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

Random Motzkin paths near boundary

Włodzimierz Bryc Włodzimierz Bryc
Department of Mathematical Sciences
University of Cincinnati
2815 Commons Way
Cincinnati, OH, 45221-0025, USA.
[email protected]
Alexey Kuznetsov Alexey Kuznetsov
Department of Mathematics and Statistics
York University, 4700 Keele Street
Toronto, Ontario, M3J 1P3, Canada
[email protected]
 and  Jacek Wesołowski Jacek Wesołowski, Faculty of Mathematics and Information Science, Warsaw University of Technology, pl. Politechniki 1, 00-661 Warszawa, Poland [email protected]

Limits of Random Motzkin paths with KPZ related asymptotics

Włodzimierz Bryc Włodzimierz Bryc
Department of Mathematical Sciences
University of Cincinnati
2815 Commons Way
Cincinnati, OH, 45221-0025, USA.
[email protected]
Alexey Kuznetsov Alexey Kuznetsov
Department of Mathematics and Statistics
York University, 4700 Keele Street
Toronto, Ontario, M3J 1P3, Canada
[email protected]
 and  Jacek Wesołowski Jacek Wesołowski, Faculty of Mathematics and Information Science, Warsaw University of Technology, pl. Politechniki 1, 00-661 Warszawa, Poland [email protected]
Abstract.

We study Motzkin paths of length LL with general weights on the edges and end points. We investigate the limit behavior of the initial and final segments of the random Motzkin path viewed as a pair of processes starting from each of the two end points as LL becomes large. We then study macroscopic limits of the resulting processes, where in two different regimes we obtain Markov processes that appeared in the description of the stationary measure for the KPZ equation on the half line and of conjectural stationary measure of the hypothetical KPZ fixed point on the half line. Our results rely on the behavior of the Al-Salam–Chihara polynomials in the neighbourhood of the upper end of their orthogonality interval and on the limiting properties of the qq-Pochhammer and qq-Gamma functions as q1q\nearrow 1.

1. Introduction and main results

In this paper we study Markov processes that arise as limits of random Motzkin paths with random endpoints. A Motzkin path of length LL is a sequence of steps on the integer lattice that starts at point (0,n0)(0,n_{0}) with the initial altitude n0n_{0} and ends at point (L,nL)(L,n_{L}) at the final altitude nLn_{L} for some non-negative integers n0,nL,Ln_{0},n_{L},L. Steps can be upward, downward, or horizontal, along the vectors (1,1)(1,1), (1,1)(1,-1) and (1,0)(1,0), respectively, and the path cannot fall below the horizontal axis. The path is uniquely determined by the sequence of altitudes 𝜸=(n0,n1,,nL){\boldsymbol{\gamma}}=(n_{0},n_{1},\dots,n_{L}) with njnj1{1,0,1}n_{j}-n_{j-1}\in\{-1,0,1\}, j=1,,Lj=1,\dots,L.

A random Motzkin path can be generated by assigning a discrete probability measure on the set of all Motzkin paths and choosing 𝜸{\boldsymbol{\gamma}} at random according to this probability measure. We will write 𝜸=(γ0,,γL){\boldsymbol{\gamma}}=(\gamma_{0},\dots,\gamma_{L}) and will use the notation 𝜸(L){\boldsymbol{\gamma}}^{(L)} if we need to explicitly indicate its dependence on the parameter LL. In our main results we are interested in the assignment of probability L(𝜸)\mathds{P}_{L}({\boldsymbol{\gamma}}) which depends on two boundary parameters ρ0,ρ1[0,1)\rho_{0},\rho_{1}\in[0,1) and two parameters q[0,1)q\in[0,1) and σ[0,1]\sigma\in[0,1] that determine the edge weights. As the probability L(𝜸)\mathds{P}_{L}({\boldsymbol{\gamma}}) of selecting a Motzkin path 𝜸{\boldsymbol{\gamma}} we take the following expression:

(1.1) L(𝜸)=1Lρ0γ0ρ1γL(2σ)H(𝜸)k=0L[γk+1]q\mathds{P}_{L}({\boldsymbol{\gamma}})=\frac{1}{\mathfrak{C}_{L}}\rho_{0}^{\gamma_{0}}\rho_{1}^{\gamma_{L}}(2\sigma)^{H({\boldsymbol{\gamma}})}\prod_{k=0}^{L}[\gamma_{k}+1]_{q}

where [n]q=1++qn1[n]_{q}=1+\dots+q^{n-1} denotes the usual qq-number, and H(𝜸)=#{j[1,L]:γj=γj1}H({\boldsymbol{\gamma}})=\#\{j\in[1,L]:\gamma_{j}=\gamma_{j-1}\} is the number of horizontal steps, and L\mathfrak{C}_{L} is the normalizing constant. This weighting of the Motzkin paths was inspired by a formula in (Barraquand and Le Doussal,, 2023, Section 2.3), where σ=1\sigma=1. A more general setup is discussed in Section 2.

The above setup differs from the most commonly studied random Motzkin paths chosen uniformly from all Motzkin paths which start and end at altitude 0. Such random Motzkin paths correspond to random walks conditioned on staying non-negative and returning to 0 at time LL; it is well known that their asymptotic fluctuations are described by the Brownian excursion Kaigh, (1976) and their behavior near boundaries is described by an explicit Markov chain, see Keener, (1992). It turns out that new phenomena and new asymptotic fluctuations arise in the presence of boundary parameters ρ0,ρ1\rho_{0},\rho_{1}, see Bryc and Wang, (2023), Bryc and Wang, (2024). Our goal is to extend Bryc and Wang, (2023) to allow more general weights that depend on parameter qq, and to study the boundary limit Markov chain in two different asymptotic regimes. As limits in the two asymptotic regimes we recover Markov processes that appeared in the description given in Bryc and Kuznetsov, (2022) of the non-Gaussian term of the stationary measure of the KPZ equation on the half-line, see Barraquand and Le Doussal, (2022), Barraquand and Corwin, (2023), and the non-Gaussian term in the conjectural stationary measure of the hypothetical KPZ fixed point on the half line in Barraquand and Le Doussal, (2022).

We use the following standard notation for the qq-Pochhammer symbols:

(a;q)n\displaystyle(a;q)_{n} =k=0n1(1aqk),\displaystyle=\prod_{k=0}^{n-1}(1-aq^{k}), (a1,,ak;q)n\displaystyle(a_{1},\dots,a_{k};q)_{n} =(a1;q)n(a2;q)n(ak;q)n,\displaystyle=(a_{1};q)_{n}(a_{2};q)_{n}\dots(a_{k};q)_{n},
(a;q)\displaystyle(a;q)_{\infty} =k=0(1aqk),\displaystyle=\prod_{k=0}^{\infty}(1-aq^{k}), (a1,,ak;q)\displaystyle(a_{1},\dots,a_{k};q)_{\infty} =(a1;q)(a2;q)(ak;q).\displaystyle=(a_{1};q)_{\infty}(a_{2};q)_{\infty}\dots(a_{k};q)_{\infty}.

Our first main result is the following limit theorem for the boundaries of the random Motzkin path. Let 𝜸(L)={γk(L)}k0{\boldsymbol{\gamma}}^{{}^{(L)}}=\{\gamma^{(L)}_{k}\}_{k\geq 0} be a sequence of the initial altitudes of a random Motzkin path of length LL, appended with 0 for k>Lk>L, and let 𝜸~(L)={γ~k(L)}k0\widetilde{{\boldsymbol{\gamma}}}^{{}^{(L)}}=\{\widetilde{\gamma}^{(L)}_{k}\}_{k\in\mathds{Z}_{\geq 0}} be a sequence of the final altitudes, γ~k(L)=γLk\widetilde{\gamma}_{k}^{(L)}=\gamma_{L-k}, k=0,1,,Lk=0,1,\dots,L, appended with 0 for k>Lk>L.

Theorem 1.1.

Suppose that 0ρ0,ρ1,q<10\leq\rho_{0},\rho_{1},q<1 and 0<σ10<\sigma\leq 1. Then

(𝜸(L),𝜸~(L))(𝑿,𝒀),\left({\boldsymbol{\gamma}}^{{}^{(L)}},\,\widetilde{{\boldsymbol{\gamma}}}^{{}^{(L)}}\right)\Rightarrow\left({\boldsymbol{X}},\,{\boldsymbol{Y}}\right),

as LL\to\infty, where on the left hand side we have random Motzkin paths with respect to measure (1.1) and on the right-hand side we have two independent Markov chains 𝐗={Xk}k0{\boldsymbol{X}}=\left\{X_{k}\right\}_{k\geq 0}, 𝐘={Yk}k0{\boldsymbol{Y}}=\left\{Y_{k}\right\}_{k\geq 0} with the same transition probabilities on 0\mathds{Z}_{\geq 0} given by

(Xk=n+δ|Xk1=n)=1qn+11+σ{sn+12sn,δ=1,σ,δ=0,sn12sn,δ=1,0,otherwise,\mathds{P}(X_{k}=n+\delta|X_{k-1}=n)=\frac{1-q^{n+1}}{1+\sigma}\cdot\begin{cases}\frac{s_{n+1}}{2s_{n}},&\delta=1,\\ \sigma,&\delta=0,\\ \frac{s_{n-1}}{2s_{n}},&\delta=-1,\\ 0,&\mbox{\em otherwise},\end{cases}

for n=0,1,n=0,1,\ldots and with the initial laws

(1.2) (X0=n)=1C0ρ0nsn,(Y0=n)=1C1ρ1nsn,n=0,1,,\mathds{P}(X_{0}=n)=\frac{1}{C_{0}}\rho_{0}^{n}s_{n},\;\mathds{P}(Y_{0}=n)=\frac{1}{C_{1}}\rho_{1}^{n}s_{n},\quad n=0,1,\ldots,

where s1=0s_{-1}=0 and

sn=k=0n(a;q)k(q;q)k(b;q)nk(q;q)nk,n=0,1,,s_{n}=\sum_{k=0}^{n}\frac{(a;q)_{k}}{(q;q)_{k}}\;\frac{(b;q)_{n-k}}{(q;q)_{n-k}},\quad n=0,1,\ldots,

with a=q(σ+i1σ2)a=-q(\sigma+i\sqrt{1-\sigma^{2}}), b=q(σi1σ2)b=-q(\sigma-i\sqrt{1-\sigma^{2}}); the normalizing constants are

Cj=(aρj,bρj;q)(ρj;q)2,j=0,1.C_{j}=\frac{(a\rho_{j},b\rho_{j};q)_{\infty}}{(\rho_{j};q)_{\infty}^{2}},\quad j=0,1.

We note that if q=0q=0 then sn=n+1s_{n}=n+1, recovering transition probabilities in (Bryc and Wang,, 2023, Theorem 1.1), who use the parameter σ\sigma that is twice our σ\sigma. It is natural to expect that as in (Bryc and Wang,, 2023, Theorem 1.6), there is a version of this result that holds also for ρ11\rho_{1}\geq 1. However, this is beyond the scope of this paper.

Theorem 1.1 will be deduced from a more general Theorem 2.4. The proof appears in Section 4 and relies on properties of the Al-Salam–Chihara polynomials, which we discuss in Section 3.

The next two theorems give macroscopic continuous-time limits of the family of Markov processes {Xk}\{X_{k}\}. In the statements of Theorems 1.2 and 1.3 below, f.d.d.\xrightarrow{\textit{f.d.d.}} denotes convergence of finite dimensional distributions.

In our first result, we take ρ01\rho_{0}\nearrow 1 at an appropriate rate but keep qq fixed. Then the normalized Markov process {Xk}\left\{X_{k}\right\} converges to the Bessel process, which for σ=1\sigma=1 appeared in the description of the non-Gaussian term in the conjectural stationary measure for the hypothetical KPZ fixed point on the half-line in (Bryc and Kuznetsov,, 2022, Theorem 2.6 ).

Theorem 1.2.

Fix 0q<10\leq q<1, 0<σ10<\sigma\leq 1 and 𝖼>0{\mathsf{c}}>0. Let {Xk(N)}k0\{X_{k}^{(N)}\}_{k\in\mathds{Z}_{\geq 0}} be a Markov process from Theorem 1.1 with the initial law that depends on NN through ρ0=e𝖼/N\rho_{0}=e^{-{\mathsf{c}}/\sqrt{N}}. Then

(1.3) 1N{XNt(N)}t0f.d.d.{ξt(𝖼)}t0 as N,\frac{1}{\sqrt{N}}\left\{X_{\left\lfloor Nt\right\rfloor}^{(N)}\right\}_{t\geq 0}\xrightarrow{\textit{f.d.d.}}\left\{\xi_{t}^{({\mathsf{c}})}\right\}_{t\geq 0}\;\mbox{ as }N\to\infty,

where {ξt(𝖼)}t0\{\xi_{t}^{({\mathsf{c}})}\}_{t\geq 0} is the 3-dimensional Bessel process with transition probabilities

(1.4) (ξt(𝖼)=dy|ξ0(𝖼)=x)=yx𝗊t1+σ(x,y)dy,\mathds{P}(\xi_{t}^{({\mathsf{c}})}=dy|\xi_{0}^{({\mathsf{c}})}=x)=\frac{y}{x}\,{\mathsf{q}}_{\frac{t}{1+\sigma}}(x,y)\,dy,

where 𝗊t{\mathsf{q}}_{t}, t>0t>0, is the transition kernel of the Brownian motion killed at hitting zero, i.e.

(1.5) 𝗊t(x,y)=12πt(exp((yx)22t)exp((y+x)22t)),x,y>0,{\mathsf{q}}_{t}(x,y)=\frac{1}{\sqrt{2\pi t}}\left(\exp(-\tfrac{(y-x)^{2}}{2t})-\exp(-\tfrac{(y+x)^{2}}{2t})\right),\quad x,y>0,

and with the initial distribution

(1.6) (ξ0(𝖼)=dx)=𝖼2xe𝖼x𝟏{x>0}dx.\mathds{P}(\xi_{0}^{({\mathsf{c}})}=dx)={\mathsf{c}}^{2}xe^{-{\mathsf{c}}x}{\bf 1}_{\{x>0\}}dx.

In our second result, we take both ρ01\rho_{0}\nearrow 1 and q1q\nearrow 1 at appropriate rates. Then, under appropriate centering and normalization, the Markov process {Xk}\left\{X_{k}\right\} converges to the Markov process on \mathds{R}, which for σ=1\sigma=1 appeared in the description of the non-Gaussian term in the representation of the stationary measure for the KPZ equation on the half line given in (Bryc and Kuznetsov,, 2022, Theorem 2.3). (This is only a one-parameter subset of the two-parameter family of such measures conjectured in Barraquand and Le Doussal, (2022) and proved rigorously in Barraquand and Corwin, (2023).) In the statement, Kν(z)K_{\nu}(z) is a modified Bessel K function (Macdonald function) of imaginary index ν\nu and positive argument zz, see e.g. (Olver et al.,, 2023, 10.32.E9).

Theorem 1.3.

Fix 0<σ10<\sigma\leq 1 and 𝖼>0{\mathsf{c}}>0. Let {Xk(N)}k0\{X_{k}^{(N)}\}_{k\in\mathds{Z}_{\geq 0}} be a Markov process from Theorem 1.1 with parameters q=e2/Nq=e^{-2/\sqrt{N}} and ρ0=e𝖼/N\rho_{0}=e^{-{\mathsf{c}}/\sqrt{N}}. Then

(1.7) 1N{XNt(N)Nlog2N(1+σ)}t0f.d.d.{ζt(𝖼)}t0 as N,\frac{1}{\sqrt{N}}\left\{X_{\left\lfloor Nt\right\rfloor}^{(N)}-\sqrt{N}\log\sqrt{2N(1+\sigma)}\right\}_{t\geq 0}\xrightarrow{\textit{f.d.d.}}\left\{\zeta_{t}^{({\mathsf{c}})}\right\}_{t\geq 0}\;\mbox{ as }N\to\infty,

where ζ(𝖼)\zeta^{({\mathsf{c}})} is a Markov process with transition probabilities

(1.8) (ζt(𝖼)=dy|ζ0(𝖼)=x)=K0(ey)K0(ex)𝗉t1+σ(x,y)dy,x,y,\mathds{P}(\zeta_{t}^{{}^{({\mathsf{c}})}}=dy|\zeta_{0}^{{}^{({\mathsf{c}})}}=x)=\frac{K_{0}(e^{-y})}{K_{0}(e^{-x})}\;{\mathsf{p}}_{\frac{t}{1+\sigma}}(x,y)\,dy,\quad x,y\in\mathds{R},

where 𝗉t{\mathsf{p}}_{t}, t>0t>0, is the Yakubovich transition kernel, Yakubovich, (2011), defined by

(1.9) 𝗉t(x,y)=2π0etu2/2Kiu(ex)Kiu(ey)du|Γ(iu)|2,{\mathsf{p}}_{t}(x,y)=\frac{2}{\pi}\int_{0}^{\infty}e^{-tu^{2}/2}K_{{\textnormal{i}}u}(e^{-x})K_{{\textnormal{i}}u}(e^{-y})\frac{du}{|\Gamma(iu)|^{2}},

and with the initial distribution

(1.10) (ζ0(𝖼)=dx)=42𝖼Γ(𝖼/2)2e𝖼xK0(ex)dx.\mathds{P}(\zeta_{0}^{({\mathsf{c}})}=dx)=\frac{4}{2^{\mathsf{c}}\Gamma({\mathsf{c}}/2)^{2}}e^{-{\mathsf{c}}x}K_{0}(e^{-x})\,dx.

Theorems 1.2 and 1.3 will follow from local limits in Theorem 4.2 and Theorem 4.3.

The paper is organized as follows. In Section 2 we introduce random Motzkin paths with general weights, we prove matrix representation, integral representation, and a general boundary limit theorem. In Section 3 we recall some properties of the Al–Salam–Chihara polynomials, and we prove new asymptotic results for their behavior near the boundary of the interval of orthogonality. In Section 4 we use these results to prove Theorem 1.1 and two local limit theorems, which we use to complete proofs of Theorems 1.2 and 1.3. In Section 5 we discuss properties of the qq-Pochhammer and qq-Gamma functions and derive the limits and bounds needed in our proofs.

2. General limit theorem for random Motzkin paths at the boundary

In this section we introduce a more general setting and prove a version of Theorem 1.1 for more general random Motzkin paths.

Recall that a Motzkin path of length L=1,2,L=1,2,\dots is a sequence of lattice points (𝒙0,,𝒙L)({\boldsymbol{x}}_{0},\dots,{\boldsymbol{x}}_{L}) such that 𝒙j=(j,nj)0×0,j=0,1,,L{\boldsymbol{x}}_{j}=(j,n_{j})\in\mathds{Z}_{\geq 0}\times\mathds{Z}_{\geq 0},\;j=0,1,\dots,L. An edge (𝒙j1,𝒙j)({\boldsymbol{x}}_{j-1},{\boldsymbol{x}}_{j}) is called a (jj-th) step, and only three types of steps are allowed: upward steps, downward steps, and horizontal steps. The edge (𝒙j1,𝒙j)({\boldsymbol{x}}_{j-1},{\boldsymbol{x}}_{j}) is an upward step if njnj1=1n_{j}-n_{j-1}=1, a downward step if njnj1=1n_{j}-n_{j-1}=-1, and a horizontal step if njnj1=0n_{j}-n_{j-1}=0, see, e.g., Flajolet and Sedgewick, (2009, Definition V.4, p. 319) or Viennot, (1985). Each such path can be identified with a sequence of non-negative integers that specify the starting point (0,n0)(0,n_{0}) and consecutive values njn_{j} along the vertical axis at step j1j\geq 1. We shall write 𝜸=(γ0,γ1,,γL){\boldsymbol{\gamma}}=(\gamma_{0},\gamma_{1},\dots,\gamma_{L}) with γj=nj\gamma_{j}=n_{j} for such a sequence and refer to 𝜸{\boldsymbol{\gamma}} as a Motzkin path. By m,n(L)\mathcal{M}_{m,n}^{(L)} we denote the family of all Motzkin paths 𝜸{\boldsymbol{\gamma}} of length LL with the initial altitude γ0=m\gamma_{0}=m and the final altitude γL=n\gamma_{L}=n. We also refer to γ0\gamma_{0} and γL\gamma_{L} as the boundary/end points of the path.

We are interested in probabilistic properties of random Motzkin paths. To introduce a probability measure on the set of the Motzkin paths, we assign weights to the edges and to the end-points of a Motzkin path. The weights for the edges arise multiplicatively from three sequences 𝒂=(aj)j0,𝒃=(bj)j0,𝒄=(cj)j0{\boldsymbol{a}}=(a_{j})_{j\geq 0},{\boldsymbol{b}}=(b_{j})_{j\geq 0},{\boldsymbol{c}}=(c_{j})_{j\geq 0} of positive real numbers. For a path 𝜸=(γ0=m,γ1,,γL1,γL=n)m,n(L){\boldsymbol{\gamma}}=(\gamma_{0}=m,\gamma_{1},\dots,\gamma_{L-1},\gamma_{L}=n)\in\mathcal{M}_{m,n}^{(L)} we define its (edge) weight

(2.1) w(𝜸)=k=1Laγk1εk+bγk1εk0cγk1εk,w({\boldsymbol{\gamma}})=\prod_{k=1}^{L}a_{\gamma_{k-1}}^{\varepsilon_{k}^{+}}b_{\gamma_{k-1}}^{\varepsilon_{k}^{0}}c_{\gamma_{k-1}}^{\varepsilon_{k}^{-}},

where

εk+(𝜸):=𝟏γk>γk1,εk(𝜸):=𝟏γk<γk1,εk0(𝜸):=𝟏γk=γk1,k=1,,L.\varepsilon_{k}^{+}({\boldsymbol{\gamma}}):={\bf 1}_{\gamma_{k}>\gamma_{k-1}},\;\varepsilon_{k}^{-}({\boldsymbol{\gamma}}):={\bf 1}_{\gamma_{k}<\gamma_{k-1}},\;\varepsilon_{k}^{0}({\boldsymbol{\gamma}}):={\bf 1}_{\gamma_{k}=\gamma_{k-1}},k=1,\dots,L.

That is, the edge weight is multiplicative in the edges, we take 𝒂{\boldsymbol{a}}, 𝒃{\boldsymbol{b}} and 𝒄{\boldsymbol{c}} as the weights of the upward steps, horizontal steps and downward steps, and the weight of a step depends on the altitude of the left-end of an edge. Note that the value of c0c_{0} does not contribute to w(𝜸)w({\boldsymbol{\gamma}}). Since i,j(L)\mathcal{M}_{i,j}^{(L)} is a finite set, the normalization constants

𝔚m,n(L)=𝜸m,n(L)w(𝜸)\mathfrak{W}^{(L)}_{m,n}=\sum_{{\boldsymbol{\gamma}}\in\mathcal{M}_{m,n}^{(L)}}w({\boldsymbol{\gamma}})

are well defined for all m,n0m,n\geq 0.

c2c_{2}b1b_{1}c1c_{1}a0a_{0}b1b_{1}a1a_{1}c2c_{2}c1c_{1}a0a_{0}112233445566778899
Figure 1. Motzkin path 𝜸=(2,1,1,0,1,1,2,1,0,1)(9){\boldsymbol{\gamma}}=(2,1,1,0,1,1,2,1,0,1)\in\mathcal{M}^{(9)} with weight contributions marked at the edges. The probability of selecting the path shown above from (9)\mathcal{M}^{(9)} is 9(𝜸)=α2β1𝜶,𝜷,9b12a02a1c12c22\mathds{P}_{9}({\boldsymbol{\gamma}})=\frac{\alpha_{2}\beta_{1}}{\mathfrak{C}_{{\boldsymbol{\alpha}},{\boldsymbol{\beta}},9}}b_{1}^{2}a_{0}^{2}a_{1}c_{1}^{2}c_{2}^{2}.

In addition to the weights of the edges, we also assign weights to the two boundary points. To this end we choose two additional non-negative sequences 𝜶=(αm)m0{\boldsymbol{\alpha}}=(\alpha_{m})_{m\geq 0} and 𝜷=(βn)n0{\boldsymbol{\beta}}=(\beta_{n})_{n\geq 0} such that

(2.2) 𝜶,𝜷,L:=m,n0αm𝔚m,n(L)βn<.\mathfrak{C}_{{\boldsymbol{\alpha}},{\boldsymbol{\beta}},L}:=\sum_{m,n\geq 0}\alpha_{m}\mathfrak{W}^{(L)}_{m,n}\beta_{n}<\infty.

