPsi}^{-1}\end{bmatrix}\mathfrak{W}(z). (6.1)

From Equations (1.11) and (3.4) one can readily determine that w(z)ω(z)w(z)\sim\omega(z) as zz\to\infty. Thus, comparing Equations (4.12) and (5.8) we find that =c01\mathcal{H}=c_{0}^{-1}. Hence, applying Equations (4.13) and (5.16) we find

limz𝔚~(z)=limt0𝒲(t).\lim_{z\to\infty}\widetilde{\mathfrak{W}}(z)=\lim_{t\to 0}\mathcal{W}(t).

We next make a slight modification to 𝒲(t)\mathcal{W}(t) given by Equation (5.15). Note that the residue condition for Wn(z)W_{n}(z) given in Equation (3.17e) is different to that for 𝒲(t)\mathcal{W}(t) given in Equation (5.1d). To resolve this issue we define the new function

𝒲~(t)=𝒲(t)(1w(z)z2ncn2nω(t))𝒲(t)[0001].\displaystyle\widetilde{\mathcal{W}}(t)=\mathcal{W}(t)-\left(1-\frac{w(z)z^{2n}c_{n}^{2n}}{\omega(t)}\right)\mathcal{W}(t)\begin{bmatrix}0&0\\ 0&1\end{bmatrix}. (6.2)

Substituting in Equation (3.8) we find that

𝒲~(t)=𝒲(t)(1(z2;q4))𝒲(t)[0001].\displaystyle\widetilde{\mathcal{W}}(t)=\mathcal{W}(t)-(1-(-z^{-2};q^{4})_{\infty})\mathcal{W}(t)\begin{bmatrix}0&0\\ 0&1\end{bmatrix}. (6.3)

Thus, the difference between 𝒲~(t)\widetilde{\mathcal{W}}(t) and 𝒲(t)\mathcal{W}(t) is bounded by O(qn/2)O(q^{n/2}) for z>qn/4z>q^{-n/4} (t>qn/4t>q^{n/4}).

Define

G(z)={𝔚~(z),forz𝒟,qn/4,𝒲~(zqn/2),forz𝒟+,qn/4.\displaystyle G(z)=\left\{\begin{array}[]{lr}\widetilde{\mathfrak{W}}(z),&\text{for}\;z\in\mathcal{D}_{-,q^{-n/4}},\\ \widetilde{\mathcal{W}}(zq^{n/2}),&\text{for}\;z\in\mathcal{D}_{+,q^{-n/4}}.\end{array}\right. (6.6)

and, furthermore

R(ζ)=Wn(ζqn/4)G(ζqn/4)1.R(\zeta)=W_{n}(\zeta q^{-n/4})G(\zeta q^{-n/4})^{-1}. (6.7)

Then, R(ζ)R(\zeta) satisfies the following RHP.

Definition 6.1 (R(ζ)R(\zeta) RHP).

A 2×22\times 2 complex matrix function R(ζ)R(\zeta), ζ\zeta\in\mathbb{C}, is a solution of the R(ζ)R(\zeta) RHP if it satisfies the following conditions:

  1. (i)

    R(ζ)R(\zeta) is analytic in Γ\mathbb{C}\setminus\Gamma.

  2. (ii)

    R(ζ)R(\zeta) has continuous boundary values R(s)R^{-}(s) and R+(s)R^{+}(s) as ζ\zeta approaches sΓs\in\Gamma from 𝒟\mathcal{D}_{-} and 𝒟+\mathcal{D}_{+} respectively, where

    R+(s)=R(s)𝔚~(sqn/4)1𝒲~(sqn/4)sΓ.\displaystyle R^{+}(s)=R^{-}(s)\widetilde{\mathfrak{W}}(sq^{-n/4})^{-1}\widetilde{\mathcal{W}}(sq^{n/4})\,\;s\in\Gamma. (6.8a)
  3. (iii)

    R(ζ)R(\zeta) satisfies

    R(ζ)=I+O(1|ζ|), as |ζ|.R(\zeta)=I+O\left(\frac{1}{|\zeta|}\right),\text{ as $|\zeta|\rightarrow\infty$}. (6.8b)

From Equations (4.13) and (5.16) we find that

𝔚~(sqn/4)1𝒲~(sqn/4)IΓ=O(qn/4).\|\widetilde{\mathfrak{W}}(sq^{-n/4})^{-1}\widetilde{\mathcal{W}}(sq^{n/4})-I\|_{\Gamma}=O(q^{n/4}).

Thus, we can apply Theorem B.2 to conclude

|R(ζ)I|=O(qn/4).|R(\zeta)-I|=O(q^{n/4}). (6.9)

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.

By the definition of R(ζ)R(\zeta), we find that

Wn(z)G(z)1\displaystyle W_{n}(z)G(z)^{-1} =\displaystyle= R(zqn/4),\displaystyle R(zq^{n/4}),
=\displaystyle= I+O(qn/4),\displaystyle I+O(q^{n/4}),
Wn(z)\displaystyle W_{n}(z) =\displaystyle= (I+O(qn/4))G(z).\displaystyle(I+O(q^{n/4}))G(z).

Looking at the (1,1)(1,1)-entry of Wn(z)W_{n}(z) we find that for zqn/4z\leq q^{-n/4},

cnnPn(z)=G(1,1)(z)+O(qn/4)G(2,1)(z).c_{n}^{n}P_{n}(z)=G_{(1,1)}(z)+O(q^{n/4})G_{(2,1)}(z).

Thus, aplying Equation (4.12) we find

(1)n/2qn2(n21)Pn(z)=μ2η2a(z)+O(qn/4)λ3cΨb(z).(-1)^{n/2}q^{\frac{n}{2}(\frac{n}{2}-1)}P_{n}(z)=\frac{\mu_{2}}{\eta_{2}}a(z)+O(q^{n/4})\frac{\lambda_{3}}{c_{\Psi}}b(z).

Repeating the arguments above for each matrix entry, Theorem 1.5 follows immediately. ∎

Proof of Theorem 1.6.

Using the transformations detailed in Section 3.1 we find that

Wn(z)=[1001]+1z[0γncn2nγn11cn2n0]+O(z2).\displaystyle W_{n}(z)=\begin{bmatrix}1&0\\ 0&1\end{bmatrix}+\frac{1}{z}\begin{bmatrix}0&\gamma_{n}c_{n}^{2n}\\ \gamma_{n-1}^{-1}c_{n}^{-2n}&0\end{bmatrix}+O(z^{-2}).

Let,

𝒲(t)=[1001]+1t[0AB0]+O(t2).\displaystyle\mathcal{W}(t)=\begin{bmatrix}1&0\\ 0&1\end{bmatrix}+\frac{1}{t}\begin{bmatrix}0&A\\ B&0\end{bmatrix}+O(t^{-2}).

Note that there is no difference in the O(1/t)O(1/t) term between 𝒲(t)\mathcal{W}(t) and 𝒲~(t)\widetilde{\mathcal{W}}(t). Using the definition of R(ζ)R(\zeta) given in Equation (6.7) and Equation (6.9) we find

γn(1)n/2q12(n21))2n\displaystyle\gamma_{n}(-1)^{n/2}q^{\frac{1}{2}(\frac{n}{2}-1)})^{2n} =\displaystyle= A(1+O(qn/2))qn/2,\displaystyle A(1+O(q^{n/2}))q^{-n/2},
γn\displaystyle\gamma_{n} =\displaystyle= A(1+O(qn/2))qn(n21)qn/2,\displaystyle A(1+O(q^{n/2}))q^{-n(\frac{n}{2}-1)}q^{-n/2},
=\displaystyle= A(1+O(qn/2))qn(n1)/2.\displaystyle A(1+O(q^{n/2}))q^{-n(n-1)/2}.

Similarly in the bottom left term we find in the limit nn\to\infty:

γn1=qn2(3n)(B1+O(qn/2)).\gamma_{n-1}=q^{\frac{n}{2}(3-n)}\left(B^{-1}+O(q^{n/2})\right).

Taking the ratio of γn\gamma_{n} and γn1\gamma_{n-1} we find that

αn=γn/γn1=ABqn(1+O(qn/2)).\alpha_{n}=\gamma_{n}/\gamma_{n-1}=ABq^{-n}(1+O(q^{n/2})).

