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

Asymptotics of the Discrete Chebyshev Polynomials

J. H. Pana{}^{\text{a}} and R. Wongb{}^{\text{b}}
Abstract

The discrete Chebyshev polynomials tn(x,N)t_{n}(x,N) are orthogonal with respect to a distribution, which is a step function with jumps one unit at the points x=0,1,,N1x=0,1,\cdots,N-1, NN being a fixed positive integer. By using a double integral representation, we have recently obtained asymptotic expansions for tn(aN,N+1)t_{n}(aN,N+1) in the double scaling limit, namely, NN\rightarrow\infty and n/Nbn/N\rightarrow b, where b(0,1)b\in(0,1) and a(,)a\in(-\infty,\infty); see [Studies in Appl. Math. 128 (2012), 337-384]. In the present paper, we continue to investigate the behaviour of these polynomials when the parameter bb approaches the endpoints of the interval (0,1)(0,1). While the case b1b\rightarrow 1 is relatively simple (since it is very much like the case when bb is fixed), the case b0b\rightarrow 0 is quite complicated. The discussion of the latter case is divided into several subcases, depending on the quantities nn, xx and xN/n2xN/n^{2}, and different special functions have been used as approximants, including Airy, Bessel and Kummer functions.

a{}^{\text{a}} Department of Mathematics, City University of Hong Kong, Tat Chee Avenue, Kowloon, Hong Kong
b{}^{\text{b}} Liu Bie Ju Centre for Mathematical Sciences, City University of Hong Kong, Tat Chee Avenue, Kowloon, Hong Kong

1 INTRODUCTION

The discrete Chebyshev polynomials tn(x,N)t_{n}(x,N) can be defined as a special case of the Hahn polynomials [1, p.174]

(1.1) Qn(x;α,β,N)=F23(n,x,n+α+β+1;N,α+1;1)=k=0n(n)k(x)k(n+α+β+1)k(N)k(α+1)kk!.\begin{split}Q_{n}(x;\alpha,\beta,N)&={{}_{3}F_{2}}(-n,-x,n+\alpha+\beta+1;-N,\alpha+1;1)\\ &=\sum_{k=0}^{n}\frac{(-n)_{k}(-x)_{k}(n+\alpha+\beta+1)_{k}}{(-N)_{k}(\alpha+1)_{k}k!}.\end{split}

With α=β=0\alpha=\beta=0 in (1.1), we have

(1.2) tn(x,N)=(1)n(Nn)nQn(x;0,0,N1)=(1)n(Nn)nk=0n(n)k(x)k(n+1)k(N+1)kk!k!;\begin{split}t_{n}(x,N)&=(-1)^{n}(N-n)_{n}Q_{n}(x;0,0,N-1)\\ &=(-1)^{n}(N-n)_{n}\sum_{k=0}^{n}\frac{(-n)_{k}(-x)_{k}(n+1)_{k}}{(-N+1)_{k}k!k!};\end{split}

see [1, p.176].

In a recent paper [8], we have studied the asymptotic behaviour of tn(x,N+1)t_{n}(x,N+1) as nn, NN\rightarrow\infty in such a way that the ratios

(1.3) a=x/Nandb=n/N,a=x/N\qquad\text{and}\qquad b=n/N,

satisfy the inequalities

(1.4) <a12and0<b<1.-\infty<a\leq\frac{1}{2}\qquad\text{and}\qquad 0<b<1.

In view of the symmetry relation [8, eq.(8)]

(1.5) tn(x,N+1)=(1)ntn(Nx,N+1),t_{n}(x,N+1)=(-1)^{n}t_{n}(N-x,N+1),

our study actually covers the entire real-axis <x<-\infty<x<\infty or, equivalently, the entire parameter range <a<-\infty<a<\infty. For a discussion of the asymptotic behaviour of the Hahn polynomials, see [4].

In the present paper, we shall investigate the behaviour of the polynomials tn(x,N+1)t_{n}(x,N+1), when the parameter bb given in (1.3) either tends to 0 or tends to 11. The case b1b\rightarrow 1 turns out to be rather simple, since the ultimate expansions and their derivations are similar to those in the case when bb is fixed; see (2.21) for 0a120\leq a\leq\frac{1}{2} and (2.28) for a<0a<0.

To derive asymptotic approximation for tn(x,N+1)t_{n}(x,N+1) when b0b\rightarrow 0, we divide our discussion into several cases, depending on the quantities nn, xx and xN/n2xN/n^{2}. A summary of our findings is given in Table 1 below.

Table 1: Asymptotic expansions in different cases when nN0\frac{n}{N}\rightarrow 0 and NN\rightarrow\infty

xx xN/n2xN/n^{2} small fixed large
large Kummer Airy Bessel
fixed Kummer Airy series(1.2)
small Kummer / series(1.2) Airy series(1.2)

Some explanation is needed for this table. By “small”, “fixed” and “large”, we mean, respectively, x=o(1)x=o(1), x[δ,M]x\in[\delta,M] and xO(1)x\gg O(1) as NN\rightarrow\infty, where δ\delta is a small positive number and MM is a large positive number. By “Airy”, we mean an asymptotic expansion whose associated approximants are the Airy functions Ai(z)\text{Ai}(z) and Bi(z)\text{Bi}(z). In the same manner, by “Bessel” and “Kummer”, we mean asymptotic expansions whose associated approximants are, respectively, the Bessel function Jv(z)J_{v}(z) and the Kummer function M(a,c,z)M(a,c,z). The Kummer function used in this paper is defined by

(1.6) 𝐌(a,c,z)=1Γ(c)M(a,c,z)=1Γ(c)s=0(a)s(c)szss!,\mathbf{M}(a,c,z)=\frac{1}{\Gamma(c)}M(a,c,z)=\frac{1}{\Gamma(c)}\sum_{s=0}^{\infty}\frac{(a)_{s}}{(c)_{s}}\frac{z^{s}}{s!},

which was introduced in [6, p.255]. The case involving Bessel functions covers an earlier result of Sharapodinov [10], where, instead of Bessel functions, Jacobi polynomials are used as an approximant. In the case when xN/n2xN/n^{2} is large and xx is either fixed or small, the series in (1.2) is itself an asymptotic expansion (in the generalized sense [11, p.10]) as NN\rightarrow\infty. Hence, there is no need to seek for another asymptotic representation.

The arrangement of the present paper is as follows. In Section 2, we recall the major results in [8]; to facilitate our presentation for later sections, we also include here brief derivations of the expansions given in [8]. In Section 3, we consider the case when xN/n2xN/n^{2} is small and show that the expansions are of Kummer-type. In Section 4, we consider the case when the quantity xN/n2xN/n^{2} is fixed; that is, when xN/n2xN/n^{2} is bounded away from zero and infinity. There are three subcases, depending on xx being large, fixed or small. In all three subcases, we will show that the expansions are of Airy-type; see Table 1. The case when both quantities xx and xN/n2xN/n^{2} are large is dealt with in Section 5, where we show that the expansion of tn(x,N+1)t_{n}(x,N+1) can be expressed in terms of Bessel functions. The final section is devoted to two remaining cases, namely, (i) when xN/n2xN/n^{2} is large and xx is either fixed or small, and (ii) x<0x<0.

2 RESULTS IN REFERENCE [8]

In [8], we have divided our discussion into two cases: (i) 0a120\leq a\leq\frac{1}{2}, and (ii) a<0a<0. As mentioned in Section 1, by virtue of the symmetry relation (1.5), these two cases cover the entire range <a<-\infty<a<\infty.

In case (i), we started with the integral representation

(2.1) tn(x,N+1)=(1)n2πiΓ(n+N+2)Γ(n+1)Γ(Nn+1)01γ11w1eNf(t,w)dwdt,t_{n}(x,N+1)=\frac{(-1)^{n}}{2\pi i}\frac{\Gamma(n+N+2)}{\Gamma(n+1)\Gamma(N-n+1)}\int_{0}^{1}\int_{\gamma_{1}}\frac{1}{w-1}e^{Nf(t,w)}\mathrm{d}w\mathrm{d}t,

where

(2.2) f(t,w)=bln(1t)+(1b)lnt+alnwaln(w1)+bln[1(1t)w],f(t,w)=b\ln(1-t)+(1-b)\ln t+a\ln w-a\ln(w-1)+b\ln[1-(1-t)w],

and the curve γ1\gamma_{1} starts at w=0w=0, runs along the lower edge of the positive real line towards w=1w=1, encircles the point w=1w=1 in the counterclockwise direction and returns to the origin along the upper edge of the positive real line. As a function of two variables, the partial derivatives f/t\partial f/\partial t and f/w\partial f/\partial w vanish at (t,w)=(t±,w±)(t,w)=(t_{\pm},w_{\pm}), where

(2.3) t+=22ab2+bb24a+4a22(1a)(1+b),w+=b+b24a+4a22b,t_{+}=\frac{2-2a-b^{2}+b\sqrt{b^{2}-4a+4a^{2}}}{2(1-a)(1+b)},\qquad w_{+}=\frac{b+\sqrt{b^{2}-4a+4a^{2}}}{2b},

and

(2.4) t=22ab2bb24a+4a22(1a)(1+b),w=bb24a+4a22b.t_{-}=\frac{2-2a-b^{2}-b\sqrt{b^{2}-4a+4a^{2}}}{2(1-a)(1+b)},\qquad w_{-}=\frac{b-\sqrt{b^{2}-4a+4a^{2}}}{2b}.

For each fixed wγ1w\in\gamma_{1}, we now find a steepest descent path of the phase function f(t,w)f(t,w) in the variable tt, which passes through a saddle point t0(w)t_{0}(w) depending on ww. (Note that not only the saddle point t0(w)t_{0}(w), but all points tt on this steepest descent path depend on ww.) The function f(t0(w),w)f(t_{0}(w),w) is a function of ww alone. To find the relevant saddle point t0(w)t_{0}(w), we solve the equation f(t,w)/t=0\partial f(t,w)/\partial t=0, and obtain

(2.5) t0(w)=2w1+1+4b2(w1)w2(1+b)w.t_{0}(w)=\frac{2w-1+\sqrt{1+4b^{2}(w-1)w}}{2(1+b)w}.

It can be shown that

(2.6) t0(w+)=t+,t0(w)=t.t_{0}(w_{+})=t_{+},\qquad t_{0}(w_{-})=t_{-}.

Define the standard transformation tτt\rightarrow\tau by

(2.7) τ2=f(t0(w),w)f(t,w);\tau^{2}=f(t_{0}(w),w)-f(t,w);

see [11, p.88]. Note that we have τ=\tau=-\infty when t=0t=0, and τ=+\tau=+\infty when t=1t=1. Furthermore, this mapping takes t=t0(w)t=t_{0}(w) to τ=0\tau=0. Coupling (2.1)(\ref{IR of DCP for x>0}) and (2.7)(\ref{mapping t to tau}) gives

(2.8) tn(x,N+1)=(1)n2πiΓ(n+N+2)Γ(n+1)Γ(Nn+1)×+γ11w1eNf(t0(w),w)eNτ2dtdτdwdτ,\begin{split}t_{n}(x,N+1)=&\frac{(-1)^{n}}{2\pi i}\frac{\Gamma(n+N+2)}{\Gamma(n+1)\Gamma(N-n+1)}\\ &\times\int_{-\infty}^{+\infty}\int_{\gamma_{1}}\frac{1}{w-1}e^{Nf({t_{0}(w)},w)}e^{-N\tau^{2}}\frac{\mathrm{d}t}{\mathrm{d}\tau}\mathrm{d}w\mathrm{d}\tau,\end{split}

where γ1\gamma_{1} is the integration path of the variable ww in (2.1). Note that the first variable of ff in (2.8) is now t0(w)t_{0}(w), instead of tt. Hence, the phase function is a function of ww alone. Setting f(t0(w),w)/w=0\partial f(t_{0}(w),w)/\partial w=0, we obtain the saddle points

(2.9) w±=b±b24a+4a22b;w_{\pm}=\frac{b\pm\sqrt{b^{2}-4a+4a^{2}}}{2b};

cf.(2.3) and (2.4). Motivated by (2.2), define the new phase function

(2.10) ψ(u)=alnualn(u1)+ηu.\psi(u)=a\ln u-a\ln(u-1)+\eta u.

The saddle points of ψ\psi are given by

(2.11) u±=η±η2+4aη2η.u_{\pm}=\frac{\eta\pm\sqrt{\eta^{2}+4a\eta}}{2\eta}.

To reduce the double integral in (2.1) into a canonical form, we define the second mapping wuw\rightarrow u by

(2.12) f(t0(w),w)=ψ(u)+γ=alnualn(u1)+ηu+γ\begin{split}f(t_{0}(w),w)&=\psi(u)+\gamma\\ &=a\ln u-a\ln(u-1)+\eta u+\gamma\end{split}

with

(2.13) u(w+)=u,u(w)=u+,u(w_{+})=u_{-},\quad\quad\quad\quad u(w_{-})=u_{+},

where η\eta and γ\gamma are real numbers depending on the parameters aa and bb in (2.2). From (2.12), we have

(2.14) dwdu=dψdu/df(t0(w),w)dw=η(uu+)(uu)w(1w)[(2a1)1+4b2w24b2w]2b2u(u1)(ww+)(ww)\begin{split}\frac{\mathrm{d}w}{\mathrm{d}u}&=\frac{\mathrm{d}\psi}{\mathrm{d}u}\left/\frac{\mathrm{d}f(t_{0}(w),w)}{\mathrm{d}w}\right.\\ &=\frac{\eta(u-u_{+})(u-u_{-})w(1-w)\left[(2a-1)-\sqrt{1+4b^{2}w^{2}-4b^{2}w}\right]}{-2b^{2}u(u-1)(w-w_{+})(w-w_{-})}\end{split}

for uu±u\neq u_{\pm}, and by L’Ho^\hat{o}pital’s rule,

(2.15) dwdu=[η2+4aη(12a)(1a)(η)b3b24a+4a2]1/2\frac{\mathrm{d}w}{\mathrm{d}u}=\left[\frac{\sqrt{\eta^{2}+4a\eta}(1-2a)(1-a)(-\eta)}{b^{3}\sqrt{b^{2}-4a+4a^{2}}}\right]^{1/2}

for u=u±u=u_{\pm}. With the change of variable wuw\rightarrow u defined in (2.12), the representation in (2.8) becomes

(2.16) tn(x,N+1)=(1)n2πiΓ(n+N+2)Γ(n+1)Γ(Nn+1)eNγ×+γuh(u,τ)u1eNψ(u)Nτ2dudτ,\begin{split}t_{n}(x,N+1)=&\frac{(-1)^{n}}{2\pi i}\frac{\Gamma(n+N+2)}{\Gamma(n+1)\Gamma(N-n+1)}e^{N\gamma}\\ &\qquad\times\int_{-\infty}^{+\infty}\int_{\gamma_{u}}\frac{h(u,\tau)}{u-1}e^{N\psi(u)-N\tau^{2}}\mathrm{d}u\mathrm{d}\tau,\end{split}

where

(2.17) h(u,τ)=u1w1dwdudtdτ,h(u,\tau)=\frac{u-1}{w-1}\frac{\mathrm{d}w}{\mathrm{d}u}\frac{\mathrm{d}t}{\mathrm{d}\tau},

and γu\gamma_{u} is the steepest descent path of ψ(u)\psi(u) in the uu-plane. An asymptotic expansion, holding uniformly for a[0,12]a\in[0,\frac{1}{2}], was then derived in [8] by an integration-by-parts technique. To state the result, we define recursively h0(u,τ)=h(u,τ)h_{0}(u,\tau)=h(u,\tau),

(2.18) hl(u,τ)=al(τ)+bl(τ)u(uu)(uu+)gl(u,τ)h_{l}(u,\tau)=a_{l}(\tau)+b_{l}(\tau)u-(u-u_{-})(u-u_{+})g_{l}(u,\tau)

and

(2.19) hl+1(u,τ)=(u1)[gl(u,τ)+ugl(u,τ)u],l=0,1,2,.h_{l+1}(u,\tau)=(u-1)\left[g_{l}(u,\tau)+u\frac{\partial g_{l}(u,\tau)}{\partial u}\right],\qquad l=0,1,2,....

Furthermore, we write

(2.20) al(τ)=j=0al,jτj,bl(τ)=j=0bl,jτj.a_{l}(\tau)=\sum_{j=0}^{\infty}a_{l,j}\tau^{j},\qquad b_{l}(\tau)=\sum_{j=0}^{\infty}b_{l,j}\tau^{j}.