With finite normalizing constant (2.2), the countable set (L)=m,n0m,n(L)\mathcal{M}^{(L)}=\bigcup_{m,n\geq 0}\mathcal{M}_{m,n}^{(L)} becomes a probability space with the discrete probability measure

(2.3) L(𝜸)𝜶,𝜷,L({𝜸})=αγ0βγL𝜶,𝜷,Lw(𝜸),𝜸(L).\mathds{P}_{L}({\boldsymbol{\gamma}})\equiv\mathds{P}_{{{\boldsymbol{\alpha}}},{{\boldsymbol{\beta}}},L}(\{{\boldsymbol{\gamma}}\})=\frac{\alpha_{\gamma_{0}}\beta_{\gamma_{L}}}{\mathfrak{C}_{{{\boldsymbol{\alpha}}},{{\boldsymbol{\beta}}},L}}\,w({\boldsymbol{\gamma}}),\quad{\boldsymbol{\gamma}}\in\mathcal{M}^{(L)}.

For an illustration, see Figure 1. We now consider a random vector (γ0(L),,γL(L))(L)(\gamma_{0}^{(L)},\dots,\gamma_{L}^{(L)})\in\mathcal{M}^{(L)} sampled from L\mathds{P}_{L}, and extend it to an infinite process denoted by

(2.4) 𝜸(L):={γk(L)}k0, with γk(L):=0 if k>L.{\boldsymbol{\gamma}}^{(L)}:=\left\{\gamma_{k}^{(L)}\right\}_{k\geq 0},\quad\mbox{ with }\quad\gamma_{k}^{(L)}:=0\mbox{ if }k>L.

Similarly, we also introduce the infinite reversed process by

(2.5) 𝜸~(L):={γ~k(L)}k0 with γ~k(L):={γLk(L), if k=0,,L,0, if k>L.\widetilde{{\boldsymbol{\gamma}}}^{(L)}:=\left\{\widetilde{\gamma}_{k}^{(L)}\right\}_{k\geq 0}\quad\mbox{ with }\quad\widetilde{\gamma}_{k}^{(L)}:=\begin{cases}\gamma_{L-k}^{(L)},&\mbox{ if }k=0,\dots,L,\\ 0,&\mbox{ if }k>L.\end{cases}

We are interested in the limit processes for both 𝜸(L){\boldsymbol{\gamma}}^{(L)} and 𝜸~(L)\widetilde{{\boldsymbol{\gamma}}}^{(L)} as LL\to\infty. In these limits, we do not scale the sequences nor the locations.

Recall that weak convergence of discrete-time processes means convergence of finite-dimensional distributions (Billingsley,, 1999). For integer-valued random variables, the latter follows from convergence of probability generating functions. We will therefore fix KK, z0,z1(0,1]z_{0},z_{1}\in(0,1] and t1,,tK,s1,,sK>0t_{1},\dots,t_{K},s_{1},\dots,s_{K}>0 and the goal is to determine two discrete time processes 𝑿={Xk}k0{\boldsymbol{X}}=\{X_{k}\}_{k\geq 0} and 𝒀={Yk}k0{\boldsymbol{Y}}=\{Y_{k}\}_{k\geq 0} such that

(2.6) limL𝔼[z0γ0(L)j=1Ktjγj(L)γj1(L)j=1KsjγLj(L)γL+1j(L)z1γL(L)]=𝔼[z0X0j=1KtjXjXj1]𝔼[z1Y0j=1KsjYjYj1].\lim_{L\to\infty}\mathds{E}\left[z_{0}^{\gamma_{0}^{(L)}}\prod_{j=1}^{K}t_{j}^{\gamma_{j}^{(L)}-\gamma_{j-1}^{(L)}}\prod_{j=1}^{K}s_{j}^{\gamma_{L-j}^{(L)}-\gamma_{L+1-j}^{(L)}}z_{1}^{\gamma_{L}^{(L)}}\right]\\ =\mathds{E}\left[z_{0}^{X_{0}}\prod_{j=1}^{K}t_{j}^{X_{j}-X_{j-1}}\right]\mathds{E}\left[z_{1}^{Y_{0}}\prod_{j=1}^{K}s_{j}^{Y_{j}-Y_{j-1}}\right].

Indeed, the above expressions uniquely determine the corresponding probability generating functions for small enough arguments. For example,

𝔼[j=0KvjXj]=𝔼[z0X0j=1KtjXjXj1]\mathds{E}\left[\prod_{j=0}^{K}v_{j}^{X_{j}}\right]=\mathds{E}\left[z_{0}^{X_{0}}\prod_{j=1}^{K}t_{j}^{X_{j}-X_{j-1}}\right]

with z0=v0vKz_{0}=v_{0}\dots v_{K} and tj=vjvj+1vKt_{j}=v_{j}v_{j+1}\dots v_{K}.

2.1. Matrix representation and integral representation

We will need a convenient representation for the left hand side of (2.6). We introduce a tri-diagonal matrix

𝑴t:=[b0a0t00c1/tb1a1t00c2/tb2a2t00c3/t],{{\boldsymbol{M}}}_{t}:=\left[\begin{matrix}b_{0}&a_{0}t&0&0&\cdots\\ c_{1}/t&b_{1}&a_{1}t&0&\\ 0&c_{2}/t&b_{2}&a_{2}t&\\ 0&0&c_{3}/t&\ddots&\ddots\\ \vdots&&&\ddots&\ddots\end{matrix}\right],

and infinite column vectors

V𝜶(z):=[αnzn,n=0,1,]TandW𝜶(z):=[βnzn,n=0,1,]T.\vec{V}_{{\boldsymbol{\alpha}}}(z):=[\alpha_{n}\,z^{n},\,n=0,1,\ldots]^{T}\quad\mbox{and}\quad\vec{W}_{{\boldsymbol{\alpha}}}(z):=[\beta_{n}\,z^{n},\,n=0,1,\ldots]^{T}.

We use the following matrix representation for the left hand side of (2.6).

Lemma 2.1 (matrix ansatz).

For every K=0,1,K=0,1,\dots such that 2KL2K\leq L, we have

(2.7) 𝔼[z0γ0(L)j=1Ktjγj(L)γj1(L)j=1KsjγLj(L)γL+1j(L)z1γL(L)]=1𝜶,𝜷,LV𝜶(z0)T𝑴t1𝑴t2𝑴tK𝑴1L2K𝑴1/sK𝑴1/s2𝑴1/s1W𝜷(z1),\mathds{E}\left[z_{0}^{\gamma_{0}^{(L)}}\prod_{j=1}^{K}t_{j}^{\gamma_{j}^{(L)}-\gamma_{j-1}^{(L)}}\prod_{j=1}^{K}s_{j}^{\gamma_{L-j}^{(L)}-\gamma_{L+1-j}^{(L)}}z_{1}^{\gamma_{L}^{(L)}}\right]\\ =\frac{1}{\mathfrak{C}_{{{\boldsymbol{\alpha}}},{{\boldsymbol{\beta}}},L}}\vec{V}_{{\boldsymbol{\alpha}}}(z_{0})^{T}{{\boldsymbol{M}}}_{t_{1}}{{\boldsymbol{M}}}_{t_{2}}\cdots{{\boldsymbol{M}}}_{t_{K}}{{\boldsymbol{M}}}_{1}^{L-2K}{{\boldsymbol{M}}}_{1/s_{K}}\cdots{{\boldsymbol{M}}}_{1/s_{2}}{{\boldsymbol{M}}}_{1/s_{1}}\vec{W}_{{\boldsymbol{\beta}}}(z_{1}),

where

(2.8) 𝜶,𝜷,L=V𝜶(1)TM1LW𝜷(1)\mathfrak{C}_{{{\boldsymbol{\alpha}}},{{\boldsymbol{\beta}}},L}=\vec{V}_{{\boldsymbol{\alpha}}}(1)^{T}M_{1}^{L}\vec{W}_{{\boldsymbol{\beta}}}(1)

is the normalization constant.

Proof.

Formula (2.7) follows from a more general formula

(2.9) m,n=0αmz0mβnz1nγm,n(L)j=1Ltjγjγj1w(𝜸)=V𝜶(z0)T𝑴t1𝑴t2𝑴tLW𝜷(z1).\sum_{m,n=0}^{\infty}\alpha_{m}z_{0}^{m}\beta_{n}z_{1}^{n}\sum_{\gamma\in\mathcal{M}_{m,n}^{{}^{(L)}}}\prod_{j=1}^{L}t_{j}^{\gamma_{j}-\gamma_{j-1}}w({\boldsymbol{\gamma}})=\vec{V}_{{\boldsymbol{\alpha}}}(z_{0})^{T}{{\boldsymbol{M}}}_{t_{1}}{{\boldsymbol{M}}}_{t_{2}}\cdots{{\boldsymbol{M}}}_{t_{L}}\vec{W}_{{\boldsymbol{\beta}}}(z_{1}).

applied to the sequence (t1,,tL)(t_{1},\dots,t_{L}) of the form (t1,,tK,1,,1,1/sK,,1/s1)(t_{1},\dots,t_{K},1,\dots,1,1/s_{K},\dots,1/s_{1}).

To prove (2.9), consider an infinite matrix 𝑾(L){{\boldsymbol{W}}}(L) with entries

(2.10) [𝑾(L)]m,n:=γm,n(L)j=1Ltjγjγj1w(𝜸).[{{\boldsymbol{W}}}(L)]_{m,n}:=\sum_{\gamma\in\mathcal{M}_{m,n}^{{}^{(L)}}}\prod_{j=1}^{L}t_{j}^{\gamma_{j}-\gamma_{j-1}}w({\boldsymbol{\gamma}}).

It is clear that the left hand side of (2.9) is

V𝜶(z0)T𝑾(L)W𝜷(z1).\vec{V}_{{\boldsymbol{\alpha}}}(z_{0})^{T}{{\boldsymbol{W}}}(L)\vec{W}_{{\boldsymbol{\beta}}}(z_{1}).

It remains to verify that

𝑾(L)=𝑴t1𝑴t2𝑴tL.{{\boldsymbol{W}}}(L)={{\boldsymbol{M}}}_{t_{1}}{{\boldsymbol{M}}}_{t_{2}}\cdots{{\boldsymbol{M}}}_{t_{L}}.

With the case L=0L=0 being trivial, we proceed by induction. In view of (2.10) we have

[𝑾(L+1)]m,n=tL+1[𝑾(L)]m,n1an1+[𝑾(L)]m,nbn+tL+11[𝑾(L)]m,n+1cn+1[{{\boldsymbol{W}}}(L+1)]_{m,n}=t_{L+1}[{{\boldsymbol{W}}}(L)]_{m,n-1}a_{n-1}+[{{\boldsymbol{W}}}(L)]_{m,n}b_{n}+t_{L+1}^{-1}[{{\boldsymbol{W}}}(L)]_{m,n+1}c_{n+1}

Thus 𝑾(L+1)=𝑾(L)𝑴tL+1{{\boldsymbol{W}}}(L+1)={{\boldsymbol{W}}}(L){{\boldsymbol{M}}}_{t_{L+1}}. ∎

The key step in our proof of boundary limit theorem is an integral representation for the left hand side of (2.6) based on the orthogonality measure of the orthogonal polynomials determined by the edge weights of Motzkin paths.

Following Viennot, (1985) and Flajolet and Sedgewick, (2009) with the sequences 𝒂,𝒃,𝒄{\boldsymbol{a}},{\boldsymbol{b}},{\boldsymbol{c}} of edge weights for the Motzkin paths that appeared in (2.1) we now associate real polynomials p1(x)=0p_{-1}(x)=0, p0(x)=1,p1(x),p_{0}(x)=1,p_{1}(x),\dots defined by the three step recurrence

(2.11) xpn(x)=anpn+1(x)+bnpn(x)+cnpn1(x),n=0,1,2,xp_{n}(x)=a_{n}p_{n+1}(x)+b_{n}p_{n}(x)+c_{n}p_{n-1}(x),\quad n=0,1,2,\dots

(With the usual conventions that p1(x)=0p_{-1}(x)=0, p0(x):=1p_{0}(x):=1 and an>0a_{n}>0, the recursion determines polynomials {pn(x)}\{p_{n}(x)\} uniquely, with p1(x)=(xb0)/a0p_{1}(x)=(x-b_{0})/a_{0}.)

By Favard’s theorem (Ismail,, 2009), polynomials {pn(x)}\{p_{n}(x)\} are orthogonal with respect to a probability measure ν\nu on the real line, which we assume to be compactly supported, and thus unique. It is well known that the L2L_{2} norm pn22:=(pn(x))2ν(dx)\|p_{n}\|_{2}^{2}:=\int_{\mathds{R}}(p_{n}(x))^{2}\nu(dx) is given by the formula

(2.12) pn22=k=1nckak1.\|p_{n}\|_{2}^{2}=\prod_{k=1}^{n}\frac{c_{k}}{a_{k-1}}.

We need some conditions on the weights of the end points 𝜶,𝜷{{\boldsymbol{\alpha}}},{{\boldsymbol{\beta}}} in (2.3).

Assumption 2.1.

We assume that

  1. (A1A_{1})

    The series

    (2.13) ϕ𝜶(x,z)=n=0αnznpn(x) and ψ𝜷(x,z)=n=0βnznpn(x)pn22\phi_{{\boldsymbol{\alpha}}}(x,z)=\sum_{n=0}^{\infty}\alpha_{n}z^{n}p_{n}(x)\quad\mbox{ and }\quad\psi_{{\boldsymbol{\beta}}}(x,z)=\sum_{n=0}^{\infty}\beta_{n}z^{n}\frac{p_{n}(x)}{\|p_{n}\|_{2}^{2}}

    converge absolutely on the support of ν\nu for all z,|z|1z\in\mathds{C},|z|\leq 1.

  2. (A2A_{2})

    The function xLϕα(x,1)ψβ(x,1)x^{L}\phi_{\alpha}(x,1)\psi_{\beta}(x,1) is integrable with respect to the measure ν\nu, and

    (2.14) m,n=0αnβmpm22|xLpn(x)pm(x)|ν(dx)<.\int_{\mathds{R}}\sum_{m,n=0}^{\infty}\alpha_{n}\frac{\beta_{m}}{\|p_{m}\|_{2}^{2}}\left|x^{L}p_{n}(x)p_{m}(x)\right|\nu(dx)<\infty.

Consider the following two infinite column vectors

P(x):=[pi(x),i=0,1,]TandQ(x):=[p~i(x),i=0,1,]T\vec{P}(x):=\left[p_{i}(x),\,i=0,1,\ldots\right]^{T}\quad\mbox{and}\quad\vec{Q}(x):=[\widetilde{p}_{i}(x),\,i=0,1,\ldots]^{T}

where

(2.15) p~n(x)=pn(x)/pn22=pn(x)k=1nak1ck.\widetilde{p}_{n}(x)=p_{n}(x)/\|p_{n}\|_{2}^{2}=p_{n}(x)\prod_{k=1}^{n}\frac{a_{k-1}}{c_{k}}.

Note that with a1:=0a_{-1}:=0, polynomials p~n(x)\widetilde{p}_{n}(x) satisfy the three step recurrence

xp~n(x)=cn+1p~n+1(x)+bnp~n(x)+an1p~n1(x).x\widetilde{p}_{n}(x)=c_{n+1}\widetilde{p}_{n+1}(x)+b_{n}\widetilde{p}_{n}(x)+a_{n-1}\widetilde{p}_{n-1}(x).

In particular,

(2.16) 𝑴1LP(x)=xLP(x) and QT(x)𝑴1L=xLQT.{{\boldsymbol{M}}}_{1}^{L}\vec{P}(x)=x^{L}\vec{P}(x)\quad\mbox{ and }\quad\vec{Q}^{T}(x){{\boldsymbol{M}}}_{1}^{L}=x^{L}\vec{Q}^{T}.

The left hand side of the expression (2.7) has the following integral representation.

Lemma 2.2.

If (2.14) holds and 2KL2K\leq L, then

(2.17) 𝔼[z0γ0(L)j=1Ktjγj(L)γj1(L)j=1KsjγLj(L)γL+1j(L)z1γL(L)]=1𝜶,𝜷,LxL2KΨ0(x)Ψ1(x)ν(dx),\mathds{E}\left[z_{0}^{\gamma_{0}^{(L)}}\prod_{j=1}^{K}t_{j}^{\gamma_{j}^{(L)}-\gamma_{j-1}^{(L)}}\prod_{j=1}^{K}s_{j}^{\gamma_{L-j}^{(L)}-\gamma_{L+1-j}^{(L)}}z_{1}^{\gamma_{L}^{(L)}}\right]=\frac{1}{\mathfrak{C}_{{{\boldsymbol{\alpha}}},{{\boldsymbol{\beta}}},L}}\int x^{L-2K}\Psi_{0}(x)\Psi_{1}(x)\nu(dx),

where

Ψ0(x)=V𝜶(z0)T𝑴t1𝑴t2𝑴tKP(x)andΨ1(x)=W𝜷(z1)T𝑴~s1𝑴~s2𝑴~sKQ(x),\Psi_{0}(x)={\vec{V}_{{\boldsymbol{\alpha}}}(z_{0})^{T}{{\boldsymbol{M}}}_{t_{1}}{{\boldsymbol{M}}}_{t_{2}}\cdots{{\boldsymbol{M}}}_{t_{K}}\vec{P}(x)}\quad\mbox{and}\quad\Psi_{1}(x)={\vec{W}_{{\boldsymbol{\beta}}}(z_{1})^{T}\widetilde{{{\boldsymbol{M}}}}_{s_{1}}\widetilde{{{\boldsymbol{M}}}}_{s_{2}}\cdots\widetilde{{{\boldsymbol{M}}}}_{s_{K}}\vec{Q}(x)},

with 𝐌~s=𝐌1/sT\widetilde{{{\boldsymbol{M}}}}_{s}={{\boldsymbol{M}}}_{1/s}^{T}. Moreover, 𝛂,𝛃,L\mathfrak{C}_{{{\boldsymbol{\alpha}}},{{\boldsymbol{\beta}}},L}, introduced in (2.8), can be written as

𝜶,𝜷,L=xL(V𝜶(1)TP(x))(W𝜷(1)TQ(x))ν(dx).\mathfrak{C}_{{{\boldsymbol{\alpha}}},{{\boldsymbol{\beta}}},L}=\int x^{L}\left(\vec{V}_{{\boldsymbol{\alpha}}}(1)^{T}\vec{P}(x)\right)\left(\vec{W}_{{\boldsymbol{\beta}}}(1)^{T}\vec{Q}(x)\right)\nu(dx).
Proof.

We use expression (2.9) with (t1,,tL)(t_{1},\dots,t_{L}) of the form (t1,,tK,1,,1,1/sK,,1/s1)(t_{1},\dots,t_{K},1,\dots,1,1/s_{K},\dots,1/s_{1}). Proceeding somewhat formally, we write the identity matrix as the integral, 𝑰=P(x)Q(x)Tν(dx){{\boldsymbol{I}}}=\int\,\vec{P}(x)\vec{Q}(x)^{T}\,\nu(dx) and rewrite the product of matrices in (2.7) as

𝑴t1𝑴t2𝑴tK𝑴1L2K𝑴1/sK𝑴1/s2𝑴1/s1\displaystyle{{\boldsymbol{M}}}_{t_{1}}{{\boldsymbol{M}}}_{t_{2}}\cdots{{\boldsymbol{M}}}_{t_{K}}{{\boldsymbol{M}}}_{1}^{L-2K}{{\boldsymbol{M}}}_{1/s_{K}}\cdots{{\boldsymbol{M}}}_{1/s_{2}}{{\boldsymbol{M}}}_{1/s_{1}}
=\displaystyle= 𝑴t1𝑴t2𝑴tK𝑴1L2K(P(x)Q(x)Tν(dx))𝑴1/sK𝑴1/s2𝑴1/s1\displaystyle{{\boldsymbol{M}}}_{t_{1}}{{\boldsymbol{M}}}_{t_{2}}\cdots{{\boldsymbol{M}}}_{t_{K}}{{\boldsymbol{M}}}_{1}^{L-2K}\left(\int\,\vec{P}(x)\vec{Q}(x)^{T}\,\nu(dx)\right)\,{{\boldsymbol{M}}}_{1/s_{K}}\cdots{{\boldsymbol{M}}}_{1/s_{2}}{{\boldsymbol{M}}}_{1/s_{1}}
=\displaystyle= 𝑴t1𝑴t2𝑴tK(𝑴1L2KP(x))Q(x)T𝑴1/sK𝑴1/s2𝑴1/s1ν(dx)\displaystyle\int{{\boldsymbol{M}}}_{t_{1}}{{\boldsymbol{M}}}_{t_{2}}\cdots{{\boldsymbol{M}}}_{t_{K}}\left({{\boldsymbol{M}}}_{1}^{L-2K}\vec{P}(x)\right)\,\vec{Q}(x)^{T}\,{{\boldsymbol{M}}}_{1/s_{K}}\cdots{{\boldsymbol{M}}}_{1/s_{2}}{{\boldsymbol{M}}}_{1/s_{1}}\nu(dx)
=(2.16)\displaystyle\stackrel{{\scriptstyle\eqref{mpq}}}{{=}} xL2K𝑴t1𝑴t2𝑴tKP(x)Q(x)T𝑴1/sK𝑴1/s2𝑴1/s1ν(dx).\displaystyle\int x^{L-2K}{{\boldsymbol{M}}}_{t_{1}}{{\boldsymbol{M}}}_{t_{2}}\cdots{{\boldsymbol{M}}}_{t_{K}}\vec{P}(x)\,\vec{Q}(x)^{T}{{\boldsymbol{M}}}_{1/s_{K}}\cdots{{\boldsymbol{M}}}_{1/s_{2}}{{\boldsymbol{M}}}_{1/s_{1}}\nu(dx).

Therefore, the right hand side of (2.7) becomes

1𝜶,𝜷,LxL2K(V𝜶(z0)T𝑴t1𝑴t2𝑴tKP(x))(Q(x)T𝑴1sK𝑴1s2𝑴1s1W𝜷(z1))ν(dx)\tfrac{1}{\mathfrak{C}_{{{\boldsymbol{\alpha}}},{{\boldsymbol{\beta}}},L}}\,\int x^{L-2K}\left(\vec{V}_{{\boldsymbol{\alpha}}}(z_{0})^{T}{{\boldsymbol{M}}}_{t_{1}}{{\boldsymbol{M}}}_{t_{2}}\cdots{{\boldsymbol{M}}}_{t_{K}}\vec{P}(x)\right)\,\left(\vec{Q}(x)^{T}{{\boldsymbol{M}}}_{\frac{1}{s_{K}}}\cdots{{\boldsymbol{M}}}_{\frac{1}{s_{2}}}{{\boldsymbol{M}}}_{\frac{1}{s_{1}}}\vec{W}_{{\boldsymbol{\beta}}}(z_{1})\right)\nu(dx)

and the result follows from the symmetry of the dot product, uTv=vTu\vec{u}^{T}\vec{v}=\vec{v}^{T}\vec{u}. (Assumption (2.14) is responsible for Fubini’s theorem, which is used to switch the integral with the infinite sums in the last expression. This avoids formal integration of the product of infinite matrices.) ∎

We will need the following elementary lemma. (See also (Bryc and Wang,, 2023, Lemma A.1).)

Lemma 2.3.

Let μ\mu be a probability measure with Bsupp(μ)[A,B]B\in{\textnormal{supp}}(\mu)\subseteq[A,B] with |A|<B<|A|<B<\infty. Let FF be a real function which is bounded on supp(μ)\mathrm{supp}(\mu) and left-continuous at BB.

Then

(2.18) limLF(x)xLμ(dx)xLμ(dx)=F(B).\lim_{L\to\infty}\frac{\int F(x)x^{L}\mu(dx)}{\int x^{L}\mu(dx)}=F(B).
Proof.

Let XX be a random variable with the law μ\mu. First we will derive some estimates on the moments of XX and |X||X|. Fix ε>0\varepsilon>0 such that Bε>|A|B-\varepsilon>|A|. Since Bsupp(μ)B\in{\textnormal{supp}}(\mu), we have C:=(X>Bε/2)>0C:=\mathds{P}(X>B-\varepsilon/2)>0. And since supp(μ)[A,B]{\textnormal{supp}}(\mu)\subseteq[A,B] we have

𝔼[XL]=𝔼[XL𝟏{X>Bε/2}]+𝔼[XL𝟏{XBε/2}]C(Bε/2)L|A|L=(Bε/2)L(C|A|L(Bε/2)L).\mathds{E}[X^{L}]=\mathds{E}[X^{L}{\bf 1}_{\{X>B-\varepsilon/2\}}]+\mathds{E}[X^{L}{\bf 1}_{\{X\leq B-\varepsilon/2\}}]\geq C(B-\varepsilon/2)^{L}-|A|^{L}\\ =(B-\varepsilon/2)^{L}\left(C-\frac{|A|^{L}}{(B-\varepsilon/2)^{L}}\right).