However, we can determine that AB=qAB=q by considering the arguments presented in Theorem 8.3. ∎

8. Recurrence coefficients and qq-discrete Painlevé

As discussed in Remark 2.5, the class of monic polynomials {Pn(c)}n=0\{P_{n}^{(c)}\}_{n=0}^{\infty} satisfying the orthogonality condition

Pn(c)(cx)Pm(c)(cx)w(cx)dqx=γn(c)δn,m,\int^{\infty}_{-\infty}P^{(c)}_{n}(cx)P^{(c)}_{m}(cx)w(cx)d_{q}x=\gamma^{(c)}_{n}\delta_{n,m}, (8.1)

where q<c1q<c\leq 1, satisfy a corresponding RHP. In this section we discuss the connection between the RHP, the asymptotic behaviour of Pn(c)P_{n}^{(c)} and uniqueness results concerning their recurrence coefficients. First, we use the RHP to show that in general Pn(c)Pn(1)P_{n}^{(c)}\neq P_{n}^{(1)}.

Lemma 8.1.

Let {Pn(c)}n=0\{P_{n}^{(c)}\}_{n=0}^{\infty} 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 c=1c=1 and c=q1/2c=q^{1/2} are the same. Furthermore,

γ1(1)γ1(q1/2)=q1/2hq(eiπ/4)hq(q1/2eiπ/4).\frac{\gamma^{(1)}_{1}}{\gamma^{(q^{1/2})}_{1}}=\frac{q^{1/2}h_{q}(e^{i\pi/4})}{h_{q}(q^{1/2}e^{i\pi/4})}.

Moreover, if c1,q1/2c\neq 1,q^{1/2} then {Pn(c)}n=0{Pn(1)}n=0\{P_{n}^{(c)}\}_{n=0}^{\infty}\neq\{P_{n}^{(1)}\}_{n=0}^{\infty}.

Proof.

Let

Y^(1,2)(1)(z)=Pn(1)(x)w(x)zxdqxPn(1)(z)w(z)hq(z).\widehat{Y}^{(1)}_{(1,2)}(z)=\int^{\infty}_{-\infty}\frac{P_{n}^{(1)}(x)w(x)}{z-x}d_{q}x-P_{n}^{(1)}(z)w(z)h_{q}(z).

From the arguments in Section 2 it follows that Y^(1,2)(1)(z)\widehat{Y}^{(1)}_{(1,2)}(z) is meromorphic with simple poles at location of the poles of w(z)w(z). At these locations

Res(Y^(1,2)(1)(z))=Res(Pn(1)(z)w(z)hq(z)).\mathrm{Res}(\widehat{Y}^{(1)}_{(1,2)}(z))=-\mathrm{Res}(P_{n}^{(1)}(z)w(z)h_{q}(z)). (8.2)

Consider the function

Fn(z)=hq(q1/2eiπ/4)hq(eiπ/4)Y^(1,2)(1)(z)+Pn(1)(z)w(z)hq(q1/2z),F_{n}(z)=\frac{h_{q}(q^{1/2}e^{i\pi/4})}{h_{q}(e^{i\pi/4})}\widehat{Y}^{(1)}_{(1,2)}(z)+P_{n}^{(1)}(z)w(z)h_{q}(q^{1/2}z),

from Equation (8.2), Lemma A.3 and Remark A.6 we conclude that Fn(z)F_{n}(z) is meromorphic with simple poles at z=qk+1/2z=q^{k+1/2} for kk\in\mathbb{Z}. The residue of these poles is given by

Res(Fn(qk+1/2))=Pn(1)(qk+1/2)w(qk+1/2)qk+1/2.\mathrm{Res}(F_{n}(q^{k+1/2}))=P_{n}^{(1)}(q^{k+1/2})w(q^{k+1/2})q^{k+1/2}.

Note that w(z)w(z) decays much faster than inverse polynomial decay and thus

limzFn(z)=hq(q1/2eiπ/4)hq(eiπ/4)limzY^(1,2)(1)(z).\lim_{z\to\infty}F_{n}(z)=\frac{h_{q}(q^{1/2}e^{i\pi/4})}{h_{q}(e^{i\pi/4})}\lim_{z\to\infty}\widehat{Y}^{(1)}_{(1,2)}(z).

Hence, the solution for the RHP corresponding to c=q1/2c=q^{1/2} (see Remark 2.5) can also be written as

Yn(q1/2)(z)={[Pn(1)(z)q1/2Fn(z)Pn1(1)(z)γn1(q1/2)Fn1(z)q1/2γn1(q1/2)],forz𝒟+,[Pn(1)(z)q1/2(Fn(z)Pn(1)(z)w(z)hq(q1/2z))Pn1(1)(z)γn1(q1/2)Fn1(z)Pn1(1)(z)w(z)hq(q1/2(z)q1/2γn1(q1/2)],forz𝒟.\displaystyle Y_{n}^{(q^{1/2})}(z)=\left\{\begin{array}[]{lr}\begin{bmatrix}P_{n}^{(1)}(z)&q^{-1/2}F_{n}(z)\\ \dfrac{P_{n-1}^{(1)}(z)}{\gamma_{n-1}^{(q^{1/2})}}&\dfrac{F_{n-1}(z)}{q^{1/2}\gamma_{n-1}^{(q^{1/2})}}\end{bmatrix},&\text{for}\,z\in\mathcal{D}_{+},\\ &\\ \begin{bmatrix}P_{n}^{(1)}(z)&q^{-1/2}(F_{n}(z)-P_{n}^{(1)}(z)w(z)h_{q}(q^{1/2}z))\\ \dfrac{P_{n-1}^{(1)}(z)}{\gamma_{n-1}^{(q^{1/2})}}&\dfrac{F_{n-1}(z)-P_{n-1}^{(1)}(z)w(z)h_{q}(q^{1/2}(z)}{q^{1/2}\gamma_{n-1}^{(q^{1/2})}}\end{bmatrix},&\text{for}\,z\in\mathcal{D}_{-}.\end{array}\right. (8.6)

Thus, we have shown that Pn(1)(z)=Pn(q1/2)(z)P_{n}^{(1)}(z)=P_{n}^{(q^{1/2})}(z). One can readily deduce from Section 2 that

limzY^(1,2)(1)(z)=γn(1)zn+1,\lim_{z\to\infty}\widehat{Y}^{(1)}_{(1,2)}(z)=\frac{\gamma_{n}^{(1)}}{z^{n+1}},

it follows from Remark 2.5,

γn(q1/2)q1/2=hq(q1/2eiπ/4)hq(eiπ/4)γn(1).\gamma_{n}^{(q^{1/2})}q^{1/2}=\frac{h_{q}(q^{1/2}e^{i\pi/4})}{h_{q}(e^{i\pi/4})}\gamma_{n}^{(1)}.

Finally, we prove that in general Pn(c)(z)Pn(1)(z)P_{n}^{(c)}(z)\neq P_{n}^{(1)}(z), if cq1/2,1c\neq q^{1/2},1. Assume to the contrary that {Pn(c)(z)}n=0={Pn(1)(z)}n=0\{P_{n}^{(c)}(z)\}_{n=0}^{\infty}=\{P_{n}^{(1)}(z)\}_{n=0}^{\infty}. Let, Yn(c)Y_{n}^{(c)} be the solution of the corresponding RHP given in Remark 2.5. Note that the first column of Yn(1)Y_{n}^{(1)} is the same as that in Yn(c)Y_{n}^{(c)}.

By Remark 2.4, we know that the second column of Y^n(c)\widehat{Y}_{n}^{(c)} must satisfy the same qq-difference equation as the second column of Y^n(1)\widehat{Y}_{n}^{(1)}, where Y^n=Yn\widehat{Y}_{n}=Y_{n} restricted to z𝒟z\in\mathcal{D}_{-}. If {Pn(c)(z)}n=0={Pn(1)(z)}n=0\{P_{n}^{(c)}(z)\}_{n=0}^{\infty}=\{P_{n}^{(1)}(z)\}_{n=0}^{\infty} then by the analyticity of Y^n\widehat{Y}_{n} and comparing even and odd terms we conclude the second column of Y^n(c)\widehat{Y}_{n}^{(c)} must satisfy

Y^n(c)(z)=Y^n(1)(z)[100C0],forz𝒟,\displaystyle\widehat{Y}_{n}^{(c)}(z)=\widehat{Y}_{n}^{(1)}(z)\begin{bmatrix}1&0\\ 0&C_{0}\end{bmatrix},\qquad\mathrm{for}\,z\in\mathcal{D}_{-},

where c0c_{0} is a constant. Denoting the (1,2)(1,2)-entry of Y^n(c)(z)\widehat{Y}_{n}^{(c)}(z), for z𝒟z\in\mathcal{D}_{-}, by Y^(1,2)(c)(z)\widehat{Y}^{(c)}_{(1,2)}(z), we conclude that Y^(1,2)(c)(z)=c0Y^(1,2)(1)(z)\widehat{Y}^{(c)}_{(1,2)}(z)=c_{0}\widehat{Y}^{(1)}_{(1,2)}(z).