The resulting expansion takes the form

(2.21) tn(x,N+1)(1)nΓ(n+N+2)eNγΓ(n+1)Γ(Nn+1)N{𝐌(aN+1,1,ηN)l=0cl(ηN)l +𝐌(aN+1,1,ηN)l=0dl(ηN)l},\begin{split}t_{n}(x,N+1)\sim\frac{(-1)^{n}\Gamma(n+N+2)e^{N\gamma}}{\Gamma(n+1)\Gamma(N-n+1)\sqrt{N}}&\left\{\mathbf{M}(aN+1,1,\eta N)\sum_{l=0}^{\infty}\frac{c_{l}}{(\eta N)^{l}}\right.\\ &\left.\text{ }+\mathbf{M}^{\prime}(aN+1,1,\eta N)\sum_{l=0}^{\infty}\frac{d_{l}}{(\eta N)^{l}}\right\},\end{split}

where 𝐌(a,c,z)\mathbf{M}(a,c,z) is the Kummer function defined in (1.6), and the coefficients clc_{l} and dld_{l} are given explicitly by

(2.22) cl=m=0lalm,2mΓ(m+12)ηm,dl=m=0lblm,2mΓ(m+12)ηm.c_{l}=\sum_{m=0}^{l}a_{l-m,2m}\Gamma(m+\frac{1}{2})\eta^{m},\qquad d_{l}=\sum_{m=0}^{l}b_{l-m,2m}\Gamma(m+\frac{1}{2})\eta^{m}.

Case (ii) was dealt with in a similar manner. We started with the double-integral representation

(2.23) tn(x,N+1)=(1)n2πiΓ(n+N+2)Γ(n+1)Γ(Nn+1)01γ21w1eNf~(t,w)dwdt,t_{n}(x,N+1)=\frac{(-1)^{n}}{2\pi i}\frac{\Gamma(n+N+2)}{\Gamma(n+1)\Gamma(N-n+1)}\int_{0}^{1}\int_{\gamma_{2}}\frac{1}{w-1}e^{N\widetilde{f}(t,w)}\mathrm{d}w\mathrm{d}t,

where

(2.24) f~(t,w)=bln(1t)+(1b)lnt+aln(w)aln(1w)+bln[1(1t)w],\widetilde{f}(t,w)=b\ln(1-t)+(1-b)\ln t+a\ln(-w)-a\ln(1-w)+b\ln[1-(1-t)w],

and the integration path γ2\gamma_{2} starts at w=1w=1, traverses along the upper edge of the positive real line towards w=0w=0, encircles the origin in the counterclockwise direction and returns to w=1w=1 along the lower edge of the positive real line.

Now we follow the same argument as given in Case (i), and use the same mapping tτt\rightarrow\tau defined in (2.7), except with f(t,w)f(t,w) replaced by f~(t,w)\widetilde{f}(t,w). The result is

(2.25) tn(x,N+1)=(1)n2πiΓ(n+N+2)Γ(n+1)Γ(Nn+1)×+γ21w1eNf~(t0(w),w)eNτ2dtdτdwdτ.\begin{split}t_{n}(x,N+1)=&\frac{(-1)^{n}}{2\pi i}\frac{\Gamma(n+N+2)}{\Gamma(n+1)\Gamma(N-n+1)}\\ &\times\int_{-\infty}^{+\infty}\int_{\gamma_{2}}\frac{1}{w-1}e^{N\widetilde{f}({t_{0}(w)},w)}e^{-N\tau^{2}}\frac{\mathrm{d}t}{\mathrm{d}\tau}\mathrm{d}w\mathrm{d}\tau.\end{split}

Because of the shape of the contour γ2\gamma_{2}, it turns out that the phase function f~(t0(w),w)\widetilde{f}(t_{0}(w),w) has only one relevant saddle point, namely ww_{-}; see (2.9). As a consequence, we define the mapping wuw\rightarrow u by

(2.26) f~(t0(w),w)=aln(u)u+γ\widetilde{f}(t_{0}(w),w)=a\ln(-u)-u+\gamma

with

(2.27) u(w)=a,u(w_{-})=a,

where γ\gamma is a constant depending on the parameters aa and bb in (1.3). The final expansion is in terms of the gamma function, and we have

(2.28) tn(x,N+1)(1)nΓ(n+N+2)NaNeNγΓ(n+1)Γ(Nn+1)Γ(aN+1)l=0clNl+12,t_{n}(x,N+1)\sim\frac{(-1)^{n}\Gamma(n+N+2)N^{-aN}e^{N\gamma}}{\Gamma(n+1)\Gamma(N-n+1)\Gamma(-aN+1)}\sum_{l=0}^{\infty}\frac{c_{l}}{N^{l+\frac{1}{2}}},

where clc_{l} are constants that can be given recursively.

3 KUMMER-TYPE EXPANSION

As b0b\rightarrow 0, some of the steps in Section 2 are no longer valid. Let us first examine the mapping tτt\rightarrow\tau given in (2.7). In this mapping, we have used the fact that for each fixed wγ1w\in\gamma_{1}, the saddle point t=t0(w)t=t_{0}(w) in (2.5) is bounded away from t=0t=0 and t=1t=1, and the steepest descent path from t=0t=0 to t=1t=1, passing through t0(w)t_{0}(w), is mapped onto (,)(-\infty,\infty) in the τ\tau-plane. However, from (2.5), we have for any fixed ww

(3.1) t0(w)=1b+O(b2)as b0;t_{0}(w)=1-b+O\left(b^{2}\right)\qquad\text{as }b\rightarrow 0;

that is, as b0b\rightarrow 0, t0(w)t_{0}(w) approaches the branch point t=1t=1 in the t-plane. Thus, the mapping (2.7) is no longer suitable in this case. Since we are more interested in the neighbourhood of t=1t=1, where the term bln(1t)b\ln(1-t) in the phase function (2.2) becomes singular, we introduce the mapping

(3.2) f(t,w)=blnττ+A,f(t,w)=b\ln\tau-\tau+A,

where the constant AA does not depend on tt or τ\tau (but may depend on ww); see (3.4) below. To make the mapping tτt\rightarrow\tau defined in (3.2) one-to-one and analytic, we prescribe t=t0(w)t=t_{0}(w) to correspond to τ=b\tau=b, which is the saddle point of blnττ+Ab\ln\tau-\tau+A; i.e.,

(3.3) τ(t0(w))=b.\tau(t_{0}(w))=b.

This gives

(3.4) A=f(t0(w),w)blnb+b.A=f(t_{0}(w),w)-b\ln b+b.

Note that we have τ=0\tau=0 when t=1t=1, and τ=+\tau=+\infty when t=0t=0. Furthermore, from (3.2), we have

(3.5) dtdτ=b/τ1f(t,w)/t=(τb)t(1t)[1(1t)w]τ(1+b)w(tt0+(w))(tt0(w)),τb,dtdτ={t0+(w)(1t0+(w))[1(1t0+(w))w]b1+4b2(w1)w}1/2,τ=b,\begin{split}\frac{\mathrm{d}t}{\mathrm{d}\tau}&=\frac{b/\tau-1}{\partial f(t,w)/\partial t}=\frac{(\tau-b)t(1-t)[1-(1-t)w]}{\tau(1+b)w(t-t^{+}_{0}(w))(t-t^{-}_{0}(w))},\quad\quad\tau\neq b,\\ \frac{\mathrm{d}t}{\mathrm{d}\tau}&=\left\{\frac{t^{+}_{0}(w)(1-t^{+}_{0}(w))[1-(1-t^{+}_{0}(w))w]}{b\sqrt{1+4b^{2}(w-1)w}}\right\}^{1/2},\hskip 45.52458pt\tau=b,\end{split}

where we have used L’Ho^\hat{o}pital’s rule for τ=b\tau=b and

t0±(w)=2w1±1+4b2(w1)w2(1+b)w;t_{0}^{\pm}(w)=\frac{2w-1\pm\sqrt{1+4b^{2}(w-1)w}}{2(1+b)w};

cf.(2.5). Coupling (2.1) and (3.2), we have, instead of (2.8),

(3.6) tn(x,N+1)=(1)n+12πiΓ(n+N+2)enbnΓ(n+1)Γ(Nn+1)×0+γ11w1eNf(t0(w),w)eN(blnττ)dtdτdwdτ.\begin{split}t_{n}(x,N+1)=&\frac{(-1)^{n+1}}{2\pi i}\frac{\Gamma(n+N+2)e^{n}b^{-n}}{\Gamma(n+1)\Gamma(N-n+1)}\\ &\times\int_{0}^{+\infty}\int_{\gamma_{1}}\frac{1}{w-1}e^{Nf({t_{0}(w)},w)}e^{N(b\ln\tau-\tau)}\frac{\mathrm{d}t}{\mathrm{d}\tau}\mathrm{d}w\mathrm{d}\tau.\end{split}

Next, let us examine the mapping (2.12). We note that this mapping still works in the present case; the only difference is that for fixed n/N(0,1)n/N\in(0,1), the constant η\eta in the mapping (2.12) is bounded away from 0, whereas when n/N0n/N\rightarrow 0, η\eta approaches 0. More precisely, we have ηb2\eta\sim-b^{2}. This can roughly be seen from the saddle points (2.9) and (2.11), together with their correspondence relation (2.13) under the mapping wuw\rightarrow u given in (2.12). To prove this rigorously, we give the following lemma.

Lemma 1.

Let η\eta be the constant defined in the mapping (2.12). If a/b2=o(1)a/b^{2}=o(1) (or, equivalently, xN/n2=o(1)xN/n^{2}=o(1)), then we have

(3.7) η=b2+o(b2) as b0.\eta=-b^{2}+o(b^{2})\quad\text{ as }b\rightarrow 0.
Proof.

From (2.5), we have for any fixed ww

(3.8) t0(w)=11+bb2(1w)1+bb4w(w1)21+b+O(b6w2(w1)3) as b0,t_{0}(w)=\frac{1}{1+b}-\frac{b^{2}(1-w)}{1+b}-\frac{b^{4}w(w-1)^{2}}{1+b}+O\left(b^{6}w^{2}(w-1)^{3}\right)\quad\text{ as }b\rightarrow 0,

where the term O(b4w(w1)2)O\left(b^{4}w(w-1)^{2}\right) holds uniformly when either ww or 1w1-w is small. If a/b20a/b^{2}\rightarrow 0, from (2.9) we have w=1w+0+w_{-}=1-w_{+}\rightarrow 0^{+} and both w±w_{\pm} bounded. Thus,

(3.9) t0(w)=1b+b2(1b)w+O(b4w2),t0(w+)=11+bb21+bw+O(b4w2).\begin{split}t_{0}(w_{-})&=1-b+b^{2}(1-b)w_{-}+O\left(b^{4}w_{-}^{2}\right),\\ t_{0}(w_{+})&=\frac{1}{1+b}-\frac{b^{2}}{1+b}w_{-}+O\left(b^{4}w_{-}^{2}\right).\end{split}

To obtain η\eta, we use (2.12) and (2.13). First, substituting (3.9) in (2.2) gives

(3.10) f(t0(w+),w+)f(t0(w),w)=2aln1wwb2+O(b3).f(t_{0}(w_{+}),w_{+})-f(t_{0}(w_{-}),w_{-})=2a\ln\frac{1-w_{-}}{w_{-}}-b^{2}+O\left(b^{3}\right).

Here, we have made use of the fact that if w±w_{\pm} lie on the upper edge of the cut along the interval 0<w<10<w<1, then we have ln(w±1)=ln(1w±)+πi\ln(w_{\pm}-1)=\ln(1-w_{\pm})+\pi i. Similarly, if w±w_{\pm} lie on the lower edge of the cut, then ln(w±1)=ln(1w±)πi\ln(w_{\pm}-1)=\ln(1-w_{\pm})-\pi i. Next, we have from (2.10) and (2.11)

(3.11) ψ(u)ψ(u+)=2aln1u+u++η1+4aη.\psi(u_{-})-\psi(u_{+})=2a\ln\frac{1-u_{+}}{u_{+}}+\eta\sqrt{1+\frac{4a}{\eta}}.

Note that u++u=1u_{+}+u_{-}=1 and η<0\eta<0. Formula (3.7) now follows from a combination of (2.12), (2.13), (3.10) and (3.11). ∎

What Lemma 1 says is that if a/b20a/b^{2}\rightarrow 0 and x=O(1)x=O(1) or xO(1)x\gg O(1), then we have |η|Nb2N=xb2/a|\eta|N\sim b^{2}N=xb^{2}/a\rightarrow\infty, i.e., ηN\eta N is large.

Coupling (3.6) and (2.12), we have

(3.12) tn(x,N+1)=(1)n+1Γ(n+N+2)enbneNγ2πiΓ(n+1)Γ(Nn+1)×0+γ1h(u,τ)u1eNψ(u)eN(blnττ)dwdτ,\begin{split}t_{n}(x,N+1)=&\frac{(-1)^{n+1}\Gamma(n+N+2)e^{n}b^{-n}e^{N\gamma}}{2\pi i\Gamma(n+1)\Gamma(N-n+1)}\\ &\times\int_{0}^{+\infty}\int_{\gamma_{1}}\frac{h(u,\tau)}{u-1}e^{N\psi(u)}e^{N(b\ln\tau-\tau)}\mathrm{d}w\mathrm{d}\tau,\end{split}

where

(3.13) h(u,τ)=u1w1dwdudtdτ,h(u,\tau)=\frac{u-1}{w-1}\frac{\mathrm{d}w}{\mathrm{d}u}\frac{\mathrm{d}t}{\mathrm{d}\tau},

dw/du\mathrm{d}w/\mathrm{d}u is given by (2.14) and (2.15) and dt/dτ\mathrm{d}t/\mathrm{d}\tau is given by (3.5). Following the same integration-by-parts procedure outlined in Section 2, we let h0(u,τ)=h(u,τ)h_{0}(u,\tau)=h(u,\tau) in (3.13), and define hl(u,τ)h_{l}(u,\tau) recursively by (2.18), (2.19) and (2.20). The final result is

tn(x,N+1)=\displaystyle t_{n}(x,N+1)= (1)n+1Γ(n+N+2)enbneNγΓ(Nn+1)Nn+1\displaystyle\frac{(-1)^{n+1}\Gamma(n+N+2)e^{n}b^{-n}e^{N\gamma}}{\Gamma(N-n+1)N^{n+1}}
(3.14) ×[𝐌(aN+1,1,ηN)l=0p1cl(ηN)l+𝐌(aN+1,1,ηN)l=0p1dl(ηN)l+εp],\displaystyle\times\Biggr{[}\mathbf{M}(aN+1,1,\eta N)\sum_{l=0}^{p-1}\frac{c_{l}}{(\eta N)^{l}}+\mathbf{M}^{\prime}(aN+1,1,\eta N)\sum_{l=0}^{p-1}\frac{d_{l}}{(\eta N)^{l}}+\varepsilon_{p}\Biggr{]},

where

(3.15) cl\displaystyle c_{l} =\displaystyle= Nn+1Γ(n+1)0al(τ)τneNτdτm=0al,mΓ(n+m+1)Γ(n+1)Nm,\displaystyle\frac{N^{n+1}}{\Gamma(n+1)}\int_{0}^{\infty}a_{l}(\tau)\tau^{n}e^{-N\tau}\mathrm{d}\tau\sim\sum_{m=0}^{\infty}a_{l,m}\frac{\Gamma(n+m+1)}{\Gamma(n+1)N^{m}},
(3.16) dl\displaystyle d_{l} =\displaystyle= Nn+1Γ(n+1)0bl(τ)τneNτdτm=0bl,mΓ(n+m+1)Γ(n+1)Nm\displaystyle\frac{N^{n+1}}{\Gamma(n+1)}\int_{0}^{\infty}b_{l}(\tau)\tau^{n}e^{-N\tau}\mathrm{d}\tau\sim\sum_{m=0}^{\infty}b_{l,m}\frac{\Gamma(n+m+1)}{\Gamma(n+1)N^{m}}

and

(3.17) εp=Nn+12πiΓ(n+1)(ηN)p0+γ1hp(u,τ)u1eNψ(u)eN(blnττ)dwdτ.\varepsilon_{p}=\frac{N^{n+1}}{2\pi i\Gamma(n+1)(\eta N)^{p}}\int_{0}^{+\infty}\int_{\gamma_{1}}\frac{h_{p}(u,\tau)}{u-1}e^{N\psi(u)}e^{N(b\ln\tau-\tau)}\mathrm{d}w\mathrm{d}\tau.