The inequalities Bϵ/2>Bϵ>|A|B-\epsilon/2>B-\epsilon>|A| imply that the term |A|L/(Bϵ/2)L|A|^{L}/(B-\epsilon/2)^{L} converges to zero as L+L\to+\infty, therefore 𝔼[XL]>12C(Bε/2)L\mathds{E}[X^{L}]>\frac{1}{2}C(B-\varepsilon/2)^{L} for large enough LL\in{\mathbb{N}}. Next, we use the inequality |x|LxL+2|A|L|x|^{L}\leq x^{L}+2|A|^{L}, which is valid for all x[A,B]x\in[A,B] and LL\in{\mathbb{N}}, and obtain

𝔼[|X|L]𝔼[XL]1+2|A|L𝔼[XL]<1+4C|A|L(Bϵ/2)L,\frac{\mathds{E}[|X|^{L}]}{\mathds{E}[X^{L}]}\leq 1+2\frac{|A|^{L}}{\mathds{E}[X^{L}]}<1+\frac{4}{C}\frac{|A|^{L}}{(B-\epsilon/2)^{L}},

which gives us an upper bound 𝔼[|X|L]<2𝔼[XL]\mathds{E}[|X|^{L}]<2\mathds{E}[X^{L}], valid for LL large enough.

Now we are ready to establish (2.18). By taking FF(B)F-F(B) instead of FF in (2.18), without loss of generality, we can assume that F(B)=0F(B)=0 (clearly, FF(B)F-F(B) remains left-continuous at BB). Now we estimate, for LL\in{\mathbb{N}} large enough

|𝔼[F(X)XL]𝔼[XL]|\displaystyle\left|\frac{\mathds{E}[F(X)X^{L}]}{\mathds{E}[X^{L}]}\right| 𝔼[|F(X)||X|L𝟏{XBε}]𝔼[XL]+𝔼[|F(X)||X|L𝟏{X>Bε}]𝔼[XL]\displaystyle\leq\frac{\mathds{E}[|F(X)||X|^{L}{\bf 1}_{\{X\leq B-\varepsilon\}}]}{\mathds{E}[X^{L}]}+\frac{\mathds{E}[|F(X)||X|^{L}{\bf 1}_{\{X>B-\varepsilon\}}]}{\mathds{E}[X^{L}]}
2CF(Bε)L(Bε/2)L+2supx[Bε,B]|F(x)|,\displaystyle\leq\frac{2}{C}\|F\|_{\infty}\frac{(B-\varepsilon)^{L}}{(B-\varepsilon/2)^{L}}+2\sup_{x\in[B-\varepsilon,B]}|F(x)|,

where we used the estimates:

𝔼[|F(X)||X|L𝟏{XBε}]F(Bε)Land𝔼[XL]12C(Bε2)L\mathds{E}[|F(X)||X|^{L}{\bf 1}_{\{X\leq B-\varepsilon\}}]\leq\|F\|_{\infty}(B-\varepsilon)^{L}\quad\mbox{and}\quad\mathds{E}[X^{L}]\geq\tfrac{1}{2}C(B-\tfrac{\varepsilon}{2})^{L}

for the first term and

𝔼[|F(X)||X|L𝟏{X>Bε}]supx[Bε,B]|F(x)|×𝔼[|X|L]and𝔼[|X|L]<2𝔼[XL]\mathds{E}[|F(X)||X|^{L}{\bf 1}_{\{X>B-\varepsilon\}}]\leq\sup_{x\in[B-\varepsilon,B]}|F(x)|\times\mathds{E}[|X|^{L}]\quad\mbox{and}\quad\mathds{E}[|X|^{L}]<2\mathds{E}[X^{L}]

for the second term. Since (Bϵ)L/(Bϵ/2)L(B-\epsilon)^{L}/(B-\epsilon/2)^{L} converges to zero when L+L\to+\infty, we obtain

lim supL+|𝔼[F(X)XL]𝔼[XL]|2supx[Bε,B]|F(x)|.\limsup\limits_{L\to+\infty}\left|\frac{\mathds{E}[F(X)X^{L}]}{\mathds{E}[X^{L}]}\right|\leq 2\sup_{x\in[B-\varepsilon,B]}|F(x)|.

Taking the limit ϵ0+\epsilon\to 0^{+}, gives us the desired result (2.18) in the case F(B)=0F(B)=0. ∎

2.2. Limit theorem

The following result is a version of (Bryc and Wang,, 2023, Theorem 1.1) for non-constant weights of edges. It describes the joint limiting behavior of processes 𝜸(L){\boldsymbol{\gamma}}^{(L)} and 𝜸~(L)\widetilde{{\boldsymbol{\gamma}}}^{(L)} introduced in (2.4) and (2.5).

Theorem 2.4.

Suppose that supp(ν)\textnormal{supp}(\nu), the support of the orthogonality measure ν\nu of the polynomials (2.11), satisfies Bsupp(ν)[A,B]B\in{\textnormal{supp}}(\nu)\subseteq[A,B] with B>|A|B>|A| and that (A1)(A_{1}) and (A2)(A_{2}) from Assumption 2.1 hold. Then πn:=pn(B)>0\pi_{n}:=p_{n}(B)>0, n=0,1,n=0,1,\ldots and

(𝜸(L),𝜸~(L))(𝑿,𝒀)\left({\boldsymbol{\gamma}}^{(L)},\,\widetilde{{\boldsymbol{\gamma}}}^{(L)}\right)\Rightarrow\left({\boldsymbol{X}},{\boldsymbol{Y}}\right)

as LL\to\infty, where 𝐗={Xk}k0{\boldsymbol{X}}=\left\{X_{k}\right\}_{k\geq 0} and 𝐘={Yk}k0{\boldsymbol{Y}}=\left\{Y_{k}\right\}_{k\geq 0} are independent Markov chains with the same transition probabilities

(2.19) 𝖰n,m=1B{anπn+1πn,m=n+1,bn,m=n,cnπn1πn,m=n1,0,otherwise,\mathsf{Q}_{n,m}=\frac{1}{B}\begin{cases}a_{n}\frac{\pi_{n+1}}{\pi_{n}},&m=n+1,\\ b_{n},&m=n,\\ c_{n}\frac{\pi_{n-1}}{\pi_{n}},&m=n-1,\\ 0,&\mbox{otherwise},\end{cases}

with m,n=0,1,m,n=0,1,\ldots, and with the initial laws given by

(2.20) (X0=n)=1C𝜶αnπn,(Y0=n)=1C𝜷βnπ~n,n=0,1,,\mathds{P}(X_{0}=n)=\frac{1}{C_{{\boldsymbol{\alpha}}}}\alpha_{n}\pi_{n},\quad\mathds{P}(Y_{0}=n)=\frac{1}{C_{{\boldsymbol{\beta}}}}\beta_{n}\widetilde{\pi}_{n},\;n=0,1,\ldots,

where C𝛂C_{{\boldsymbol{\alpha}}}, C𝛃C_{{\boldsymbol{\beta}}} are the normalizing constants and π~n=p~n(B)\widetilde{\pi}_{n}=\widetilde{p}_{n}(B) with p~n\widetilde{p}_{n} given in (2.15).

Proof.

We first verify that transition probabilities (2.19) and (2.24) are well defined, i.e. πn>0\pi_{n}>0, n=0,1,n=0,1,\dots. Clearly, π0=1\pi_{0}=1. For n1n\geq 1, the polynomial pn(x)p_{n}(x) has leading term a01a11an11xna_{0}^{-1}a_{1}^{-1}\dots a_{n-1}^{-1}x^{n}, thus pn(x)+p_{n}(x)\to+\infty as x+x\to+\infty. If πn=pn(B)\pi_{n}=p_{n}(B) were negative, this would imply that pnp_{n} has a zero outside of the support of the measure of orthogonality, which is impossible, see e.g. Ismail, (2009). The case pn(B)=0p_{n}(B)=0 is also impossible, since the interlacing property of the zeroes of orthogonal polynomials would imply that pn+1p_{n+1} has a zero in the interval (B,)(B,\infty).

To determine the limit as LL\to\infty of the left hand side of (2.6), we re-write the right hand side of (2.17) as

xLν(dx)xL(V𝜶(1)TP(x))(W𝜷(1)TQ(x))ν(dx)xL2Kν(dx)xLν(dx)xL2KΨ0(x)Ψ1(x)ν(dx)xL2Kν(dx)\frac{\int x^{L}\nu(dx)}{\int x^{L}\left(\vec{V}_{{\boldsymbol{\alpha}}}(1)^{T}\vec{P}(x)\right)\left(\vec{W}_{{\boldsymbol{\beta}}}(1)^{T}\vec{Q}(x)\right)\nu(dx)}\cdot\frac{\int x^{L-2K}\nu(dx)}{\int x^{L}\nu(dx)}\cdot\frac{\int x^{L-2K}\Psi_{0}(x)\Psi_{1}(x)\nu(dx)}{\int x^{L-2K}\nu(dx)}

Using Lemma 2.3, each of the factors above converges as LL\to\infty and we get

(2.21) limL𝔼[z0γ0(L)j=1Ktjγj(L)γj1(L)j=1KsjγLj(L)γL+1j(L)z1γL(L)]=1(V𝜶(1)TP(1))(Q(1)TW𝜷(1))1B2KΨ0(B)Ψ1(B)=Ψ0(B)BKC𝜶Ψ1(B)BKC𝜷.\lim_{L\to\infty}\mathds{E}\left[z_{0}^{\gamma_{0}^{(L)}}\prod_{j=1}^{K}t_{j}^{\gamma_{j}^{(L)}-\gamma_{j-1}^{(L)}}\prod_{j=1}^{K}s_{j}^{\gamma_{L-j}^{(L)}-\gamma_{L+1-j}^{(L)}}z_{1}^{\gamma_{L}^{(L)}}\right]\\ =\frac{1}{\left(\vec{V}_{{\boldsymbol{\alpha}}}(1)^{T}\vec{P}(1)\right)\left(\vec{Q}(1)^{T}\vec{W}_{{\boldsymbol{\beta}}}(1)\right)}\cdot\frac{1}{B^{2K}}\cdot\Psi_{0}(B)\Psi_{1}(B)=\frac{\Psi_{0}(B)}{B^{K}C_{{\boldsymbol{\alpha}}}}\cdot\frac{\Psi_{1}(B)}{B^{K}C_{{\boldsymbol{\beta}}}}.

We will show that the latter expressions matches the product on the right hand side of (2.6).

To this end it suffices to prove

(2.22) 1BKV𝜶(z)T𝑴t1𝑴t2𝑴tKP(B)=n0αnπnzn𝔼(T1:K|X0=n),\frac{1}{B^{K}}\,\vec{V}_{{\boldsymbol{\alpha}}}(z)^{T}{{\boldsymbol{M}}}_{t_{1}}{{\boldsymbol{M}}}_{t_{2}}\cdots{{\boldsymbol{M}}}_{t_{K}}\vec{P}(B)=\sum_{n\geq 0}\alpha_{n}\pi_{n}z^{n}\,\mathds{E}(T_{1:K}|X_{0}=n),

where Ti:K=j=iKtjXjXj1T_{i:K}=\prod_{j=i}^{K}t_{j}^{X_{j}-X_{j-1}}, as well as

(2.23) 1BKW𝜷(z)T𝑴~s1𝑴~s2𝑴~sKQ(B)=n0βnπ~nzn𝔼(S1:K|Y0=n),\frac{1}{B^{K}}\,\vec{W}_{{\boldsymbol{\beta}}}(z)^{T}\widetilde{{{\boldsymbol{M}}}}_{s_{1}}\widetilde{{{\boldsymbol{M}}}}_{s_{2}}\cdots\widetilde{{{\boldsymbol{M}}}}_{s_{K}}\vec{Q}(B)=\sum_{n\geq 0}\beta_{n}\widetilde{\pi}_{n}z^{n}\,\mathds{E}(S_{1:K}|Y_{0}=n),

where Si:K=j=iKsjYjYj1S_{i:K}=\prod_{j=i}^{K}s_{j}^{Y_{j}-Y_{j-1}}, and

(2.24) (Yk+1=n+δ|Yk=n)=1B{cn+1π~n+1π~n,δ=1,bn,δ=0an1π~n1π~n,δ=1.\mathds{P}(Y_{k+1}=n+\delta|Y_{k}=n)=\frac{1}{B}\begin{cases}c_{n+1}\frac{\widetilde{\pi}_{n+1}}{\widetilde{\pi}_{n}},&\delta=1,\\ b_{n},&\delta=0\\ a_{n-1}\frac{\widetilde{\pi}_{n-1}}{\widetilde{\pi}_{n}},&\delta=-1.\end{cases}

Note that the processes 𝑿{\boldsymbol{X}} and 𝒀{\boldsymbol{Y}} have the same transition probabilities, since

π~n+1π~n=πn+1πnancn+1.\frac{\widetilde{\pi}_{n+1}}{\widetilde{\pi}_{n}}=\frac{\pi_{n+1}}{\pi_{n}}\frac{a_{n}}{c_{n+1}}.

We prove only (2.22) since the proof of (2.23) follows along the same lines. We use induction with respect to KK.

The case of K=0K=0 is immediate since the left hand side of (2.22) is V𝜶(z)TP(B)=n0αnπnzn\vec{V}_{{\boldsymbol{\alpha}}}(z)^{T}\vec{P}(B)=\sum_{n\geq 0}\,\alpha_{n}\pi_{n}\,z^{n}. For K>0K>0 we first note that the left hand side of (2.22) can be written as

RK:=1BK1Vα~(z,t1)(z)T𝑴t2𝑴tKP(B),R_{K}:=\frac{1}{B^{K-1}}\,\vec{V}_{\widetilde{\alpha}(z,t_{1})}(z)^{T}{{\boldsymbol{M}}}_{t_{2}}\cdots{{\boldsymbol{M}}}_{t_{K}}\vec{P}(B),

where α~n(z,t)=(αn1an1tz+αnbn+αn+1cn+1zt)/B\widetilde{\alpha}_{n}(z,t)=(\alpha_{n-1}a_{n-1}\frac{t}{z}+\alpha_{n}b_{n}+\alpha_{n+1}c_{n+1}\frac{z}{t})/B, n0n\geq 0. Consequently, by induction assumption for K1K-1 we get

RK=n0α~n(z,t1)πn𝔼(T2:K|X1=n)=n0(αn1an1t1z+αnbn+αn+1cn+1zt1)πnB𝔼(T2:K|X1=n)=n0αnπn(t1z𝔼(T2:K|X1=n+1)anπn+1Bπn+𝔼(T2:K|X1=n)bnB+zt1𝔼(T2:K|X1=n1)cnπn1Bπn)=n0αnπn𝔼((t1z)X1nT2:K|X0=n)=n0αnπnzn𝔼(T1:K|X0=n).R_{K}=\sum_{n\geq 0}\,\widetilde{\alpha}_{n}(z,t_{1})\pi_{n}\,\mathds{E}(T_{2:K}|X_{1}=n)\\ =\sum_{n\geq 0}\left(\alpha_{n-1}a_{n-1}\frac{t_{1}}{z}+\alpha_{n}b_{n}+\alpha_{n+1}c_{n+1}\frac{z}{t_{1}}\right)\frac{\pi_{n}}{B}\,\mathds{E}(T_{2:K}|X_{1}=n)\\ =\sum_{n\geq 0}\,\alpha_{n}\pi_{n}\Big{(}\tfrac{t_{1}}{z}\,\mathds{E}(T_{2:K}|X_{1}=n+1)\,\tfrac{a_{n}\pi_{n+1}}{B\pi_{n}}+\mathds{E}(T_{2:K}|X_{1}=n)\,\tfrac{b_{n}}{B}+\tfrac{z}{t_{1}}\,\mathds{E}(T_{2:K}|X_{1}=n-1)\,\tfrac{c_{n}\pi_{n-1}}{B\pi_{n}}\Big{)}\\ =\sum_{n\geq 0}\alpha_{n}\pi_{n}\,\mathds{E}\left(\left(\frac{t_{1}}{z}\right)^{X_{1}-n}T_{2:K}|X_{0}=n\right)=\sum_{n\geq 0}\alpha_{n}\pi_{n}z^{n}\,\mathds{E}(T_{1:K}|X_{0}=n).

Formula (2.21) shows that the limit processes 𝑿{\boldsymbol{X}} and 𝒀{\boldsymbol{Y}} are independent. This ends the proof. ∎

For proofs of local limit theorems we need to determine the limit of N(XNt=yN|X0=xN)\sqrt{N}\mathds{P}\left(X_{\left\lfloor Nt\right\rfloor}=y_{N}\middle|X_{0}=x_{N}\right) as NN\to\infty for suitably chosen sequences (yN)(y_{N}) and (xN)(x_{N}). The following formula is useful in computing such limits.

Proposition 2.5.

For a Markov process {Xk}\{X_{k}\}, we have

(2.25) (Xk=n|X0=m)=πnπm1Bk[A,B]xkpm(x)p~n(x)ν(dx).\mathds{P}(X_{k}=n|X_{0}=m)=\frac{\pi_{n}}{\pi_{m}}\frac{1}{B^{k}}\int_{[A,B]}x^{k}p_{m}(x)\widetilde{p}_{n}(x)\,\nu(dx).

(This resembles (Karlin and McGregor,, 1959, p. 67).)

Proof.

Recall notation (2.1). Fix 𝜸m,n(k){\boldsymbol{\gamma}}\in\mathcal{M}_{m,n}(k). Due to cancellations,

(X1=γ1,X2=γ2,,Xk=γk|X0=γ0)=w(𝜸)Bkπγkπγ0,\mathds{P}(X_{1}=\gamma_{1},X_{2}=\gamma_{2},\dots,X_{k}=\gamma_{k}|X_{0}=\gamma_{0})=\frac{w({\boldsymbol{\gamma}})}{B^{k}}\frac{\pi_{\gamma_{k}}}{\pi_{\gamma_{0}}},

so

(Xk=n|X0=m)=πnBkπm𝜸m,n(k)w(𝜸).\mathds{P}(X_{k}=n|X_{0}=m)=\frac{\pi_{n}}{B^{k}\pi_{m}}\sum_{{\boldsymbol{\gamma}}\in\mathcal{M}_{m,n}^{(k)}}w({\boldsymbol{\gamma}}).

By Viennot, (1985, (5)) (see also (Bryc and Wang,, 2023, Proposition 2.1)) we have

pm(x)pn(x)xLν(dx)=pn22γm,n(L)w(𝜸).\int_{\mathds{R}}p_{m}(x)p_{n}(x)x^{L}\,\nu(dx)=\|p_{n}\|_{2}^{2}\sum_{\gamma\in\mathcal{M}_{m,n}^{(L)}}w({\boldsymbol{\gamma}}).

We get

(Xk=n|X0=m)=πnBkπmpn22xkpm(x)pn(x)ν(dx).\mathds{P}(X_{k}=n|X_{0}=m)=\frac{\pi_{n}}{B^{k}\pi_{m}\|p_{n}\|_{2}^{2}}\int_{\mathds{R}}\,x^{k}p_{m}(x)p_{n}(x)\,\nu(dx).

3. Properties of the Al-Salam–Chihara polynomials

The general theory developed in Section 2 can be advanced further when it is specialized to the setting considered in Section 1. To proceed, we need to recall the definition and properties of the Al-Salam–Chihara polynomials.

We will be working with Al-Salam-Chihara polynomials {Qn}\{Q_{n}\} in real variable xx, defined by the three-step recursion

(3.1) 2xQn(x;a,b|q)=Qn+1(x;a,b|q)+(a+b)qnQn(x;a,b|q)+(1qn)(1abqn1)Qn1(x;a,b|q),n=0,1,,2xQ_{n}(x;a,b|q)\\ =Q_{n+1}(x;a,b|q)+(a+b)q^{n}Q_{n}(x;a,b|q)+(1-q^{n})(1-abq^{n-1})Q_{n-1}(x;a,b|q),\;n=0,1,\dots,

where parameters a,ba,b are either real or complex conjugate, |ab|<1|ab|<1, and q[0,1)q\in[0,1). As usual, we initialize the recursion with Q10Q_{-1}\equiv 0 and Q01Q_{0}\equiv 1.

It is known, see Ismail, (2009); Koekoek et al., (2010) or Koekoek and Swarttouw, (1994), that if |a|<1,|b|<1|a|<1,|b|<1, then the polynomials {Qn}\{Q_{n}\} are orthogonal with respect to the unique probability density supported on [1,1][-1,1] and given by

(3.2) g(x)=(q,ab;q)2π1x2|(e2iθ;q)|2|(aeiθ,beiθ;q)|2,x=cosθ.g(x)=\frac{(q,ab;q)_{\infty}}{2\pi\sqrt{1-x^{2}}}\frac{\left|(e^{2i\theta};q)_{\infty}\right|^{2}}{\left|(ae^{i\theta},be^{i\theta};q)_{\infty}\right|^{2}},\quad x=\cos\theta.

It is also known that the generating function of {Qn}\{Q_{n}\} is well defined for complex |t|<1|t|<1, and is given by the infinite product:

(3.3) n=0Qn(cosθ;a,b|q)(q;q)ntn=(at,bt;q)(eiθt,eiθt;q).\sum_{n=0}^{\infty}\frac{Q_{n}(\cos\theta;a,b|q)}{(q;q)_{n}}t^{n}=\frac{(at,bt;q)_{\infty}}{(e^{i\theta}t,e^{-i\theta}t;q)_{\infty}}.

Invoking the qq-binomial theorem, see e.g. (Gasper and Rahman,, 2004, (1.3.2)), the latter implies the following formula, which can also be recalculated from (Berg and Ismail,, 1996, p. 50) or (Ismail,, 2009, (15.1.12)):

(3.4) Qn(cosθ;a,b|q)(q;q)n=k=0n(beiθ;q)k(q;q)keikθ(aeiθ;q)nk(q;q)nkei(nk)θ.\frac{Q_{n}(\cos\theta;a,b|q)}{(q;q)_{n}}=\sum_{k=0}^{n}\frac{(be^{i\theta};q)_{k}}{(q;q)_{k}}e^{-ik\theta}\frac{(ae^{-i\theta};q)_{n-k}}{(q;q)_{n-k}}e^{i(n-k)\theta}.

In particular, at the boundary of the interval of orthogonality, we have

(3.5) Qn(1;a,b|q)(q;q)n=k=0n(a;q)k(b;q)nk(q;q)k(q;q)nk.\frac{Q_{n}(1;a,b|q)}{(q;q)_{n}}=\sum_{k=0}^{n}\frac{(a;q)_{k}(b;q)_{n-k}}{(q;q)_{k}(q;q)_{n-k}}.

3.1. Maximum of the Al-Salam–Chihara polynomials

Chebyshev polynomials UnU_{n} and TnT_{n} on the interval [1,1][-1,1] are bounded in absolute value by their value at x=1x=1. Similar result holds for the qq-Hermite polynomials, see (Ismail,, 2009, Theorem 13.1.2). For our proofs we need to extend this property to the Al-Salam–Chihara polynomials.

Proposition 3.1.

Let q[0,1)q\in[0,1), and a,ba,b be real or a complex conjugate pair such that 0ab10\leq ab\leq 1 and a+b0a+b\leq 0. Then for all x[1,1]x\in[-1,1] we have

(3.6) |Qn(x;a,b|q)|Qn(1;a,b|q).|Q_{n}(x;a,b|q)|\leq Q_{n}(1;a,b|q).
Remark 3.1.

A similar bound seems to be implied by (Ismail,, 2009, (15.1.4)) without explicit assumption that a+b0a+b\leq 0. We note that (3.6) does not hold for arbitrary a,ba,b:

Q2(cosθ;qeiα,qeiα;q)=2q3cos(2α)+4(q+1)qcos(α)cos(u)+2cos(2u)+[4]q.Q_{2}(\cos\theta;-qe^{i\alpha},-qe^{-i\alpha};q)=2q^{3}\cos(2\alpha)+4(q+1)q\cos(\alpha)\cos(u)+2\cos(2u)+[4]_{q}.

In particular, Q2(cosθ;q,q;q)=3q3+q24(q+1)qcos(θ)+q+2cos(2θ)+1Q_{2}(\cos\theta;q,q;q)=3q^{3}+q^{2}-4(q+1)q\cos(\theta)+q+2\cos(2\theta)+1 has a strict maximum at θ=π\theta=\pi which exceeds the value at θ=0\theta=0 by 8q(1+q)8q(1+q).