By the analyticity of Yn(c)(z)Y_{n}^{(c)}(z), we deduce that at the poles of w(z)w(z), in particular at z=eiπ/4,e3iπ/4,eiπ/4,e3iπ/4z=e^{i\pi/4},e^{3i\pi/4},e^{-i\pi/4},e^{-3i\pi/4}, we have

c0Y^(1,2)(1)(z)+w(z)Pn(c)(z)h(z/c)=0.c_{0}\widehat{Y}^{(1)}_{(1,2)}(z)+w(z)P_{n}^{(c)}(z)h(z/c)=0. (8.7)

Furthermore,

Y^(1,2)(1)(z)+w(z)Pn(1)(z)h(z)=0.\widehat{Y}^{(1)}_{(1,2)}(z)+w(z)P_{n}^{(1)}(z)h(z)=0. (8.8)

Equations (8.7) and (8.8) can only be satisfied if the ratio h(z/c)/h(z)h(z/c)/h(z) is equal at the four points z=eiπ/4,e3iπ/4,eiπ/4,e3iπ/4z=e^{i\pi/4},e^{3i\pi/4},e^{-i\pi/4},e^{-3i\pi/4}. By Lemma A.3 and Remark A.6, it follows that the equality can only hold at these four points if c=1,q1/2c=1,q^{1/2}, which is the desired result. ∎

Theorem 8.2.

Suppose that the sequence of monic polynomials {Pn(c)(x)}n=0\{P^{(c)}_{n}(x)\}_{n=0}^{\infty} satisfies the orthogonality condition

Pn(c)(cx)Pm(c)(cx)w(cx)dqx=γn(c)δn,m,\int^{\infty}_{-\infty}P^{(c)}_{n}(cx)P^{(c)}_{m}(cx)w(cx)d_{q}x=\gamma^{(c)}_{n}\delta_{n,m},

where w(x)w(x) is given by Equation (1.11) and q<c1q<c\leq 1. Then, the recurrence coefficients {αn(c)}n=1\{\alpha^{(c)}_{n}\}_{n=1}^{\infty}, which occur in the recurrence relation

xPn(c)(x)=Pn+1(c)(x)+αn(c)Pn1(c)(x),xP^{(c)}_{n}(x)=P^{(c)}_{n+1}(x)+\alpha^{(c)}_{n}P^{(c)}_{n-1}(x), (8.9)

solve the equation:

αn(αn+1+qn1αn+q2αn1q2n3αn+1αnαn1)=(qn1)q1n,\alpha_{n}(\alpha_{n+1}+q^{n-1}\alpha_{n}+q^{-2}\alpha_{n-1}-q^{2n-3}\alpha_{n+1}\alpha_{n}\alpha_{n-1})=(q^{-n}-1)q^{1-n}, (8.10)

with initial conditions αn=0\alpha_{n}=0 for n0n\leq 0. Furthermore,

Dq1Pn(c)=[n]q1Pn1(c)+qn3q11αn(c)αn1(c)αn2(c)Pn3(c).D_{q^{-1}}P^{(c)}_{n}=[n]_{q^{-1}}P^{(c)}_{n-1}+\frac{q^{n-3}}{q^{-1}-1}\alpha^{(c)}_{n}\alpha^{(c)}_{n-1}\alpha^{(c)}_{n-2}P^{(c)}_{n-3}. (8.11)
Proof.

In order to show Equation (8.11), we observe that

k=Dq1(Pn(c)(cqk))Pm(c)(cqk)w(cqk)qk=k=Pn(c)(cqk)(Pm(c)(cqk+1)Pm(c)(cqk))cqk(q11)w(cqk)qk+k=c4q4kPn(c)(cqk)Pm(c)(cqk+1)w(cqk)c(q11),\sum^{\infty}_{k=-\infty}D_{q^{-1}}(P^{(c)}_{n}(cq^{k}))P^{(c)}_{m}(cq^{k})w(cq^{k})q^{k}=\\ \sum^{\infty}_{k=-\infty}P^{(c)}_{n}(cq^{k})\frac{(P^{(c)}_{m}(cq^{k+1})-P^{(c)}_{m}(cq^{k}))}{cq^{k}(q^{-1}-1)}w(cq^{k})q^{k}\\ +\sum^{\infty}_{k=-\infty}\frac{c^{4}q^{4k}P^{(c)}_{n}(cq^{k})P^{(c)}_{m}(cq^{k+1})w(cq^{k})}{c(q^{-1}-1)}, (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

Dq1Pn(c)=An(c)Pn1(c)+Bn(c)Pn3(c),D_{q^{-1}}P^{(c)}_{n}=A^{(c)}_{n}P^{(c)}_{n-1}+B^{(c)}_{n}P^{(c)}_{n-3}, (8.13)

where An(c)A^{(c)}_{n}, Bn(c)B^{(c)}_{n} are constants depending on nn. We can immediately find An(c)A^{(c)}_{n} explicitly by using the identity

Dq1(xn)=[n]q1xn1.D_{q^{-1}}(x^{n})=[n]_{q^{-1}}x^{n-1}.

Since the sequence {Pn(c)(x)}n=0\{P^{(c)}_{n}(x)\}_{n=0}^{\infty} consists of monic polynomials, we obtain

An(c)=[n]q1.A^{(c)}_{n}=[n]_{q^{-1}}.

In order to derive Bn(c)B^{(c)}_{n}, we first note that:

Dq1(xPn(c))=xq1Dq1Pn(c)(x)+Pn(c)(x).D_{q^{-1}}(xP^{(c)}_{n})=xq^{-1}D_{q^{-1}}P^{(c)}_{n}(x)+P^{(c)}_{n}(x). (8.14)

We now take the q1q^{-1}-derivative of both sides of Equation (8.9). After gathering linearly independent terms in Equation (8.14) we find

(q1[n]q1+1[n+1]q1)Pn(c)(x)+(q1[n]q1αn1(c)+q1Bn(c)Bn+1(c)αn(c)[n1]q1)Pn2(c)(x)+(q1Bn(c)αn3(c)αn(c)Bn1(c))Pn4(c)(x)=0,(q^{-1}[n]_{q^{-1}}+1-[n+1]_{q^{-1}})P^{(c)}_{n}(x)\\ +(q^{-1}[n]_{q^{-1}}\alpha^{(c)}_{n-1}+q^{-1}B^{(c)}_{n}-B^{(c)}_{n+1}-\alpha^{(c)}_{n}[n-1]_{q^{-1}})P^{(c)}_{n-2}(x)\\ +(q^{-1}B^{(c)}_{n}\alpha^{(c)}_{n-3}-\alpha^{(c)}_{n}B^{(c)}_{n-1})P^{(c)}_{n-4}(x)=0, (8.15)

which leads to two equations for BiB_{i} and αj(c)\alpha^{(c)}_{j}:

q1[n]q1αn1(c)+q1Bn(c)\displaystyle q^{-1}[n]_{q^{-1}}\alpha^{(c)}_{n-1}+q^{-1}B^{(c)}_{n} =\displaystyle= Bn+1(c)+αn(c)[n1]q1,\displaystyle B^{(c)}_{n+1}+\alpha^{(c)}_{n}[n-1]_{q^{-1}}, (8.16)
q1Bn(c)αn3(c)\displaystyle q^{-1}B^{(c)}_{n}\alpha^{(c)}_{n-3} =\displaystyle= αn(c)Bn1(c).\displaystyle\alpha^{(c)}_{n}B^{(c)}_{n-1}. (8.17)

Equation (8.17) implies Bn(c)=c^qnαn(c)αn1(c)αn2(c)B^{(c)}_{n}=\hat{c}q^{n}\alpha^{(c)}_{n}\alpha^{(c)}_{n-1}\alpha^{(c)}_{n-2}, for some constant c^\hat{c}.