Note that since b=n/N0b=n/N\rightarrow 0, the series in (3.15) and (3.16) are asymptotic. Moreover, since ηN\eta N is large in this case, (3) is indeed a compound asymptotic expansion as NN\rightarrow\infty.

Now, we consider the subcase in which a/b20a/b^{2}\rightarrow 0 and x=o(1)x=o(1). In this case, the expansion in (3) is still asymptotic as long as |η|Nb2N|\eta|N\sim b^{2}N\rightarrow\infty or, equivalently, nO(N)n\gg O(\sqrt{N}). If n=O(N)n=O(\sqrt{N}), then the series in (1.2) is itself an asymptotic expansion as NN\rightarrow\infty. This completes our discussion of all three cases listed under the condition “xN/n2xN/n^{2} small” in Table 1.

4 AIRY-TYPE EXPANSION

For the case a/b2[γ,M]a/b^{2}\in[\gamma,M], where γ\gamma is a small positive number and MM is a large positive number, the saddle points w±w_{\pm} in (2.9) are bounded away from the singularities w=0w=0, w=1w=1 and w=w=\infty, and coalesce with each other when a/b214(1a)a/b^{2}\rightarrow\frac{1}{4(1-a)}. Therefore, to derive an asymptotic expansion uniformly for a/b2[γ,M]a/b^{2}\in[\gamma,M], we need the cubic transformation (see (4.6) below) introduced by Chester, Friedman and Ursell [2]. The resulting expansion is in terms of the Airy function Ai()\text{Ai}(\cdot).

Following the same argument as in Section 3, we again use the mapping from tτt\rightarrow\tau in (3.2), and start with the integral representation (3.6)

(4.1) tn(x,N+1)=(1)n+12πiΓ(n+N+2)enbnΓ(n+1)Γ(Nn+1)×0+γ11w1eNf(t0(w),w)eN(blnττ)dtdτdwdτ,\begin{split}t_{n}(x,N+1)=&\frac{(-1)^{n+1}}{2\pi i}\frac{\Gamma(n+N+2)e^{n}b^{-n}}{\Gamma(n+1)\Gamma(N-n+1)}\\ &\times\int_{0}^{+\infty}\int_{\gamma_{1}}\frac{1}{w-1}e^{Nf({t_{0}(w)},w)}e^{N(b\ln\tau-\tau)}\frac{\mathrm{d}t}{\mathrm{d}\tau}\mathrm{d}w\mathrm{d}\tau,\end{split}

where the integration path is described in the line below (2.2). To proceed further, we divide γ1\gamma_{1} into two parts, and denote the part in the upper half of the plane by γ(+)\gamma^{(+)}, and the other part in the lower half of the plane by γ()\gamma^{(-)}. Recall from (2.2) that the phase function in (4.1) is given by

(4.2) f(t0(w),w)=bln(1t0(w))+(1b)lnt0(w)+alnwaln(w1)+bln[1(1t0(w))w],\begin{split}f(t_{0}(w),w)=&b\ln(1-t_{0}(w))+(1-b)\ln t_{0}(w)+a\ln w\\ &-a\ln(w-1)+b\ln[1-(1-t_{0}(w))w],\end{split}

which has a cut (,1](-\infty,1] in ww-plane. Put

(4.3) f¯(t0(w),w)=bln(1t0(w))+(1b)lnt0(w)+alnwaln(1w)+bln[1(1t0(w))w],\begin{split}\overline{f}(t_{0}(w),w)=&b\ln(1-t_{0}(w))+(1-b)\ln t_{0}(w)+a\ln w\\ &-a\ln(1-w)+b\ln[1-(1-t_{0}(w))w],\end{split}

which has cuts (,0](-\infty,0] and [1,)[1,\infty). From (4.2) and (4.3), we have

(4.4) f(t0(w),w)=f¯(t0(w),w)+aπif(t_{0}(w),w)=\overline{f}(t_{0}(w),w)+a\pi i

for wγ(+)w\in\gamma^{(+)}, and

(4.5) f(t0(w),w)=f¯(t0(w),w)aπif(t_{0}(w),w)=\overline{f}(t_{0}(w),w)-a\pi i

for wγ()w\in\gamma^{(-)}. We now make the standard transformation

(4.6) f¯(t0(w),w)=13u3ζu+A,\overline{f}(t_{0}(w),w)=\frac{1}{3}u^{3}-\zeta u+A,

with the correspondence between the critical points of the two sides prescribed by

(4.7) w+ζ,wζ.w_{+}\leftrightarrow\sqrt{\zeta},\qquad\quad w_{-}\leftrightarrow-\sqrt{\zeta}.

If ζ>0\zeta>0, then ±ζ\pm\sqrt{\zeta} are both real; if ζ<0\zeta<0, then ±ζ\pm\sqrt{\zeta} are complex conjugates and purely imaginary. The values of AA and ζ\zeta can be obtained by using (4.6) and (4.7). We also have

(4.8) dwdu=(uζ)(u+ζ)w(1w)[(2a1)1+4b2w24b2w]2b2(ww+)(ww),u±ζ,dwdu={2ζa(12a)(1a)b3b24a+4a2}1/2,u=±ζ,\begin{split}\frac{\mathrm{d}w}{\mathrm{d}u}&=\frac{(u-\sqrt{\zeta})(u+\sqrt{\zeta})w(1-w)\left[(2a-1)-\sqrt{1+4b^{2}w^{2}-4b^{2}w}\right]}{-2b^{2}(w-w_{+})(w-w_{-})},\quad u\neq\pm\sqrt{\zeta},\\ \frac{\mathrm{d}w}{\mathrm{d}u}&=\left\{\frac{2\sqrt{\zeta}a(1-2a)(1-a)}{b^{3}\sqrt{b^{2}-4a+4a^{2}}}\right\}^{1/2},\hskip 45.52458ptu=\pm\sqrt{\zeta},\end{split}

where we have again used L’Ho^\hat{o}pital’s rule for u=±ζu=\pm\sqrt{\zeta}. Let us first consider the case ζ<0\zeta<0, and deform the image of the contour γ(+)\gamma^{(+)} under the mapping wuw\rightarrow u defined in (4.6) to the steepest descent path of 13u3ζu+A\frac{1}{3}u^{3}-\zeta u+A in the uu-plane which passes through ζ\sqrt{\zeta}. We denote the path by C2C_{2}. Similarly, we deform the image of the contour γ()\gamma^{(-)}, and denote the steepest descent path passing through ζ-\sqrt{\zeta} by C3C_{3}; see Figure 1. Next, we consider the case ζ>0\zeta>0. Note that here the saddle points ±ζ\pm\sqrt{\zeta} are real, and that the contours C2C_{2} and C3C_{3} pass through both of them; see Figure 2. Clearly, in both cases, C3C_{3} is the reflection of C2C_{2} with respect to the real axis in the uu-plane.

Let C1C_{1} denote the dotted curve shown in Figures 1 and 2, and recall the identities eiaNπ=cosaNπ+isinaNπe^{iaN\pi}=\cos aN\pi+i\sin aN\pi and eiaNπ=cosaNπisinaNπe^{-iaN\pi}=\cos aN\pi-i\sin aN\pi. A combination of (4.1), (4.4), (4.5) and (4.6) then gives

(4.9) tn(x,N+1)=(1)nΓ(n+N+2)enbneNAΓ(n+1)Γ(Nn+1)×{cosaNπ2πi0+C1h0(u,τ)eN(13u3ζu)eN(blnττ)dudτ+sinaNπ2π0+C3C2h0(u,τ)eN(13u3ζu)eN(blnττ)dudτ},\begin{split}t_{n}(x,N+1)=&\frac{(-1)^{n}\Gamma(n+N+2)e^{n}b^{-n}e^{NA}}{\Gamma(n+1)\Gamma(N-n+1)}\\ &\times\left\{\frac{\cos{aN\pi}}{2\pi i}\int_{0}^{+\infty}\int_{C_{1}}h_{0}(u,\tau)e^{N\left(\frac{1}{3}u^{3}-\zeta u\right)}e^{N(b\ln\tau-\tau)}\mathrm{d}u\mathrm{d}\tau\right.\\ &\quad\left.+\frac{\sin{aN\pi}}{2\pi}\int_{0}^{+\infty}\int_{C_{3}-C_{2}}h_{0}(u,\tau)e^{N\left(\frac{1}{3}u^{3}-\zeta u\right)}e^{N(b\ln\tau-\tau)}\mathrm{d}u\mathrm{d}\tau\right\},\end{split}

where

(4.10) h0(u,τ)=1w1dtdτdwdu,h_{0}(u,\tau)=\frac{1}{w-1}\frac{\mathrm{d}t}{\mathrm{d}\tau}\frac{\mathrm{d}w}{\mathrm{d}u},

dt/dτ\mathrm{d}t/\mathrm{d}\tau, dw/du\mathrm{d}w/\mathrm{d}u are given respectively by (3.5) and (4.8).

Refer to caption
Figure 1: C2C_{2} and C3C_{3} (ζ<0)(\zeta<0).
Refer to caption
Figure 2: C2C_{2} and C3C_{3} (ζ>0)(\zeta>0).

4.1 When ζ\zeta is bounded

Following the standard integration-by-parts procedure [11, p.368], we define recursively

(4.11) hl(u,τ)\displaystyle h_{l}(u,\tau) =\displaystyle= al(τ)+bl(τ)u(u2ζ)gl(u,τ),\displaystyle a_{l}(\tau)+b_{l}(\tau)u-(u^{2}-\zeta)g_{l}(u,\tau),
(4.12) hl+1(u,τ)\displaystyle h_{l+1}(u,\tau) =\displaystyle= gl(u,τ)u.\displaystyle\frac{\partial g_{l}(u,\tau)}{\partial u}.

Furthermore, expand

(4.13) al(τ)=j=0al,jτj,bl(τ)=j=0bl,jτj.a_{l}(\tau)=\sum_{j=0}^{\infty}a_{l,j}\tau^{j},\qquad b_{l}(\tau)=\sum_{j=0}^{\infty}b_{l,j}\tau^{j}.

From (4.9), (4.11), (4.12) and (4.13), it follows

tn(x,N+1)=\displaystyle t_{n}(x,N+1)= (1)nΓ(n+N+2)enbneNAΓ(Nn+1)Nn+1\displaystyle\frac{(-1)^{n}\Gamma(n+N+2)e^{n}b^{-n}e^{NA}}{\Gamma(N-n+1)N^{n+1}}
(4.14) ×[cosaNπ(Ai(N2/3ζ)l=0p1clNl+1/3Ai(N2/3ζ)l=0p1dlNl+2/3)\displaystyle\times\left[\cos{aN\pi}\left(\text{Ai}(N^{2/3}\zeta)\sum_{l=0}^{p-1}\frac{c_{l}}{N^{l+1/3}}-\text{Ai}^{\prime}(N^{2/3}\zeta)\sum_{l=0}^{p-1}\frac{d_{l}}{N^{l+2/3}}\right)\right.
+sinaNπ(Bi(N2/3ζ)l=0p1clNl+1/3Bi(N2/3ζ)l=0p1dlNl+2/3)+εp],\displaystyle\left.\quad+\sin{aN\pi}\left(\text{Bi}(N^{2/3}\zeta)\sum_{l=0}^{p-1}\frac{c_{l}}{N^{l+1/3}}-\text{Bi}^{\prime}(N^{2/3}\zeta)\sum_{l=0}^{p-1}\frac{d_{l}}{N^{l+2/3}}\right)+\varepsilon_{p}\right],

where

(4.15) εp=cos(aNπ)Nn+12πiNpΓ(n+1)0+C1hp(u,τ)eN(13u3ζu)eN(blnττ)dudτ+sin(aNπ)Nn+12πiNpΓ(n+1)0+C3C2hp(u,τ)eN(13u3ζu)eN(blnττ)dudτ,\begin{split}\varepsilon_{p}=&\frac{\cos{(aN\pi)}N^{n+1}}{2\pi iN^{p}\Gamma{(n+1)}}\int_{0}^{+\infty}\int_{C_{1}}h_{p}(u,\tau)e^{N\left(\frac{1}{3}u^{3}-\zeta u\right)}e^{N(b\ln\tau-\tau)}\mathrm{d}u\mathrm{d}\tau\\ &+\frac{\sin{(aN\pi)}N^{n+1}}{2\pi iN^{p}\Gamma{(n+1)}}\int_{0}^{+\infty}\int_{C_{3}-C_{2}}h_{p}(u,\tau)e^{N\left(\frac{1}{3}u^{3}-\zeta u\right)}e^{N(b\ln\tau-\tau)}\mathrm{d}u\mathrm{d}\tau,\end{split}

and the coefficients are given by

(4.16) cl=Nn+1Γ(n+1)0al(τ)τneNτdτm=0al,mΓ(n+m+1)Γ(n+1)Nm,dl=Nn+1Γ(n+1)0bl(τ)τneNτdτm=0bl,mΓ(n+m+1)Γ(n+1)Nm.\begin{split}c_{l}&=\frac{N^{n+1}}{\Gamma(n+1)}\int_{0}^{\infty}a_{l}(\tau)\tau^{n}e^{-N\tau}\mathrm{d}\tau\sim\sum_{m=0}^{\infty}a_{l,m}\frac{\Gamma(n+m+1)}{\Gamma(n+1)N^{m}},\\ d_{l}&=\frac{N^{n+1}}{\Gamma(n+1)}\int_{0}^{\infty}b_{l}(\tau)\tau^{n}e^{-N\tau}\mathrm{d}\tau\sim\sum_{m=0}^{\infty}b_{l,m}\frac{\Gamma(n+m+1)}{\Gamma(n+1)N^{m}}.\end{split}

In the present case, a/b2[δ,M]a/b^{2}\in[\delta,M] or, equivalently, w±w_{\pm} are bounded. Thus ζ\zeta is also bounded in view of (4.7). Therefore, it is easy to prove that (4.1) is asymptotic; for details, see [11, p.371-372].

4.2 When ζ\zeta\rightarrow-\infty

When a/b2a/b^{2}\rightarrow\infty, by (2.9) the saddle points w±w_{\pm} approach 12±i\frac{1}{2}\pm i\infty, respectively, along the line Re w=12\text{Re }w=\frac{1}{2}. In view of (4.7), this is equivalent to saying that ζ\zeta\rightarrow-\infty. To prove that (4.1) is also asymptotic in this case, we rewrite the expansion in (4.1) as

tn(x,N+1)=\displaystyle t_{n}(x,N+1)= (1)nΓ(n+N+2)enbneNAΓ(Nn+1)Nn+1\displaystyle\frac{(-1)^{n}\Gamma(n+N+2)e^{n}b^{-n}e^{NA}}{\Gamma(N-n+1)N^{n+1}}
(4.17) ×[cosaNπ(Ai(N2/3ζ)N1/3l=0p1c~l(aN)lAi(N2/3ζ)N2/3l=0p1d~l(aN)l)\displaystyle\times\left[\cos{aN\pi}\left(\frac{\text{Ai}(N^{2/3}\zeta)}{N^{1/3}}\sum_{l=0}^{p-1}\frac{\widetilde{c}_{l}}{(aN)^{l}}-\frac{\text{Ai}^{\prime}(N^{2/3}\zeta)}{N^{2/3}}\sum_{l=0}^{p-1}\frac{\widetilde{d}_{l}}{(aN)^{l}}\right)\right.
+sinaNπ(Bi(N2/3ζ)N1/3l=0p1c~l(aN)lBi(N2/3ζ)N2/3l=0p1d~l(aN)l)+εp],\displaystyle\quad\left.\ +\sin{aN\pi}\left(\frac{\text{Bi}(N^{2/3}\zeta)}{N^{1/3}}\sum_{l=0}^{p-1}\frac{\widetilde{c}_{l}}{(aN)^{l}}-\frac{\text{Bi}^{\prime}(N^{2/3}\zeta)}{N^{2/3}}\sum_{l=0}^{p-1}\frac{\widetilde{d}_{l}}{(aN)^{l}}\right)+\varepsilon_{p}\right],

where εp\varepsilon_{p} is the same as in (4.15) and

(4.18) c~l=alcl and d~l=aldl.\widetilde{c}_{l}=a^{l}c_{l}\quad\text{ and }\quad\widetilde{d}_{l}=a^{l}d_{l}.