Corollary 3.1.

Let q[0,1)q\in[0,1), r[0,1]r\in[0,1], α[π/2,π/2]\alpha\in[-\pi/2,\pi/2]. Denote a=reiαa=-re^{i\alpha} and b=reiαb=-re^{-i\alpha}. Then for all x[1,1]x\in[-1,1] bound (3.6) holds.

Proof.

Indeed, a+b=rcosα0a+b=-r\cos\alpha\leq 0 and ab=r2[0,1]ab=r^{2}\in[0,1]

Proof of Proposition 3.1.

We will write QnQ_{n} as a trigonometric polynomial

Qn(x;a,b|q)=m=0na(n,m)Tm(x),Q_{n}(x;a,b|q)=\sum\limits_{m=0}^{n}a(n,m)T_{m}(x),

where Tm(cosθ)=cos(mθ)T_{m}(\cos\theta)=\cos(m\theta) is a Chebyshev polynomial of the first kind, m=0,1,m=0,1,\dots. We are going to show that a(n,m)0a(n,m)\geq 0 for all 0mn0\leq m\leq n. Since for x=cosθ[1,1]x=\cos\theta\in[-1,1] we have |Tm(x)|1=Tm(1)|T_{m}(x)|\leq 1=T_{m}(1), this will imply the bound (3.6).

To use (Szwarc,, 1992, Theorem 1), we rewrite the recursion for the Al-Salam–Chihara polynomials into his notation as

xQn(x)=γnQn+1(x)+βnQn(x)+αnQn1(x),xQ_{n}(x)=\gamma_{n}Q_{n+1}(x)+\beta_{n}Q_{n}(x)+\alpha_{n}Q_{n-1}(x),

where

γn=1/2,βn=a+b2qn,αn=(1qn)(1abqn1)2.\gamma_{n}=1/2,\quad\beta_{n}=\frac{a+b}{2}q^{n},\quad\alpha_{n}=\frac{(1-q^{n})(1-abq^{n-1})}{2}.

The Chebyshev polynomials satisfy recursion

xTn(x)=γnTn+1(x)+βnTn(x)+αnTn1(x),xT_{n}(x)=\gamma_{n}^{\prime}T_{n+1}(x)+\beta_{n}^{\prime}T_{n}(x)+\alpha_{n}^{\prime}T_{n-1}(x),

where

γn=12(1+δn=0),βn=0,αn=12δn>0.\gamma_{n}^{\prime}=\frac{1}{2}(1+\delta_{n=0}),\quad\beta_{n}^{\prime}=0,\quad\alpha_{n}^{\prime}=\frac{1}{2}\delta_{n>0}.

It is then clear that the assumptions of (Szwarc,, 1992, Theorem 1) are satisfied. Thus a(m,n)0a(m,n)\geq 0 for all 0mn0\leq m\leq n, n=0,1,n=0,1,\dots, ending the proof. ∎

3.2. Pointwise asymptotics near the end of the interval of orthogonality

We will need pointwise asymptotics for the Al-Salam–Chihara polynomials at the upper endpoint of the orthogonality interval. Such pointwise limits have been studied for orthogonal polynomials both on finite intervals and on unbounded domains Aptekarev, (1993); Baik et al., (2003); Deift et al., (2001, 1999); Ismail and Li, (2013); Ismail, (1986, 2005); Ismail and Wilson, (1982); Ismail et al., (2022); Kuijlaars et al., (2004); Lubinsky, (2020); Nevai, (1984). In particular, (Aptekarev,, 1993, Theorem 1) gives general conditions on the orthogonality measure and on the three-step recursion of the orthonormal polynomials {qn(x)}\{q_{n}(x)\} which imply

(3.7) limn1nα+1/2qn(1u22n2)=Jα(u)uα\lim_{n\to\infty}\frac{1}{n^{\alpha+1/2}}\,q_{n}\left(1-\tfrac{u^{2}}{2n^{2}}\right)=\frac{J_{\alpha}(u)}{u^{\alpha}}

uniformly on compact subsets of complex plane, where JαJ_{\alpha} is the Bessel function.

Lubinsky (Lubinsky,, 2020, Theorem 1.1) proves a version of (3.7) which can be re-written as

(3.8) limnqn(1u22n2)qn(1)=2αΓ(α+1)Jα(u)uα\lim_{n\to\infty}\frac{q_{n}\left(1-\frac{u^{2}}{2n^{2}}\right)}{q_{n}(1)}=2^{\alpha}\Gamma(\alpha+1)\frac{J_{\alpha}(u)}{u^{\alpha}}

uniformly on compact subsets of complex plane. For our density function (3.2) the assumptions in (Lubinsky,, 2020, Theorem 1.1) are satisfied with α=1/2\alpha=1/2. Indeed, one can re-write (3.2) as

(3.9) g(x)=21x2(q,ab;q)|(qe2iθ;q)|2π|(aeiθ,beiθ;q)|2,x=cosθ.g(x)=\frac{2\sqrt{1-x^{2}}(q,ab;q)_{\infty}\left|(qe^{2i\theta};q)_{\infty}\right|^{2}}{\pi\left|(ae^{i\theta},be^{i\theta};q)_{\infty}\right|^{2}},\;x=\cos\theta.

Note that for α=1/2\alpha=1/2, the right hand side of (3.8) simplifies to u1sinuu^{-1}\sin u and that the limit under normalization (3.7) is discussed in (Lubinsky,, 2020, Theorem 1.3).

Pointwise asymptotics for the Al-Salam–Chihara polynomials at the endpoints of the interval of orthogonality follows from the above results. However, our proof is quite direct, and the approach extends to the case of varying orthogonality measures, so we include it for completeness. Our asymptotic result for the Al-Salam-Chihara polynomials with varying orthogonality measure as q1q\nearrow 1 seem to be new and does not follow from Levin and Lubinsky, (2020).

The following asymptotics holds for fixed q<1q<1.

Theorem 3.2.

Fix a,ba,b real or complex conjugate with |a|,|b|<1|a|,|b|<1, q[0,1)q\in[0,1) and u>0u>0. Let {Qn}\{Q_{n}\} be the Al-Salam–Chihara polynomials (3.1). If uMuu_{M}\to u, then

(3.10) limM1MQM(1uM22M2;a,b|q)=sinuu(a,b;q)(q;q).\lim_{M\to\infty}\tfrac{1}{M}\,Q_{M}\left(1-\tfrac{u_{M}^{2}}{2M^{2}};a,b|q\right)=\frac{\sin u}{u}\;\frac{(a,b;q)_{\infty}}{(q;q)_{\infty}}.
Remark 3.2.

We will also need the case u=0u=0, with

(3.11) limn1nQn(1;a,b|q)=(a,b;q)(q;q).\lim_{n\to\infty}\frac{1}{n}Q_{n}(1;a,b|q)=\frac{(a,b;q)_{\infty}}{(q;q)_{\infty}}.

We also note the following extension of bound (3.6): there is a constant C=C(a,b,q)C=C(a,b,q) such that for all n=0,1n=0,1\dots and x[1,1]x\in[-1,1] we have

(3.12) |Qn(x;a,b|q)|C(n+1).|Q_{n}(x;a,b|q)|\leq C(n+1).
Proof of Theorem 3.2.

The use of MM in (3.10) is solely to avoid notation clash when we use it in the proof of Theorem 4.2, so for the proof, we replace MM by nn.

The series (3.3) converges for all |t|<1|t|<1, thus we can write

(3.13) Qn(x;a,b|q)(q;q)n=12πi|t|=rtn1(at,bt;q)(eiθt,eiθt;q)𝑑t,\frac{Q_{n}(x;a,b|q)}{(q;q)_{n}}=\frac{1}{2\pi{\textnormal{i}}}\oint_{|t|=r}t^{-n-1}\frac{(at,bt;q)_{\infty}}{(e^{{\textnormal{i}}\theta}t,e^{-{\textnormal{i}}\theta}t;q)_{\infty}}dt,

where r(0,1)r\in(0,1) and the integration is done in the counter-clockwise direction. We fix now x=cosθ(1,1)x=\cos\theta\in(-1,1), so that θ(0,π)\theta\in(0,\pi) and we note that the integrand

F(t):=tn1(at,bt;q)(eiθt,eiθt;q)=tn1(at,bt;q)(eiθtq,eiθtq;q)×1(1eiθt)(1eiθt)F(t):=t^{-n-1}\frac{(at,bt;q)_{\infty}}{(e^{{\textnormal{i}}\theta}t,e^{-{\textnormal{i}}\theta}t;q)_{\infty}}=t^{-n-1}\frac{(at,bt;q)_{\infty}}{(e^{{\textnormal{i}}\theta}tq,e^{-{\textnormal{i}}\theta}tq;q)_{\infty}}\times\frac{1}{(1-e^{{\textnormal{i}}\theta}t)(1-e^{-{\textnormal{i}}\theta}t)}

has two simple poles at t=e±iθt=e^{\pm{\textnormal{i}}\theta} inside the disk |t|<q1|t|<q^{-1}. By the Cauchy Residue Theorem we can write

(3.14) Qn(x;a,b|q)(q;q)n=Res(F;eiθ)Res(F;eiθ)+12πi|t|=Rtn1(at,bt;q)(eiθt,eiθt;q)𝑑t,\frac{Q_{n}(x;a,b|q)}{(q;q)_{n}}=-{\textnormal{Res}}(F;e^{{\textnormal{i}}\theta})-{\textnormal{Res}}(F;e^{-{\textnormal{i}}\theta})+\frac{1}{2\pi i}\oint_{|t|=R}t^{-n-1}\frac{(at,bt;q)_{\infty}}{(e^{{\textnormal{i}}\theta}t,e^{-{\textnormal{i}}\theta}t;q)_{\infty}}dt,

where RR is in the interval (1,q1)(1,q^{-1}).

Recalling that |1|z|||1z|1+|z|\left|1-|z|\right|\leq|1-z|\leq 1+|z| and |t|=R>1|t|=R>1 we get

|(at,bt;q)(eiθt,eiθt;q)|(|a|R,|b|R;q)(R;q)2.\left|\frac{(at,bt;q)_{\infty}}{(e^{{\textnormal{i}}\theta}t,e^{-{\textnormal{i}}\theta}t;q)_{\infty}}\right|\leq\frac{(-|a|R,-|b|R;q)_{\infty}}{(R;\,q)^{2}_{\infty}}.

Therefore, the third term in (3.14) is bounded by

(|a|R,|b|R;q)(R;q)2Rnn0.\frac{(-|a|R,-|b|R;q)_{\infty}}{(R;\,q)^{2}_{\infty}\,R^{n}}\stackrel{{\scriptstyle n\to\infty}}{{\longrightarrow}}0.

Since under our assumptions θ=θn\theta=\theta_{n} with cosθn=1un2/(2n2)\cos\theta_{n}=1-u_{n}^{2}/(2n^{2}) is such that nθnun\theta_{n}\to u, we see that

1nRes(F;e±iθn)=ei(n+1)θnn(e±iθneiθn)(ae±iθn,be±iθn;q)(e±2iθnq;q)(q;q)neiu(a,b;q)±2iu(q;q)2.\frac{1}{n}{\textnormal{Res}}(F;e^{\pm{\textnormal{i}}\theta_{n}})=\frac{e^{\mp{\textnormal{i}}(n+1)\theta_{n}}}{n(e^{\pm{\textnormal{i}}\theta_{n}}-e^{\mp{\textnormal{i}}\theta_{n}})}\frac{(ae^{\pm{\textnormal{i}}\theta_{n}},be^{\pm{\textnormal{i}}\theta_{n}};q)_{\infty}}{(e^{\pm 2{\textnormal{i}}\theta_{n}}q;q)_{\infty}(q;q)_{\infty}}\stackrel{{\scriptstyle n\to\infty}}{{\longrightarrow}}\frac{e^{\mp iu}(a,b;q)_{\infty}}{\pm 2{\textnormal{i}}u\,(q;\,q)_{\infty}^{2}}.

Consequently, returning to (3.14) we get

1nQn(1un22n2;a,b|q)n(eiueiu)(a,b;q)2iu(q;q).\tfrac{1}{n}\,Q_{n}\left(1-\frac{u_{n}^{2}}{2n^{2}};a,b\middle|q\right)\stackrel{{\scriptstyle n\to\infty}}{{\longrightarrow}}\frac{(e^{iu}-e^{-iu})(a,b;q)_{\infty}}{2{\textnormal{i}}u(q;q)_{\infty}}.

Proof of Remark 3.2.

The limit (3.11) follows from (3.5) and an elementary lemma that if αnα\alpha_{n}\to\alpha, βnβ\beta_{n}\to\beta then n1k=1nαnβnkαβn^{-1}\sum_{k=1}^{n}\alpha_{n}\beta_{n-k}\to\alpha\beta. (To verify the latter, write 1nk=1nαkβnk=1nk=1n(αkα)βnk+αnk=1nβk0+αβ\tfrac{1}{n}\sum_{k=1}^{n}\alpha_{k}\beta_{n-k}=\tfrac{1}{n}\sum_{k=1}^{n}(\alpha_{k}-\alpha)\beta_{n-k}+\frac{\alpha}{n}\sum_{k=1}^{n}\beta_{k}\to 0+\alpha\beta.) To prove (3.12) we use (3.4), which by triangle inequality gives a bound with explicit constant:

supx[1,1]|Qn(x;a,b|q)|(|a|,|b|;q)n(q;q)n(n+1).\sup_{x\in[-1,1]}|Q_{n}(x;a,b|q)|\leq\frac{(-|a|,-|b|;q)_{n}}{(q;q)_{n}}(n+1).

The following asymptotics holds for q1q\nearrow 1. The statement is somewhat cumbersome as we will need to apply this result to a,ba,b that vary with qq.

Theorem 3.3.

If a~,b~\widetilde{a},\widetilde{b} are real in [0,)[0,\infty), or a complex conjugate pair with Re(a~)0\textnormal{Re}(\widetilde{a})\geq 0, then with m=Mx+Mlog(M(1+a~)(1+b~))m=\left\lfloor Mx\right\rfloor+\left\lfloor M\log\left(M\sqrt{(1+\widetilde{a})(1+\widetilde{b})}\right)\right\rfloor, qM=e2/Mq_{M}=e^{-2/M} and aM=qMa~a_{M}=-q_{M}\widetilde{a}, bM:=qMb~b_{M}:=-q_{M}\widetilde{b} we have

(3.15) limM(qM;qM)2M(aM,bM;qM)Qm(cos(u/M);aM,bM|qM)(qM;qM)m=Ki|u|(ex),u.\lim_{M\to\infty}\frac{(q_{M};q_{M})_{\infty}^{2}}{M(a_{M},b_{M};q_{M})_{\infty}}\;\frac{Q_{m}(\cos(u/M);a_{M},b_{M}|q_{M})}{(q_{M};q_{M})_{m}}=K_{{\textnormal{i}}|u|}(e^{-x}),\;u\in\mathds{R}.
Proof.

The proof relies on two technical estimates that we will prove in Section 5. In the proof, for simplicity of notation we suppress the dependence of qq on MM. Fix uu\in{\mathbb{R}} and let a=qa~a=-q\widetilde{a}, b=qb~b=-q\widetilde{b}, where q=e2/Mq=e^{-2/M}.

In view of (3.13) with r=qr=q, we can write

(3.16) Qm(cos(u/M);a,b|q)(q;q)m=12πi|t|=q(at,bt;q)(qiu/2t,qiu/2t;q)tm1𝑑t,\frac{Q_{m}(\cos(u/M);a,b|q)}{(q;q)_{m}}=\frac{1}{2\pi{\textnormal{i}}}\oint_{|t|=q}\frac{(at,bt;q)_{\infty}}{(q^{{\textnormal{i}}u/2}t,q^{-{\textnormal{i}}u/2}t;q)_{\infty}}t^{-m-1}dt,

where we are integrating counterclockwise along a circle of radius q<1q<1. Next we change the variable of integration t=qzt=q^{z}, so that dt=ln(q)qzdz=(2/M)tdzdt=\ln(q)q^{z}dz=-(2/M)tdz, and obtain

(3.17) Qm(cos(u/M);a,b|q)(q;q)m=1Mπi1Mπi/21+Mπi/2(aqz,bqz;q)(qiu/2+z,qiu/2+z;q)e2mz/M𝑑z.\frac{Q_{m}(\cos(u/M);a,b|q)}{(q;q)_{m}}=\frac{1}{M\pi{\textnormal{i}}}\int_{1-M\pi{\textnormal{i}}/2}^{1+M\pi{\textnormal{i}}/2}\frac{(aq^{z},bq^{z};q)_{\infty}}{(q^{{\textnormal{i}}u/2+z},q^{-{\textnormal{i}}u/2+z};q)_{\infty}}e^{2mz/M}dz.

From (3.17), in view of (5.2), we get

(3.18) (q;q)2M(a,b;q)Qm(cos(u/M);a,b|q)(q;q)m=1πi1Mπi/21+Mπi/2Γq(iu/2+z)Γq(iu/2+z)M2(1q)22z(a~q1+z,b~q1+z;q)(a~q,b~q;q)e2mz/M𝑑z=1πi1i1+ifM(z)𝑑z,\frac{(q;q)_{\infty}^{2}}{M(a,b;q)_{\infty}}\;\frac{Q_{m}(\cos(u/M);a,b|q)}{(q;q)_{m}}\\ =\frac{1}{\pi{\textnormal{i}}}\int_{1-M\pi{\textnormal{i}}/2}^{1+M\pi{\textnormal{i}}/2}\tfrac{\Gamma_{q}({\textnormal{i}}u/2+z)\Gamma_{q}(-{\textnormal{i}}u/2+z)}{M^{2}(1-q)^{2-2z}}\;\tfrac{(-\widetilde{a}q^{1+z},-\widetilde{b}q^{1+z};q)_{\infty}}{(-\widetilde{a}q,-\widetilde{b}q;q)_{\infty}}\;e^{2mz/M}dz=\frac{1}{\pi{\textnormal{i}}}\int_{1-{\textnormal{i}}\infty}^{1+{\textnormal{i}}\infty}f_{M}(z)dz,

where

(3.19) fM(z)=𝟏{|Im(z)|Mπ/2}Γq(iu/2+z)Γq(iu/2+z)(a~q1+z,b~q1+z;q)M22z(1q)22z(a~q,b~q;q)e2mz/M2logM.f_{M}(z)={\mathbf{1}}_{\{|\textnormal{Im}(z)|\leq M\pi/2\}}\frac{\Gamma_{q}({\textnormal{i}}u/2+z)\Gamma_{q}(-{\textnormal{i}}u/2+z)(-\widetilde{a}q^{1+z},-\widetilde{b}q^{1+z};q)_{\infty}}{M^{2-2z}(1-q)^{2-2z}(-\widetilde{a}q,-\widetilde{b}q;q)_{\infty}}\;e^{2mz/M-2\log M}.

We now write z=1+isz=1+{\textnormal{i}}s. In order to use the dominated convergence theorem, we verify that for every x,ux,u\in\mathds{R} (recall that mm depends on xx) there are constants A,B,MA,B,M_{*} such that for all M>MM>M_{*} and all real ss we have

(3.20) |fM(1+is)|AeB|s|.|f_{M}(1+{\textnormal{i}}s)|\leq Ae^{-B|s|}.

To verify (3.20) we write |fM(1+is)|=f(1)f(2)f(3)|f_{M}(1+{\textnormal{i}}s)|=f^{(1)}\cdot f^{(2)}\cdot f^{(3)}, where

f(1)\displaystyle f^{(1)} =𝟏{|s|Mπ/2}|Γq(1+is+iu/2)Γq(1+isiu/2)|,\displaystyle={\mathbf{1}}_{\{|s|\leq M\pi/2\}}|\Gamma_{q}(1+{\textnormal{i}}s+{\textnormal{i}}u/2)\Gamma_{q}(1+{\textnormal{i}}s-{\textnormal{i}}u/2)|,
f(2)\displaystyle f^{(2)} =|(a~q2+is,b~q2+is;q)||(a~q,b~q;q)|,\displaystyle=\frac{|(-\widetilde{a}q^{2+{\textnormal{i}}s},-\widetilde{b}q^{2+{\textnormal{i}}s};q)_{\infty}|}{|(-\widetilde{a}q,-\widetilde{b}q;q)_{\infty}|},
f(3)\displaystyle f^{(3)} =|e2(1+is)m/MM2(1q)22(1+is)|=e2(m/MlogM).\displaystyle=\left|\frac{e^{2(1+{\textnormal{i}}s)m/M}}{M^{2}(1-q)^{2-2(1+{\textnormal{i}}s)}}\right|=e^{2(m/M-\log M)}.

From (5.24) we see that there exist constants A,B,M>0A,B,M_{*}>0 such that f(1)AeB|s|f^{(1)}\leq Ae^{-B|s|} for all real ss. From (5.26) we see that f(2)1f^{(2)}\leq 1. With

mM=Mx+Mlog(M(1+a~)(1+b~))Mx+log(1+a~)(1+b~)+logM,\frac{m}{M}=\frac{\left\lfloor Mx\right\rfloor+\left\lfloor M\log(M\sqrt{(1+\widetilde{a})(1+\widetilde{b})})\right\rfloor}{M}\leq x+\log\,\sqrt{(1+\widetilde{a})(1+\widetilde{b})}+\log\,M,

we see that

f(3)e2(x+log(1+a~)(1+b~)).f^{(3)}\leq e^{2\left(x+\log\sqrt{(1+\widetilde{a})(1+\widetilde{b})}\right)}.

Thus (3.20) holds.

We can now apply the dominated convergence theorem and pass to the limit inside the integral (3.18). We fix z=1+isz=1+{\textnormal{i}}s and compute limMfM(z)\lim_{M\to\infty}f_{M}(z) by computing the limit for the factors in (3.19).

Clearly, 𝟏{|Im(z)|<Mπ/2}1{\mathbf{1}}_{\{|\textnormal{Im}(z)|<M\pi/2\}}\to 1 as MM\to\infty and by (5.3) we get

limq1Γq(iu/2+z)Γq(iu/2+z)=Γ(iu/2+z)Γ(iu/2+z).\lim_{q\to 1^{-}}\Gamma_{q}({\textnormal{i}}u/2+z)\Gamma_{q}(-{\textnormal{i}}u/2+z)=\Gamma({\textnormal{i}}u/2+z)\Gamma(-{\textnormal{i}}u/2+z).

Since limMM(1qM)=2\lim_{M\to\infty}\,M(1-q_{M})=2, from (5.1) we get

limM(a~qM1+z,b~qM1+z;qM)M22z(1qM)22z(a~qM,b~qM;qM)=1422z((1+a~)(1+b~))z.\lim_{M\to\infty}\frac{(-\widetilde{a}q_{M}^{1+z},-\widetilde{b}q_{M}^{1+z};q_{M})_{\infty}}{M^{2-2z}(1-q_{M})^{2-2z}(-\widetilde{a}q_{M},-\widetilde{b}q_{M};q_{M})_{\infty}}=\tfrac{1}{4}\frac{2^{2z}}{\left((1+\widetilde{a})(1+\widetilde{b})\right)^{z}}.

Finally, since

mM=x+logM+log(1+a~)(1+b~)+O(1M),\frac{m}{M}=x+\log M+\log\sqrt{(1+\widetilde{a})(1+\widetilde{b})}+O(\tfrac{1}{M}),

we get

e2mz/M2zlogMe2zx+2zlogM+2zlog(1+a~)(1+b~)2zlogM=((1+a~)(1+b~))ze2zx.e^{2mz/M-2z\log M}\sim e^{2zx+2z\log M+2z\log\sqrt{(1+\widetilde{a})(1+\widetilde{b})}-2z\log M}=\left((1+\widetilde{a})(1+\widetilde{b})\right)^{z}e^{2zx}.

Putting all the factors together,

limMfM(z)=14Γ(iu/2+z)Γ(iu/2+z)22ze2zx.\lim_{M\to\infty}f_{M}(z)=\frac{1}{4}\Gamma({\textnormal{i}}u/2+z)\Gamma(-{\textnormal{i}}u/2+z)2^{2z}e^{2zx}.

Thus by the dominated convergence theorem,

limM1πi1i1+ifM(z)𝑑z=14πi1i1+iΓ(iu/2+z)Γ(iu/2+z)22ze2zx𝑑z.\lim_{M\to\infty}\frac{1}{\pi{\textnormal{i}}}\int_{1-{\textnormal{i}}\infty}^{1+{\textnormal{i}}\infty}f_{M}(z)dz=\frac{1}{4\pi{\textnormal{i}}}\int_{1-{\textnormal{i}}\infty}^{1+{\textnormal{i}}\infty}\Gamma({\textnormal{i}}u/2+z)\Gamma(-{\textnormal{i}}u/2+z)2^{2z}e^{2zx}dz.