We now proceed to show that αn(c)\alpha^{(c)}_{n} satisfies Equation (8.10). Substituting Bn(c)B_{n}^{(c)} into Equation (8.16) we find

qn2[n]q1αn(c)qn1[n1]q1αn1(c)=c^(αn+1(c)q2αn2(c)).\frac{q^{-n-2}[n]_{q^{-1}}}{\alpha^{(c)}_{n}}-\frac{q^{-n-1}[n-1]_{q^{-1}}}{\alpha^{(c)}_{n-1}}=\hat{c}\bigl{(}\alpha^{(c)}_{n+1}-q^{-2}\alpha^{(c)}_{n-2}\bigr{)}. (8.18)

We will rearrange Equation (8.18) with the goal of obtaining a telescoping sum. Multiplying Equation (8.18) by 1+dαn(c)αn1(c)q2n31+d\alpha^{(c)}_{n}\alpha^{(c)}_{n-1}q^{2n-3}, for some constant dd, we have

qn2[n]q1αn(c)qn1[n1]q1αn1(c)+dαn1(c)qn5[n]q1dαn(c)qn4[n1]q1=c^(αn+1(c)q2αn2(c)+dq2n3αn+1(c)αn(c)αn1(c)dq2n5αn(c)αn1(c)αn2(c)).\frac{q^{-n-2}[n]_{q^{-1}}}{\alpha^{(c)}_{n}}-\frac{q^{-n-1}[n-1]_{q^{-1}}}{\alpha^{(c)}_{n-1}}+d\alpha^{(c)}_{n-1}q^{n-5}[n]_{q^{-1}}-d\alpha^{(c)}_{n}q^{n-4}[n-1]_{q^{-1}}\\ =\hat{c}(\alpha^{(c)}_{n+1}-q^{-2}\alpha^{(c)}_{n-2}+dq^{2n-3}\alpha^{(c)}_{n+1}\alpha^{(c)}_{n}\alpha^{(c)}_{n-1}-dq^{2n-5}\alpha^{(c)}_{n}\alpha^{(c)}_{n-1}\alpha^{(c)}_{n-2}). (8.19)

Therefore, we find

qn2[n]q1αn(c)qn1[n1]q1αn1(c)\displaystyle\frac{q^{-n-2}[n]_{q^{-1}}}{\alpha^{(c)}_{n}}-\frac{q^{-n-1}[n-1]_{q^{-1}}}{\alpha^{(c)}_{n-1}} =\displaystyle= c^αn+1(c)+dqn41q1αn(c)+dq51q1αn1(c)\displaystyle\hat{c}\alpha^{(c)}_{n+1}+d\frac{q^{n-4}}{1-q^{-1}}\alpha^{(c)}_{n}+d\frac{q^{-5}}{1-q^{-1}}\alpha^{(c)}_{n-1}
+c^dq2n3αn+1(c)αn(c)αn1(c)\displaystyle+\hat{c}dq^{2n-3}\alpha^{(c)}_{n+1}\alpha^{(c)}_{n}\alpha^{(c)}_{n-1}
dq31q1αn(c)dqn51q1αn1(c)c^q2αn2(c)\displaystyle-d\frac{q^{-3}}{1-q^{-1}}\alpha^{(c)}_{n}-d\frac{q^{n-5}}{1-q^{-1}}\alpha^{(c)}_{n-1}-\hat{c}q^{-2}\alpha^{(c)}_{n-2}
c^dq2n5αn(c)αn1(c)αn2(c).\displaystyle-\hat{c}dq^{2n-5}\alpha^{(c)}_{n}\alpha^{(c)}_{n-1}\alpha^{(c)}_{n-2}.

Letting c^=q3c~1q1\hat{c}=\frac{q^{-3}\tilde{c}}{1-q^{-1}}, we are led to

qn2[n]q1αn(c)qn1[n1]q1αn1(c)=11q1(c~q3αn+1(c)+dqn4αn(c)+dq5αn1(c)+c~dq2n6αn+1(c)αn(c)αn1(c))11q1(dq3αn(c)+dqn5αn1(c)+c~q5αn2(c)+c~dq2n8αn(c)αn1(c)αn2(c)).\frac{q^{-n-2}[n]_{q^{-1}}}{\alpha^{(c)}_{n}}-\frac{q^{-n-1}[n-1]_{q^{-1}}}{\alpha^{(c)}_{n-1}}\\ =\frac{1}{1-q^{-1}}(\tilde{c}q^{-3}\alpha^{(c)}_{n+1}+dq^{n-4}\alpha^{(c)}_{n}+dq^{-5}\alpha^{(c)}_{n-1}+\tilde{c}dq^{2n-6}\alpha^{(c)}_{n+1}\alpha^{(c)}_{n}\alpha^{(c)}_{n-1})\\ -\frac{1}{1-q^{-1}}(dq^{-3}\alpha^{(c)}_{n}+dq^{n-5}\alpha^{(c)}_{n-1}+\tilde{c}q^{-5}\alpha^{(c)}_{n-2}+\tilde{c}dq^{2n-8}\alpha^{(c)}_{n}\alpha^{(c)}_{n-1}\alpha^{(c)}_{n-2}).

Thus, if we choose d=c~d=\tilde{c} and take the sum from 22 to nn we find

q1n(1qn)αn(c)1q1α1(c)=c~(αn+1(c)+qn1αn(c)+q2αn1(c)+c~q2n3αn+1(c)αn(c)αn1(c)(α2(c)+α1(c))).\frac{q^{1-n}(1-q^{-n})}{\alpha^{(c)}_{n}}-\frac{1-q^{-1}}{\alpha^{(c)}_{1}}\\ =\tilde{c}(\alpha^{(c)}_{n+1}+q^{n-1}\alpha^{(c)}_{n}+q^{-2}\alpha^{(c)}_{n-1}+\tilde{c}q^{2n-3}\alpha^{(c)}_{n+1}\alpha^{(c)}_{n}\alpha^{(c)}_{n-1}-(\alpha^{(c)}_{2}+\alpha^{(c)}_{1})).

It remains to determine c~\tilde{c} and (1q1)/α1(c)c~(α2(c)+α1(c))(1-q^{-1})/\alpha^{(c)}_{1}-\tilde{c}(\alpha^{(c)}_{2}+\alpha^{(c)}_{1}). Observe that

k=Dq1((cqk)n)w(cqk)qk\displaystyle\sum_{k=-\infty}^{\infty}D_{q^{-1}}((cq^{k})^{n})w(cq^{k})q^{k} =\displaystyle= k=(cqk)n+3w(cqk)qkq11,\displaystyle\sum_{k=-\infty}^{\infty}\frac{(cq^{k})^{n+3}w(cq^{k})q^{k}}{q^{-1}-1}, (8.20)

where we have used Equations (1.1) and (3.6).

We now use Equation (8.20) to determine c~\tilde{c} and (1q1)/α1(c)c~(α2(c)+α1(c))(1-q^{-1})/\alpha^{(c)}_{1}-\tilde{c}(\alpha^{(c)}_{2}+\alpha^{(c)}_{1}). Without loss of generality assume that w(cx)dqx=1\int_{-\infty}^{\infty}w(cx)d_{q}x=1. From Equation (8.20) we find that

(cx)4w(cx)dqx\displaystyle\int_{-\infty}^{\infty}(cx)^{4}w(cx)d_{q}x =\displaystyle= (q11).\displaystyle(q^{-1}-1). (8.21)
(cx)6w(cx)dqx\displaystyle\int_{-\infty}^{\infty}(cx)^{6}w(cx)d_{q}x =\displaystyle= (q31)(cx)2w(cx)dqx.\displaystyle(q^{-3}-1)\int_{-\infty}^{\infty}(cx)^{2}w(cx)d_{q}x. (8.22)

Using Equation (8.9) we determine

α1(c)\displaystyle\alpha^{(c)}_{1} =\displaystyle= (cx)2w(cx)dqx.\displaystyle\int_{-\infty}^{\infty}(cx)^{2}w(cx)d_{q}x. (8.23)
α3(c)\displaystyle\alpha^{(c)}_{3} =\displaystyle= P3(c)(cx)2w(cx)dqxP2(c)(cx)2w(cx)dqx,\displaystyle\frac{\int_{-\infty}^{\infty}P^{(c)}_{3}(cx)^{2}w(cx)d_{q}x}{\int_{-\infty}^{\infty}P^{(c)}_{2}(cx)^{2}w(cx)d_{q}x},
=\displaystyle= ((cx)62(α1(c)+α2(c))(cx)4+(α1(c)+α2(c))2(cx)2)w(cx)dqx((cx)42α1(c)(cx)2+(α1(c))2)w(cx)dqx.\displaystyle\frac{\int_{-\infty}^{\infty}\bigl{(}(cx)^{6}-2(\alpha^{(c)}_{1}+\alpha^{(c)}_{2})(cx)^{4}+(\alpha^{(c)}_{1}+\alpha^{(c)}_{2})^{2}(cx)^{2})w(cx)d_{q}x}{\int_{-\infty}^{\infty}((cx)^{4}-2\alpha^{(c)}_{1}(cx)^{2}+(\alpha^{(c)}_{1})^{2}\bigr{)}w(cx)d_{q}x}.