Thus, it is sufficient to first prove the boundedness of the coefficients c~l\widetilde{c}_{l} and d~l\widetilde{d}_{l}, and then establish the asymptotic nature of the error term in (4.2), namely, to prove that there exist positive constants ApA_{p}, ApA_{p}^{\prime}, BpB_{p} and BpB_{p}^{\prime} such that

(4.19) |εp1|Ap(aN)pAi~(N2/3ζ)N1/3+Ap(aN)pAi~(N2/3ζ)N2/3,|εp2|Bp(aN)pBi~(N2/3ζ)N1/3+Bp(aN)pBi~(N2/3ζ)N2/3,\begin{split}\left|\varepsilon_{p}^{1}\right|\leq&\frac{A_{p}}{(aN)^{p}}\frac{\widetilde{\text{Ai}}(N^{2/3}\zeta)}{N^{1/3}}+\frac{A_{p}^{\prime}}{(aN)^{p}}\frac{\widetilde{\text{Ai}}^{\prime}(N^{2/3}\zeta)}{N^{2/3}},\\ \left|\varepsilon_{p}^{2}\right|\leq&\frac{B_{p}}{(aN)^{p}}\frac{\widetilde{\text{Bi}}(N^{2/3}\zeta)}{N^{1/3}}+\frac{B_{p}^{\prime}}{(aN)^{p}}\frac{\widetilde{\text{Bi}}^{\prime}(N^{2/3}\zeta)}{N^{2/3}},\end{split}

where

(4.20) εp1=cos(aNπ)Nn+12πiNpΓ(n+1)0+C1hp(u,τ)eN(13u3ζu)eN(blnττ)dudτ,εp2=sin(aNπ)Nn+12πiNpΓ(n+1)0+C3C2hp(u,τ)eN(13u3ζu)eN(blnττ)dudτ,\begin{split}\varepsilon_{p}^{1}=&\frac{\cos(aN\pi)N^{n+1}}{2\pi iN^{p}\Gamma(n+1)}\int_{0}^{+\infty}\int_{C_{1}}h_{p}(u,\tau)e^{N\left(\frac{1}{3}u^{3}-\zeta u\right)}e^{N(b\ln\tau-\tau)}\mathrm{d}u\mathrm{d}\tau,\\ \varepsilon_{p}^{2}=&\frac{\sin(aN\pi)N^{n+1}}{2\pi iN^{p}\Gamma(n+1)}\int_{0}^{+\infty}\int_{C_{3}-C_{2}}h_{p}(u,\tau)e^{N\left(\frac{1}{3}u^{3}-\zeta u\right)}e^{N(b\ln\tau-\tau)}\mathrm{d}u\mathrm{d}\tau,\end{split}

and Ai~(z)=Bi~(z)=[Ai2(z)+Bi2(z)]1/2\widetilde{\text{Ai}}(z)=\widetilde{\text{Bi}}(z)=\left[\text{Ai}^{2}(z)+\text{Bi}^{2}(z)\right]^{1/2} for z<0z<0. Note that in the present case, ζ<0\zeta<0.

From (4.8), we recall that dw/du\mathrm{d}w/\mathrm{d}u depends on ζ\zeta. The value of ζ\zeta is obtained by solving the two equations gotten from (4.6) with ww and uu replaced respectively by w±w_{\pm} and ±ζ\pm\sqrt{\zeta}; see [11, p.367]. Thus, ζ\zeta and hence dw/du\mathrm{d}w/\mathrm{d}u both depend on the parameters aa and bb in (1.3). As aa and bb approach zero, ζ\zeta may tend to infinity. When the saddle points u±=±ζu_{\pm}=\pm\sqrt{\zeta} are bounded, the mapping wuw\rightarrow u defined by the cubic transformation (4.6) is analytic, and its derivative dw/du\mathrm{d}w/\mathrm{d}u is bounded. However, in the present case, the saddle points u±=±ζiu_{\pm}=\pm\sqrt{-\zeta}i go to infinity as ζ\zeta\rightarrow-\infty. Hence, the coefficients al(τ)a_{l}(\tau), bl(τ)b_{l}(\tau) and the function hl(u,τ)h_{l}(u,\tau) given recursively in (4.11) and (4.12) may all blow up as ζ\zeta\rightarrow-\infty, where h0(u,τ)h_{0}(u,\tau) is defined by (4.10). Since the coefficients clc_{l} and dld_{l} in (4.1) are related to al(τ)a_{l}(\tau) and bl(τ)b_{l}(\tau) via (4.16) and (4.13), to prove that the expansion in (4.2) is asymptotic when ζ\zeta\rightarrow-\infty, we must first give estimates for the coefficient functions al(τ)a_{l}(\tau) and bl(τ)b_{l}(\tau). To this end, we shall adopt a method introduced by Olde Daalhuis and Temme [5]. To begin with, we define

(4.21) A0(u,ζ)=uu2ζ,B0(u,ζ)=1u2ζ.A_{0}(u,\zeta)=\frac{u}{u^{2}-\zeta},\qquad B_{0}(u,\zeta)=\frac{1}{u^{2}-\zeta}.

Using (4.11) and the Cauchy residue theorem, it can be verified that

(4.22) a0(τ)=12πiΓh0(u,τ)A0(u,ζ)du,b0(τ)=12πiΓh0(u,τ)B0(u,ζ)du,\begin{split}a_{0}(\tau)=&\frac{1}{2\pi i}\int_{\Gamma}h_{0}(u,\tau)A_{0}(u,\zeta)\mathrm{d}u,\\ b_{0}(\tau)=&\frac{1}{2\pi i}\int_{\Gamma}h_{0}(u,\tau)B_{0}(u,\zeta)\mathrm{d}u,\end{split}

where Γ\Gamma is the contour consisting of two circles, centering at ±ζi\pm\sqrt{-\zeta}i, both with radius RR, where RR could be as large as possible until the circles reach the singularities of h0(u,τ)h_{0}(u,\tau) in the uu-plane.

We note that w=w±w=w_{\pm} are removable singularities of dw/du\mathrm{d}w/\mathrm{d}u. However, dw/du\mathrm{d}w/\mathrm{d}u blows up at the points w=w±e2kπiw=w_{\pm}e^{2k\pi i}, where k0k\neq 0 is any integer; see (4.8). To find their image points in the uu-plane under the mapping (4.6), we take wk=w+e2kπiw_{k}=w_{+}e^{2k\pi i} as an example and denote its image point by uku_{k}. The image point of we2kπiw_{-}e^{2k\pi i} can be treated in a similar manner. By (4.6), we have

(4.23) f¯(t0(wk),wk)=13uk3ζuk+A.\overline{f}(t_{0}(w_{k}),w_{k})=\frac{1}{3}u_{k}^{3}-\zeta u_{k}+A.

From (4.6) and (4.7), we also have

(4.24) f¯(t0(w+),w+)=13u+3ζu++A.\overline{f}(t_{0}(w_{+}),w_{+})=\frac{1}{3}u_{+}^{3}-\zeta u_{+}+A.

Subtracting (4.24) from (4.23) gives

(4.25) 2kπai=i(ζ)3/2(13zk3zk2).2k\pi ai=i(-\zeta)^{3/2}\left(-\frac{1}{3}z_{k}^{3}-z_{k}^{2}\right).

Since 0<a120<a\leq\frac{1}{2} and ζO(1)-\zeta\gg O(1), from (4.25) it follows that |zk|2kπa1/2(ζ)3/4|z_{k}|\sim\sqrt{2k\pi}a^{1/2}(-\zeta)^{-3/4}. Therefore, there exists a constant 0<c0<10<c_{0}<1, independent of aa, bb and ζ\zeta, such that the interior of the two circles with centres at u=±ζiu=\pm\sqrt{-\zeta}i and radius R=c0a(ζ)1/4R=c_{0}\sqrt{a}(-\zeta)^{-1/4} is free of the singularities of h0(u,τ)h_{0}(u,\tau) in the uu-plane. Since h0(u,τ)h_{0}(u,\tau) is now analytic inside the contour Γ\Gamma, there exists a constant chc_{h} such that

(4.26) |h0(u,τ)|chh(τ)\left|h_{0}(u,\tau)\right|\leq c_{h}h(\tau)

for uu in the domain enclosed by the contour Γ\Gamma, where h(τ)h(\tau) denotes the maximum of the two functions |h0(±ζi,τ)|\left|h_{0}(\pm\sqrt{-\zeta}i,\tau)\right|. Note that as functions of τ\tau, h0(±ζi,τ)h_{0}(\pm\sqrt{-\zeta}i,\tau) are analytic in the neighbourhood of steepest descent path in the τ\tau-plane.

We further introduce rational functions AkA_{k} and BkB_{k}, k=0,1,2k=0,1,2..., defined recursively by

(4.27) Ak+1(u,ζ)=1u2ζdduAk(u,ζ),Bk+1(u,ζ)=1u2ζdduBk(u,ζ),\begin{split}A_{k+1}(u,\zeta)=&\frac{1}{u^{2}-\zeta}\frac{\mathrm{d}}{\mathrm{d}u}A_{k}(u,\zeta),\\ B_{k+1}(u,\zeta)=&\frac{1}{u^{2}-\zeta}\frac{\mathrm{d}}{\mathrm{d}u}B_{k}(u,\zeta),\end{split}

where A0A_{0} and B0B_{0} are given in (4.21). By induction, we can show that AkA_{k} and BkB_{k} are expressible as

(4.28) Ak(u,ζ)=i=0[(k+1)/2]ck,i1uk+12i(u2ζ)2k+1i,Bk(u,ζ)=i=0[k/2]ck,i2uk2i(u2ζ)2ki,\begin{split}A_{k}(u,\zeta)=&\sum_{i=0}^{\left[(k+1)/2\right]}\frac{c^{1}_{k,i}u^{k+1-2i}}{\left(u^{2}-\zeta\right)^{2k+1-i}},\\ B_{k}(u,\zeta)=&\sum_{i=0}^{\left[k/2\right]}\frac{c^{2}_{k,i}u^{k-2i}}{\left(u^{2}-\zeta\right)^{2k-i}},\end{split}

where ck,i1c^{1}_{k,i} and ck,i2c^{2}_{k,i} are constants independent of uu and ζ\zeta. As in (4.22), by Cauchy’s theorem, we have from equations (4.11) and (4.12)

(4.29) ak(τ)=12πiΓhk(u,τ)A0(u,ζ)du=12πiΓhk1(u,τ)A1(u,ζ)du12πiΓ(ak1(τ)+bk1(τ)u)A1(u,ζ)du=12πiΓhk1(u,τ)A1(u,ζ)du=12πiΓh0(u,τ)Ak(u,ζ)du,\begin{split}a_{k}(\tau)=&\frac{1}{2\pi i}\int_{\Gamma}h_{k}(u,\tau)A_{0}(u,\zeta)\mathrm{d}u\\ =&\frac{1}{2\pi i}\int_{\Gamma}h_{k-1}(u,\tau)A_{1}(u,\zeta)\mathrm{d}u-\frac{1}{2\pi i}\int_{\Gamma}\left(a_{k-1}(\tau)+b_{k-1}(\tau)u\right)A_{1}(u,\zeta)\mathrm{d}u\\ =&\frac{1}{2\pi i}\int_{\Gamma}h_{k-1}(u,\tau)A_{1}(u,\zeta)\mathrm{d}u\\ \vdots&\\ =&\frac{1}{2\pi i}\int_{\Gamma}h_{0}(u,\tau)A_{k}(u,\zeta)\mathrm{d}u,\end{split}

where we have used integration-by-parts to derive the second equality. The second term in the second equality vanishes because (ak1(τ)+bk1(τ)u)A1(u,ζ)\left(a_{k-1}(\tau)+b_{k-1}(\tau)u\right)A_{1}(u,\zeta) is O(u2)O\left(u^{-2}\right) as |u||u|\rightarrow\infty and all poles of that function lie inside Γ\Gamma; see (4.28). Similarly, we also have

(4.30) bk(τ)=12πiΓh0(u,τ)Bk(u,ζ)du.b_{k}(\tau)=\frac{1}{2\pi i}\int_{\Gamma}h_{0}(u,\tau)B_{k}(u,\zeta)\mathrm{d}u.

Using (4.28), it is easy to obtain the estimates

(4.31) |Ak(u,ζ)|Ckak1/2(ζ)1/4and|Bk(u,ζ)|Ckak,\left|A_{k}(u,\zeta)\right|\leq C_{k}a^{-k-1/2}(-\zeta)^{1/4}\quad\text{and}\quad\left|B_{k}(u,\zeta)\right|\leq C_{k}a^{-k},

for uu on and inside the contour Γ\Gamma and ζ-\zeta\rightarrow\infty. Here and thereafter, CkC_{k} is used as a generic symbol for constants independent of uu, ζ\zeta, aa and bb. Substituting (4.31) into (4.29) and (4.30) gives

(4.32) |ak(τ)|Ckh(τ)ak,\left|a_{k}(\tau)\right|\leq C_{k}h(\tau)a^{-k},

and

(4.33) |bk(τ)|Ckh(τ)ak+1/2(ζ)1/4.\left|b_{k}(\tau)\right|\leq C_{k}h(\tau)a^{-k+1/2}(-\zeta)^{-1/4}.

Therefore, alal(τ)a^{l}a_{l}(\tau) and albl(τ)a^{l}b_{l}(\tau) are both bounded for τ\tau in the neighbourhood of steepest descent path in the τ\tau-plane, and of course for τ\tau in the neighbourhood of τ=0\tau=0. Thus we have the boundedness of the coefficients c~l\widetilde{c}_{l} and d~l\widetilde{d}_{l}.

To estimate hk(u,τ)h_{k}(u,\tau), we use the rational functions Rk(u,w,ζ)R_{k}(u,w,\zeta), k=0,1,2k=0,1,2..., defined recursively by

(4.34) R0(u,w,ζ)=1uw,Rk+1(u,w,ζ)=1u2ζdduRk(u,w,ζ).\begin{split}R_{0}(u,w,\zeta)=&\frac{1}{u-w},\\ R_{k+1}(u,w,\zeta)=&\frac{1}{u^{2}-\zeta}\frac{\mathrm{d}}{\mathrm{d}u}R_{k}(u,w,\zeta).\end{split}

These functions were also introduced by Olde Daalhuis and Temme [5]. They showed by induction that Rk(u,w,ζ)R_{k}(u,w,\zeta) can be written as

(4.35) Rk(u,w,ζ)=i=0k1j=0min{i,k1i}Cijuij(uw)k+1ij(u2ζ)k+i,k=1,2,R_{k}(u,w,\zeta)=\sum_{i=0}^{k-1}\sum_{j=0}^{\min\{i,k-1-i\}}\frac{C_{ij}u^{i-j}}{(u-w)^{k+1-i-j}\left(u^{2}-\zeta\right)^{k+i}},\qquad\qquad k=1,2...,

where CijC_{ij} do not depend on uu, ww and ζ\zeta. Similar to (4.29), we have

(4.36) hk(w,τ)=12πiΓh0(u,τ)Rk(u,w,ζ)du,h_{k}(w,\tau)=\frac{1}{2\pi i}\int_{\Gamma}h_{0}(u,\tau)R_{k}(u,w,\zeta)\mathrm{d}u,

where Γ\Gamma is the same contour used in (4.29) and ww lies inside two disks centered at ±ζi\pm\sqrt{-\zeta}i and with radius 12c0a(ζ)1/4\frac{1}{2}c_{0}\sqrt{a}(-\zeta)^{-1/4}. It is easy to verify from (4.35) that

(4.37) |Rk(u,w,ζ)|Ckak1/2(ζ)1/4,\left|R_{k}(u,w,\zeta)\right|\leq C_{k}a^{-k-1/2}(-\zeta)^{1/4},

and from (4.36) that

(4.38) |hk(w,τ)|Ckakh(τ).\left|h_{k}(w,\tau)\right|\leq C_{k}a^{-k}h(\tau).

Substituting (4.38) into (4.20) gives (4.19); for details, see [5, p.311-312]. Note that to make the expansion (4.2) asymptotic, we require x=aNx=aN to be large.