This completes the proof by the Mellin-Barnes type formula for the Bessel K function:

Kiu(ex)=14πicic+iΓ(s+iu/2)Γ(siu/2)22se2xs𝑑s,K_{{\textnormal{i}}u}(e^{-x})=\frac{1}{4\pi{\textnormal{i}}}\int_{c-{\textnormal{i}}\infty}^{c+{\textnormal{i}}\infty}\Gamma(s+{\textnormal{i}}{u}/{2})\Gamma(s-{\textnormal{i}}{u}/{2})2^{2s}e^{2xs}ds,

where c>0c>0, u,xu,x\in\mathds{R}. (This is (Olver et al.,, 2023, 10.32.13) [10.32.13] https://dlmf.nist.gov/10.32.E13 with ν=iu\nu={\textnormal{i}}u, z=ex>0z=e^{-x}>0, and shifted variable of integration t=s+iu/2t=s+{\textnormal{i}}u/2.) ∎

4. Motzkin paths with qq-number weights

We now return to the setting from Section 1, providing additional details and further results for the case of edge weights an=[n+2]qa_{n}=[n+2]_{q}, bn=2σ[n+1]qb_{n}=2\sigma[n+1]_{q}, cn=[n]qc_{n}=[n]_{q}.

We note that with the above choice of the edge weights, probability measure (1.1) is the same as (2.3) with the boundary weights αn=ρ0n[n+1]q\alpha_{n}=\rho_{0}^{n}[n+1]_{q} and βn=ρ1n\beta_{n}=\rho_{1}^{n}. Indeed, we note that for an upward step we have aγk1=[γk1+2]q=[γk+1]qa_{\gamma_{k-1}}=[\gamma_{k-1}+2]_{q}=[\gamma_{k}+1]_{q}, for a horizontal step we have bγk1=2σ[γk1+1]q=2σ[γk+1]qb_{\gamma_{k-1}}=2\sigma[\gamma_{k-1}+1]_{q}=2\sigma[\gamma_{k}+1]_{q} and for a downward step we have cγk1=[γk1]q=[γk+1]qc_{\gamma_{k-1}}=[\gamma_{k-1}]_{q}=[\gamma_{k}+1]_{q}. So formula (2.1) gives w(𝜸)=(2σ)H(𝜸)k=1L[γk+1]qw({\boldsymbol{\gamma}})=(2\sigma)^{H({\boldsymbol{\gamma}})}\prod_{k=1}^{L}[\gamma_{k}+1]_{q} and thus (2.3) gives (1.1).

Recursion (2.11) becomes

(4.1) xpn(x)=[n+2]qpn+1(x)+2σ[n+1]qpn(x)+[n]qpn1(x).xp_{n}(x)=[n+2]_{q}p_{n+1}(x)+2\sigma[n+1]_{q}p_{n}(x)+[n]_{q}p_{n-1}(x).

Then (2.12) gives

(4.2) pn22=k=1n[k]q[k+1]q=1[n+1]q, hence p~n(x)=[n+1]qpn(x).\|p_{n}\|_{2}^{2}=\prod_{k=1}^{n}\frac{[k]_{q}}{[k+1]_{q}}=\frac{1}{[n+1]_{q}},\mbox{ hence $\widetilde{p}_{n}(x)=[n+1]_{q}p_{n}(x)$.}
Lemma 4.1.

Suppose 0σ10\leq\sigma\leq 1. Polynomials {pn}\{p_{n}\} are orthogonal with respect to the density supported on the interval [A,B][A,B] with

(4.3) A=21σ1q,B=21+σ1q.A=-2\frac{1-\sigma}{1-q},\quad B=2\frac{1+\sigma}{1-q}.

Moreover,

(4.4) πn=pn(B)=1[n+1]qk=0n(a;q)k(q;q)k(b;q)nk(q;q)nk,\pi_{n}=p_{n}(B)=\frac{1}{[n+1]_{q}}\sum_{k=0}^{n}\frac{(a;q)_{k}}{(q;q)_{k}}\;\frac{(b;q)_{n-k}}{(q;q)_{n-k}},

where

(4.5) a=q(σ+i1σ2),b=q(σi1σ2).a=-{q}\left(\sigma+i\sqrt{1-\sigma^{2}}\right),\;b=-{q}\left(\sigma-i\sqrt{1-\sigma^{2}}\right).

We note that a,ba,b are the solutions of the system of equations

(4.6) a+b=2σq,ab=q2.a+b=-2\sigma q,\quad ab=q^{2}.
Proof.

To determine the orthogonality measure as well as the sequence πn=pn(B)\pi_{n}=p_{n}(B), we establish a connection between polynomials {pn}\{p_{n}\} and monic Al-Salam-Chihara polynomials {Qn}\{Q_{n}\}, defined by the three step recursion (3.1) with complex conjugate parameters (4.5).

It is straightforward to verify that if {pn}\{p_{n}\} satisfy recursion (4.1) then polynomials

Qn(x;a,b|q):=(1q)n[n+1]q!pn(2x+σ1q),n=0,1,Q_{n}(x;a,b|q):=\left(1-q\right)^{n}[n+1]_{q}!\;p_{n}\left(2\;\frac{x+\sigma}{1-q}\right),\;n=0,1,\dots

satisfy recursion (3.1) with parameters (4.5). Conversely, with y=2x+σ1qy=2\frac{x+\sigma}{1-q}, we have

(4.7) pn(y)=1[n+1]q(q,q)nQn(x;a,b|q).p_{n}(y)=\frac{1}{[n+1]_{q}(q,q)_{n}}Q_{n}(x;a,b|q).

Therefore, with gg defined by (3.2) (recall, that gg is supported on [1,1][-1,1]), polynomials {pn}\{p_{n}\} are orthogonal with respect to the weight function (probability density function)

(4.8) 1q2g((1q)x2σ)\frac{1-q}{2}\;g\left(\tfrac{(1-q)x}{2}-\sigma\right)

which is supported on the interval [A,B][A,B] given by (4.3). Combining (3.5) and (4.7) we see that

(4.9) πn=pn(B)=Qn(1;a,b|q)[n+1]q(q;q)n\pi_{n}=p_{n}(B)=\frac{Q_{n}(1;a,b|q)}{[n+1]_{q}(q;q)_{n}}

is given by (4.4), see (3.5). ∎

With the above preparations, we can now prove Theorem 1.1.

Proof of Theorem 1.1.

We apply Theorem 2.4. With αn=ρ0n[n+1]q\alpha_{n}=\rho_{0}^{n}[n+1]_{q} and βn=ρ1n\beta_{n}=\rho_{1}^{n}, conditions (A1)(A_{1}) and (A2)(A_{2}) from Assumption 2.1 hold by (4.7), (4.2) and (3.12). The assumptions on the support of the orthogonality measure follow from Lemma 4.1. The probabilities πn=sn[n+1]q\pi_{n}=\frac{s_{n}}{[n+1]_{q}} are given by (4.4), so the transition matrix is just a recalculation of (2.19). Recalling that π~n=[n+1]qπn\widetilde{\pi}_{n}=[n+1]_{q}\pi_{n}, see (4.2), the initial laws (1.2) arise from (2.20). Finally, the formulas for the normalization constants are calculated from the relation (4.7) by applying (3.3) with t=ρjt=\rho_{j} and θ=0\theta=0. ∎

4.1. Local limit theorems

We will deduce Theorem 1.2 and Theorem 1.3 from two local limit theorems. These theorems use (2.25) to study convergence of the transition probabilities for the Markov process {Xk}\{X_{k}\} from Theorem 1.1 under the 1:2 scaling of space and time.

We have the following local limit theorem for fixed qq.

Theorem 4.2.

Let {Xk}\{X_{k}\} be the Markov process introduced in Theorem 1.1. If 0q<10\leq q<1, and 0<σ10<\sigma\leq 1 then for x,y,t>0x,y,t>0 we have

(4.10) limNN(XNt=yN|X0=xN)=yx𝗊t1+σ(x,y),\lim_{N\to\infty}\sqrt{N}\,\mathds{P}\left(X_{\left\lfloor Nt\right\rfloor}=\left\lfloor y\sqrt{N}\right\rfloor\middle|X_{0}=\left\lfloor x\sqrt{N}\right\rfloor\right)=\frac{y}{x}\,{\mathsf{q}}_{\frac{t}{1+\sigma}}(x,y),

where 𝗊t{\mathsf{q}}_{t}, t>0t>0, is given in (1.5).

Furthermore, if ρ0=1𝖼/N\rho_{0}=1-{\mathsf{c}}/\sqrt{N} varies with NN for some fixed constant 𝖼>0{\mathsf{c}}>0, then denoting by X0(N)X_{0}^{(N)} a random variable with the initial law (1.2), for x>0x>0 we have

(4.11) limNN(X0(N)=xN)=𝖼2xe𝖼x.\lim_{N\to\infty}\sqrt{N}\,\mathds{P}\left(X_{0}^{(N)}=\left\lfloor x\sqrt{N}\right\rfloor\right)={\mathsf{c}}^{2}xe^{-{\mathsf{c}}x}.
Proof.

We use (2.25) with

(4.12) k=Nt,m=xN,n=yN.k=\left\lfloor Nt\right\rfloor,\quad m=\left\lfloor x\sqrt{N}\right\rfloor,\quad n=\left\lfloor y\sqrt{N}\right\rfloor.

(To avoid notation clash, we shall use ww and w~\widetilde{w} instead of x,yx,y for the variables of integration, and in formulas such as (4.7).)

In view of (3.11) and (4.7), it is clear that for x,y>0x,y>0 we have

(4.13) limNπyNπxN=yx.\lim_{N\to\infty}\frac{\pi_{\left\lfloor y\sqrt{N}\right\rfloor}}{\pi_{\left\lfloor x\sqrt{N}\right\rfloor}}=\frac{y}{x}.

Indeed, the constant on the right hand side of (3.11) is non-zero, as our a,ba,b have modulus q<1q<1, see (4.5). So, see (4.8) and (4.3), we need to find the limit of the expression

1q2NBkABw~kpm(w~)p~n(w~)g(1q2w~σ)𝑑w~=N(1+σ)k[m+1]q11(w+σ)kQm(w)(q;q)mQn(w)(q;q)ng(w)𝑑w.\tfrac{1-q}{2}\tfrac{\sqrt{N}}{B^{k}}\int_{A}^{B}\widetilde{w}^{k}p_{m}(\widetilde{w})\widetilde{p}_{n}(\widetilde{w})g\left(\tfrac{1-q}{2}\widetilde{w}-\sigma\right)d\widetilde{w}=\tfrac{\sqrt{N}}{(1+\sigma)^{k}[m+1]_{q}}\int_{-1}^{1}(w+\sigma)^{k}\frac{Q_{m}(w)}{(q;q)_{m}}\frac{Q_{n}(w)}{(q;q)_{n}}g(w)dw.

(Here, we used (4.7) and identity/substitution w~/B=(w+σ)/(1+σ)\widetilde{w}/B=(w+\sigma)/(1+\sigma).)

Next we observe that the integral over the interval [1,0][-1,0] is negligible. Indeed, on this interval |w+σ|σ(1σ)=r(1+σ)|w+\sigma|\leq\sigma\vee(1-\sigma)=r(1+\sigma), which together with (3.12) shows that

N[m+1]q|10(w+σ1+σ)kQm(w)(q;q)mQn(w)(q;q)ng(w)𝑑w|CN(m+1)(n+1)rk10g(w)𝑑w0\tfrac{\sqrt{N}}{[m+1]_{q}}\left|\int_{-1}^{0}\left(\tfrac{w+\sigma}{1+\sigma}\right)^{k}\tfrac{Q_{m}(w)}{(q;q)_{m}}\tfrac{Q_{n}(w)}{(q;q)_{n}}g(w)dw\right|\leq C\sqrt{N}(m+1)(n+1)r^{k}\int_{-1}^{0}g\left(w\right)dw\to 0

as NN\to\infty, since k,m,nk,m,n are given by (4.12) and r[0,1)r\in[0,1).

It remains to analyze the case of the integral over [0,1][0,1]. Changing the variable of integration by setting u=2N(1w)u=\sqrt{2N(1-w)}, i.e. w=1u22Nw=1-\frac{u^{2}}{2N}, we get

(4.14) N[m+1]q01(w+σ1+σ)kQm(w)(q;q)mQn(w)(q;q)ng(w)𝑑w1[m+1]q02N(1u22(1+σ)N)kQm(1u22N)N(q;q)mQn(1u22N)N(q;q)nNg(1u22N)u𝑑u.\frac{\sqrt{N}}{[m+1]_{q}}\int_{0}^{1}\left(\frac{w+\sigma}{1+\sigma}\right)^{k}\,\frac{Q_{m}(w)}{(q;q)_{m}}\frac{Q_{n}(w)}{(q;q)_{n}}g(w)dw\\ \frac{1}{[m+1]_{q}}\int_{0}^{\sqrt{2N}}\left(1-\tfrac{u^{2}}{2(1+\sigma)N}\right)^{k}\frac{Q_{m}\left(1-\tfrac{u^{2}}{2N}\right)}{\sqrt{N}(q;q)_{m}}\frac{Q_{n}\left(1-\tfrac{u^{2}}{2N}\right)}{\sqrt{N}(q;q)_{n}}\sqrt{N}g\left(1-\tfrac{u^{2}}{2N}\right)udu.

Clearly, 1[m+1]q1q\frac{1}{[m+1]_{q}}\to 1-q. The next observation is that for u[0,2N]u\in[0,\sqrt{2N}] and k=Ntk=\left\lfloor Nt\right\rfloor, we have

(4.15) (1u22(1+σ)N)keu2t2(1+σ) as N.\left(1-\frac{u^{2}}{2(1+\sigma)N}\right)^{k}\sim e^{-\frac{u^{2}t}{2(1+\sigma)}}\mbox{ as $N\to\infty$}.

Now we refer to (3.9) for g(1u22N)g\left(1-\tfrac{u^{2}}{2N}\right). Noting that θuN1(1u22N)2\theta\sim\frac{u}{\sqrt{N}}\sim\sqrt{1-(1-\frac{u^{2}}{2N})^{2}}, we get

(4.16) Ng(1u22N)2uπ(q;q)3(ab;q)(a,b;q)2=2uπ(1q)(q;q)4(a,b;q)2\sqrt{N}g\left(1-\frac{u^{2}}{2N}\right)\sim\frac{2\,u}{\pi}(q;q)_{\infty}^{3}\frac{(ab;q)_{\infty}}{(a,b;q)^{2}_{\infty}}=\frac{2u}{\pi(1-q)}\;\frac{(q;q)_{\infty}^{4}}{(a,b;q)^{2}_{\infty}}

for u[0,2N]u\in[0,\sqrt{2N}], where the last equality is by (4.6).

We now determine the asymptotics for the remaining factors of the integrand on the right hand side of (4.14). We apply Theorem 3.2 with M=NxM=\left\lfloor\sqrt{N}x\right\rfloor\to\infty and with uM/M:=u/Nu_{M}/M:=u/\sqrt{N} so that uM=uM/N=ux+O(1/M)u_{M}=uM/\sqrt{N}=ux+O(1/M). Recalling (4.12), we see that m=Mm=M and thus

(4.17) Qm(1u22N)N=NxNQM(1uM22M2)Mxsin(ux)ux(a,b;q)(q;q)=sin(ux)u(a,b;q)(q;q)\frac{Q_{m}\left(1-\frac{u^{2}}{2N}\right)}{\sqrt{N}}=\frac{\left\lfloor\sqrt{N}x\right\rfloor}{\sqrt{N}}\;\frac{Q_{M}\left(1-\frac{u_{M}^{2}}{2M^{2}}\right)}{M}\to x\,\frac{\sin(ux)}{ux}\;\frac{(a,b;q)_{\infty}}{(q;q)_{\infty}}=\frac{\sin(ux)}{u}\;\frac{(a,b;q)_{\infty}}{(q;q)_{\infty}}

as NN\to\infty. The same argument with nn from (4.12) gives

(4.18) Qn(1u22N)Nsin(uy)u(a,b;q)(q;q) as N.\frac{Q_{n}\left(1-\frac{u^{2}}{2N}\right)}{\sqrt{N}}\to\frac{\sin(uy)}{u}\;\frac{(a,b;q)_{\infty}}{(q;q)_{\infty}}\mbox{ as $N\to\infty$}.

To verify uniform integrability, we apply elementary bounds

(q;q)|(z;q)|(q;q),|z|q(q;q)_{\infty}\leq|(z;q)_{\infty}|\leq(-q;q)_{\infty},\quad|z|\leq q

within (3.9). We get

Ng(1u22N)N1(1u22N)22(q;q)2π(1q)(q;q)C(q)|u|.\sqrt{N}\,g\left(1-\tfrac{u^{2}}{2N}\right)\leq\sqrt{N}\sqrt{1-\left(1-\tfrac{u^{2}}{2N}\right)^{2}}\;\frac{2(-q;q)_{\infty}^{2}}{\pi(1-q)(q;q)_{\infty}}\;\leq C(q)|u|.

Together with an elementary bound (1x/N)kext(1-x/N)^{k}\leq e^{-xt} for 0xN0\leq x\leq N and the fact that [m+1]q1/(1q)[m+1]_{q}\to 1/(1-q) this show that the integrand on the right hand side of (4.14) is bounded by an integrable function Cu2ecu2Cu^{2}e^{-cu^{2}}, so we can use dominated convergence theorem to pass to the limit under the integral. Combining (4.13), (4.14), (4.15), (4.16), (4.17) and (4.18) we conclude that the left hand side of (4.10) becomes

yx(q;q)4(a,b;q)22π0limN(𝟏u<2N(1u22(1+σ)N)kQm(1u22N)N(q;q)mQn(1u22N)N(q;q)n)u2du=yx2π0eu2t2(1+σ)sin(ux)sin(uy)𝑑u=yx1+σ2πt(e(yx)2(1+σ)2te(y+x)2(1+σ)2t)\frac{y}{x}\,\frac{(q;q)_{\infty}^{4}}{(a,b;q)_{\infty}^{2}}\;\frac{2}{\pi}\int_{0}^{\infty}\lim_{N\to\infty}\left({\bf 1}_{u<\sqrt{2N}}\left(1-\frac{u^{2}}{2(1+\sigma)N}\right)^{k}\frac{Q_{m}\left(1-\frac{u^{2}}{2N}\right)}{\sqrt{N}(q;q)_{m}}\frac{Q_{n}\left(1-\frac{u^{2}}{2N}\right)}{\sqrt{N}(q;q)_{n}}\right)\,u^{2}\,du\\ =\,\frac{y}{x}\,\frac{2}{\pi}\int_{0}^{\infty}e^{-\frac{u^{2}t}{2(1+\sigma)}}\sin(ux)\sin(uy)\,du=\frac{y}{x}\,\sqrt{\frac{1+\sigma}{2\pi t}}\,\left(e^{-\frac{(y-x)^{2}(1+\sigma)}{2t}}-e^{-\frac{(y+x)^{2}(1+\sigma)}{2t}}\right)

as required.

To prove (4.11) we use (3.11). Using notation (4.12), from (1.2) and (3.5) we get

(4.19) N(X0(N)=m)=N(ρ0;q)2(aρ0,bρ0;q)ρ0mQm(1;a,b|q)(q;q)m\displaystyle\sqrt{N}\mathds{P}(X_{0}^{(N)}=m)=\sqrt{N}\frac{(\rho_{0};q)_{\infty}^{2}}{(a\rho_{0},b\rho_{0};q)_{\infty}}\rho_{0}^{m}\frac{Q_{m}(1;a,b|q)}{(q;q)_{m}}
=N(1𝖼N)2xNN(1𝖼N)xN(qρ0;q)2(q;q)mQm(1;a,b|q)m(aρ0,bρ0;q)𝖼2xe𝖼x\displaystyle=N\left(1-\frac{{\mathsf{c}}}{\sqrt{N}}\right)^{2}\,\frac{\left\lfloor x\sqrt{N}\right\rfloor}{N}\,\left(1-\frac{{\mathsf{c}}}{\sqrt{N}}\right)^{\left\lfloor x\sqrt{N}\right\rfloor}\,\frac{(q\rho_{0};q)_{\infty}^{2}}{(q;q)_{m}}\;\frac{Q_{m}(1;a,b|q)}{m\,(a\rho_{0},b\rho_{0};q)_{\infty}}\to{\mathsf{c}}^{2}x\,e^{-{\mathsf{c}}x}

as NN\to\infty

Next we prove a local limit theorem with q1q\nearrow 1. Note that if parameter qq varies with NN, then the law of the Markov process {Xk}\{X_{k}\} from Theorem 1.1 also varies with NN, so we denote this process by {Xk(N)}\{X_{k}^{(N)}\}.

For zz\in\mathds{R} we denote

(4.20) mN(z)=zN+NlogN 2log(1+σ).m_{N}(z)=\left\lfloor z\sqrt{N}\right\rfloor+\left\lfloor\sqrt{N}\,\log\,\sqrt{N\,2\log(1+\sigma)}\right\rfloor.

With this notation we have the following version of Theorem 4.2 for q1q\nearrow 1:

Theorem 4.3.

If 0<σ10<\sigma\leq 1 is fixed and q=e2/Nq=e^{-2/\sqrt{N}} then for x,yx,y\in\mathds{R}, t>0t>0 we have

(4.21) limNN(XNt(N)=mN(y)|X0(N)=mN(x))=K0(ey)K0(ex)𝗉t1+σ(x,y),\lim_{N\to\infty}\sqrt{N}\,\mathds{P}\Big{(}X_{\left\lfloor Nt\right\rfloor}^{(N)}=m_{N}(y)\Big{|}\;X_{0}^{(N)}=m_{N}(x)\Big{)}=\frac{K_{0}(e^{-y})}{K_{0}(e^{-x})}\;{\mathsf{p}}_{\frac{t}{1+\sigma}}(x,y),

where 𝗉t{\mathsf{p}}_{t}, t>0t>0, is given in (1.9).

In addition, if ρ0=e𝖼/N=q𝖼/2\rho_{0}=e^{-{\mathsf{c}}/\sqrt{N}}=q^{{\mathsf{c}}/2} for some fixed constant 𝖼>0{\mathsf{c}}>0, then

(4.22) limNN(X0(N)=mN(x))=42𝖼Γ(𝖼/2)2K0(ex).\lim_{N\to\infty}\sqrt{N}\mathds{P}\left(X_{0}^{(N)}=m_{N}(x)\right)=\frac{4}{2^{\mathsf{c}}\Gamma({\mathsf{c}}/2)^{2}}K_{0}(e^{-x}).
Proof.

We write the parameters (4.5) in trigonometric form a=qeiαa=-qe^{{\textnormal{i}}\alpha}, b=qeiαb=-qe^{-{\textnormal{i}}\alpha} with σ=cosα\sigma=\cos\alpha, α(π/2,π/2)\alpha\in(-\pi/2,\pi/2).

First, we prove (4.22). As in (4.19) we have

N(X0(N)=mN(x))=N(ρ0;q)2(aρ0,bρ0;q)ρ0mN(x)QmN(x)(1;a,b|q)(q;q)mN(x).\sqrt{N}\,\mathds{P}(X_{0}^{(N)}=m_{N}(x))=\sqrt{N}\frac{(\rho_{0};q)_{\infty}^{2}}{(a\rho_{0},b\rho_{0};q)_{\infty}}\rho_{0}^{m_{N}(x)}\frac{Q_{m_{N}(x)}(1;a,b|q)}{(q;q)_{m_{N}(x)}}.

Denote M=NM=\sqrt{N}. Then with a~=eiα\widetilde{a}=e^{i\alpha} and b~=eiα\widetilde{b}=e^{-i\alpha} we see that mm as defined in Theorem 3.3 assumes the form m=mN(x)m=m_{N}(x). Consequently, we can rewrite this expression as follows

N(X0(N)=mN(x))=(MemM)𝖼M2𝖼((ρ0;q)(q;q))2(a,b;q)(aρ0,bρ0;q)×JM,\sqrt{N}\,\mathds{P}(X_{0}^{(N)}=m_{N}(x))=\left(Me^{-\frac{m}{M}}\right)^{{\mathsf{c}}}\,M^{2-{\mathsf{c}}}\,\left(\frac{(\rho_{0};q)_{\infty}}{(q;q)_{\infty}}\right)^{2}\frac{(a,b;q)_{\infty}}{(a\rho_{0},b\rho_{0};q)_{\infty}}\times J_{M},

where

(4.23) JM=(q;q)2M(qa~,qb~;q)Qm(1;qa~,qb~|q)(q;q)mK0(ex)as NJ_{M}=\frac{(q;q)_{\infty}^{2}}{M(-q\widetilde{a},-q\widetilde{b};q)_{\infty}}\frac{Q_{m}(1;-q\widetilde{a},-q\widetilde{b}|q)}{(q;q)_{m}}\to K_{0}(e^{-x})\quad\mbox{as }\;N\to\infty

with the limit deduced from (3.15). Clearly,

(4.24) MemM=NemN(x)Nex2(1+σ)as N.Me^{-\frac{m}{M}}=\sqrt{N}e^{-\frac{m_{N}(x)}{\sqrt{N}}}\to\frac{e^{-x}}{\sqrt{2(1+\sigma)}}\quad\mbox{as }\;N\to\infty.