Furthermore, by the orthogonality of {Pn(c)(x)}n=0\{P^{(c)}_{n}(x)\}_{n=0}^{\infty} we find

P1(c)(cx)P3(c)(cx)w(cx)dqx\displaystyle\int_{-\infty}^{\infty}P^{(c)}_{1}(cx)P^{(c)}_{3}(cx)w(cx)d_{q}x =\displaystyle= ((cx)4(α1(c)+α2(c))(cx)2)w(cx)dqx,\displaystyle\int_{-\infty}^{\infty}\bigl{(}(cx)^{4}-(\alpha^{(c)}_{1}+\alpha^{(c)}_{2})(cx)^{2}\bigr{)}w(cx)d_{q}x, (8.25)
=\displaystyle= 0,\displaystyle 0,

where we have used the recurrence relation, Equation (8.9), to determine P3(c)(x)P^{(c)}_{3}(x). Applying Equations (8.21) to (8.25) we deduce

c~\displaystyle\tilde{c} =\displaystyle= 1.\displaystyle-1.
q11\displaystyle q^{-1}-1 =\displaystyle= α1(c)(α1(c)+α2(c)).\displaystyle\alpha^{(c)}_{1}(\alpha^{(c)}_{1}+\alpha^{(c)}_{2}).

These values give Equations (8.10) and (8.11) and Theorem 8.2 follows immediately. ∎

As a consequence of Theorem 8.2 the sequences {αn(c)}n=1\{\alpha^{(c)}_{n}\}_{n=1}^{\infty} all provide positive solutions of Equation (8.10). However, we will show that the limits of these sequences, as nn\to\infty, are in general not the same. In particular, we show that the asymptotic limit as nn\to\infty of αn(1)\alpha^{(1)}_{n} does not equal the limit of αn(c)\alpha^{(c)}_{n} if c1,q1/2c\neq 1,q^{1/2}.

Theorem 8.3.

Suppose the sequence of orthogonal polynomials {Pn(c)(z)}n=0\{P_{n}^{(c)}(z)\}_{n=0}^{\infty} and recurrence coefficients {αn(c)}n=1\{\alpha^{(c)}_{n}\}_{n=1}^{\infty} are defined as in Theorem 8.2. Then, it follows that

limnαn(c)αn(1)q1nαn(1)0,\lim_{n\to\infty}\frac{\alpha^{(c)}_{n}-\alpha^{(1)}_{n}}{q^{1-n}-\alpha^{(1)}_{n}}\neq 0, (8.26)

if c1,q1/2c\neq 1,q^{1/2}.

Proof.

From Lemma 8.1, we have {αn(1)}n=0={αn(q1/2)}n=0\{\alpha^{(1)}_{n}\}_{n=0}^{\infty}=\{\alpha^{(q^{1/2})}_{n}\}_{n=0}^{\infty}. It remains to study the case c1,q1/2c\neq 1,q^{1/2}. Let y1(z)y_{1}(z) and y2(z)y_{2}(z) be two solutions of Equation (4.2). Then, Y(z)=[y1(z),y2(z)]TY(z)=[y_{1}(z),y_{2}(z)]^{T} is a solution of the matrix equation

Y(q1z)=[1q2z2q2zq2z1q2z2]Y(z).\displaystyle Y(q^{-1}z)=\begin{bmatrix}1-q^{-2}z^{2}&q^{-2}z\\ -q^{-2}z&1-q^{-2}z^{2}\end{bmatrix}Y(z). (8.27)

Using the definition of the qq-derivative, and Equation (8.9) to write Pn3(x)P_{n-3}(x) in terms of Pn(x)P_{n}(x) and Pn1(x)P_{n-1}(x), Equation (8.11) can be re-written as

Pn(c)(q1z)=(1qn3z2αn(c))Pn(c)(z)+((qn1)zzαn(c)αn1(c)qn3+z3αn(c)qn3)Pn1(c)(z).P^{(c)}_{n}(q^{-1}z)=(1-q^{n-3}z^{2}\alpha^{(c)}_{n})P_{n}^{(c)}(z)\\ +((q^{-n}-1)z-z\alpha^{(c)}_{n}\alpha^{(c)}_{n-1}q^{n-3}+z^{3}\alpha^{(c)}_{n}q^{n-3})P^{(c)}_{n-1}(z). (8.28)

Taking nn1n\mapsto n-1 and again using Equation (8.9), Equation (8.28) also allows us to express Pn1(c)(q1z)P^{(c)}_{n-1}(q^{-1}z) in terms of Pn(c)(z)P^{(c)}_{n}(z) and Pn1(c)(z)P^{(c)}_{n-1}(z). This results in the matrix difference equation

[Pn(c)(q1z)qn/2Pn1(c)(q1z)]=\displaystyle\begin{bmatrix}P^{(c)}_{n}(q^{-1}z)\\ q^{-n/2}P^{(c)}_{n-1}(q^{-1}z)\end{bmatrix}= ([1qn3αn(c)z2lnzknz1qn4αn1(c)z2]+r(z))\displaystyle\left(\begin{bmatrix}1-q^{n-3}\alpha^{(c)}_{n}z^{2}&l_{n}z\\ k_{n}z&1-q^{n-4}\alpha^{(c)}_{n-1}z^{2}\end{bmatrix}+r(z)\right)\cdot
[Pn(c)(z)qn/2Pn1(c)(z)],\displaystyle\qquad\qquad\cdot\begin{bmatrix}P^{(c)}_{n}(z)\\ q^{-n/2}P^{(c)}_{n-1}(z)\end{bmatrix},

where

ln\displaystyle l_{n} =\displaystyle= qn/2αn(c)αn1(c)q3n23,\displaystyle q^{-n/2}-\alpha^{(c)}_{n}\alpha^{(c)}_{n-1}q^{\frac{3n}{2}-3},
kn\displaystyle k_{n} =\displaystyle= q13n2/αn1(c)αn2(c)qn24,\displaystyle q^{1-\frac{3n}{2}}/\alpha^{(c)}_{n-1}-\alpha^{(c)}_{n-2}q^{\frac{n}{2}-4},

and r(z)r(z) is given by

r(z)=z[0z2anq3n23qn/2qn24z2qn/2(αn1(c))1zq1n/αn1(c)zαn2(c)qn4+z3qn4].\displaystyle r(z)=z\begin{bmatrix}0&z^{2}a_{n}q^{\frac{3n}{2}-3}-q^{n/2}\\ q^{\frac{n}{2}-4}z^{2}-q^{-n/2}(\alpha^{(c)}_{n-1})^{-1}&zq^{1-n}/\alpha^{(c)}_{n-1}-z\alpha^{(c)}_{n-2}q^{n-4}+z^{3}q^{n-4}\end{bmatrix}. (8.29)