5 BESSEL-TYPE EXPANSION

In the case of Hahn Polynomials Qn(x;α,β,N)Q_{n}(x;\alpha,\beta,N) given in (1.1), Sharapodinov [10] has given an asymptotic formula involving Jacobi polynomials when the parameters satisfy α\alpha, β12\beta\geq\frac{1}{2} and 2ncN2\leq n\leq c\sqrt{N}, where cc is a positive constant. The values of the variable xx are also required to be large; more precisely, xεNx\geq\varepsilon N and ε>0\varepsilon>0 is a small number. Although discrete Chebyshev polynomials tn(x,N)t_{n}(x,N) given in (1.2) is a special case of the Hahn polynomials, the values of the parameters are α=β=0\alpha=\beta=0; that is, Sharapodinov’s result does not include our case. However, since the leading term in the uniform asymptotic expansion of the Jacobi polynomials is a Bessel function (see [7, p.451]), the work of Sharapodinov did inspire us to look for an asymptotic expansion for tn(x,N+1)t_{n}(x,N+1) involving Bessel functions, when the parameters aa and bb in (1.3) satisfy a/b2a/b^{2}\rightarrow\infty and the variable xx is large. Our method differs completely from that of Sharapodinov.

Returning to (2.1), and making the change of variable v=1/wv=1/w, we have

(5.1) tn(x,N+1)=(1)n2πiΓ(n+N+2)Γ(n+1)Γ(Nn+1)01γ31v(1v)eNf^(t,v)dvdt,t_{n}(x,N+1)=\frac{(-1)^{n}}{2\pi i}\frac{\Gamma(n+N+2)}{\Gamma(n+1)\Gamma(N-n+1)}\int_{0}^{1}\int_{\gamma_{3}}\frac{1}{v(1-v)}e^{N\widehat{f}(t,v)}\mathrm{d}v\mathrm{d}t,

where

(5.2) f^(t,v)=bln(1t)+(1b)lntaln(1v)blnv+bln[t+v1];\widehat{f}(t,v)=b\ln(1-t)+(1-b)\ln t-a\ln(1-v)-b\ln v+b\ln[t+v-1];

the curve γ3\gamma_{3} starts at v=+v=+\infty, runs along the lower edge of the positive real line towards v=1v=1, encircles the point v=1v=1 in the clockwise direction and returns to ++\infty along the upper edge of the positive real line. Note that since eNblnv=vne^{Nb\ln v}=v^{n}, there is no need to have a cut from the origin to infinity.

The saddle point of f^(t,v)\widehat{f}(t,v) in the tt-plane is given by

(5.3) t0(v)=2v+4b24b2v+v22(1+b);t_{0}(v)=\frac{2-v+\sqrt{4b^{2}-4b^{2}v+v^{2}}}{2(1+b)};

cf.(2.5), where the vv-plane is cut along two line segments joining 0 to the two conjugate points 2b2±2b1b2i2b^{2}\pm 2b\sqrt{1-b^{2}}i, and the branch of the square root is chosen so that 4b24b2v+v2v\sqrt{4b^{2}-4b^{2}v+v^{2}}\sim v as vv\rightarrow\infty. From (5.3), we have

(5.4) t0(v)=1+O(b)as |v|cb,t0(v)1bas vb,\begin{split}t_{0}(v)&=1+O(b)\qquad\hskip-5.69046pt\text{as }|v|\sim cb,\\ t_{0}(v)&\sim 1-b\qquad\hskip 14.22636pt\text{as }v\gg b,\end{split}

where cc is a positive constant, and for vbv\ll b,

(5.5) t0(v)112vas v0+,t0(v)12bas v0,\begin{split}t_{0}(v)&\sim 1-\frac{1}{2}v\qquad\text{as }v\rightarrow 0^{+},\\ t_{0}(v)&\rightarrow 1-2b\qquad\text{as }v\rightarrow 0^{-},\end{split}

where 0+0^{+} and 00^{-} mean limits approaching 0 from the right-hand side of the cut and the left-hand side of the cut, respectively. Moreover, easy calculation shows that for any vv, t0(v)t_{0}(v) is not a real number on the cut (1,)(1,\infty) in the tt-plane. Following the same argument given prior to (3.2), we introduce the mapping

(5.6) f^(t,v)=blnττ+A\widehat{f}(t,v)=b\ln\tau-\tau+A

with the correspondence between the saddle points t=t0(v)t=t_{0}(v) and τ=b\tau=b given by

(5.7) τ(t0(v))=b.\tau(t_{0}(v))=b.

Coupling (5.6) and (5.7) yields

(5.8) A=f^(t0(v),v)blnb+b.A=\widehat{f}(t_{0}(v),v)-b\ln b+b.

The zeros of f^(t,v)/t=0\partial\widehat{f}(t,v)/\partial t=0 are given by

(5.9) t0±(v)=2v±4b24b2v+v22(1+b),t_{0}^{\pm}(v)=\frac{2-v\pm\sqrt{4b^{2}-4b^{2}v+v^{2}}}{2(1+b)},

where t0+(v)t_{0}^{+}(v) is the relavant saddle point t0(v)t_{0}(v) given in (5.3)(\ref{saddle point of t in case 3}). By straightforward calculation, we have

(5.10) dtdτ=b/τ1f^(t,v)/t=(τb)t(1t)(t+v1)τ(1+b)(tt0+(v))(tt0(v)),τb,dtdτ={t0+(v)(1t0+(v))(t0+(v)+v1)b4b24b2v+v2}1/2,τ=b.\begin{split}\frac{\mathrm{d}t}{\mathrm{d}\tau}&=\frac{b/\tau-1}{\partial\widehat{f}(t,v)/\partial t}=\frac{(\tau-b)t(1-t)(t+v-1)}{\tau(1+b)(t-t^{+}_{0}(v))(t-t^{-}_{0}(v))},\quad\quad\tau\neq b,\\ \frac{\mathrm{d}t}{\mathrm{d}\tau}&=\left\{\frac{t^{+}_{0}(v)(1-t^{+}_{0}(v))(t^{+}_{0}(v)+v-1)}{b\sqrt{4b^{2}-4b^{2}v+v^{2}}}\right\}^{1/2},\hskip 45.52458pt\tau=b.\end{split}

Note that we have τ=0\tau=0 when t=1t=1, and τ=+\tau=+\infty when t=0t=0. For any fixed vv, we can deform the original interval of integration 0t10\leq t\leq 1 into a steepest descent path Γv\Gamma_{v}, passing through t0(v)t_{0}(v). Also note that t0(v)t_{0}^{-}(v) is not on Γv\Gamma_{v}, unless 0<v<10<v<1. Moreover, if 0<v<10<v<1, then 0<t0(v)<t0+(v)<10<t_{0}^{-}(v)<t_{0}^{+}(v)<1 and Γv\Gamma_{v} is the real interval [0,1][0,1]. In our case, there is only one point on the path γ3\gamma_{3} in the vv-plane (see (5.1)), where it crosses the real line. Let us denote this point by v0v_{0}. When v=v0v=v_{0}, the mapping which we have introduced in (5.6) becomes singular at the point t0(v)t_{0}^{-}(v), since dt/dτ\mathrm{d}t/\mathrm{d}\tau in (5.10) blows up. However for this particular case, we only need to slightly modify the path by replacing part of the original path near this point by a small half circle as shown in Figure 3.

Refer to caption
Figure 3: The indentation of the integration path in the tt-plane for 0<v<10<v<1.

The corresponding integration path in the τ\tau-plane following the mapping (5.6) also needs to be modified. However, this small modification on the integration path will not affect the following argument and calculation. With this in mind, we will simply ignore this particular case, and proceed with the assumption that the integration path in the τ\tau-plane is always [0,+)[0,+\infty) for all ww, and that dt/dτ\mathrm{d}t/\mathrm{d}\tau will not blow up in the neighbourhood of the path.

Thus, from (5.1) and (5.6), we have

(5.11) tn(x,N+1)=(1)n+12πiΓ(n+N+2)enbnΓ(n+1)Γ(Nn+1)×0+γ31v(1v)eNf^(t0(v),v)eN(blnττ)dtdτdvdτ.\begin{split}t_{n}(x,N+1)=&\frac{(-1)^{n+1}}{2\pi i}\frac{\Gamma(n+N+2)e^{n}b^{-n}}{\Gamma(n+1)\Gamma(N-n+1)}\\ &\times\int_{0}^{+\infty}\int_{\gamma_{3}}\frac{1}{v(1-v)}e^{N\widehat{f}({t_{0}(v)},v)}e^{N(b\ln\tau-\tau)}\frac{\mathrm{d}t}{\mathrm{d}\tau}\mathrm{d}v\mathrm{d}\tau.\end{split}

Here, we rewrite the phase function f^(t0(v),v)\widehat{f}(t_{0}(v),v) as

(5.12) f^(t0(v),v)=(1b)lnt0(v)+bln(1t0(v))aln(1v)bln(vt0(v)+v1).\widehat{f}(t_{0}(v),v)=(1-b)\ln t_{0}(v)+b\ln(1-t_{0}(v))-a\ln(1-v)-b\ln(\frac{v}{t_{0}(v)+v-1}).

Recalling the statements following (5.2) and (5.3), we know that there are only two cuts in the vv-plane: one along the infinite interval [1,+)[1,+\infty) and the other along the bent line joining the conjugate points 2b2±2b1b2i2b^{2}\pm 2b\sqrt{1-b^{2}}i and passing through the origin. To find the saddle points of f^(t0(v),v)\widehat{f}(t_{0}(v),v), we set

f^(t0(v),v)v=0,\frac{\partial\widehat{f}(t_{0}(v),v)}{\partial v}=0,

and obtain

(5.13) v±=b2bi4a4a2b22a(1a).v_{\pm}=\frac{b^{2}\mp bi\sqrt{4a-4a^{2}-b^{2}}}{2a(1-a)}.

Since (5.1) is obtained from (2.1) by making the change of variable v=1/wv=1/w, (5.13) can also be derived from (2.9). Note that in this case, b2/a0b^{2}/a\rightarrow 0. Furthermore, since a<12a<\frac{1}{2}, the quantity inside the square root is positive. Hence, v±v_{\pm} are distinct, and approach v=0v=0.

Define

g(u)=m(u1u),g(u)=m(u-\frac{1}{u}),

where m>0m>0 is some constant to be determined. The saddle points of g(u)g(u) are

(5.14) u±=±i.u_{\pm}=\pm i.

Make the transformation

(5.15) f^(t0(v),v)=g(u)+γ=m(u1u)+γ\begin{split}\widehat{f}(t_{0}(v),v)&=g(u)+\gamma\\ &=m(u-\frac{1}{u})+\gamma\end{split}

with

(5.16) u(v+)=u,u(v)=u+.u(v_{+})=u_{-},\quad\quad\quad\quad u(v_{-})=u_{+}.

Note the fact that t0(v±)=t±t_{0}(v_{\pm})=t_{\pm}, where t±t_{\pm} are given in (2.3) and (2.4). This can be seen from (2.6) and the change of variable v=1/wv=1/w that we have made. Thus, substituting (5.16) into (5.15) gives

(5.17) γ=12ln1b1+b+12blnb21b2=blnbb+O(b3)\begin{split}\gamma&=\frac{1}{2}\ln\frac{1-b}{1+b}+\frac{1}{2}b\ln\frac{b^{2}}{1-b^{2}}\\ &=b\ln b-b+O(b^{3})\end{split}

and

(5.18) m=12{(1b)arctanb4a4a2b222ab2aarctanb4a4a2b22a2a2b22barctan4a4a2b22+b2a}barctana1a,\begin{split}m=-\frac{1}{2}\biggr{\{}&(1-b)\arctan\frac{b\sqrt{4a-4a^{2}-b^{2}}}{2-2a-b^{2}}-a\arctan\frac{b\sqrt{4a-4a^{2}-b^{2}}}{2a-2a^{2}-b^{2}}\\ &\left.-2b\arctan\frac{\sqrt{4a-4a^{2}-b^{2}}}{2+b-2a}\right\}\\ \sim\ b&\arctan\sqrt{\frac{a}{1-a}},\end{split}

when b2/a0b^{2}/a\rightarrow 0 and b0b\rightarrow 0. Moreover, from (5.15) we have

(5.19) dvdu=m(ui)(u+i)v(1v)[(12a)v+4b24b2v+v2]2a(1a)u2(vv+)(vv),uu±,dvdu|u=u±={2m(v)2(1v)(12a)b4a4a2b2}1/2,u=u±.\begin{split}&\frac{\mathrm{d}v}{\mathrm{d}u}=\frac{m(u-i)(u+i)v(1-v)\left[(1-2a)v+\sqrt{4b^{2}-4b^{2}v+v^{2}}\right]}{2a(1-a)u^{2}(v-v_{+})(v-v_{-})},\quad\quad u\neq u_{\pm},\\ &\left.\frac{\mathrm{d}v}{\mathrm{d}u}\right|_{u=u_{\pm}}=\left\{\frac{-2m(v_{\mp})^{2}(1-v_{\mp})(1-2a)}{b\sqrt{4a-4a^{2}-b^{2}}}\right\}^{1/2},\qquad\qquad u=u_{\pm}.\end{split}

Here we have made use of the equality 4b24b2v+v2=(12a)v\sqrt{4b^{2}-4b^{2}v_{\mp}+v_{\mp}^{2}}=(1-2a)v_{\mp}.

Coupling (5.11) and (5.15), the integral representation of tn(x,N+1)t_{n}(x,N+1) becomes

(5.20) tn(x,N+1)=(1)n+12πiΓ(n+N+2)enbneNγΓ(n+1)Γ(Nn+1)×0+γ^h(u,τ)ueN[m(u1u)]eN(blnττ)dudτ,\begin{split}t_{n}(x,N+1)=&\frac{(-1)^{n+1}}{2\pi i}\frac{\Gamma(n+N+2)e^{n}b^{-n}e^{N\gamma}}{\Gamma(n+1)\Gamma(N-n+1)}\\ &\times\int_{0}^{+\infty}\int_{\widehat{\gamma}}\frac{h(u,\tau)}{u}e^{N\left[m\left(u-\frac{1}{u}\right)\right]}e^{N(b\ln\tau-\tau)}\mathrm{d}u\mathrm{d}\tau,\end{split}

where

(5.21) h(u,τ)=uv(1v)dtdτdvdu,h(u,\tau)=\frac{u}{v(1-v)}\frac{\mathrm{d}t}{\mathrm{d}\tau}\frac{\mathrm{d}v}{\mathrm{d}u},

and the contour γ^\widehat{\gamma} starts from -\infty, encircles the point u=0u=0 in the counterclockwise direction and returns to -\infty. For l=0,1,2,l=0,1,2,\cdot\cdot\cdot, we define recursively

(5.22) hl(u,τ)\displaystyle h_{l}(u,\tau) =\displaystyle= al(τ)+bl(τ)u+(1+1u2)gl(u,τ),\displaystyle a_{l}(\tau)+\frac{b_{l}(\tau)}{u}+\left(1+\frac{1}{u^{2}}\right)g_{l}(u,\tau),
(5.23) hl+1(u,τ)\displaystyle h_{l+1}(u,\tau) =\displaystyle= umddu{gl(u,τ)u},\displaystyle-\frac{u}{m}\frac{\mathrm{d}}{\mathrm{d}u}\left\{\frac{g_{l}(u,\tau)}{u}\right\},

where h0(u,τ)=h(u,τ)h_{0}(u,\tau)=h(u,\tau) given in (5.21). Furthermore, we expand al(τ)a_{l}(\tau) and bl(τ)b_{l}(\tau) at τ=0\tau=0, and write

(5.24) al(τ)=j=0al,jτj,bl(τ)=j=0bl,jτj.a_{l}(\tau)=\sum_{j=0}^{\infty}a_{l,j}\tau^{j},\qquad b_{l}(\tau)=\sum_{j=0}^{\infty}b_{l,j}\tau^{j}.