Moreover, recalling (5.2), we can write

(4.25) M2𝖼((ρ0;q)(q;q))2=M2𝖼((q𝖼2;q)(q;q))2=(M(1q))2𝖼Γq(𝖼/2)222𝖼Γ(𝖼/2)2M^{2-{\mathsf{c}}}\,\left(\frac{(\rho_{0};q)_{\infty}}{(q;q)_{\infty}}\right)^{2}=M^{2-{\mathsf{c}}}\,\left(\frac{(q^{\frac{{\mathsf{c}}}{2}};q)_{\infty}}{(q;q)_{\infty}}\right)^{2}=\frac{(M(1-q))^{2-{\mathsf{c}}}}{\Gamma_{q}\left({\mathsf{c}}/2\right)^{2}}\to\frac{2^{2-{\mathsf{c}}}}{\Gamma\left({\mathsf{c}}/2\right)^{2}}

as NN\to\infty, where the last limit follows from (5.3).

Finally, relying on (5.1), we get

(4.26) (a,b;q)(aρ0,bρ0;q)=(a~q,b~q;q)(a~q1+𝖼/2,b~q1+𝖼/2;q)=(1+a~q𝖼2)(1+b~q𝖼2)(1+a~)(1+b~)(a~;q)(a~q𝖼2;q)(b~;q)(b~q𝖼2;q)(1+eiα)𝖼/2(1+eiα)𝖼/2=2𝖼/2(1+σ)𝖼/2\frac{(a,b;q)_{\infty}}{(a\rho_{0},b\rho_{0};q)_{\infty}}=\frac{(-\widetilde{a}q,-\widetilde{b}q;q)_{\infty}}{(-\widetilde{a}q^{1+{\mathsf{c}}/2},-\widetilde{b}q^{1+{\mathsf{c}}/2};q)_{\infty}}\\ =\frac{(1+\widetilde{a}q^{\frac{{\mathsf{c}}}{2}})(1+\widetilde{b}q^{\frac{{\mathsf{c}}}{2}})}{(1+\widetilde{a})(1+\widetilde{b})}\,\frac{(-\widetilde{a};q)_{\infty}}{(-\widetilde{a}q^{\frac{{\mathsf{c}}}{2}};q)_{\infty}}\,\frac{(-\widetilde{b};q)_{\infty}}{(-\widetilde{b}q^{\frac{{\mathsf{c}}}{2}};q)_{\infty}}\to(1+e^{{\textnormal{i}}\alpha})^{{\mathsf{c}}/2}(1+e^{-{\textnormal{i}}\alpha})^{{\mathsf{c}}/2}=2^{{\mathsf{c}}/2}(1+\sigma)^{{\mathsf{c}}/2}

as (1+eiα)𝖼/2(1+eiα)𝖼/2=(2+2cosα)𝖼/2(1+e^{{\textnormal{i}}\alpha})^{{\mathsf{c}}/2}(1+e^{-{\textnormal{i}}\alpha})^{{\mathsf{c}}/2}=(2+2\cos\alpha)^{{\mathsf{c}}/2}. Combining (4.23), (4.24), (4.25) and (4.26) we get (4.22).

Second, we prove (4.21). Our starting point is formula (2.25) which we use with

(4.27) m:=mN(x),n:=mN(y),k=Nt.m:=m_{N}(x),\quad n:=m_{N}(y),\quad k=\left\lfloor Nt\right\rfloor.

Even though x,yx,y\in\mathds{R}, note that m,nm,n are non-negative integers for large enough NN. In particular, we note that m/n1m/n\to 1 and m/k0m/k\to 0 as NN\to\infty. It is also clear that [n+1]q[m+1]qN/2[n+1]_{q}\sim[m+1]_{q}\sim\sqrt{N}/2. (To avoid notation clash, we use ww and w~\widetilde{w} instead of x,yx,y for the variables of integration, or in formulas like (4.7).)

From (2.25), we get

(4.28) N(Xk=mN(y)|X0=mN(x))=πnπmNBkABw~kpm(w~)[n+1]qpn(w~)g(w~)𝑑w~=Qn(1;a,b|q)(q;q)n(q;q)mQm(1;a,b|q)N[n+1]q11(w+σ)k(1+σ)kQm(w;a,b|q)(q;q)mQn(w;a,b|q)(q;q)ng(w)𝑑w,\sqrt{N}\mathds{P}(X_{k}=m_{N}(y)|X_{0}=m_{N}(x))=\frac{\pi_{n}}{\pi_{m}}\frac{\sqrt{N}}{B^{k}}\int_{A}^{B}\widetilde{w}^{k}p_{m}(\widetilde{w})[n+1]_{q}p_{n}(\widetilde{w})g(\widetilde{w})d\widetilde{w}\\ =\frac{Q_{n}(1;a,b|q)}{(q;q)_{n}}\;\frac{(q;q)_{m}}{Q_{m}(1;a,b|q)}\;\frac{\sqrt{N}}{[n+1]_{q}}\int_{-1}^{1}\frac{\left(w+\sigma\right)^{k}}{(1+\sigma)^{k}}\frac{Q_{m}\left(w;a,b|q\right)}{(q;q)_{m}}\,\frac{Q_{n}\left(w;a,b|q\right)}{(q;q)_{n}}\,g(w)\,dw,

where the latter equation follows from (4.9) and (4.7).

By Theorem 3.3 used with M=NM=\sqrt{N}, u=0u=0, and m=mN(y)m=m_{N}(y) or m=mN(x)m=m_{N}(x) we get

(4.29) limNQn(1;a,b|q)(q;q)n(q;q)mQm(1;a,b|q)=limN(q;q)2QmN(y)(1;a,b|q)N(a,b;q)(q;q)mN(y)×limNN(a,b;q)(q;q)mN(x)(q;q)2QmN(x)(1;a,b|q)=K0(ey)K0(ex).\lim_{N\to\infty}\frac{Q_{n}(1;a,b|q)}{(q;q)_{n}}\;\frac{(q;q)_{m}}{Q_{m}(1;a,b|q)}\\ =\lim_{N\to\infty}\frac{(q;q)_{\infty}^{2}Q_{m_{N}(y)}\,(1;a,b|q)}{\sqrt{N}(a,b;q)_{\infty}\,(q;q)_{m_{N}(y)}}\times\lim_{N\to\infty}\frac{\sqrt{N}(a,b;q)_{\infty}(q;q)_{m_{N}(x)}}{(q;q)_{\infty}^{2}Q_{m_{N}(x)}(1;a,b|q)}=\frac{K_{0}(e^{-y})}{K_{0}(e^{-x})}.

This gives the pair of Bessel functions K0K_{0} on the right-hand side of (4.21).

Next, we show that the contribution from the integral in (4.28) over [1,1/2][-1,1/2] is negligible. Noting that for ww from this interval |w+σ|<(1/2+σ)(1σ)=r(1+σ)|w+\sigma|<(1/2+\sigma)\vee(1-\sigma)=r(1+\sigma), similarly as in the proof of Theorem 4.2, the contribution of this integral is bounded by

CrkN[n+1]q11/2|Qm(w;a,b|q)|(q;q)m|Qn(w;a,b|q)|(q;q)ng(w)𝑑wCrkN2[n+1]q11(Qm2(w;a,b|q)(q;q)m2+Qn2(w;a,b|q)(q;q)n2)g(w)𝑑w=CrkN2[n+1]q([m+1]q+[n+1]q)0.Cr^{k}\,\frac{\sqrt{N}}{[n+1]_{q}}\int_{-1}^{1/2}\frac{|Q_{m}\left(w;a,b|q\right)|}{(q;q)_{m}}\frac{|Q_{n}\left(w;a,b|q\right)|}{(q;q)_{n}}g\left(w\right)dw\leq\\ Cr^{k}\,\frac{\sqrt{N}}{2[n+1]_{q}}\int_{-1}^{1}\left(\frac{Q_{m}^{2}\left(w;a,b|q\right)}{(q;q)_{m}^{2}}+\frac{Q_{n}^{2}\left(w;a,b|q\right)}{(q;q)_{n}^{2}}\right)g\left(w\right)dw\\ =Cr^{k}\,\frac{\sqrt{N}}{2[n+1]_{q}}\left([m+1]_{q}+[n+1]_{q}\right)\to 0.

Here CC is a constant from boundedness of the convergent sequence in (4.29), and the last equality follows from (4.2) and (4.7). To compute the limit we used the observation that [n+1]q[m+1]qN/2[n+1]_{q}\sim[m+1]_{q}\sim\sqrt{N}/2 and r[0,1)r\in[0,1).

Substituting w=cos(u/N)w=\cos(u/\sqrt{N}) into (4.28) and discarding the non-contributing part of the integral, we have

(4.30) N(Xk=mN(y)|X0=mN(x))K0(ey)K0(ex)0 1(0,π3N)(u)fN(u)𝑑u,\sqrt{N}\mathds{P}(X_{k}=m_{N}(y)|X_{0}=m_{N}(x))\sim\frac{K_{0}(e^{-y})}{K_{0}(e^{-x})}\int_{0}^{\infty}\,\mathbf{1}_{(0,\,\frac{\pi}{3}\sqrt{N})}(u)f_{N}(u)\,du,

where

fN(u)=1[n+1]q(σ+cosuN1+σ)kQm(cosuN;a,b|q)(q;q)mQn(cosuN;a,b|q)(q;q)ng(cos(uN))sin(uN).f_{N}(u)=\frac{1}{[n+1]_{q}}\left(\tfrac{\sigma+\cos\frac{u}{\sqrt{N}}}{1+\sigma}\right)^{k}\frac{Q_{m}\left(\cos\frac{u}{\sqrt{N}};a,b|q\right)}{(q;q)_{m}}\,\frac{Q_{n}\left(\cos\frac{u}{\sqrt{N}};a,b|q\right)}{(q;q)_{n}}g\left(\cos(\tfrac{u}{\sqrt{N}})\right)\sin(\tfrac{u}{\sqrt{N}}).

Referring first to (3.2) and then using (1q)(ab;q)=(q;q)(1-q)(ab;q)_{\infty}=(q;q)_{\infty} and (5.2) we get

g(cos(uN))sin(uN)=(q,ab;q)2π|(qiu;q)|2|(aqiu/2,bqiu/2;q)|2=(q;q)4(1q)2π|Γq(iu)|2|(aqiu/2,bqiu/2;q)|2.g\left(\cos(\tfrac{u}{\sqrt{N}})\right)\sin(\tfrac{u}{\sqrt{N}})=\frac{(q,ab;q)_{\infty}}{2\pi}\,\frac{|(q^{{\textnormal{i}}u};q)_{\infty}|^{2}}{|(aq^{{\textnormal{i}}u/2},bq^{{\textnormal{i}}u/2};q)_{\infty}|^{2}}=\frac{(q;q)_{\infty}^{4}(1-q)}{2\pi|\Gamma_{q}({\textnormal{i}}u)|^{2}\,|(aq^{{\textnormal{i}}u/2},bq^{{\textnormal{i}}u/2};q)_{\infty}|^{2}}.

Thus fNf_{N} can be written as

(4.31) fN(u)=N(1q)[n+1]q(σ+cosuN1+σ)Nt12π|Γq(iu)|2×(q;q)2QmN(x)(cosuN;a,b|q)N(a,b;q)(q;q)mN(x)(q;q)2QmN(y)(cosuN;a,b|q)N(a,b;q)(q;q)mN(y)(a,b;q)2|(aqiu/2,bqiu/2;q)|2.f_{N}(u)=\frac{N(1-q)}{[n+1]_{q}}\,\left(\frac{\sigma+\cos\frac{u}{\sqrt{N}}}{1+\sigma}\right)^{\left\lfloor Nt\right\rfloor}\,\frac{1}{2\pi\,|\Gamma_{q}({\textnormal{i}}u)|^{2}}\\ \times\frac{(q;q)_{\infty}^{2}Q_{m_{N}(x)}(\cos\frac{u}{\sqrt{N}};a,b|q)}{\sqrt{N}(a,b;q)_{\infty}\,(q;q)_{m_{N}(x)}}\,\frac{(q;q)_{\infty}^{2}Q_{m_{N}(y)}(\cos\frac{u}{\sqrt{N}};a,b|q)}{\sqrt{N}(a,b;q)_{\infty}\,(q;q)_{m_{N}(y)}}\,\frac{(a,b;q)_{\infty}^{2}}{|(aq^{{\textnormal{i}}u/2},bq^{{\textnormal{i}}u/2};q)_{\infty}|^{2}}.

Using [n+1]qN/2[n+1]_{q}\sim\sqrt{N}/2 again, we see that the first factor is asymptotically constant, N(1q)/[n+1]q4{N(1-q)}/{[n+1]_{q}}\sim 4. Furthermore,

(σ+cosuN1+σ)Nteu2t2(1+σ).\left(\frac{\sigma+\cos\frac{u}{\sqrt{N}}}{1+\sigma}\right)^{\left\lfloor Nt\right\rfloor}\to e^{-\frac{u^{2}t}{2(1+\sigma)}}.

By Lemma 5.1, Γq(iu)Γ(iu)\Gamma_{q}({\textnormal{i}}u)\to\Gamma({\textnormal{i}}u). From Theorem 3.3, we see that for real zz, recall (4.20),

(4.32) (q;q)2N(a,b;q)QmN(z)(cosuN;a,b|q)(q;q)mN(z)Kiu(ez).\frac{(q;q)_{\infty}^{2}}{\sqrt{N}(a,b;q)_{\infty}}\,\frac{Q_{m_{N}(z)}(\cos\frac{u}{\sqrt{N}};a,b|q)}{(q;q)_{m_{N}(z)}}\to K_{{\textnormal{i}}u}(e^{-z}).

In view of Lemma 5.1

(a,b;q)(aqiu/2,bqiu/2;q)=(eiαq;q)(eiαq1+iu/2;q)(eiαq;q)(eiαq1+iu/2;q)(1+eiα)iu/2(1+eiα)iu/2=(2+2σ)iu2.\frac{(a,b;q)_{\infty}}{(aq^{{\textnormal{i}}u/2},bq^{{\textnormal{i}}u/2};q)_{\infty}}=\frac{(-e^{{\textnormal{i}}\alpha}q;q)_{\infty}}{(-e^{{\textnormal{i}}\alpha}q^{1+{\textnormal{i}}u/2};q)_{\infty}}\frac{(-e^{-{\textnormal{i}}\alpha}q;q)_{\infty}}{(-e^{-{\textnormal{i}}\alpha}q^{1+{\textnormal{i}}u/2};q)_{\infty}}\\ \to(1+e^{{\textnormal{i}}\alpha})^{{\textnormal{i}}u/2}(1+e^{-{\textnormal{i}}\alpha})^{{\textnormal{i}}u/2}=(2+2\sigma)^{{\textnormal{i}}\frac{u}{2}}.

Note that |(2+2σ)iu2|=1|(2+2\sigma)^{{\textnormal{i}}\frac{u}{2}}|=1, so this factor does not contribute to the limit of fNf_{N}.

To summarize, we see that

(4.33) limN 1(0,π3N)(u)fN(u)=𝟏(0,)(u)2πeu2t2(1+σ)Kiu(ex)Kiu(ey)1|Γ(iu)|2.\lim_{N\to\infty}\,\mathbf{1}_{(0,\,\frac{\pi}{3}\sqrt{N})}(u)\,f_{N}(u)=\mathbf{1}_{(0,\infty)}(u)\;\frac{2}{\pi}\,e^{-\frac{u^{2}t}{2(1+\sigma)}}\,K_{{\textnormal{i}}u}(e^{-x})\,K_{{\textnormal{i}}u}(e^{-y})\,\frac{1}{|\Gamma({\textnormal{i}}u)|^{2}}.

We now justify that one can pass to the limit under the integral (4.30). For 0<u<π3N0<u<\frac{\pi}{3}\,\sqrt{N} we have

(cos(u/N)+σ1+σ)k(1u22(1+σ)N)NtCeu2t2(1+σ)\left(\frac{\cos(u/\sqrt{N})+\sigma}{1+\sigma}\right)^{k}\leq\left(1-\frac{u^{2}}{2(1+\sigma)N}\right)^{Nt}\leq Ce^{-\frac{u^{2}t}{2(1+\sigma)}}

for some C<2.5C<2.5. Recall that t,σt,\sigma are fixed.

By Proposition 3.1 and Theorem 3.3 we know that for real zz, recall (4.32), we have

|(q;q)2N(a,b;q)QmN(z)(cosuM;a,b|q)(q;q)mN(z)|(q;q)2N(a,b;q)QmN(z)(1;a,b|q)(q;q)mN(z)K0(ez),\left|\frac{(q;q)_{\infty}^{2}}{\sqrt{N}(a,b;q)_{\infty}}\,\frac{Q_{m_{N}(z)}(\cos\frac{u}{M};a,b|q)}{(q;q)_{m_{N}(z)}}\right|\leq\frac{(q;q)_{\infty}^{2}}{\sqrt{N}(a,b;q)_{\infty}}\,\frac{Q_{m_{N}(z)}(1;a,b|q)}{(q;q)_{m_{N}(z)}}\to K_{0}(e^{-z}),

therefore the expression on the left-hand side above is uniformly bounded in u,Nu,N. By Lemma 5.9, there exist A>0A>0, B>0B>0 and MM^{*} such that

𝟏{|u|<πN/2}(a,b;q)2|(aqiu/2,bqiu/2;q)|2AeB|u|{\mathbf{1}}_{\{|u|<\pi\sqrt{N}/2\}}\frac{(a,b;q)_{\infty}^{2}}{|(aq^{{\textnormal{i}}u/2},bq^{{\textnormal{i}}u/2};q)_{\infty}|^{2}}\leq Ae^{B|u|}

for all NM2N\geq M_{*}^{2} and all uu\in\mathds{R}. By Lemma 5.7, 1/|Γq(iu)|2Au2eB|u|1/|\Gamma_{q}({\textnormal{i}}u)|^{2}\leq Au^{2}e^{B|u|} for |u|<Nπ/2|u|<\sqrt{N}\pi/2. Moreover, recalling that n=mN(y)n=m_{N}(y), the convergent sequence N(1q)/[n+1]qN(1-q)/[n+1]_{q} is bounded. Combining all these bounds, we see that there are constants A,B,M>0A,B,M_{*}>0 such that for all N>M2N>M_{*}^{2} function fN(u)𝟏(0,π3N)(u)f_{N}(u)\mathbf{1}_{(0,\frac{\pi}{3}\sqrt{N})}(u) is bounded on [0,)[0,\infty) by the integrable function

Au2eB|u|eu2t2(1+σ),Au^{2}e^{B|u|}e^{-\frac{u^{2}t}{2(1+\sigma)}},

so we can invoke the dominated convergence theorem. Taking the limit (4.33) inside the integral in (4.30), we see that

limNN(Xk=mN(y)|X0=mN(x))=K0(ey)K0(ex)2π0eu2t2(1+σ)Kiu(ex)Kiu(ey)du|Γ(iu)|2,\lim_{N\to\infty}\sqrt{N}\mathds{P}(X_{k}=m_{N}(y)|X_{0}=m_{N}(x))=\frac{K_{0}(e^{-y})}{K_{0}(e^{-x})}\,\frac{2}{\pi}\int_{0}^{\infty}\,e^{-\frac{u^{2}t}{2(1+\sigma)}}\,K_{{\textnormal{i}}u}(e^{-x})\,K_{{\textnormal{i}}u}(e^{-y})\,\frac{du}{|\Gamma({\textnormal{i}}u)|^{2}},

which ends the proof by (1.9). ∎

4.2. Proofs of Theorems 1.2 and 1.3

Both proofs are very similar.

Proof of Theorem 1.2.

Denote by {Xk(N)}\{X_{k}^{(N)}\} the Markov process from the conclusion of Theorem 1.1 with ρ0=e𝖼/N\rho_{0}=e^{-{\mathsf{c}}/\sqrt{N}}. For fixed t0=0<t1<<tdt_{0}=0<t_{1}<\dots<t_{d} and x0,,xd>0x_{0},\dots,x_{d}>0 by (4.11), (4.10) of Theorem 4.2 and the Markov property, we have

limNN(d+1)/2(XtiN(N)N=xiNN,i=0,1,,d)=𝖼2e𝖼x0xdj=1d𝗊tjtj11+σ(xj1,xj),\lim_{N\to\infty}N^{(d+1)/2}\,\mathds{P}\left(\frac{X_{\left\lfloor t_{i}N\right\rfloor}^{(N)}}{\sqrt{N}}=\tfrac{\left\lfloor x_{i}\sqrt{N}\right\rfloor}{\sqrt{N}},\,i=0,1,\ldots,d\right)={\mathsf{c}}^{2}e^{-{\mathsf{c}}x_{0}}\,x_{d}\,\prod_{j=1}^{d}\,{\mathsf{q}}_{\frac{t_{j}-t_{j-1}}{1+\sigma}}(x_{j-1},x_{j}),

where we recall that 𝗊t{\mathsf{q}}_{t} is defined in (1.5). This proves local convergence to the joint density of the vector (ξ0,ξt1,,ξtd)(\xi_{0},\xi_{t_{1}},\dots,\xi_{t_{d}}), see (1.4) and (1.6).

By (Billingsley,, 1999, Theorem 3.3), the convergence of finite-dimensional distributions as stated in (1.3) follows. ∎

Proof of Theorem 1.3.

Denote by {Xk(N)}\{X_{k}^{(N)}\} the Markov process from the conclusion of Theorem 1.1 with ρ0=e𝖼/N\rho_{0}=e^{-{\mathsf{c}}/\sqrt{N}} and q=e2/Nq=e^{-2/\sqrt{N}}.

Clearly, the centering sequence at the left-hand side of (1.7) in Theorem 1.3 can be replaced by Nlog2N(1+σ)\left\lfloor\sqrt{N}\log\sqrt{2N(1+\sigma)}\right\rfloor, N1N\geq 1.

For fixed t0=0<t1<<tdt_{0}=0<t_{1}<\dots<t_{d} and x0,,xdx_{0},\dots,x_{d}\in\mathds{R} by (4.22), (4.21) of Theorem 4.3 and the Markov property we have

limNNd+12(XtiN(N)Nlog2N(1+σ)N=xiNN,i=0,1,,d)=(4.20)limNNd+12(XtiN(N)=mN(xi),i=0,1,,d)=42𝖼Γ(𝖼/2)2K0(ex)j=1d𝗉tjtj11+σ(xj1,xj),\lim_{N\to\infty}N^{\frac{d+1}{2}}\mathds{P}\Big{(}\frac{X_{\left\lfloor t_{i}N\right\rfloor}^{(N)}-\left\lfloor\sqrt{N}\log\sqrt{2N(1+\sigma)}\right\rfloor}{\sqrt{N}}=\frac{\left\lfloor x_{i}\sqrt{N}\right\rfloor}{\sqrt{N}},\,i=0,1,\ldots,d\Big{)}\\ \stackrel{{\scriptstyle\eqref{zN}}}{{=}}\lim_{N\to\infty}N^{\frac{d+1}{2}}\mathds{P}\Big{(}X_{\left\lfloor t_{i}N\right\rfloor}^{(N)}=m_{N}(x_{i}),\,i=0,1,\ldots,d\Big{)}\\ =\frac{4}{2^{\mathsf{c}}\Gamma({\mathsf{c}}/2)^{2}}K_{0}(e^{-x})\prod_{j=1}^{d}{\mathsf{p}}_{\frac{t_{j}-t_{j-1}}{1+\sigma}}(x_{j-1},x_{j}),

where we recall that 𝗉t{\mathsf{p}}_{t} is defined in (1.9). This proves local convergence to the joint density of the vector (ζ0,ζt1,,ζtd)(\zeta_{0},\zeta_{t_{1}},\dots,\zeta_{t_{d}}) , see (1.8) and (1.10).