For the case c=1c=1 we know that as nn\to\infty Equation (8) approaches Equation (8.27) (this follows from the the proof that Wn(z)G(z)W_{n}(z)\sim G(z) as nn\to\infty , see Section 7). Assume that αn(c)\alpha^{(c)}_{n} has the same asymptotic behaviour as αn(1)\alpha^{(1)}_{n}. Then Equation (8) must similarly approach Equation (8.27). In particular, r(z)=o(1)r(z)=o(1) as nn\to\infty and, lnl_{n}, knk_{n} approach non-zero constants. Note that the condition r(z)=o(1)r(z)=o(1) as nn\to\infty follows from the first order asymptotic behaviour i.e. αn(c)q1n\alpha^{(c)}_{n}\sim q^{1-n} as nn\to\infty. In order for lnl_{n} and knk_{n} to approach the same non-zero constant as the case c=1c=1 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 z𝒟+z\in\mathcal{D}_{+}, Y2n(c)(z)Y^{(c)}_{2n}(z) approaches the matrix

Y2n(c)(z)[a(z)w(z)(hq(z/c)a(z)λ1b(z))b(z)w(z)(hq(z/c)b(z)λ2a(z))],asn,\displaystyle Y^{(c)}_{2n}(z)\sim\begin{bmatrix}a(z)&w(z)(h_{q}(z/c)a(z)-\lambda_{1}b(z))\\ b(z)&w(z)(h_{q}(z/c)b(z)-\lambda_{2}a(z))\end{bmatrix},\;\text{as}\,n\to\infty, (8.30)

for some real constants λ1,2\lambda_{1,2} (recall a(z)a(z) and b(z)b(z) are defined in Lemma 4.2). By the meromorphicity of Y2n(c)(z)Y^{(c)}_{2n}(z) at the poles of w(z)w(z) we find

hq(z/c)a(z)λ1b(z)=hq(z/c)b(z)λ2a(z)=0.h_{q}(z/c)a(z)-\lambda_{1}b(z)=h_{q}(z/c)b(z)-\lambda_{2}a(z)=0. (8.31)

Thus,

hq(z/c)2=λ1λ2.h_{q}(z/c)^{2}=\lambda_{1}\lambda_{2}. (8.32)

Hence, hq(z/c)h_{q}(z/c) is either real or imaginary. We conclude from Lemma A.5 and Remark A.6 that Equation (8.32) is satisfied iff c=1,q1/2c=1,q^{1/2}. It follows that αn(c)\alpha^{(c)}_{n} has the same asymptotic behaviour as αn(1)\alpha^{(1)}_{n} (i.e. Equation (8.26) is zero) iff c=1,q1/2c=1,q^{1/2}. ∎

9. Conclusion

In this paper, we described the asymptotic behaviour of a class of qFIIqF_{II} polynomials, defined in Section 1.3, by using the qq-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 qFIIqF_{II} polynomials. In Theorem 1.7, we detailed the implications of our analysis for the qq-Painlevé equation satisfied by the recurrence coefficients, {αn}n=1\{\alpha_{n}\}_{n=1}^{\infty}, of qFIIqF_{II} polynomials (see Equation (1.5)).

Perhaps the most unexpected results of this paper concern the effect of the properties of hq(z)h_{q}(z) on the class of orthogonal polynomials that arise when the lattice was varied. The values of hq(z)h_{q}(z) at the poles of the weight function w(z)w(z) play an important role in determining the behaviour of orthogonal polynomials supported on the shifted lattice cqkcq^{k}, kk\in\mathbb{Z}. This observation enabled us to determine whether the class of orthogonal polynomials were invariant as cc 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 (x4;q4)1(-x^{4};q^{4})_{\infty}^{-1}. But, the methodology can readily be extended to describe discrete qq-Hermite II polynomials [3, Chapter 18.27] with weight (x2;q2)1(-x^{2};q^{2})_{\infty}^{-1}. However, generalising the results to higher order weights i.e. (x2m;q2m)1(-x^{2m};q^{2m})_{\infty}^{-1}, for m>2m>2, 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 nn\to\infty. In this paper we described the asymptotic behaviour of one particular solution, which satisfies the limit

limnqnαn(1)=q.\lim_{n\to\infty}q^{n}\alpha^{(1)}_{n}=q. (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 qq-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 qFII(c)qF^{(c)}_{II} polynomials, defined in Section 1.3.

Appendix A Properties of hq(z)h_{q}(z)

In this section we prove some properties of the function hq(z)h_{q}(z) defined in Equation (1.2), which are used in this paper. Before discussing hq(z)h_{q}(z) we first prove a necessary Lemma.

Lemma A.1.

Let C(z)C(z) be a function defined on {0}\mathbb{C}\setminus\{0\}, which is analytic everywhere except for simple poles at qkq^{k} for kk\in\mathbb{Z}. Then, C(qz)C(z)C(qz)\neq C(z).

Proof.

We prove the result by contradiction. Assume C(qz)=C(z)C(qz)=C(z). Define

G(z)=(z,qz1;q).G(z)=(-z,-qz^{-1};q)_{\infty}.

By direct calculation one can show G(qz)=z1G(z)G(qz)=z^{-1}G(z). Furthermore, by definition, G(z)G(z) is zero on the qq-lattice qkq^{k}, kk\in\mathbb{Z}. Let

F(z)=C(z)G(z),F(z)=C(z)G(z),

then it follows F(z)F(z) is analytic in {0}\mathbb{C}\setminus\{0\} and satisfies the difference equation

F(qz)=z1F(z).F(qz)=z^{-1}F(z). (A.1)

As F(z)F(z) is analytic in {0}\mathbb{C}\setminus\{0\} we can write F(z)F(z) as the Laurent series

F(z)=k=Fkzk.F(z)=\sum_{k=-\infty}^{\infty}F_{k}z^{k}.

Comparing the coefficients of zz in Equation (A.1), one can readily determine

Fk=c0qk(k1)/2.F_{k}=c_{0}q^{k(k-1)/2}. (A.2)

However, there is only one solution with zz coefficients given by Equation (A.2) (up to scaling by a constant) and it follows that F(z)=c0G(z)F(z)=c_{0}G(z). Thus, if C(qz)=C(z)C(qz)=C(z), then C(z)=c0C(z)=c_{0}, and C(z)C(z) has no poles. ∎

Corollary A.2.

Let C(z)C(z) be a function defined on {0}\mathbb{C}\setminus\{0\}, which is analytic everywhere except for simple poles at ±qk\pm q^{k} for kk\in\mathbb{Z}. Furthermore, suppose C(z)C(z) satisfies C(qz)=C(z)C(qz)=C(z). Then, C(z)=c1hq(z)+c0C(z)=c_{1}h_{q}(z)+c_{0}, where c0c_{0} and c1c_{1} are constants and hq(z)h_{q}(z) is as defined in Definition 1.4.

Proof.

As both C(z)C(z) and hq(z)h_{q}(z) have simple poles at z=1z=-1 we conclude that there exists a c10c_{1}\neq 0 such that

Res(C(1))=c1Res(hq(1)).\mathrm{Res}(C(-1))=c_{1}\mathrm{Res}(h_{q}(-1)).

Furthermore, both C(z)C(z) and hq(z)h_{q}(z) are invariant under the transformation zqzz\to qz, hence for all kk\in\mathbb{Z}

Res(C(qk))=c1Res(hq(qk)).\mathrm{Res}(C(-q^{k}))=c_{1}\mathrm{Res}(h_{q}(-q^{k})).

Thus, the function

D(z)=C(z)c1hq(z),D(z)=C(z)-c_{1}h_{q}(z),

is meromorphic in {0}\mathbb{C}\setminus\{0\}, with possible simple poles at qkq^{k} for kk\in\mathbb{Z}, and satisfies D(qz)=D(z)D(qz)=D(z). However, by Lemma A.1, D(z)D(z) can not have simple poles at qkq^{k} for kk\in\mathbb{Z}. Hence, D(z)D(z) is analytic in {0}\mathbb{C}\setminus\{0\} and it follows that D(z)D(z) can be written as a convergent Laurent series. Thus,

D(z)=j=djzj.D(z)=\sum_{j=-\infty}^{\infty}d_{j}z^{j}.

Substituting this into the qq-difference equation D(qz)=D(z)D(qz)=D(z), we conclude D(z)=d0D(z)=d_{0} (=c0)(=c_{0}) and Corollary A.2 follows immediately. ∎

Lemma A.3.

The function hq(z)h_{q}(z) has zero real part along the circles |z|=1|z|=1 and |z|=q1/2|z|=q^{1/2}.

Proof.