It is easy to see that

(5.25) al(τ)=12[hl(i,τ)+hl(i,τ)],bl(τ)=i2[hl(i,τ)hl(i,τ)].\begin{split}a_{l}(\tau)&=\frac{1}{2}\left[h_{l}(i,\tau)+h_{l}(-i,\tau)\right],\\ b_{l}(\tau)&=\frac{i}{2}\left[h_{l}(i,\tau)-h_{l}(-i,\tau)\right].\end{split}

From (5.20), (5.22) and (5.23), we have

tn(x,N+1)=\displaystyle t_{n}(x,N+1)= (1)n+12πiΓ(n+N+2)enbneNγΓ(n+1)Γ(Nn+1)\displaystyle\frac{(-1)^{n+1}}{2\pi i}\frac{\Gamma(n+N+2)e^{n}b^{-n}e^{N\gamma}}{\Gamma(n+1)\Gamma(N-n+1)}
(5.26) ×0+(0+)[a0(τ)+b0(τ)u+(1+1u2)g0(u,τ)]eNm(u1u)uτneNτdudτ\displaystyle\times\int_{0}^{+\infty}\int_{-\infty}^{\left(0^{+}\right)}\left[a_{0}(\tau)+\frac{b_{0}(\tau)}{u}+\left(1+\frac{1}{u^{2}}\right)g_{0}(u,\tau)\right]\frac{e^{Nm\left(u-\frac{1}{u}\right)}}{u}\tau^{n}e^{-N\tau}\mathrm{d}u\mathrm{d}\tau
=\displaystyle= (1)n+1Γ(n+N+2)enbneNγΓ(Nn+1)Nn+1[c0J0(2Nm)+d0J1(2Nm)+ε1+],\displaystyle\frac{(-1)^{n+1}\Gamma(n+N+2)e^{n}b^{-n}e^{N\gamma}}{\Gamma(N-n+1)N^{n+1}}\biggr{[}c_{0}J_{0}(2Nm)+d_{0}J_{1}(2Nm)+\varepsilon_{1}^{+}\biggr{]},

where

(5.27) c0=Nn+1Γ(n+1)0a0(τ)τneNτdτm=0a0,mΓ(n+m+1)Γ(n+1)Nm,d0=Nn+1Γ(n+1)0b0(τ)τneNτdτm=0b0,mΓ(n+m+1)Γ(n+1)Nm\begin{split}c_{0}&=\frac{N^{n+1}}{\Gamma(n+1)}\int_{0}^{\infty}a_{0}(\tau)\tau^{n}e^{-N\tau}\mathrm{d}\tau\sim\sum_{m=0}^{\infty}a_{0,m}\frac{\Gamma(n+m+1)}{\Gamma(n+1)N^{m}},\\ d_{0}&=\frac{N^{n+1}}{\Gamma(n+1)}\int_{0}^{\infty}b_{0}(\tau)\tau^{n}e^{-N\tau}\mathrm{d}\tau\sim\sum_{m=0}^{\infty}b_{0,m}\frac{\Gamma(n+m+1)}{\Gamma(n+1)N^{m}}\end{split}

and

(5.28) ε1+=Nn+12Γ(n+1)πi0+(0+)(1+1u2)g0(u,τ)ueNm(u1u)τneNτdudτ.\begin{split}\varepsilon_{1}^{+}=\frac{N^{n+1}}{2\Gamma(n+1)\pi i}\int_{0}^{+\infty}\int_{-\infty}^{\left(0^{+}\right)}\left(1+\frac{1}{u^{2}}\right)\frac{g_{0}(u,\tau)}{u}e^{Nm\left(u-\frac{1}{u}\right)}\tau^{n}e^{-N\tau}\mathrm{d}u\mathrm{d}\tau.\end{split}

In (5), we have made use of the integral representations of Bessel function and Gamma function

Jv(z)=12πi(0+)uv1exp{z2(u1u)}duJ_{v}(z)=\frac{1}{2\pi i}\int_{-\infty}^{\left(0^{+}\right)}u^{-v-1}\exp\left\{\frac{z}{2}\left(u-\frac{1}{u}\right)\right\}\mathrm{d}u

and

Γ(n+1)=Nn+10+eNττndτ;\Gamma(n+1)=N^{n+1}\int_{0}^{+\infty}e^{-N\tau}\tau^{n}\mathrm{d}\tau;

see [7, (10.9.19) and (5.9.1)]. Using (5.23) and integration by parts, we can rewrite ε1+\varepsilon_{1}^{+} in (5.28) as

(5.29) ε1+=Nn2Γ(n+1)πi0+(0+)h1(u,τ)ueNm(u1u)τneNτdudτ.\begin{split}\varepsilon_{1}^{+}=\frac{N^{n}}{2\Gamma(n+1)\pi i}\int_{0}^{+\infty}\int_{-\infty}^{\left(0^{+}\right)}\frac{h_{1}(u,\tau)}{u}e^{Nm\left(u-\frac{1}{u}\right)}\tau^{n}e^{-N\tau}\mathrm{d}u\mathrm{d}\tau.\end{split}

Repeating the procedure above, we obtain

(5.30) tn(x,N+1)=(1)n+1Γ(n+N+2)enbneNγΓ(Nn+1)Nn+1[J0(2Nm)l=0p1clNl+J1(2Nm)l=0p1dlNl+εp+],\begin{split}t_{n}(x,N+1)=\frac{(-1)^{n+1}\Gamma(n+N+2)e^{n}b^{-n}e^{N\gamma}}{\Gamma(N-n+1)N^{n+1}}\biggr{[}&J_{0}(2Nm)\sum_{l=0}^{p-1}\frac{c_{l}}{N^{l}}\\ &\left.+J_{1}(2Nm)\sum_{l=0}^{p-1}\frac{d_{l}}{N^{l}}+\varepsilon_{p}^{+}\right],\end{split}

where

(5.31) cl\displaystyle c_{l} =\displaystyle= Nn+1Γ(n+1)0al(τ)τneNτdτm=0al,mΓ(n+m+1)Γ(n+1)Nm,\displaystyle\frac{N^{n+1}}{\Gamma(n+1)}\int_{0}^{\infty}a_{l}(\tau)\tau^{n}e^{-N\tau}\mathrm{d}\tau\sim\sum_{m=0}^{\infty}a_{l,m}\frac{\Gamma(n+m+1)}{\Gamma(n+1)N^{m}},
(5.32) dl\displaystyle d_{l} =\displaystyle= Nn+1Γ(n+1)0bl(τ)τneNτdτm=0bl,mΓ(n+m+1)Γ(n+1)Nm,\displaystyle\frac{N^{n+1}}{\Gamma(n+1)}\int_{0}^{\infty}b_{l}(\tau)\tau^{n}e^{-N\tau}\mathrm{d}\tau\sim\sum_{m=0}^{\infty}b_{l,m}\frac{\Gamma(n+m+1)}{\Gamma(n+1)N^{m}},
(5.33) εp+\displaystyle\varepsilon_{p}^{+} =\displaystyle= Nn+1p2Γ(n+1)πi0+(0+)hp(u,τ)ueNm(u1u)τneNτdudτ.\displaystyle\frac{N^{n+1-p}}{2\Gamma(n+1)\pi i}\int_{0}^{+\infty}\int_{-\infty}^{\left(0^{+}\right)}\frac{h_{p}(u,\tau)}{u}e^{Nm\left(u-\frac{1}{u}\right)}\tau^{n}e^{-N\tau}\mathrm{d}u\mathrm{d}\tau.

Since h0(u,τ)h_{0}(u,\tau) involves dv/du{\mathrm{d}v}/{\mathrm{d}u} and dv/du{\mathrm{d}v}/{\mathrm{d}u} depends on the parameters aa and bb in (1.3), the coefficients clc_{l} and dld_{l} in (5.31) and (5.32) also depend on aa and bb. For an estimate on these coefficients, see (5.57) below.

To facilitate the application of expansion (5.30), we recall that the constants γ\gamma and mm are explicitly given in (5.17) and (5.18), and note that the leading coefficient can be (asymptotically) calculated by using (5.24) and (5.25). Indeed, we have

c0a0,0=12[h0(i,0)+h(i,0)]c_{0}\sim a_{0,0}=\frac{1}{2}\left[h_{0}(i,0)+h(-i,0)\right]

and

d0b0,0=i2[h0(i,0)h(i,0)]d_{0}\sim b_{0,0}=\frac{i}{2}\left[h_{0}(i,0)-h(-i,0)\right]

as b0b\rightarrow 0 and NN\rightarrow\infty. Furthermore, equation (5.39) gives

(5.34) h0(±i,0)=(1a)vb2(12a)mb4a4a2b2(1v)e±Δi,\begin{split}h_{0}(\pm i,0)=\frac{-(1-a)v_{\mp}}{b}\sqrt{\frac{2(1-2a)m}{b\sqrt{4a-4a^{2}-b^{2}}(1-v_{\mp})}}e^{\pm\Delta i},\end{split}

where Δ=2arctana1aa1a\Delta=2\arctan\sqrt{\frac{a}{1-a}}-\sqrt{\frac{a}{1-a}}.

5.1 The mapping vuv\rightarrow u in (5.15)

In the case under discussion, a/b2a/b^{2}\rightarrow\infty and xx\rightarrow\infty. Thus, since 0<a<120<a<\frac{1}{2}, we have 4a4a2b2>04a-4a^{2}-b^{2}>0, and the saddle points in (5.13) are complex.

Theorem 1.

When a(0,12]a\in(0,\frac{1}{2}], b0+b\rightarrow 0^{+} and b2/a0b^{2}/a\rightarrow 0, the mapping vuv\rightarrow u defined in (5.15) is one-to-one and analytic for vDvv\in D_{v} in the vv-plane and uDuu\in D_{u} in the uu-plane, where the image of the boundary of DuD_{u} in the ZZ-plane is given in (5.36), and DvD_{v} is the image of DuD_{u} under the mapping (5.15).

Proof.

As in the cases of Charlier polynomials [9] and Meixner polynomials [3], we introduce an intermediate variable ZZ defined by

(5.35) bln(1t0(v))+(1b)lnt0(v)aln(1v)bln(vt0(v)+v1)γ=Z=m(u1u),\begin{split}b\ln(1-t_{0}(v))+&(1-b)\ln t_{0}(v)-a\ln(1-v)-b\ln(\frac{v}{t_{0}(v)+v-1})-\gamma\\ =&Z=m\left(u-\frac{1}{u}\right),\end{split}

where t0(v)t_{0}(v) is the relevant saddle point given in (5.3).

Refer to caption
Figure 4: The upper half of the vv-plane.
Refer to caption
Figure 5: The image of Region I in the ZZ-plane.
Refer to caption
Figure 6: The image of Region II in the ZZ-plane.
Refer to caption
Figure 7: The upper half of the uu-plane.

We consider the upper half of the vv-plane. Since the functions f^(t0(v),v)γ\widehat{f}(t_{0}(v),v)-\gamma and g(u)g(u) are both symmetric with respect to the real line, the case for the lower half of the vv-plane can be handled in the same manner. To avoid multi-valuedness around the saddle point, we divide the upper half of the vv-plane into two parts by using the steepest descent path through vv_{-}, i.e., the point denoted by DD; see Figure 4. Call the two parts region I and region III. Note that there are two branch cuts: one along the infinite interval [1,)[1,\infty), and the other along the line segment joining v=0v=0 to v=2b2+i2b1b2v=2b^{2}+i2b\sqrt{1-b^{2}}, which was introduced in (5.3). We denoted the point 2b2+i2b1b22b^{2}+i2b\sqrt{1-b^{2}} by the letter TT in Figure 4.

As vv traverses along the boundary ABCDEFGAABCDEFGA of region I, the image point ZZ traverses along the corresponding boundary of a region in the ZZ-plane; see Figure 5. In Figure 7, we draw the boundary of the region, corresponding to region I, in the uu-plane. The image point Z=ψ(u)+γZ=\psi(u)+\gamma traverses along the boundary of the same region shown in Figure 5. Note that we have used an arc GABCGABC to avoid the cut OTOT. The boundary curves EFEF and GABCGABC in the vv-plane are rather arbitrary; for convenience, we choose them to be the ones whose images in the ZZ-plane are as indicated in Figure 5.

From (5.15), it is readily seen that the mapping vuv\rightarrow u is the composite function of g1g^{-1} and f^(t0(v),v)γ\widehat{f}(t_{0}(v),v)-\gamma. We have just verified that this mapping is one-to-one on the boundary of region I. By the same argument, one can prove that this mapping is also one-to-one on the boundary of region III. For the image of region III in the intermediate ZZ-plane, see Figure 6. In the uu-plane, we let DuD_{u} denote the union of the regions I to IV, with outer boundary FEEKJIVUTNNMFEE^{\prime}KJIVUTN^{\prime}NM, and inner boundary GABCCHWQQRSLGABCC^{\prime}HWQ^{\prime}QRSL; see Figure 7. Under the mapping Z=g(u)Z=g(u), we express the images of parts of the outer and inner boundaries of DuD_{u} in the ZZ-plane as follows:

(5.36) GA:Re Z=bM0Im Zθbπ;AB:bMRe ZbMIm Z=θbπ;BC and CH:Re Z=bM0Im Zθbπ;IJ:Re Z=aM0Im Zaπ;JK:aMRe ZaMIm Z=aπ;KF:Re Z=aM0Im Zaπ,\begin{split}GA:\qquad&\text{Re }Z=bM\hskip 42.67912pt\qquad 0\leq\text{Im }Z\leq\theta b\pi;\\ AB:\qquad&-bM\leq\text{Re }Z\leq bM\hskip 34.14322pt\text{Im }Z=\theta b\pi;\\ BC\text{ and }C^{\prime}H:\qquad&\text{Re }Z=-bM\hskip 54.06006pt0\leq\text{Im }Z\leq\theta b\pi;\\ IJ:\qquad&\text{Re }Z=aM\hskip 63.16515pt0\leq\text{Im }Z\leq a\pi;\\ JK:\qquad&-aM\leq\text{Re }Z\leq aM\hskip 34.14322pt\text{Im }Z=a\pi;\\ KF:\qquad&\text{Re }Z=-aM\hskip 56.9055pt0\leq\text{Im }Z\leq a\pi,\end{split}

where M>0M>0 is any fixed constant, and θ\theta is a generic symbol for a constant in (0,1)(0,1). Furthermore, we denote by DvD_{v} the corresponding region of DuD_{u} in the vv-plane.

By Theorem 1.2.2 of [13, p.12], the mapping vuv\rightarrow u is one-to-one in the interior of both regions I and III in the upper half of the vv-plane. As explained earlier, the one-to-one property of this mapping in the lower half of the vv-plane can be established by using the symmetry of the functions with respect to the real axis. Note that the only possible singular points of the mapping vuv\rightarrow u in DvD_{v} are at v=vv=v_{\mp}. Since the images of these points in the uu-plane (i.e., u=±iu=\pm i) are bounded, the mapping is in fact one-to-one and analytic in DvD_{v}.

5.2 Analyticity of h0(u,τ)h_{0}(u,\tau)

We now investigate the function h0(u,τ)=h(u,τ)h_{0}(u,\tau)=h(u,\tau) given in (5.21). By (5.10) and (5.19), we have

(5.37) h0(u,τ)=m(τb)t(1t)(t+v1)(ui)(u+i)[(12a)v+4b24b2v+v2]2a(1a)τ(1+b)(tt0+(v))(tt0(v))u(vv+)(vv)h_{0}(u,\tau)=\frac{m(\tau-b)t(1-t)(t+v-1)(u-i)(u+i)\left[(1-2a)v+\sqrt{4b^{2}-4b^{2}v+v^{2}}\right]}{2a(1-a)\tau(1+b)(t-t^{+}_{0}(v))(t-t^{-}_{0}(v))u(v-v_{+})(v-v_{-})}

when τb\tau\neq b and u±iu\neq\pm i, and

(5.38) h0(u,τ)=m(ui)(u+i)[(12a)v+4b24b2v+v2]2a(1a)u(vv+)(vv)×t0(v)(1t0(v))(t0(v)+v1)b4b24b2v+v2\begin{split}h_{0}(u,\tau)=&\frac{m(u-i)(u+i)\left[(1-2a)v+\sqrt{4b^{2}-4b^{2}v+v^{2}}\right]}{2a(1-a)u(v-v_{+})(v-v_{-})}\\ &\times\sqrt{\frac{t_{0}(v)(1-t_{0}(v))(t_{0}(v)+v-1)}{b\sqrt{4b^{2}-4b^{2}v+v^{2}}}}\end{split}

when τ=b\tau=b and u±iu\neq\pm i, where uDuu\in D_{u} and DuD_{u} is defined in the previous subsection. Note that u=0u=0 is excluded from DuD_{u}. It is easy to see that (t1)/τ(t-1)/\tau is bounded by a constant, independent of bb and aa, as τ0\tau\rightarrow 0 for vDvv\in D_{v} and uDuu\in D_{u}. Therefore, h0(u,τ)h_{0}(u,\tau) is analytic in τ\tau for τ\tau in the neighbourhood of [0,[0,\infty). Moreover, h0(u,τ)h_{0}(u,\tau) is also analytic in uu for uDuu\in D_{u}, since the only singularities u=±iu=\pm i in DuD_{u} are both removable; indeed, we have

(5.39) h0(u,τ)=(τb)t(1t)(t+v1)τ(1+b)(tt0+(v))(tt0(v))2(12a)m(1v)b4a4a2b2h_{0}(u,\tau)=\frac{(\tau-b)t(1-t)(t+v_{\mp}-1)}{\tau(1+b)(t-t^{+}_{0}(v_{\mp}))(t-t^{-}_{0}(v_{\mp}))}\sqrt{\frac{2(1-2a)m}{(1-v_{\mp})b\sqrt{4a-4a^{2}-b^{2}}}}

when τb\tau\neq b and u=±iu=\pm i, and by the equation following (5.19) we also have

(5.40) h0(u,τ)=2mt(1t)(t+v1)b2v(1v)4a4a2b2h_{0}(u,\tau)=\sqrt{\frac{2mt_{\mp}(1-t_{\mp})(t_{\mp}+v_{\mp}-1)}{b^{2}v_{\mp}(1-v_{\mp})\sqrt{4a-4a^{2}-b^{2}}}}

when τ=b\tau=b and u=±iu=\pm i. To reach (5.40), we have written u±u_{\pm} as u±=u±2u_{\pm}=\sqrt{u_{\pm}^{2}} and move u±2u_{\pm}^{2} inside the square root in dv/du\mathrm{d}v/\mathrm{d}u at u=u±u=u_{\pm}; see (5.19).