By (Billingsley,, 1999, Theorem 3.3) the convergence of finite-dimensional distributions as stated in (1.7) follows. ∎

5. Auxiliary results on special functions

5.1. Some useful limits for q1q\nearrow 1

The following result appears in Ch. 16 of Ramanujan’s notebook, see (Adiga et al.,, 1985, Entry 1), (Berndt,, 1991, p. 13, Entry 1(i)) or (Gasper and Rahman,, 2004, (I34)) for the statement. The proof in (Koornwinder,, 1990, Proposition A.2) is for real λ\lambda, which is not enough for our purposes. The proof in (Berndt,, 1991, p. 13) uses analytic continuation from |z|<1|z|<1. We note that we always use principal branch of the logarithm and of the power function.

Lemma 5.1 (Ramanujan).

For all λ\lambda\in{\mathbb{C}} and zz in cut complex plane [1,){\mathbb{C}}\setminus[1,\infty) we have

(5.1) (z;q)(zqλ;q)(1z)λ\frac{(z;q)_{\infty}}{(zq^{\lambda};q)_{\infty}}\to(1-z)^{\lambda}

as q1q\to 1^{-}. The convergence is uniform in zz on compact subsets of [1,){\mathbb{C}}\setminus[1,\infty).

Proof.

Denote the left-hand side of (5.1) by fq(z)f_{q}(z). Then

hq(z):=fq(z)fq(z)=n0[qλ+n1zqλ+nqn1zqn]=1qλ1q×(1q)n0qn(1zqλ+n)(1zqn).h_{q}(z):=\frac{f_{q}^{\prime}(z)}{f_{q}(z)}=\sum\limits_{n\geq 0}\bigg{[}\frac{q^{\lambda+n}}{1-zq^{\lambda+n}}-\frac{q^{n}}{1-zq^{n}}\bigg{]}=-\frac{1-q^{\lambda}}{1-q}\times(1-q)\sum\limits_{n\geq 0}\frac{{q^{n}}}{(1-zq^{\lambda+n})(1-zq^{n})}.

Thus as q1q\to 1^{-} we have

hq(z)λ01dy(1zy)2=λ1z.h_{q}(z)\to-\lambda\int_{0}^{1}\frac{dy}{(1-zy)^{2}}=-\frac{\lambda}{1-z}.

This convergence clearly holds pointwise, for all z[1,)z\in{\mathbb{C}}\setminus[1,\infty). To prove uniform convergence on compact subsets, we use Montel’s theorem, which requires us to show that the functions hqh_{q} are uniformly bounded on compact subsets of [1,){\mathbb{C}}\setminus[1,\infty) as q1q\to 1^{-}. For small ϵ>0\epsilon>0, we define DϵD_{\epsilon} to be the closed domain

Dϵ={z:|arg(z)|ϵ}{z:|z|<1ϵ}.D_{\epsilon}=\{z\in{\mathds{C}}:|\arg(z)|\leq\epsilon\}\setminus\{z\in{\mathds{C}}:|z|<1-\epsilon\}.

The domain DϵD_{\epsilon} is shown on Figure 2.

ϵ\epsiloniy{\textnormal{i}}yxxDϵD_{\epsilon}1ϵ1-\epsilon1ϵ1-\epsilon11
Figure 2. Domain DϵD_{\epsilon}

Now we take an arbitrary compact set K[1,)K\subset{\mathbb{C}}\setminus[1,\infty). There exists ϵ>0\epsilon>0 small enough such that KDϵK\subset{\mathbb{C}}\setminus D_{\epsilon}. Since q[0,1)q\in[0,1), for any zKz\in K (in fact, for any zDεz\not\in D_{\varepsilon}) and any n=0,1,n=0,1,\dots we have qnzDϵDϵ/2q^{n}z\in{\mathbb{C}}\setminus D_{\epsilon}\subset{\mathbb{C}}\setminus D_{\epsilon/2}.

Next, noting that limq1qλ=1\lim_{q\to 1^{-}}q^{\lambda}=1, there is q0(0,1)q_{0}\in(0,1) such that |qλ|<1ϵ/2|q^{\lambda}|<1-\epsilon/2 and |arg(qλ)|<ϵ/2|\arg(q^{\lambda})|<\epsilon/2 for all q(q0,1)q\in(q_{0},1). This implies that for any zDϵz\not\in D_{\epsilon} we have qλzDϵ/2q^{\lambda}z\in{\mathbb{C}}\setminus D_{\epsilon/2}. Indeed, if zDϵz\not\in D_{\epsilon} then either |z|<1ϵ|z|<1-\epsilon or |arg(z)|>ϵ|\arg(z)|>\epsilon. In the first case we have |zqλ|<(1ϵ)(1+ϵ/2)<1ϵ/2|zq^{\lambda}|<(1-\epsilon)(1+\epsilon/2)<1-\epsilon/2. In the second case, |arg(qλz)||arg(z)||arg(qλ)|>ϵ/2|\arg(q^{\lambda}z)|\geq|\arg(z)|-|\arg(q^{\lambda})|>\epsilon/2.

We have thus shown that there exists q0q_{0} such that if zKz\in K then, with zqnDϵzq^{n}\not\in D_{\epsilon}, both qnzq^{n}z and qn+λzq^{n+\lambda}z are in Dϵ/2{\mathbb{C}}\setminus D_{\epsilon/2} for all n=0,1,n=0,1,\dots and all q(q0,1)q\in(q_{0},1). But the distance from 11 to the set Dϵ/2{\mathbb{C}}\setminus D_{\epsilon/2} is strictly positive. This means that there is δ>0\delta>0 such that for all zKz\in K, q(q0,1)q\in(q_{0},1) and every n0n\geq 0 we have

|1zqλ+n|δ,|1zqn|δ,|1-zq^{\lambda+n}|\geq\delta,\;\;\;|1-zq^{n}|\geq\delta,

which implies an upper bound

|hq(z)|<|1qλ||1q|×(1q)n0qnδ2=|1qλ||1q|×δ2<C|λ|δ2,|h_{q}(z)|<\frac{|1-q^{\lambda}|}{|1-q|}\times(1-q)\sum\limits_{n\geq 0}q^{n}\delta^{-2}=\frac{|1-q^{\lambda}|}{|1-q|}\times\delta^{-2}<C|\lambda|\delta^{-2},

for some constant C>0C>0.

Thus we have proved that the functions hq(z)h_{q}(z) converge to λ/(1z)-\lambda/(1-z) as q1q\to 1^{-}, uniformly on compact subsets of [1,){\mathbb{C}}\setminus[1,\infty). Then, integrating over the segment from 0 to zz

fq(z)=exp(0zhq(w)𝑑w)(1z)λf_{q}(z)=\exp\Big{(}\int_{0}^{z}h_{q}(w)dw\Big{)}\to(1-z)^{\lambda}

as q1q\to 1^{-}, also uniformly on compact subsets of [1,){\mathbb{C}}\setminus[1,\infty). ∎

Recall (Gasper and Rahman,, 2004, Section 1.10) that for z{0,1,2,}z\in\mathds{C}\setminus\{0,-1,-2,\dots\} and q[0,1)q\in[0,1), the qq-Gamma function is defined by

(5.2) Γq(z):=(1q)1z(q;q)(qz;q).\Gamma_{q}(z):=(1-q)^{1-z}\;\frac{(q;q)_{\infty}}{(q^{z};q)_{\infty}}.

The following result appears Ch. 16 of Ramanujan notebook. For the statement, see (Adiga et al.,, 1985, Entry 1), or (Berndt,, 1991, p. 13, Entry 1(ii)). For the proof, see (Koornwinder,, 1990, Theorem B.2).

Lemma 5.2.

If z0,1,2,z\neq 0,-1,-2,\dots then

(5.3) limq1Γq(z)=Γ(z).\lim_{q\nearrow 1}\Gamma_{q}(z)=\Gamma(z).

We will also need the following.

Lemma 5.3.

Let q=e2/Mq=e^{-2/M}. For any c>0c>0 the function

t|Γq(c+it)|t\mapsto|\Gamma_{q}(c+{\textnormal{i}}t)|

is periodic with period MπM\pi, it is increasing for t[Mπ/2,0]t\in[-M\pi/2,0] and decreasing for t[0,Mπ/2]t\in[0,M\pi/2].

Proof.

Referring to (5.2) we note that

|Γq(c+it)|=(1q)1c(q;q)n01|1qn+c+it||\Gamma_{q}(c+{\textnormal{i}}t)|=(1-q)^{1-c}(q;q)_{\infty}\prod\limits_{n\geq 0}\frac{1}{|1-q^{n+c+{\textnormal{i}}t}|}

As tt changes over the interval [Mπ/2,Mπ/2][-M\pi/2,M\pi/2], the point w=w(t)=qn+c+itw=w(t)=q^{n+c+{\textnormal{i}}t} goes around a circle |w|=qn+c|w|=q^{n+c}, thus 1w1-w goes around a circle centered at 11 and having radius r=qn+cr=q^{n+c}. Thus, |1w(t)|=1+q2(n+c)2qn+ccos(2tM)|1-w(t)|=\sqrt{1+q^{2(n+c)}-2q^{n+c}\cos(\frac{2t}{M})} is periodic with period MπM\pi, has minimum at t=0t=0 and maximum at t=±Mπ/2t=\pm M\pi/2 and it is clearly monotone between these points, which means that for every n0n\geq 0, the function

t[Mπ/2,Mπ/2]1|1qn+c+it|t\in[-M\pi/2,M\pi/2]\mapsto\frac{1}{|1-q^{n+c+{\textnormal{i}}t}|}

is increasing on [Mπ/2,0][-M\pi/2,0] and decreasing on [0,Mπ/2][0,M\pi/2]. Taking the product of positive increasing/decreasing functions gives us an increasing/decreasing function. ∎

5.2. Bounds on Γq\Gamma_{q}

Following numerous references, such as Corwin and Knizel, (2021); Zhang, (2008) we will use the Jacobi theta functions to derive the bounds we need.

We assume that vv\in{\mathds{C}}, Im(τ)>0\textnormal{Im}(\tau)>0 and set q:=eπiτq:=e^{\pi{\textnormal{i}}\tau}. The Jacobi theta functions are defined as

(5.4) θ1(v|τ)\displaystyle{\theta}_{1}(v|\tau) =2q1/4n0(1)nqn(n+1)sin((2n+1)πv),\displaystyle=2q^{1/4}\sum\limits_{n\geq 0}(-1)^{n}q^{n(n+1)}\sin((2n+1)\pi v),
(5.5) θ4(v|τ)\displaystyle{\theta}_{4}(v|\tau) =1+2n1(1)nqn2cos(2nπv).\displaystyle=1+2\sum\limits_{n\geq 1}(-1)^{n}q^{n^{2}}\cos(2n\pi v).

We will need the following modular transformations of theta functions θ1(v|τ){\theta}_{1}(v|\tau) and θ4(v|τ){\theta}_{4}(v|\tau).

(5.6) θ1(v|τ)\displaystyle{\theta}_{1}(v|\tau) =iq1/4eπivθ4(vτ2|τ),\displaystyle={\textnormal{i}}q^{1/4}e^{-\pi{\textnormal{i}}v}{\theta}_{4}(v-\frac{\tau}{2}|\tau),
(5.7) θ1(v|τ)\displaystyle{\theta}_{1}(v|\tau) =iiτeπiv2/τθ1(vτ|1τ).\displaystyle=i\sqrt{\frac{i}{\tau}}e^{-\pi{\textnormal{i}}v^{2}/\tau}{\theta}_{1}(\frac{v}{\tau}|-\frac{1}{\tau}).

The Jacobi’s triple product identity for θ1(v|τ){\theta}_{1}(v|\tau) is

(5.8) θ1(v|τ)=2q1/4sin(πv)n1(1q2n)(1q2ne2πiv)(1q2ne2πiv).{\theta}_{1}(v|\tau)=2q^{1/4}\sin(\pi v)\prod\limits_{n\geq 1}(1-q^{2n})(1-q^{2n}e^{2\pi{\textnormal{i}}v})(1-q^{2n}e^{-2\pi{\textnormal{i}}v}).

That is,

(5.9) θ1(v|τ)=2q1/4sin(πv)(q2,q2e2πiv,q2e2πiv;q2).{\theta}_{1}(v|\tau)=2q^{1/4}\sin(\pi v)(q^{2},q^{2}e^{2\pi{\textnormal{i}}v},q^{2}e^{-2\pi{\textnormal{i}}v};q^{2})_{\infty}.\\

All of these results can be found in (Rademacher,, 1973, Chapter 10).

The next results also require the following elementary bounds: for all uu\in{\mathds{R}}

(5.10) e|u|usinh(u)10e4|u|/5 and |u1eu|1+|u|.e^{-|u|}\leq\frac{u}{\sinh(u)}\leq 10e^{-4|u|/5}\;\;{\textnormal{ and }}\;\;\Big{|}\frac{u}{1-e^{u}}\Big{|}\leq 1+|u|.

We leave their proofs to the reader.

Lemma 5.4.

Let q=e2/Mq=e^{-2/M}. There exists MM^{*} and a universal constant 0<C<0<C<\infty such that

(5.11) 1Ceπ2|x|<|Γq(1+ix)|<Ce3π20|x|\frac{1}{C}e^{-\frac{\pi}{2}|x|}<|\Gamma_{q}(1+{\textnormal{i}}x)|<Ce^{-\frac{3\pi}{20}|x|}

for all MMM\geq M^{*} and all x[πM/2,πM/2]x\in[-\pi M/2,\pi M/2].

Proof.

In view of the product formula (5.8) combined with (5.2) we have the following representation

(5.12) |Γq(1+ix)|2=2sin(xM)e14M(q;q)3θ1(xMπ|iMπ).|\Gamma_{q}(1+{\textnormal{i}}x)|^{2}=2\sin(\tfrac{x}{M})\frac{e^{-\frac{1}{4M}}(q;q)_{\infty}^{3}}{\theta_{1}\left(\frac{x}{M\pi}|\frac{{\textnormal{i}}}{M\pi}\right)}.

We use (5.7) to rewrite (5.12) as follows

|Γq(1+ix)|2=((q;q)exp(π2M12)Mπ)3e14Msin(x/M)x/M2sinh(πx)v(x,M)πxsinh(πx)ex2M,|\Gamma_{q}(1+{\textnormal{i}}x)|^{2}=\left(\frac{(q;q)_{\infty}\,\exp(\frac{\pi^{2}M}{12})}{\sqrt{M\pi}}\right)^{3}\,e^{-\frac{1}{4M}}\,\frac{\sin(x/M)}{x/M}\,\frac{2\sinh(\pi x)}{v(x,M)}\,\frac{\pi x}{\sinh(\pi x)}\,e^{\frac{x^{2}}{M}},

where

(5.13) v(x,M):=ieπ2M/4θ1(ix|iπM)=2sinh(πx)(1+S(x,M)),v(x,M):=-{\textnormal{i}}e^{\pi^{2}M/4}\theta_{1}({\textnormal{i}}x|{\textnormal{i}}\pi M)=2\sinh(\pi x)\left(1+S(x,M)\right),

with

S(x,M):=n1(1)n×sinh((2n+1)πx)sinh(πx)eπ2Mn(n+1).S(x,M):=\sum\limits_{n\geq 1}(-1)^{n}\times\frac{\sinh((2n+1)\pi x)}{\sinh(\pi x)}\;e^{-\pi^{2}Mn(n+1)}.

In view of formula (35) in Zhang, (2008) with γ=π\gamma=\pi, n=M2n=M^{2} and a=1/2a=1/2 (or (Corwin and Knizel,, 2021, Proposition 2.3)) we have

(q;q)exp(π2M12)Mπ1as M.\frac{(q;q)_{\infty}\,\exp(\frac{\pi^{2}M}{12})}{\sqrt{M\pi}}\to 1\quad\mbox{as $M\to\infty$}.

Consequently, for δ>0\delta>0 and MM large enough we have

(5.14) 1δ<((q;q)exp(π2M12)Mπ)3e14M<1+δ.1-\delta<\left(\frac{(q;q)_{\infty}\,\exp(\tfrac{\pi^{2}M}{12})}{\sqrt{M\pi}}\right)^{3}\,e^{-\frac{1}{4M}}<1+\delta.

For x[πM2,πM2]x\in[-\frac{\pi M}{2},\frac{\pi M}{2}], applying bounds (5.10) we obtain:

|sinh((2n+1)πx)sinh(πx)|=(2n+1)|sinh((2n+1)πx)(2n+1)πx|×|πxsinh(πx)|\displaystyle\bigg{|}\frac{\sinh((2n+1)\pi x)}{\sinh(\pi x)}\bigg{|}=(2n+1)\bigg{|}\frac{\sinh((2n+1)\pi x)}{(2n+1)\pi x}\bigg{|}\times\bigg{|}\frac{\pi x}{\sinh(\pi x)}\bigg{|}
(2n+1)×e(2n+1)π|x|×10e4π|x|/510(2n+1)eπ2M(n+1/2).\displaystyle\;\;\;\leq(2n+1)\times e^{(2n+1)\pi|x|}\times 10e^{-4\pi|x|/5}\leq 10(2n+1)e^{\pi^{2}M(n+1/2)}.

Consequently,

|S(x,M)|\displaystyle|S(x,M)| 10n1(2n+1)eπ2M(n21/2)\displaystyle\leq 10\sum\limits_{n\geq 1}(2n+1)e^{-\pi^{2}M(n^{2}-1/2)}
10eπ2M/2×[n1(2n+1)eπ2(n21)]=C1eπ2M/2\displaystyle\leq 10e^{-\pi^{2}M/2}\times\Big{[}\sum\limits_{n\geq 1}(2n+1)e^{-\pi^{2}(n^{2}-1)}\Big{]}=C_{1}e^{-\pi^{2}M/2}

for an absolute constant C1>0C_{1}>0. Therefore, for ε>0\varepsilon>0 and for MM large enough we have

(5.15) 1εv(x,M)2sinh(πx)1+ε,1-\varepsilon\leq\frac{v(x,M)}{2\sinh(\pi x)}\leq 1+\varepsilon,

for all x[πM/2,πM/2]x\in[-\pi M/2,\pi M/2]. In other words, we proved that

(5.16) ieπ2M/4θ1(ix|iπM)=2sinh(πx)(1+o(1)),M,-{\textnormal{i}}e^{\pi^{2}M/4}\theta_{1}({\textnormal{i}}x|{\textnormal{i}}\pi M)=2\sinh(\pi x)(1+o(1)),\;\;\;M\to\infty,\;\;

uniformly in x[πM/2,πM/2]x\in[-\pi M/2,\pi M/2].

Taking into account (5.14), (5.15) and the trivial estimate

2πsin(x/M)x/M1,x[πM2,πM2],\frac{2}{\pi}\leq\frac{\sin(x/M)}{x/M}\leq 1,\quad x\in[-\tfrac{\pi M}{2},\tfrac{\pi M}{2}],

we get

2π1δ1+εex2/Mπ|x||sinh(πx)||Γq(1+ix)|21+δ1εex2/Mπ|x||sinh(πx)|\frac{2}{\pi}\frac{1-\delta}{1+\varepsilon}\,e^{x^{2}/M}\frac{\pi|x|}{|\sinh(\pi x)|}\leq|\Gamma_{q}(1+{\textnormal{i}}x)|^{2}\leq\frac{1+\delta}{1-\varepsilon}\,e^{x^{2}/M}\frac{\pi|x|}{|\sinh(\pi x)|}

which holds for MM large enough and for all x[πM/2,πM/2]x\in[-{\pi M}/2,{\pi M}/2]. Setting ε=δ=1/2\varepsilon=\delta=1/2 we get

(5.17) 23ππ|x||sinh(πx)|<23πex2/Mπ|x||sinh(πx)|<|Γq(1+ix)|2<3ex2/Mπ|x||sinh(πx)|.\frac{2}{3\pi}\frac{\pi|x|}{|\sinh(\pi x)|}<\frac{2}{3\pi}e^{x^{2}/M}\frac{\pi|x|}{|\sinh(\pi x)|}<|\Gamma_{q}(1+{\textnormal{i}}x)|^{2}<3e^{x^{2}/M}\frac{\pi|x|}{|\sinh(\pi x)|}.

In the range x[πM/2,πM/2]x\in[-\pi M/2,\pi M/2] we have x2/Mπ|x|/2x^{2}/M\leq\pi|x|/2. This fact, the bounds (5.10) and (5.17) imply that for all MM large enough

23πeπ|x|<|Γq(1+ix)|2<30e310π|x|,\frac{2}{3\pi}e^{-\pi|x|}<|\Gamma_{q}(1+{\textnormal{i}}x)|^{2}<30e^{-\frac{3}{10}\pi|x|},

from which the desired result (5.11) follows by taking the square root. ∎

Lemma 5.5.

Let q=e2/Mq=e^{-2/M} and assume that β(π/4,3π/4)\beta\in(\pi/4,3\pi/4). There exist A>0A>0 and MM^{*} (which may depend on β\beta) such that

(5.18) |Γq(1iβM+ix)||Γq(1iβM)|<Aeπx/2\frac{|\Gamma_{q}(1-{\textnormal{i}}\beta M+{\textnormal{i}}x)|}{|\Gamma_{q}(1-{\textnormal{i}}\beta M)|}<Ae^{\pi x/2}

for all MMM\geq M^{*} and x[0,πM/2]x\in[0,\pi M/2].

Proof.

Expression (5.12) with modular transformation (5.7) applied twice gives

(5.19) |Γq(1iβM+ix)|2|Γq(1iβM)|2=sin(βx/M)sin(β)θ1(iβM|iπM)θ1(i(βMx)|iπM)ex2/M2βx.\frac{|\Gamma_{q}(1-{\textnormal{i}}\beta M+{\textnormal{i}}x)|^{2}}{|\Gamma_{q}(1-{\textnormal{i}}\beta M)|^{2}}=\frac{\sin(\beta-x/M)}{\sin(\beta)}\,\frac{\theta_{1}({\textnormal{i}}\beta M|{\textnormal{i}}\pi M)}{\theta_{1}({\textnormal{i}}(\beta M-x)|{\textnormal{i}}\pi M)}\,\,e^{x^{2}/M-2\beta x}.

We consider two cases. If β(π/4,π/2]\beta\in(\pi/4,\pi/2], then 1/sinβ21/\sin\beta\leq\sqrt{2} and the first arguments in both θ1\theta_{1} functions arguments satisfy the assumptions |βM|Mπ/2|\beta M|\leq M\pi/2 and |βMx|Mπ/2|\beta M-x|\leq M\pi/2. Using estimate (5.16) we obtain

θ1(iβM|iπM)θ1(i(βMx)|iπM)=sinh(πβM)sinh(π(βMx))(1+o(1)),as M,\frac{\theta_{1}({\textnormal{i}}\beta M|{\textnormal{i}}\pi M)}{\theta_{1}({\textnormal{i}}(\beta M-x)|{\textnormal{i}}\pi M)}=\frac{\sinh(\pi\beta M)}{\sinh(\pi(\beta M-x))}(1+o(1)),\;\;\;\mbox{as }\;M\to\infty,

uniformly in x[Mπ2,Mπ2]x\in[-\frac{M\pi}{2},\,\frac{M\pi}{2}]. We also write

sin(βxM)sinh(πβM)sinh(π(βMx))=12πM×sin(βx/M)βx/M×2π(xβM)1e2π(xβM)×(1e2πβM)×eπx.\sin(\beta-\tfrac{x}{M})\frac{\sinh(\pi\beta M)}{\sinh(\pi(\beta M-x))}=\frac{-1}{2\pi M}\times\frac{\sin(\beta-x/M)}{\beta-x/M}\times\frac{2\pi(x-\beta M)}{1-e^{2\pi(x-\beta M)}}\times(1-e^{-2\pi\beta M})\times e^{\pi x}.

The second and fourth factors in the right hand side are bounded by one in absolute value, and using the last bound in (5.10) we see that the product of the first and third factors is bounded by 1+π1+\pi. Using these results and the bound x2/Mπx/2x^{2}/M\leq\pi x/2 on x[0,πM/2]x\in[0,\pi M/2] we see that in the case β(π/4,π/2]\beta\in(\pi/4,\pi/2] we have

|Γq(1iβM+ix)|2|Γq(1iβM)|2<A1eπx/22βx+πx=A1ex(3π/22β)A1eπx\frac{|\Gamma_{q}(1-{\textnormal{i}}\beta M+{\textnormal{i}}x)|^{2}}{|\Gamma_{q}(1-{\textnormal{i}}\beta M)|^{2}}<A_{1}e^{\pi x/2-2\beta x+\pi x}=A_{1}e^{x(3\pi/2-2\beta)}\leq A_{1}e^{\pi x}

for some absolute constant A1A_{1}, and this implies (5.18).