Let z=reiθz=re^{i\theta}. Substituting this into Equation (1.2) and determining the real part we find

Re(hq(reiθ))=k=2rqkcos(θ)(r2q2k)r4+q4k2r2q2kcos(2θ).\mathrm{Re}(h_{q}(re^{i\theta}))=\sum_{k=-\infty}^{\infty}\frac{2rq^{k}cos(\theta)(r^{2}-q^{2k})}{r^{4}+q^{4k}-2r^{2}q^{2k}cos(2\theta)}. (A.3)

First we consider the case r=1r=1. The RHS of Equation (A.3) becomes

k=2qkcos(θ)(1q2k)1+q4k2q2kcos(2θ)\displaystyle\sum_{k=-\infty}^{\infty}\frac{2q^{k}cos(\theta)(1-q^{2k})}{1+q^{4k}-2q^{2k}cos(2\theta)} =\displaystyle= k=12qkcos(θ)(1q2k)1+q4k2q2kcos(2θ)\displaystyle\sum_{k=1}^{\infty}\frac{2q^{k}cos(\theta)(1-q^{2k})}{1+q^{4k}-2q^{2k}cos(2\theta)}
+k=12qkcos(θ)(1q2k)1+q4k2q2kcos(2θ),\displaystyle\,+\sum_{k=-\infty}^{-1}\frac{2q^{k}cos(\theta)(1-q^{2k})}{1+q^{4k}-2q^{2k}cos(2\theta)},
=\displaystyle= k=12qkcos(θ)(1q2k)1+q4k2q2kcos(2θ)\displaystyle\sum_{k=1}^{\infty}\frac{2q^{k}cos(\theta)(1-q^{2k})}{1+q^{4k}-2q^{2k}cos(2\theta)}
+k=12qkcos(θ)(1q2k)1+q4k2q2kcos(2θ),\displaystyle\,+\sum_{k=1}^{\infty}\frac{2q^{-k}cos(\theta)(1-q^{-2k})}{1+q^{-4k}-2q^{-2k}cos(2\theta)},
=\displaystyle= 0.\displaystyle 0.

Next we consider the case r=q1/2r=q^{1/2}. The RHS of Equation (A.3) can be written as

k=2qk+1/2cos(θ)(qq2k)q2+q4k2q2k+1cos(2θ)\displaystyle\sum_{k=-\infty}^{\infty}\frac{2q^{k+1/2}cos(\theta)(q-q^{2k})}{q^{2}+q^{4k}-2q^{2k+1}cos(2\theta)} =\displaystyle= k=12qk+1/2cos(θ)(qq2k)q2+q4k2q2k+1cos(2θ)\displaystyle\sum_{k=1}^{\infty}\frac{2q^{k+1/2}cos(\theta)(q-q^{2k})}{q^{2}+q^{4k}-2q^{2k+1}cos(2\theta)}
+k=02qk+1/2cos(θ)(qq2k)q2+q4k2q2k+1cos(2θ),\displaystyle\,+\sum_{k=-\infty}^{0}\frac{2q^{k+1/2}cos(\theta)(q-q^{2k})}{q^{2}+q^{4k}-2q^{2k+1}cos(2\theta)},
=\displaystyle= k=12qk+1/2cos(θ)(qq2k)q2+q4k2q2k+1cos(2θ)\displaystyle\sum_{k=1}^{\infty}\frac{2q^{k+1/2}cos(\theta)(q-q^{2k})}{q^{2}+q^{4k}-2q^{2k+1}cos(2\theta)}
+k=02qk+1/2cos(θ)(qq2k)q2+q4k2q2k+1cos(2θ),\displaystyle\,+\sum_{k=0}^{\infty}\frac{2q^{-k+1/2}cos(\theta)(q-q^{-2k})}{q^{2}+q^{-4k}-2q^{-2k+1}cos(2\theta)},
=\displaystyle= k=12qk+1/2cos(θ)(qq2k)q2+q4k2q2k+1cos(2θ)\displaystyle\sum_{k=1}^{\infty}\frac{2q^{k+1/2}cos(\theta)(q-q^{2k})}{q^{2}+q^{4k}-2q^{2k+1}cos(2\theta)}
+j=12qj+1/2cos(θ)(q2jq)q2+q4j2q2j+1cos(2θ),\displaystyle\,+\sum_{j=1}^{\infty}\frac{2q^{j+1/2}cos(\theta)(q^{2j}-q)}{q^{2}+q^{4j}-2q^{2j+1}cos(2\theta)},
=\displaystyle= 0.\displaystyle 0.

Corollary A.4.

The function hq(z)h_{q}(z) defined in Equation (1.2) can also be written as

hq(z)=c1z(qz2,qz2;q2)(z2,q2z2;q2),h_{q}(z)=c_{1}\frac{z(qz^{2},qz^{-2};q^{2})_{\infty}}{(z^{2},q^{2}z^{-2};q^{2})_{\infty}}, (A.4)

for some constant c1c_{1}.

Proof.

We observe that the function

h(z)=z(qz2,qz2;q2)(z2,q2z2;q2),h(z)=\frac{z(qz^{2},qz^{-2};q^{2})_{\infty}}{(z^{2},q^{2}z^{-2};q^{2})_{\infty}},

is meromorphic with simple poles at ±qk\pm q^{k} for kk\in\mathbb{Z}, and satisfies h(qz)=h(z)h(qz)=h(z). By Corollary A.2 we conclude h(z)=c1hq(z)+c0h(z)=c_{1}h_{q}(z)+c_{0}. By definition, h(z)h(z) has zeros at z=q1/2+kz=q^{1/2+k}, for kk\in\mathbb{Z}, and by Lemma A.3, hq(z)h_{q}(z) also has zeros at z=q1/2+kz=q^{1/2+k}, for kk\in\mathbb{Z}. Thus, we conclude

hq(z)=c1z(qz2,qz2;q2)(z2,q2z2;q2),h_{q}(z)=c_{1}\frac{z(qz^{2},qz^{-2};q^{2})_{\infty}}{(z^{2},q^{2}z^{-2};q^{2})_{\infty}},

for some constant c1c_{1}. ∎

Lemma A.5.

Along the ray z=reiπ/4z=re^{i\pi/4}, (r0)(r\in\mathbb{R}_{\geq 0}) the real part of hq(z)h_{q}(z) is non-zero except at r=qk/2r=q^{k/2}, for kk\in\mathbb{Z} where hq(z)h_{q}(z) is complete imaginary.

Proof.

From Equation (A.3) we determine that the real part of hq(reiπ/4)h_{q}(re^{i\pi/4}) is given by

Re(hq(reiπ/4))=k=2rqk(r2q2k)r4+q4k.\mathrm{Re}(h_{q}(re^{i\pi/4}))=\sum_{k=-\infty}^{\infty}\frac{\sqrt{2}rq^{k}(r^{2}-q^{2k})}{r^{4}+q^{4k}}.

Define the function

F(u)=k=uqk(u2q2k)u4+q4k,F(u)=\sum_{k=-\infty}^{\infty}\frac{uq^{k}(u^{2}-q^{2k})}{u^{4}+q^{4k}},

where uu is a complex variable. We note that the sum is well defined and converges for all uqke(iπ+2nπ)/4u\neq q^{k}e^{(i\pi+2n\pi)/4}, where kk\in\mathbb{Z} and n=0,1,2,3n=0,1,2,3. Furthermore, F(u)F(u) satisfies the qq-difference equation F(qu)=F(u)F(qu)=F(u). We now multiply F(u)F(u) by a function with zeros at the poles of F(u)F(u) to give us a function analytic in {0}\mathbb{C}\setminus\{0\}. Define

g(u)=F(u)(u4,q4u4;q4).g(u)=F(u)(-u^{4},q^{4}u^{-4};q^{4})_{\infty}.

Note that (u4,q4u4;q4)(-u^{4},q^{4}u^{-4};q^{4})_{\infty} is an even function and satisfies ((qu)4,q4(qu)4;q4)=u4(u4,q4u4;q4)(-(qu)^{4},q^{4}(qu)^{-4};q^{4})_{\infty}=u^{-4}(-u^{4},q^{4}u^{-4};q^{4})_{\infty}. Thus, g(u)g(u) is analytic in {0}\mathbb{C}\setminus\{0\}, g(u)g(u) satisfies the qq-difference equation g(qu)=u4g(u)g(qu)=u^{-4}g(u), and g(u)g(u) is an odd function. Let us represent g(u)g(u) with the convergent Laurent series

g(u)=j=gjuj.g(u)=\sum_{j=-\infty}^{\infty}g_{j}u^{j}.