For the analysis to be used in the next subsection, we now give an estimate for h0(u,0)h_{0}(u,0) when uu lies on the boundaries of DuD_{u}, i.e, the outer boundary FEEKJIVUTNNMFEE^{\prime}KJIVUTN^{\prime}NM and the inner boundary GABCCHWQQRSLGABCC^{\prime}HWQ^{\prime}QRSL shown in Figure 7. For convenience, let us denote the outer boundary by COC_{O} and the inner boundary by CIC_{I}.

First, we show that for uu on the inner boundary CIC_{I},

(5.41) C1mb<|u|<C2mb,C_{1}\frac{m}{b}<|u|<C_{2}\frac{m}{b},

where C1<C2C_{1}<C_{2} and CC, C1C_{1}, C2C_{2} are used, here and thereafter, as generic symbols for constants independent of uu, vv, mm, aa and bb. Recall from (5.18) that mbarctana1am\sim b\arctan\sqrt{\frac{a}{1-a}} as b0b\rightarrow 0. To prove (5.41), we take the part ABAB of the inner boundary CIC_{I} as an illustration. Using (5.36), we obtain

(5.42) |m(11|u|2)Re u|bMandm(1+1|u|2)Im u=θbπ.\left|m\left(1-\frac{1}{|u|^{2}}\right)\text{Re }u\right|\leq bM\quad\text{and}\quad m\left(1+\frac{1}{|u|^{2}}\right)\text{Im }u=\theta b\pi.

Since |Im u||u|\left|\text{Im }u\right|\leq\left|u\right|, from the equality in (5.42) we have |u|2θbπm|u|+10|u|^{2}-\frac{\theta b\pi}{m}|u|+1\geq 0. Therefore,

(5.43) |u|θbπ2m(θbπ2m)21Cmbπ.|u|\leq\frac{\theta b\pi}{2m}-\sqrt{\left(\frac{\theta b\pi}{2m}\right)^{2}-1}\leq\frac{Cm}{b\pi}.

Rewriting (5.42) gives

(5.44) m|u|2|Re u|bM+m|Re u|andm|u|2|Im u|θbπ+m|Im u|,\frac{m}{|u|^{2}}\left|\text{Re }u\right|\leq bM+m\left|\text{Re }u\right|\quad\text{and}\quad\frac{m}{|u|^{2}}\left|\text{Im }u\right|\leq\theta b\pi+m\left|\text{Im }u\right|,

which lead to m|u|bM+θbπ+m|u|\frac{m}{|u|}\leq bM+\theta b\pi+m|u|. Substituting (5.43) into the right-hand side of the last inequality, and combining the resulting inequality with (5.43), we obtain (5.41) immediately. Similarly, for uu on the outer boundary COC_{O}, we have the estimates

(5.45) C1am<|u|<C2am.C_{1}\frac{a}{m}<|u|<C_{2}\frac{a}{m}.

Hence, it follows from (5.41), (5.45) and (5.37) that

(5.46) |h(u,0)|C for uCICO.|h(u,0)|\leq C\qquad\text{ for }u\in C_{I}\cup C_{O}.

5.3 Error bounds for the remainder

To prove the asymptotic nature of the expansion in (5.30), we need to give precise estimates for its coefficients clc_{l}, dld_{l} in (5.31) and (5.32), and the error term εp+\varepsilon_{p}^{+} given in (5.33), since the derivative dv/du\mathrm{d}v/\mathrm{d}u in (5.19) may blow up as v±v_{\pm} approach 0, just like the case in Section 4. Because the coefficients clc_{l} and dld_{l} are related to ap(τ)a_{p}(\tau) and bp(τ)b_{p}(\tau) by (5.24), (5.31) and (5.32), let us first estimate ap(τ)a_{p}(\tau) and bp(τ)b_{p}(\tau). To this end, we define recursively

A0(u)=u1+u2,B0(u)=11+u2,A_{0}(u)=\frac{u}{1+u^{2}},\qquad B_{0}(u)=-\frac{1}{1+u^{2}},

and

Ap+1(u)\displaystyle A_{p+1}(u) =\displaystyle= 1m(1+1u2)1{1u+ddu}Ap(u),\displaystyle\frac{1}{m}\left(1+\frac{1}{u^{2}}\right)^{-1}\left\{\frac{1}{u}+\frac{\mathrm{d}}{\mathrm{d}u}\right\}A_{p}(u),
Bp+1(u)\displaystyle B_{p+1}(u) =\displaystyle= 1m(1+1u2)1{1u+ddu}Bp(u)\displaystyle\frac{1}{m}\left(1+\frac{1}{u^{2}}\right)^{-1}\left\{\frac{1}{u}+\frac{\mathrm{d}}{\mathrm{d}u}\right\}B_{p}(u)

for p=0,1,2,p=0,1,2,.... By induction, it can be shown that

(5.47) Ap(u)=(12m)p(1u)p+1(u21+u2)p+1l=0pcp,l(u21+u2)l,Bp(u)=(12m)p(1u)p+2(u21+u2)p+1l=0pdp,l(u21+u2)l\begin{split}A_{p}(u)&=\left(\frac{1}{2m}\right)^{p}\left(\frac{1}{u}\right)^{p+1}\left(\frac{u^{2}}{1+u^{2}}\right)^{p+1}\sum_{l=0}^{p}c_{p,l}\left(\frac{u^{2}}{1+u^{2}}\right)^{l},\\ B_{p}(u)&=\left(\frac{1}{2m}\right)^{p}\left(\frac{1}{u}\right)^{p+2}\left(\frac{u^{2}}{1+u^{2}}\right)^{p+1}\sum_{l=0}^{p}d_{p,l}\left(\frac{u^{2}}{1+u^{2}}\right)^{l}\end{split}

for k=0,1,2,k=0,1,2,..., where ck,lc_{k,l} and dk,ld_{k,l} are constants independent of mm and uu (but dependent of τ\tau); see [12]. Furthermore, by (5.25), (5.22) and (5.23), it can be proved that

(5.48) ap(τ)=12πiCOCIh0(u,τ)Ap(u)dua_{p}(\tau)=\frac{1}{2\pi i}\int_{C_{O}\bigcup C_{I}}h_{0}(u,\tau)A_{p}(u)\mathrm{d}u

and

(5.49) bp(τ)=12πiCOCIh0(u,τ)Bp(u)du;b_{p}(\tau)=\frac{1}{2\pi i}\int_{C_{O}\bigcup C_{I}}h_{0}(u,\tau)B_{p}(u)\mathrm{d}u;

see also [12]. Define q:=min{a,b}q:=\min\{a,b\}; since in this case both nn and xx\rightarrow\infty, and b2/a0b^{2}/a\rightarrow 0, we have qNqN\rightarrow\infty, m/q=O(b/a)0m/q=O(b/\sqrt{a})\rightarrow 0 if aba\leq b, and m/qarctana/(1a)m/q\sim\arctan\sqrt{a/(1-a)} if a>ba>b.

Using (5.47), (5.41) and (5.45), it is easily verified that

(5.50) |Ap(u)|Cmbp+1,|Bp(u)|Cbp\left|A_{p}(u)\right|\leq\frac{Cm}{b^{p+1}},\qquad\left|B_{p}(u)\right|\leq\frac{C}{b^{p}}

for uu on the inner boundary CIC_{I}, and

(5.51) |Ap(u)|Cmap+1,|Bp(u)|Cm2ap+2\left|A_{p}(u)\right|\leq\frac{Cm}{a^{p+1}},\qquad\left|B_{p}(u)\right|\leq\frac{Cm^{2}}{a^{p+2}}

for uu on the outer boundary COC_{O}, where CC is used again as a generic symbol for constants independent of uu, aa and bb. Also we have the estimates

(5.52) CI|ds|CmbandCO|ds|Cam.\int_{C_{I}}\left|\mathrm{d}s\right|\leq\frac{Cm}{b}\quad\text{and}\quad\int_{C_{O}}\left|\mathrm{d}s\right|\leq\frac{Ca}{m}.

To see this, we take one part of CIC_{I}, namely, ABAB as an example; see (5.35) and (5.36), where ZZ in (5.35) satisfies Im Z=θbπ\text{Im }Z=\theta b\pi. Using the second equality in (5.36), and in polar coordinates, u=reiβu=re^{i\beta}, the last equation is equivalent to

(5.53) (r2+1)rsinβ=θbπm.\frac{(r^{2}+1)}{r}\sin\beta=\frac{\theta b\pi}{m}.

Thus

(5.54) dr=(1+r2)2m(1r2)θbπcosβdβ.\mathrm{d}r=\frac{(1+r^{2})^{2}m}{(1-r^{2})\theta b\pi}\cos\beta\mathrm{d}\beta.

Since in this case r<1r<1, we have |dr|Cmb|dβ|\left|\mathrm{d}r\right|\leq\frac{Cm}{b}\left|\mathrm{d}\beta\right|. By (5.41), |r|C2m/b|r|\leq C_{2}m/b and

(5.55) |ds|=(dr)2+r2(dβ)2Cmb|dβ|,\left|\mathrm{d}s\right|=\sqrt{(\mathrm{d}r)^{2}+r^{2}(\mathrm{d}\beta)^{2}}\leq\frac{Cm}{b}\left|\mathrm{d}\beta\right|,

from which the first inequality in (5.52) follows. The second inequality can be established in a similar manner.

Since h0(u,0)0h_{0}(u,0)\neq 0 and h0(u,τ)h_{0}(u,\tau) is analytic in τ\tau, we have |h0(u,τ)|C|h0(u,0)||h_{0}(u,\tau)|\leq C|h_{0}(u,0)| for some constant C>0C>0 and for τ\tau in the neighbourhood of the origin. By a combination of (5.48), (5.49), (5.50), (5.51), (5.52) and (5.46), we obtain

(5.56) |ap(τ)|Cqpand|bp(τ)|Cqp|a_{p}(\tau)|\leq\frac{C}{q^{p}}\quad\text{and}\quad|b_{p}(\tau)|\leq\frac{C}{q^{p}}

for τ\tau near τ=0\tau=0. Furthermore, by (5.24), (5.31), (5.32) and (5.56), we have

(5.57) |cp|Cqpand|dp|Cqp.\left|c_{p}\right|\leq\frac{C}{q^{p}}\quad\text{and}\quad\left|d_{p}\right|\leq\frac{C}{q^{p}}.

To estimate the remainder in (5.33), we split the loop contour into two parts; see Figure 8. The bounded part of the contour, denoted by Γ1\Gamma_{1}, is contained in a subdomain D~u\widetilde{D}_{u} of DuD_{u}, which has a distance c1a/mc_{1}a/m from the outer boundary of DuD_{u} (i.e., COC_{O}), and has a distance c2m/bc_{2}m/b from the inner boundary of DuD_{u} (i.e., CIC_{I}), c1c_{1} and c2c_{2} being two constants independent of uu, aa and bb. Note that c1a/mc_{1}a/m may become large, whereas c2m/bc_{2}m/b may approach zero. The unbounded part of the contour, denoted by Γ2\Gamma_{2}, is the rest of the loop outside the subdomain D~u\widetilde{D}_{u}. Put

(5.58) εp,1+\displaystyle\varepsilon_{p,1}^{+} =\displaystyle= Nn+1p2Γ(n+1)πi0+Γ1hp(u,τ)ueNm(u1u)τneNτdudτ,\displaystyle\frac{N^{n+1-p}}{2\Gamma(n+1)\pi i}\int_{0}^{+\infty}\int_{\Gamma_{1}}\frac{h_{p}(u,\tau)}{u}e^{Nm\left(u-\frac{1}{u}\right)}\tau^{n}e^{-N\tau}\mathrm{d}u\mathrm{d}\tau,
(5.59) εp,2+\displaystyle\varepsilon_{p,2}^{+} =\displaystyle= Nn+1p2Γ(n+1)πi0+Γ2hp(u,τ)ueNm(u1u)τneNτdudτ.\displaystyle\frac{N^{n+1-p}}{2\Gamma(n+1)\pi i}\int_{0}^{+\infty}\int_{\Gamma_{2}}\frac{h_{p}(u,\tau)}{u}e^{Nm\left(u-\frac{1}{u}\right)}\tau^{n}e^{-N\tau}\mathrm{d}u\mathrm{d}\tau.
Refer to caption
Figure 8: The contour in uu-plane.

As in [5] and [12], it can be shown that εp,2+\varepsilon_{p,2}^{+} is exponentially small in comparison with εp,1+\varepsilon_{p,1}^{+}. To estimate εp,1+\varepsilon_{p,1}^{+}, we also follow [5] and [12] by first deriving an integral representation for hk(u,τ)h_{k}(u,\tau). Define recursively

(5.60) Q0(u,w)=1uw,Qp+1(u,w)=1m(1+1u2)1{1u+ddu}Qp(u,w)\begin{split}Q_{0}(u,w)&=\frac{1}{u-w},\\ Q_{p+1}(u,w)&=\frac{1}{m}\left(1+\frac{1}{u^{2}}\right)^{-1}\left\{\frac{1}{u}+\frac{\mathrm{d}}{\mathrm{d}u}\right\}Q_{p}(u,w)\end{split}

for p=0,1,2,p=0,1,2,\cdots. Using induction, one can easily verify that

(5.61) Qp(u,w)=1(2m)pi=0p1j=0piqp,i,j(11+u2)p+iu2p+ij(1uw)p+1ij,Q_{p}(u,w)=\frac{1}{(2m)^{p}}\sum_{i=0}^{p-1}\sum_{j=0}^{p-i}q_{p,i,j}\left(\frac{1}{1+u^{2}}\right)^{p+i}u^{2p+i-j}\left(\frac{1}{u-w}\right)^{p+1-i-j},

where qp,i,jq_{p,i,j} are some constants depending only on ii, jj and pp. Similar to (4.36), for wD~uw\in\widetilde{D}_{u} we have from (5.22), (5.23) and (5.60) the integral representation

(5.62) hp(w,τ)=12πiCICOh0(u,τ)Qp(u,w)du.h_{p}(w,\tau)=\frac{1}{2\pi i}\int_{C_{I}\cup C_{O}}h_{0}(u,\tau)Q_{p}(u,w)\mathrm{d}u.

From (5.61), it is easy to see that

(5.63) |Qp(u,w)|Cmap+1 for uCO and wD~u,|Qp(u,w)|C1mbp1 for uCI and wD~u.\begin{split}|Q_{p}(u,w)|&\leq C\frac{m}{a^{p+1}}\quad\text{ for }u\in C_{O}\text{ and }w\in\widetilde{D}_{u},\\ |Q_{p}(u,w)|&\leq C\frac{1}{mb^{p-1}}\quad\text{ for }u\in C_{I}\text{ and }w\in\widetilde{D}_{u}.\end{split}

Coupling the above inequalities with (5.52) and (5.62), we obtain

(5.64) |hp(w,τ)|Cqp|h_{p}(w,\tau)|\leq\frac{C}{q^{p}}

for wD~uw\in\tilde{D}_{u} and τ\tau in the neighbourhood of τ=0\tau=0.