Next we consider the case β(π/2,3π/4)\beta\in(\pi/2,3\pi/4). We use formula (5.6) which together with (5.19) gives

(5.20) |Γq(1iβM+ix)|2|Γq(1iβM)|2=sin(βx/M)sin(β)θ4(iβ~M|iπM)θ4(i(β~Mx)|iπM)ex2/M+(π2β)x\frac{|\Gamma_{q}(1-{\textnormal{i}}\beta M+{\textnormal{i}}x)|^{2}}{|\Gamma_{q}(1-{\textnormal{i}}\beta M)|^{2}}=\frac{\sin(\beta-x/M)}{\sin(\beta)}\,\frac{\theta_{4}({\textnormal{i}}\tilde{\beta}M|{\textnormal{i}}\pi M)}{\theta_{4}({\textnormal{i}}(\tilde{\beta}M-x)|{\textnormal{i}}\pi M)}\,e^{x^{2}/M+(\pi-2\beta)x}

with β~=βπ/2(0,π/4)\tilde{\beta}=\beta-\pi/2\in(0,\pi/4).

We use the same method as in the proof of Lemma 5.4. In view of (5.5) we write

(5.21) θ4(iw|iπM)=1+S~(w,M),\theta_{4}({\textnormal{i}}w|{\textnormal{i}}\pi M)=1+\tilde{S}(w,M),

and estimate

S~(w,M):=2n1(1)neπ2Mn2cosh(2nπw)\tilde{S}(w,M):=2\sum\limits_{n\geq 1}(-1)^{n}e^{-\pi^{2}Mn^{2}}\cosh(2n\pi w)

as follows: for any small ε>0\varepsilon>0, M1M\geq 1 and |w|(1ε)Mπ/2|w|\leq(1-\varepsilon){M\pi}/2 we have

|S~(w,M)|2n1eπ2Mn2+2nπ|w|2n1eπ2Mn2+nπ2(1ε)M==2eπ2Mεn1eπ2M(n2n)(n1)π2εM2eπ2Mεn1eπ2M(n2n)C2eπ2Mε.|\tilde{S}(w,M)|\leq 2\sum\limits_{n\geq 1}e^{-\pi^{2}Mn^{2}+2n\pi|w|}\leq 2\sum\limits_{n\geq 1}e^{-\pi^{2}Mn^{2}+n\pi^{2}(1-\varepsilon)M}=\\ =2e^{-\pi^{2}M\varepsilon}\sum\limits_{n\geq 1}e^{-\pi^{2}M(n^{2}-n)-(n-1)\pi^{2}\varepsilon M}\leq 2e^{-\pi^{2}M\varepsilon}\sum\limits_{n\geq 1}e^{-\pi^{2}M(n^{2}-n)}\leq\,C_{2}e^{-\pi^{2}M\varepsilon}.

for an absolute constant C2>0C_{2}>0,

From the above estimate and (5.21) it follows that for arbitrary ε(0,1)\varepsilon\in(0,1), as M+M\to+\infty,

(5.22) θ4(iw|iπM)=1+o(1),uniformly inw[(1ε)πM2,(1ε)πM2].\theta_{4}({\textnormal{i}}w|{\textnormal{i}}\pi M)=1+o(1),\quad\mbox{uniformly in}\;\;w\in[-(1-\varepsilon)\tfrac{\pi M}{2},(1-\varepsilon)\tfrac{\pi M}{2}].

Since β~(0,π/4)\tilde{\beta}\in(0,\pi/4) and x[0,πM/2]x\in[0,\pi M/2], it is easy to see that

β~Mx,β~M[(1ε)πM2,(1ε)πM2]withε=2β~/π.\tilde{\beta}M-x,\;\tilde{\beta}M\in\left[-(1-\varepsilon)\tfrac{\pi M}{2},(1-\varepsilon)\tfrac{\pi M}{2}\right]\quad\mbox{with}\;\;\varepsilon=2\tilde{\beta}/\pi.

Therefore, applying (5.22) for w=β~Mxw=\tilde{\beta}M-x and w=β~Mw=\tilde{\beta}M to the second factor in (5.20) we conclude that

(5.23) |Γq(1iβM+ix)|2|Γq(1iβM)|2\displaystyle\frac{|\Gamma_{q}(1-{\textnormal{i}}\beta M+{\textnormal{i}}x)|^{2}}{|\Gamma_{q}(1-{\textnormal{i}}\beta M)|^{2}} =sin(βx/M)sin(β)ex2/M+(π2β)x×(1+o(1))\displaystyle=\frac{\sin(\beta-x/M)}{\sin(\beta)}e^{x^{2}/M+(\pi-2\beta)x}\times(1+o(1))

as MM\to\infty uniformly in x[0,πM/2]x\in[0,\pi M/2]. Since β(π/2,3π/4)\beta\in(\pi/2,3\pi/4), the right-hand side in (5.23) can be bounded by Cexp(πx)C\exp(\pi x) for x[0,πM/2]x\in[0,\,\pi M/2] and we again conclude that (5.18) holds. ∎

Lemma 5.6.

Fix uu\in\mathds{R} and let qM=e2/Mq_{M}=e^{-2/M}, M=1,2M=1,2\dots. Then there exist constants A,B,M>0A,B,M_{*}>0 that depend only on uu such that for all M>MM>M_{*} we have

(5.24) 𝟏{|s|Mπ/2}|ΓqM(1+i(s+u))|AeB|s|{\mathbf{1}}_{\{|s|\leq M\pi/2\}}\left|\Gamma_{q_{M}}(1+{\textnormal{i}}(s+u))\right|\leq Ae^{-B|s|}

for all ss\in\mathds{R}.

Proof.

We will show that (5.24) holds with A=Ce3π|u|/20A=Ce^{3\pi|u|/20}, B=3π/40B=3\pi/40 and M=max{M,4|u|/π}M_{*}=\max\{M^{*},4|u|/\pi\}, where MM^{*} and CC are from Lemma 5.4. Let M>MM>M_{*} and |s|Mπ/2|s|\leq M\pi/2.

We consider two cases. If |s+u|Mπ/2|s+u|\leq M\pi/2, then (5.24) follows from Lemma 5.4. On the other hand, if |s+u|>Mπ/2|s+u|>M\pi/2, then with M>M=max{M,4|u|/π}4|u|/πM>M_{*}=\max\{M^{*},4|u|/\pi\}\geq 4|u|/\pi we have |u|Mπ/4|u|\leq M\pi/4. Recalling that |s|Mπ/2|s|\leq M\pi/2, we get Mπ/2<|s+u|Mπ/2+Mπ/4=3Mπ/4M\pi/2<|s+u|\leq M\pi/2+M\pi/4=3M\pi/4.

By Lemma 5.3, function x|Γq(1+ix)|x\mapsto|\Gamma_{q}(1+{\textnormal{i}}x)| is periodic with period MπM\pi and is increasing on the interval [Mπ/2,Mπ/4][-M\pi/2,-M\pi/4]. With x=s+u[Mπ/2,3Mπ/4]x=s+u\in[M\pi/2,3M\pi/4] we have xMπ[Mπ/2,Mπ/4]x-M\pi\in[-M\pi/2,-M\pi/4] so we get

|Γq(1+ix)|=|Γq(1+i(xMπ))||Γq(1iMπ/4)|Ce3π280MCe3π40|s|,\left|\Gamma_{q}(1+{\textnormal{i}}x)\right|=\left|\Gamma_{q}(1+{\textnormal{i}}(x-M\pi))\right|\leq\left|\Gamma_{q}(1-{\textnormal{i}}M\pi/4)\right|\leq Ce^{-\frac{3\pi^{2}}{80}M}\leq Ce^{-\frac{3\pi}{40}|s|},

where we used Lemma 5.4 and then the bound M2π|s|M\geq\tfrac{2}{\pi}|s|. Thus in this case (5.24) follows, too.

We also need the following result.

Lemma 5.7.

There exist constants A,B,M>0A,B,M_{*}>0 such that with q=e2/Mq=e^{-2/M}, for all M>MM>M_{*} and all uu\in\mathds{R} we have

(5.25) 𝟏{|u|<Mπ/2}1|Γq(iu)|2Au2eB|u|.{\mathbf{1}}_{\{|u|<M\pi/2\}}\frac{1}{|\Gamma_{q}({\textnormal{i}}u)|^{2}}\leq Au^{2}e^{B|u|}.
Proof.

Definition (5.2) yields

1|Γq(iu)|2=|1qiu|2(1q)21|Γq(1+iu)|2.\frac{1}{|\Gamma_{q}({\textnormal{i}}u)|^{2}}=\frac{|1-q^{{\textnormal{i}}u}|^{2}}{(1-q)^{2}}\frac{1}{|\Gamma_{q}(1+{\textnormal{i}}u)|^{2}}.

Since |1qiu|2=4sin2(u/M)|1-q^{{\textnormal{i}}u}|^{2}=4\sin^{2}(u/M). we get

|1qiu|2(1q)2=4u21(M(1e2/M))2(sin(u/M)u/M)2Cu2\frac{|1-q^{{\textnormal{i}}u}|^{2}}{(1-q)^{2}}=4u^{2}\,\frac{1}{(M(1-e^{-2/M}))^{2}}\,\left(\frac{\sin(u/M)}{u/M}\right)^{2}\leq Cu^{2}

Thus (5.25) follows from Lemma 5.4. ∎

5.3. Additional technical lemmas

The following technical Lemma is needed in proof of Theorem 3.3.

Lemma 5.8.

Let a~,b~\widetilde{a},\widetilde{b} be either real or complex conjugates. If Re(a~)0\textnormal{Re}(\widetilde{a})\geq 0, Re(b~)0\textnormal{Re}(\widetilde{b})\geq 0, c0c\geq 0 then for ss\in\mathds{R} and all 0q<10\leq q<1 we have

(5.26) |(a~q1+c+is,b~q1+c+is;q)(a~q,b~q;q)|1.\left|\frac{(-\widetilde{a}q^{1+c+{\textnormal{i}}s},-\widetilde{b}q^{1+c+{\textnormal{i}}s};q)_{\infty}}{(-\widetilde{a}q,-\widetilde{b}q;q)_{\infty}}\right|\leq 1.
Proof.

We first prove the bound for the case of complex-conjugate parameters. Write a~=reiα\widetilde{a}=re^{{\textnormal{i}}\alpha}, b~=reiα\widetilde{b}=re^{-{\textnormal{i}}\alpha} with α[π/2,π/2]\alpha\in[-\pi/2,\pi/2] and r>0r>0. (The case r=0r=0 is trivial.) Consider one factor of the infinite product:

|(1+rqn+c+iseiα)(1+rqn+c+iseiα)||(1+rqneiα)(1+rqneiα)|=|eiα+rqn+c+is||eiα+rqn+c+is||eiα+rqn||eiα+rqn|,n1.\frac{|(1+rq^{n+c+{\textnormal{i}}s}e^{{\textnormal{i}}\alpha})(1+rq^{n+c+{\textnormal{i}}s}e^{-{\textnormal{i}}\alpha})|}{|(1+rq^{n}e^{{\textnormal{i}}\alpha})(1+rq^{n}e^{-{\textnormal{i}}\alpha})|}=\frac{|e^{{\textnormal{i}}\alpha}+rq^{n+c+{\textnormal{i}}s}|\cdot|e^{-{\textnormal{i}}\alpha}+rq^{n+c+{\textnormal{i}}s}|}{|e^{{\textnormal{i}}\alpha}+rq^{n}|\cdot|e^{-{\textnormal{i}}\alpha}+rq^{n}|},\;n\geq 1.

As ss varies through the real line, the point z=rqn+c+isz=-rq^{n+c+{\textnormal{i}}s} moves on the circle |z|=rqn+c|z|=rq^{n+c}. One readily checks that the product in the numerator is largest when Re(z)\textnormal{Re}(z) is minimal, i.e., when s=0s=0. Thus

|(1+rqn+c+iseiα)(1+rqn+c+iseiα)||(1+rqneiα)(1+rqneiα)||(1+rqn+ceiα)(1+rqn+ceiα)||(1+rqneiα)(1+rqneiα)|.\frac{|(1+rq^{n+c+{\textnormal{i}}s}e^{{\textnormal{i}}\alpha})(1+rq^{n+c+{\textnormal{i}}s}e^{-{\textnormal{i}}\alpha})|}{|(1+rq^{n}e^{{\textnormal{i}}\alpha})(1+rq^{n}e^{-{\textnormal{i}}\alpha})|}\leq\frac{|(1+rq^{n+c}e^{{\textnormal{i}}\alpha})(1+rq^{n+c}e^{-{\textnormal{i}}\alpha})|}{|(1+rq^{n}e^{{\textnormal{i}}\alpha})(1+rq^{n}e^{-{\textnormal{i}}\alpha})|}.

To end the proof, we note that

|1+rqn+ceiα|2|1+rqneiα|21.\frac{|1+rq^{n+c}e^{{\textnormal{i}}\alpha}|^{2}}{|1+rq^{n}e^{{\textnormal{i}}\alpha}|^{2}}\leq 1.

The proof for the case of real a~,b~[0,)\widetilde{a},\widetilde{b}\in[0,\infty) requires some minor modifications and is omitted. (This case is not used in the proof of Theorem 3.3.) ∎

The proof of Theorem 4.3 uses the following technical estimate:

Lemma 5.9.

Let q=e2/Mq=e^{-2/M}. If a=qeiαa=-qe^{{\textnormal{i}}\alpha}, b=qeiαb=-qe^{-{\textnormal{i}}\alpha} and |α|<π/2|\alpha|<\pi/2 then there exist A>0A>0, B>0B>0 and MM_{*} such that

(5.27) 𝟏{|u|<πM/2}(a,b;q)2|(aqiu/2,bqiu/2;q)|2AeB|u|{\mathbf{1}}_{\{|u|<\pi M/2\}}\frac{(a,b;q)_{\infty}^{2}}{|(aq^{{\textnormal{i}}u/2},bq^{{\textnormal{i}}u/2};q)_{\infty}|^{2}}\leq Ae^{B|u|}

for all MMM\geq M_{*} and all uu\in\mathds{R}.

Proof.

Replacing uu by u-u if needed, without loss of generality we may assume u0u\geq 0. In view of (5.2) the left hand side of (5.27) can be rewritten in terms of qq-Gamma functions. We have

|(a;q)||(aqiu/2;q)|=|Γq(1iβM+iu/2)||Γq(1iβM)|,\frac{|(a;q)_{\infty}|}{|(aq^{{\textnormal{i}}u/2};q)_{\infty}|}=\frac{|\Gamma_{q}(1-{\textnormal{i}}\beta M+{\textnormal{i}}u/2)|}{|\Gamma_{q}(1-{\textnormal{i}}\beta M)|},

where β=(α+π)/2(π/4,3π/4)\beta=(\alpha+\pi)/2\in(\pi/4,3\pi/4). Similarly,

|(b;q)||(bqiu/2;q)|=|Γq(1iβM+iu/2)||Γq(1iβM)|,\frac{|(b;q)_{\infty}|}{|(bq^{{\textnormal{i}}u/2};q)_{\infty}|}=\frac{|\Gamma_{q}(1-{\textnormal{i}}\beta^{\prime}M+{\textnormal{i}}u/2)|}{|\Gamma_{q}(1-{\textnormal{i}}\beta^{\prime}M)|},

where β=(πα)/2(π/4,3π/4)\beta^{\prime}=(\pi-\alpha)/2\in(\pi/4,3\pi/4).

Therefore inequality (5.27) follows from Lemma 5.5.∎

Acknowledgements

The first author was supported in part by Simons Foundation/SFARI Award Number: 703475, US. The second author was supported in part by Natural Sciences and Engineering Research Council of Canada. The authors acknowledge support from the Taft Research Center at the University of Cincinnati.

References

  • Adiga et al., (1985) Adiga, C., Berndt, B. C., Bhargava, S., and Watson, G. N. (1985). Ch 16 of Ramanujan’s Second Notebook: Theta-Functions and qq-Series: Theta Functions and Q-series, volume 315. American Mathematical Soc.
  • Aptekarev, (1993) Aptekarev, A. I. (1993). Asymptotics of orthogonal polynomials in a neighborhood of the endpoints of the interval of orthogonality. Sbornik: Mathematics, 76(1):35.
  • Baik et al., (2003) Baik, J., Kriecherbauer, T., McLaughlin, K.-R., and Miller, P. D. (2003). Uniform asymptotics for polynomials orthogonal with respect to a general class of discrete weights and universality results for associated ensembles: announcement of results. International Mathematics Research Notices, 2003(15):821–858.
  • Barraquand and Corwin, (2023) Barraquand, G. and Corwin, I. (2023). Stationary measures for the log-gamma polymer and KPZ equation in half-space. Ann. Probab., 51(5):1830–1869.
  • Barraquand and Le Doussal, (2022) Barraquand, G. and Le Doussal, P. (2022). Steady state of the KPZ equation on an interval and Liouville quantum mechanics. Europhysics Letters, 137(6):61003. ArXiv preprint with Supplementary material: https://arxiv.org/abs/2105.15178.
  • Barraquand and Le Doussal, (2023) Barraquand, G. and Le Doussal, P. (2023). Stationary measures of the KPZ equation on an interval from Enaud-Derrida’s matrix product ansatz representation. J. Phys. A, 56(14):Paper No. 144003, 14.
  • Berg and Ismail, (1996) Berg, C. and Ismail, M. E. H. (1996). qq-Hermite polynomials and classical orthogonal polynomials. Canad. J. Math., 48(1):43–63.
  • Berndt, (1991) Berndt, B. C. (1991). Ramanujan’s Notebooks. Part III. Springer-Verlag, New York.
  • Billingsley, (1999) Billingsley, P. (1999). Convergence of Probability Measures. Wiley Series in Probability and Statistics: Probability and Statistics. John Wiley & Sons Inc., New York, second edition. A Wiley-Interscience Publication.
  • Bryc and Kuznetsov, (2022) Bryc, W. and Kuznetsov, A. (2022). Markov limits of steady states of the KPZ equation on an interval. ALEA, Lat. Am. J. Probab. Math. Stat., 19:1329–1351. (http://arxiv.org/abs/2109.04462).
  • Bryc and Wang, (2023) Bryc, W. and Wang, Y. (2023). Limit theorems for random Motzkin paths near boundary. Bernoulli (to appear). http://arxiv.org/abs/2304.00924.
  • Bryc and Wang, (2024) Bryc, W. and Wang, Y. (2024). Fluctuations of random Motzkin paths II. ALEA, Lat. Am. J. Probab. Math. Stat., 21:73–94. http://arxiv.org/abs/2304.12975.
  • Corwin and Knizel, (2021) Corwin, I. and Knizel, A. (2021). Stationary measure for the open KPZ equation. Communications on Pure and Applied Mathematics, to appear. arXiv preprint arXiv:2103.12253, DOI: 10.1002/cpa.22174.
  • Deift et al., (2001) Deift, P., Kriecherbauer, T., McLaughlin, K.-R., Venakides, S., and Zhou, X. (2001). A Riemann–Hilbert approach to asymptotic questions for orthogonal polynomials. Journal of Computational and Applied Mathematics, 133(1-2):47–63.
  • Deift et al., (1999) Deift, P., Kriecherbauer, T., McLaughlin, K. T.-R., Venakides, S., and Zhou, X. (1999). Strong asymptotics of orthogonal polynomials with respect to exponential weights. Communications on Pure and Applied Mathematics, 52(12):1491–1552.
  • Flajolet and Sedgewick, (2009) Flajolet, P. and Sedgewick, R. (2009). Analytic Combinatorics. Cambridge University Press, Cambridge.
  • Gasper and Rahman, (2004) Gasper, G. and Rahman, M. (2004). Basic Hypergeometric Series, volume 96. Cambridge university press.
  • Ismail, (2005) Ismail, M. E. (2005). Asymptotics of qq-orthogonal polynomials and a qq-Airy function. International Mathematics Research Notices, 2005(18):1063–1088.
  • Ismail and Li, (2013) Ismail, M. E. and Li, X. (2013). Plancherel–Rotach asymptotics for qq-orthogonal polynomials. Constructive Approximation, 37:341–356.
  • Ismail et al., (2022) Ismail, M. E., Zhang, R., and Zhou, K. (2022). Orthogonal polynomials of Askey-Wilson type. arXiv preprint arXiv:2205.05280.
  • Ismail, (1986) Ismail, M. E. H. (1986). Asymptotics of the Askey-Wilson and qq-Jacobi polynomials. SIAM J. Math. Anal., 17(6):1475–1482.
  • Ismail, (2009) Ismail, M. E. H. (2009). Classical and Quantum Orthogonal Polynomials in One Variable, volume 98 of Encyclopedia of Mathematics and its Applications. Cambridge University Press, Cambridge.
  • Ismail and Wilson, (1982) Ismail, M. E. H. and Wilson, J. A. (1982). Asymptotic and generating relations for the qq-Jacobi and φ34{}_{4}\varphi_{3} polynomials. J. Approx. Theory, 36(1):43–54.
  • Kaigh, (1976) Kaigh, W. D. (1976). An invariance principle for random walk conditioned by a late return to zero. The Annals of Probability, pages 115–121.
  • Karlin and McGregor, (1959) Karlin, S. and McGregor, J. (1959). Random walks. Illinois Journal of Mathematics, 3(1):66–81.
  • Keener, (1992) Keener, R. W. (1992). Limit theorems for random walks conditioned to stay positive. The Annals of Probability, pages 801–824.
  • Koekoek et al., (2010) Koekoek, R., Lesky, P. A., Swarttouw, R. F., Koekoek, R., Lesky, P. A., and Swarttouw, R. F. (2010). Hypergeometric Orthogonal Polynomials. Springer.
  • Koekoek and Swarttouw, (1994) Koekoek, R. and Swarttouw, R. F. (1994). The askey-scheme of hypergeometric orthogonal polynomials and its qq-analogue. https://homepage.tudelft.nl/11r49/askey/.
  • Koornwinder, (1990) Koornwinder, T. H. (1990). Jacobi functions as limit cases of qq-ultraspherical polynomials. Journal of Mathematical Analysis and Applications, 148(1):44–54.
  • Kuijlaars et al., (2004) Kuijlaars, A. B. J., McLaughlin, K. T.-R., Van Assche, W., and Vanlessen, M. (2004). The Riemann-Hilbert approach to strong asymptotics for orthogonal polynomials on [1,1][-1,1]. Adv. Math., 188(2):337–398.
  • Levin and Lubinsky, (2020) Levin, E. and Lubinsky, D. S. (2020). Local limits for orthogonal polynomials for varying measures via universality. Journal of Approximation Theory, 254:105394.
  • Lubinsky, (2020) Lubinsky, D. S. (2020). Pointwise asymptotics for orthonormal polynomials at the endpoints of the interval via universality. International Mathematics Research Notices, 2020(4):961–982.
  • Nevai, (1984) Nevai, P. (1984). Asymptotics for orthogonal polynomials associated with exp(x4)\exp(-x^{4}). SIAM journal on mathematical analysis, 15(6):1177–1187.
  • Olver et al., (2023) Olver, F. W. J., Daalhuis, A. B. O., Lozier, D. W., Schneider, B. I., Boisvert, R. F., Clark, C. W., Millerand, B. R., Saunders, B. V., Cohl, H. S., and McClain, M. A. (2023). NIST digital library of mathematical functions. https://dlmf.nist.gov/, 2023-12-15 DLMF Update; Version 1.1.12.
  • Rademacher, (1973) Rademacher, H. (1973). Topics in Analytic Number Theory. Die Grundlehren der mathematischen Wissenschaften, Band 169. Springer-Verlag, New York-Heidelberg. Edited by E. Grosswald, J. Lehner and M. Newman.
  • Szwarc, (1992) Szwarc, R. (1992). Connection coefficients of orthogonal polynomials. Canadian Mathematical Bulletin, 35(4):548–556.
  • Viennot, (1985) Viennot, G. (1985). A combinatorial theory for general orthogonal polynomials with extensions and applications. In Brezinski, C., Draux, A., Magnus, A. P., Maroni, P., and Ronveaux, A., editors, Polynômes Orthogonaux et Applications, pages 139–157, Berlin, Heidelberg. Springer Berlin Heidelberg.
  • Yakubovich, (2011) Yakubovich, S. (2011). The heat kernel and Heisenberg inequalities related to the Kontorovich-Lebedev transform. Commun. Pure Appl. Anal., 10(2):745–760.
  • Zhang, (2008) Zhang, R. (2008). On asymptotics of qq-gamma functions. J. Math. Anal. Appl., 339(2):1313–1321.