As g(u)g(u) satisfies the difference equation g(qu)=u4g(u)g(qu)=u^{-4}g(u) we find

gj+4=qjgj.g_{j+4}=q^{j}g_{j}.

Hence, we conclude that there are four linearly independent solutions, which can be chosen such that two are odd and two are even. As g(u)g(u) is odd we conclude it is the sum of two linearly independent odd solutions. Consider the two functions

G1(u)=u(u2,q2u2;q2)(qu2,qu2;q2),G_{1}(u)=u(u^{2},q^{2}u^{-2};q^{2})_{\infty}(qu^{2},qu^{-2};q^{2})_{\infty},
G2(u)=u(u2,q2u2;q2)(qu2,qu2;q2).G_{2}(u)=u(-u^{2},-q^{2}u^{-2};q^{2})_{\infty}(-qu^{2},-qu^{-2};q^{2})_{\infty}.

We note that G1(u)G_{1}(u) has zeros at u=±qk/2u=\pm q^{k/2}, for kk\in\mathbb{Z} and G2(u)G_{2}(u) has zeros at u=±iqk/2u=\pm iq^{k/2}, for kk\in\mathbb{Z}. Furthermore, both G1(u)G_{1}(u) and G2(u)G_{2}(u) are odd and satisfy the difference equation G(qu)=u4G(u)G(qu)=u^{-4}G(u). Thus,

g(u)=c1G1(u)+c2G2(u),g(u)=c_{1}G_{1}(u)+c_{2}G_{2}(u),

for some constants c1c_{1} and c2c_{2}. By Lemma A.3 we conclude that c2=0c_{2}=0 and g(u)g(u) only has zeros at the zeros of G1(u)G_{1}(u) which occur at u=±qk/2u=\pm q^{k/2}, for kk\in\mathbb{Z}. Hence, this is where the zeros of F(u)F(u) are and Lemma A.5 follows immediately. ∎

Remark A.6.

From Equation (A.3) we deduce that

Re(hq(reiπ/4))=Re(hq(re3iπ/4))=Re(hq(re3iπ/4))=Re(hq(reiπ/4)).\mathrm{Re}(h_{q}(re^{i\pi/4}))=-\mathrm{Re}(h_{q}(re^{3i\pi/4}))=-\mathrm{Re}(h_{q}(re^{-3i\pi/4}))=\mathrm{Re}(h_{q}(re^{-i\pi/4})).

Using an analogous expression to Equation (A.3), for the imaginary part of hq(z)h_{q}(z), one can also show by direct calculation that

Im(hq(reiπ/4))=Im(hq(re3iπ/4))=Im(hq(re3iπ/4))=Im(hq(reiπ/4))0.\mathrm{Im}(h_{q}(re^{i\pi/4}))=\mathrm{Im}(h_{q}(re^{3i\pi/4}))=-\mathrm{Im}(h_{q}(re^{-3i\pi/4}))=-\mathrm{Im}(h_{q}(re^{-i\pi/4}))\neq 0.

Appendix B RHP theory

For completeness we recall a well known result from RHP theory and prove it below. Let R(z)R(z) be a solution of the following RHP:

Definition B.1.

Let Γ\Gamma be an appropriate curve (see Definition 1.3) with interior 𝒟\mathcal{D}_{-} and exterior 𝒟+\mathcal{D}_{+}. A 2×22\times 2 complex matrix function R(z)R(z), zz\in\mathbb{C}, is a solution of the RHP (B.1) if it satisfies the following conditions:

  1. (i)

    R(z)R(z) is analytic in Γ\mathbb{C}\setminus\Gamma.

  2. (ii)

    R(z)R(z) has continuous boundary values R(s)R^{-}(s) and R+(s)R^{+}(s) as zz approaches sΓs\in\Gamma from 𝒟\mathcal{D}_{-} and 𝒟+\mathcal{D}_{+} respectively, where

    R+(s)=R(s)J(s),sΓ,\displaystyle R^{+}(s)=R^{-}(s)J(s),\;s\in\Gamma, (B.1a)

    for a 2×22\times 2 matrix J(s)J(s).

  3. (iii)

    R(z)R(z) satisfies

    R(z)=[1001]+O(1z),as|z|.\displaystyle R(z)=\begin{bmatrix}1&0\\ 0&1\end{bmatrix}+O\left(\frac{1}{z}\right),\;as\,|z|\to\infty. (B.1b)
Theorem B.2.

Suppose that J(s)J(s) can be analytically extended to a neighbourhood of Γ\Gamma. Furthermore, for a given 0<ϵ<10<\epsilon<1 suppose that

J(s)I<ϵ,||J(s)-I||<\epsilon,

in this neighbourhood (where ||.||||.|| is the matrix norm). Then, the solution of the RHP given by Definition B.1 satisfies

R(z)I<O(ϵ),||R(z)-I||<O(\epsilon),

for all zz\in\mathbb{C}.

Proof.

Multiple sources give a proof of Theorem B.2 with various conditions on the jump matrix JJ \citesDeift1999strongkuijlaars2003riemann. For our setting the jump matrix is quite well behaved and satisfies all these constraints. We include a brief proof for completeness. Let

Δ(s)=R(s)I,\Delta(s)=R(s)-I,

substituting Δ(s)\Delta(s) into Equation (B.1a) gives

R+(s)=R(s)(I+ϵΔ(s)).R^{+}(s)=R^{-}(s)(I+\epsilon\Delta(s)).

By the asymptotic condition, Equation (B.1b), we conclude that

R(z)=I+ϵ2πiΓR(s)Δ(s)zs𝑑s.R(z)=I+\frac{\epsilon}{2\pi i}\oint_{\Gamma}\frac{R_{-}(s)\Delta(s)}{z-s}ds. (B.2)

Let LL be defined as L=supz𝒟(|R(z)|)L=\sup_{z\in\mathcal{D}_{-}}(|R(z)|), and let the maximum be at zLz_{L}. As R(z)R(z) is analytic in 𝒟\mathcal{D}_{-} it follows |R(z)||R(z)| achieves its maximum on the boundary (i.e. on Γ\Gamma). By assumption R(z)R(z) and Δ(z)\Delta(z) are also analytic for some fixed distance rr from Γ\Gamma, let us call this curve Γi\Gamma_{i}. Therefore,

R(zL)=(I+ϵ2πiΓiR(s)Δ(s)zs𝑑s)(I+ϵΔ(s))1,R(z_{L})=\left(I+\frac{\epsilon}{2\pi i}\oint_{\Gamma_{i}}\frac{R_{-}(s)\Delta(s)}{z-s}ds\right)\left(I+\epsilon\Delta(s)\right)^{-1}, (B.3)

where (I+ϵΔ(s))1\left(I+\epsilon\Delta(s)\right)^{-1} can be determined using the Nuemann series

(I+ϵΔ(s))1=j=0(ϵΔ(s))j.\left(I+\epsilon\Delta(s)\right)^{-1}=\sum_{j=0}^{\infty}(-\epsilon\Delta(s))^{j}.

We conclude from Equation (B.3)

L<|I+ϵΔΓilen(Γi)2πr||j=0(ϵΔΓ)j|,L<\left|I+\frac{\epsilon\|\Delta\|_{\Gamma_{i}}\mathrm{len}(\Gamma_{i})}{2\pi r}\right|\left|\sum_{j=0}^{\infty}(-\epsilon\|\Delta\|_{\Gamma})^{j}\right|, (B.4)

and hence |L1|=O(ϵ)|L-1|=O(\epsilon). Thus we find that,

|R(z)I|<cΓO(ϵ)|R(z)-I|<c_{\Gamma}O(\epsilon) (B.5)

For some constant

cΓ=ΔΓilen(Γi)2πr+ΔΓ,c_{\Gamma}=\frac{\|\Delta\|_{\Gamma_{i}}\mathrm{len}(\Gamma_{i})}{2\pi r}+\|\Delta\|_{\Gamma},

which is independent of ϵ\epsilon. ∎

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 qq-Discrete Painlevé equations for recurrence coefficients of modified qq-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 qq-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 qq-orthogonal polynomials and the qq-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