If NmNm is bounded, one can easily show from (5.58) and (5.64) that

(5.65) |εp,1+|=O(1qpNp).|\varepsilon_{p,1}^{+}|=O\left(\frac{1}{q^{p}N^{p}}\right).

If NmNm\rightarrow\infty we divide the integration path Γ1\Gamma_{1} into three pieces: the steepest descent path L1L_{1} through u=u+=+iu=u_{+}=+i in the upper half of the uu-plane, the steepest descent path L2L_{2} through u=u=iu=u_{-}=-i in the lower half of the uu-plane, and the circular arc |u|=1δ0|u|=1-\delta_{0}, denoted by L3L_{3}, which joins the steepest descent paths L1L_{1} and L2L_{2} at u=uδ+u=u_{\delta}^{+} and u=uδu=u_{\delta}^{-}, respectively; see Figure 8. Hence,

(5.66) εp,1+=I1+I2+I3,\varepsilon_{p,1}^{+}=I_{1}+I_{2}+I_{3},

where IiI_{i}, i=1,2,3i=1,2,3, denotes the integral over the subcontour LiL_{i}. Applying the steepest descent method [11, p.84], and using (5.64), it can be easily verified that

(5.67) |I1,2|=O(1qpNp1Nm).|I_{1,2}|=O\left(\frac{1}{q^{p}N^{p}}\sqrt{\frac{1}{Nm}}\right).

For I3I_{3}, since

Re(u1u)=(2δ)δ(1δ)2Re u(2δ)δ(1δ)2Re uδ±=Re(u1u)|u=uδ±<Re(u1u)|u=u±,\begin{split}\text{Re}\left(u-\frac{1}{u}\right)&=-\frac{(2-\delta)\delta}{(1-\delta)^{2}}\text{Re }u\leq-\frac{(2-\delta)\delta}{(1-\delta)^{2}}\text{Re }u_{\delta}^{\pm}\\ &=\left.\text{Re}\left(u-\frac{1}{u}\right)\right|_{u=u_{\delta}^{\pm}}<\text{Re}\left.\left(u-\frac{1}{u}\right)\right|_{u=u_{\pm}},\end{split}

I2I_{2} is exponentially small in comparison with I1,2I_{1,2}. Therefore,

(5.68) |εp,1+|=O(1qpNp1Nm).|\varepsilon_{p,1}^{+}|=O\left(\frac{1}{q^{p}N^{p}}\sqrt{\frac{1}{Nm}}\right).

It is well-known that the Bessel function Jα(t)J_{\alpha}(t) is bounded when tt is bounded, and that as tt\rightarrow\infty,

(5.69) Jα(t)2πtcos(t12απ14π).J_{\alpha}(t)\sim\sqrt{\frac{2}{\pi t}}\cos\left(t-\frac{1}{2}\alpha\pi-\frac{1}{4}\pi\right).

Coupling the estimates (5.65) and (5.68), and together with (5.69) and (5.57), we can find a constant MM independent of aa and bb such that

(5.70) |εp+|MqN(|J0(2Nm)||cp1|Np1+|J1(2Nm)||dp1|Np1),|\varepsilon_{p}^{+}|\leq\frac{M}{qN}\left(|J_{0}(2Nm)|\frac{|c_{p-1}|}{N^{p-1}}+|J_{1}(2Nm)|\frac{|d_{p-1}|}{N^{p-1}}\right),

which establishes the asymptotic nature of the expansion in (5.30), given that qNqN\rightarrow\infty. In conclusion, in the case n2/xN0n^{2}/xN\rightarrow 0 and xx\rightarrow\infty, (5.30) is an asymptotic expansion.

6 REMAINING CASES

We are now left with only two easy cases to consider.

6.1 a>0a>0, a/b2a/b^{2} large and xx bounded

In this case, the series in (1.2) is itself an asymptotic expansion, since

φk(N):=(n)k(x)k(n+1)k(N+1)kk!k!,k=0,1,,n,\varphi_{k}(N):=\frac{(-n)_{k}(-x)_{k}(n+1)_{k}}{(-N+1)_{k}k!k!},\qquad\qquad k=0,1,\cdots,n,

is an asymptotic sequence when xN/n2xN/n^{2}\rightarrow\infty; see Table 1 and also [11, p.10].

6.2 a<0a<0

The case a<0a<0 is relatively easy, compared with the case a>0a>0. Let us start with the integral representation (2.23). For fixed wγ2w\in\gamma_{2}, the phase function f~(t,w)\widetilde{f}(t,w) in (2.23) has a saddle point

(6.1) t0(w)=2w1+1+4b2(w1)w2(1+b)w;t_{0}(w)=\frac{2w-1+\sqrt{1+4b^{2}(w-1)w}}{2(1+b)w};

cf.(2.5). Note that the only difference between f(t,w)f(t,w) in (2.2) and f~(t,w)\widetilde{f}(t,w) in (2.24) is the choice of branch cuts in the ww-plane due to the term alnwaln(w1)a\ln w-a\ln(w-1) in f(t,w)f(t,w) and the term aln(w)aln(1w)a\ln(-w)-a\ln(1-w) in f~(t,w)\widetilde{f}(t,w). Similar to (3.1), for any fixed wγ2w\in\gamma_{2}, we have t0(w)=1b+O(b2)t_{0}(w)=1-b+O(b^{2}) as b0b\rightarrow 0; in particular, t0(0)=1bt_{0}(0)=1-b. By the same reasoning given for (3.2), we introduce the mapping tτt\rightarrow\tau defined by

(6.2) blnττ+A=f~(t,w)f~(t0(w),w)b\ln\tau-\tau+A=\widetilde{f}(t,w)-\widetilde{f}(t_{0}(w),w)

with

(6.3) τ(t0(w))=b.\tau(t_{0}(w))=b.

Coupling (6.3) and (6.2) yields A=bblnbA=b-b\ln b. From (6.2), we have

(6.4) dtdτ=b/τ1f~(t,w)/t=(τb)t(1t)[1(1t)w]τ(1+b)w(tt0+(w))(tt0(w)),τb,dtdτ={t0+(w)(1t0+(w))[1(1t0+(w))w]b1+4b2(w1)w}1/2,τ=b,\begin{split}\frac{\mathrm{d}t}{\mathrm{d}\tau}&=\frac{b/\tau-1}{\partial\widetilde{f}(t,w)/\partial t}=\frac{(\tau-b)t(1-t)[1-(1-t)w]}{\tau(1+b)w(t-t^{+}_{0}(w))(t-t^{-}_{0}(w))},\quad\quad\tau\neq b,\\ \frac{\mathrm{d}t}{\mathrm{d}\tau}&=\left\{\frac{t^{+}_{0}(w)(1-t^{+}_{0}(w))[1-(1-t^{+}_{0}(w))w]}{b\sqrt{1+4b^{2}(w-1)w}}\right\}^{1/2},\hskip 45.52458pt\tau=b,\end{split}

where we have used L’Ho^\hat{o}spital’s rule for τ=b\tau=b and

(6.5) t0±(w)=2w1±1+4b2(w1)w2(1+b)w;t_{0}^{\pm}(w)=\frac{2w-1\pm\sqrt{1+4b^{2}(w-1)w}}{2(1+b)w};

cf.(3.5). Furthermore, it follows from (6.2) and (2.23) that

(6.6) tn(x,N+1)=(1)n2πiΓ(n+N+2)enbnΓ(n+1)Γ(Nn+1)×0+γ21w1eNf~(t0(w),w)eN(blnττ)dtdτdwdτ;\begin{split}t_{n}(x,N+1)=&\frac{(-1)^{n}}{2\pi i}\frac{\Gamma(n+N+2)e^{n}b^{-n}}{\Gamma(n+1)\Gamma(N-n+1)}\\ &\times\int_{0}^{+\infty}\int_{\gamma_{2}}\frac{1}{w-1}e^{N\widetilde{f}({t_{0}(w)},w)}e^{N(b\ln\tau-\tau)}\frac{\mathrm{d}t}{\mathrm{d}\tau}\mathrm{d}w\mathrm{d}\tau;\end{split}

cf.(3.6). Solving the equation f(t0(w),w)/w=0\partial f(t_{0}(w),w)/\partial w=0, we obtain the saddle points

(6.7) w±=b±b24a+4a22b=12±1214a(1a)b2;\begin{split}w_{\pm}=&\frac{b\pm\sqrt{b^{2}-4a+4a^{2}}}{2b}\\ =&\frac{1}{2}\pm\frac{1}{2}\sqrt{1-\frac{4a(1-a)}{b^{2}}};\end{split}

cf.(2.9). The relevant saddle point on the integral path γ2\gamma_{2} is the negative saddle point ww_{-}.

The following procedure is the same as that given in [8]. Recall the Hankel integral for the Gamma function

(6.8) 1Γ(z)=12πi(0+)euuzdu,\frac{1}{\Gamma(z)}=\frac{1}{2\pi i}\int^{(0+)}_{-\infty}e^{u}u^{-z}\mathrm{d}u,

where the contour is a loop starting at -\infty, encircling the origin in the counterclockwise direction and returning to -\infty. With z=aN+1z=-aN+1 and uu replaced by u-u, we obtain

(6.9) 1Γ(aN+1)=NaN2πi(0+)1ueNψ~(u)du,\frac{1}{\Gamma(-aN+1)}=\frac{N^{aN}}{2\pi i}\int^{(0+)}_{\infty}\frac{1}{u}e^{N\widetilde{\psi}(u)}\mathrm{d}u,

where ψ~(u)=aln(u)u\widetilde{\psi}(u)=a\ln(-u)-u. Make the transformation wuw\rightarrow u defined by

(6.10) f~(t0(w),w)=aln(u)u+γ\begin{split}\widetilde{f}(t_{0}(w),w)&=a\ln(-u)-u+\gamma\end{split}

with

(6.11) u(w)=a,u(w_{-})=a,

where γ\gamma is a constant to be determined. We have from (6.10)

(6.12) dwdu=(au)w(w1)[(12a)+4b2w24b2w+1]2ub2(ww+)(ww),ua,dwdu={(1a)(12a)b3b24a+4a2}1/2,u=a,\begin{split}\frac{\mathrm{d}w}{\mathrm{d}u}&=\frac{(a-u)w(w-1)\left[(1-2a)+\sqrt{4b^{2}w^{2}-4b^{2}w+1}\right]}{2ub^{2}(w-w_{+})(w-w_{-})},\quad\quad u\neq a,\\ \frac{\mathrm{d}w}{\mathrm{d}u}&=\left\{\frac{(1-a)(1-2a)}{b^{3}\sqrt{b^{2}-4a+4a^{2}}}\right\}^{1/2},\qquad u=a,\end{split}

where we again have used L’Ho^\hat{o}spital’s rule for u=au=a.

By (6.6) and (6.10), we have

(6.13) tn(x,N+1)=(1)n2πiΓ(n+N+2)enbneNγΓ(n+1)Γ(Nn+1)×0+(0+)h(u,τ)ueN(aln(u)u)eN(blnττ)dudτ,\begin{split}t_{n}(x,N+1)=&\frac{(-1)^{n}}{2\pi i}\frac{\Gamma(n+N+2)e^{n}b^{-n}e^{N\gamma}}{\Gamma(n+1)\Gamma(N-n+1)}\\ &\times\int_{0}^{+\infty}\int_{\infty}^{\left(0^{+}\right)}\frac{h(u,\tau)}{u}e^{N\left(a\ln(-u)-u\right)}e^{N(b\ln\tau-\tau)}\mathrm{d}u\mathrm{d}\tau,\end{split}

where

(6.14) h(u,τ)=uw1dtdτdwdu.h(u,\tau)=\frac{u}{w-1}\frac{\mathrm{d}t}{\mathrm{d}\tau}\frac{\mathrm{d}w}{\mathrm{d}u}.

Define recursively

(6.15) hl(u,τ)\displaystyle h_{l}(u,\tau) =\displaystyle= al(τ)+(ua)gl(u,τ),\displaystyle a_{l}(\tau)+(u-a)g_{l}(u,\tau),
(6.16) hl+1(u,τ)\displaystyle h_{l+1}(u,\tau) =\displaystyle= ugl(u,τ)u,\displaystyle u\frac{\partial g_{l}(u,\tau)}{\partial u},

where h0(u,τ)=h(u,τ)h_{0}(u,\tau)=h(u,\tau), and expand al(τ)a_{l}(\tau) into a Maclaurin series

(6.17) al(τ)=j=0al,jτj.a_{l}(\tau)=\sum_{j=0}^{\infty}a_{l,j}\tau^{j}.

By an integration-by-parts procedure, we obtain

(6.18) tn(x,N+1)=(1)nΓ(n+N+2)enbneNγΓ(Nn+1)Γ(aN+1)NaN+n+1[l=0p1clNl+εp],t_{n}(x,N+1)=\frac{(-1)^{n}\Gamma(n+N+2)e^{n}b^{-n}e^{N\gamma}}{\Gamma(N-n+1)\Gamma(-aN+1)N^{aN+n+1}}\Biggr{[}\sum_{l=0}^{p-1}\frac{c_{l}}{N^{l}}+\varepsilon_{p}^{-}\Biggr{]},

where

(6.19) cl=Nn+1Γ(n+1)0al(τ)τneNτdτm=0al,mΓ(n+m+1)Γ(n+1)Nmc_{l}=\frac{N^{n+1}}{\Gamma(n+1)}\int_{0}^{\infty}a_{l}(\tau)\tau^{n}e^{-N\tau}\mathrm{d}\tau\sim\sum_{m=0}^{\infty}a_{l,m}\frac{\Gamma(n+m+1)}{\Gamma(n+1)N^{m}}

and

(6.20) εp=NaN+n+1pΓ(aN+1)2Γ(n+1)πi0+(0+)hp(u,τ)ueNψ~(u)τneNτdudτ.\varepsilon_{p}^{-}=\frac{N^{aN+n+1-p}\Gamma(-aN+1)}{2\Gamma(n+1)\pi i}\int_{0}^{+\infty}\int_{-\infty}^{\left(0^{+}\right)}\frac{h_{p}(u,\tau)}{u}e^{N\widetilde{\psi}(u)}\tau^{n}e^{-N\tau}\mathrm{d}u\mathrm{d}\tau.

Estimation of the error term given in (6.20) is exactly the same as what we have done in [8].

References

  • [1] R. Beals and R. Wong, Special Functions, A Graduate Text, Cambridge University Press, Cambridge, 2010.
  • [2] C. Chester, B. Friedman and F. Ursell, An extension of the method of steepest descents, Proc. Cambridge Philos. Soc. 53 (1957), 599-611.
  • [3] X.-S. Jin and R. Wong, Uniform asymptotic expansions for Meixner polynomials, Constr. Approx. 14 (1998), 113-150.
  • [4] Y. Lin and R. Wong, Global Asymptotics of the Hahn Polynomials, Analysis and Applications, to appear.
  • [5] A. B. Olde Daalhuis and N. M. Temme, Uniform Airy-type expansions of integrals, SIAM J. Math. Anal. 25 (1994), 304-321.
  • [6] F. W. J. Olver, Asymptotics and Special Functions, Academic Press, New York, 1974. (Reprinted by A. K. Peters, Wellesley, MA, 1997.)
  • [7] F. W. J. Olver, D. W. Lozier, C. W. Clark and R. F. Boisvert, NIST Handbook of Mathematical Functions, Cambridge University Press, Cambridge, 2010.
  • [8] J. H. Pan and R. Wong, Uniform Asymptotic Expansions for the Discrete Chebyshev Polynomials, Studies in Applied Mathematics 128 (2012), 337-384.
  • [9] Bo, Rui and R. Wong, Uniform asymptotic expansion of Charlier polynomials. Methods and Applications of Analysis 1 (1994), 294-313.
  • [10] I. I. Sharapudinov, Asymptotic Properties of Orthogonal Hahn Polynomials in a Discrete Variable, Sbornik: Mathematics, Vol. 68, 1 (1991), 111-132.
  • [11] R. Wong, Asymptotic Approximations of Integrals, Academic Press, Boston, 1989. (Reprinted by SIAM, Philadelphia, PA, 2001)
  • [12] R. Wong and Y. Q. Zhao, Estimates for the error term in a uniform asymptotic expansion of the Jacobi polynomials. Anal. Appl. (Singap.) 1 (2003), no. 2, 213-241.
  • [13] R. Wong, Lecture Notes on Applied Analysis, World Scientific, Singapore, 2010.