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

Gaussian unitary ensemble with jump discontinuities and the coupled Painlevé II and IV systems

Xiao-Bo Wu111School of Mathematics and Computer Science, Shangrao Normal University, Shangrao 334001, China. E-mail: [email protected]  and Shuai-Xia Xu222Institut Franco-Chinois de l’Energie Nucléaire, Sun Yat-sen University, Guangzhou 510275, China. E-mail: [email protected]

 

Abstract We study the orthogonal polynomials and the Hankel determinants associated with Gaussian weight with two jump discontinuities. When the degree nn is finite, the orthogonal polynomials and the Hankel determinants are shown to be connected to the coupled Painlevé IV system. In the double scaling limit as the jump discontinuities tend to the edge of the spectrum and the degree nn grows to infinity, we establish the asymptotic expansions for the Hankel determinants and the orthogonal polynomials, which are expressed in terms of solutions of the coupled Painlevé II system. As applications, we re-derive the recently found Tracy-Widom type expressions for the gap probability of there being no eigenvalues in a finite interval near the the extreme eigenvalue of large Gaussian unitary ensemble and the limiting conditional distribution of the largest eigenvalue in Gaussian unitary ensemble by considering a thinned process.


2010 Mathematics Subject Classification: 33E17; 34M55; 41A60

Keywords and phrases: Random matrices; Gaussian unitary ensemble; Tracy-Widom distribution; Painlevé equations; Hankel determinants; orthogonal polynomials; Riemann-Hilbert approach.

 

1 Introduction and statement of results

Consider the Gaussian Unitary Ensemble (GUE), where the joint probability density function of the eigenvalues is given by

ρn(λ1,,λn)=1Zni=1neλi21i<jnn|λiλj|2;\rho_{n}(\lambda_{1},...,\lambda_{n})=\frac{1}{Z_{n}}\prod_{i=1}^{n}e^{-\lambda_{i}^{2}}\prod_{1\leq i<j\leq n}^{n}|\lambda_{i}-\lambda_{j}|^{2}; (1.1)

see [15, 23]. Here ZnZ_{n}, known as the partition function, is a normalization constant. It is well known that the density function (1.1) can be expressed in the determinantal form

ρn(λ1,,λn)=det[Kn(λi,λj)]1i,jn,\rho_{n}(\lambda_{1},...,\lambda_{n})=\det[K_{n}(\lambda_{i},\lambda_{j})]_{1\leq i,j\leq n}, (1.2)

where

Kn(x,y)=e12(x2+y2)k=0n1Hk(x)Hk(y).K_{n}(x,y)=e^{-\frac{1}{2}(x^{2}+y^{2})}\sum_{k=0}^{n-1}H_{k}(x)H_{k}(y). (1.3)

The polynomial Hk(x)H_{k}(x) therein is the kk-th degree monic orthogonal polynomial with respect to the Gaussian weight ex2e^{-x^{2}}, which is the Hermite polynomial except for a constant [24, Eq. (18.5.13)].

Introduce the Hankel determinant

Dn(s1,s2;ω1,ω2)\displaystyle D_{n}(s_{1},s_{2};\omega_{1},\omega_{2}) =det(𝚁xj+kw(x;s1,s2;ω1,ω2)𝑑x)j,k=0n1\displaystyle=\det\left(\int_{\mathtt{R}}x^{j+k}w(x;s_{1},s_{2};\omega_{1},\omega_{2})dx\right)_{j,k=0}^{n-1}
=1n!i=1nw(xj;s1,s2;ω1,ω2)1i<jnn(xixj)2dx1dxn,\displaystyle=\frac{1}{n!}\int_{-\infty}^{\infty}\cdots\int_{-\infty}^{\infty}\prod_{i=1}^{n}w(x_{j};s_{1},s_{2};\omega_{1},\omega_{2})\prod_{1\leq i<j\leq n}^{n}(x_{i}-x_{j})^{2}dx_{1}\cdots dx_{n}, (1.4)

where

w(x;s1,s2;ω1,ω2)=ex2{1x<s1,ω1s1<x<s2,ω2x>s2,w(x;s_{1},s_{2};\omega_{1},\omega_{2})=e^{-x^{2}}\left\{\begin{array}[]{cc}1&x<s_{1},\\ \omega_{1}&s_{1}<x<s_{2},\\ \omega_{2}&x>s_{2},\end{array}\right. (1.5)

with the constant ωk0,k=1,2\omega_{k}\geqslant 0,k=1,2. If ω1=ω2=1\omega_{1}=\omega_{2}=1, the Hankel determinant Dn(s1,s2;1,1)D_{n}(s_{1},s_{2};1,1) is corresponding to the pure Gaussian weight ex2e^{-x^{2}} and can be evaluated explicitly

DnGUE=Dn(s1,s2;1,1)=(2π)n/22n2/2k=1n1k!;D_{n}^{\texttt{GUE}}=D_{n}(s_{1},s_{2};1,1)=(2\pi)^{n/2}2^{-n^{2}/2}\prod_{k=1}^{n-1}k!~{}; (1.6)

see [23, Equation (4.1.5)]. There is a system of monic orthogonal polynomials πn(x)=πn(x;s1,s2)=xn+\pi_{n}(x)=\pi_{n}(x;s_{1},s_{2})=x^{n}+\cdots, orthogonal with respect to the weight function w(x)=w(x;s1,s2;ω1,ω2)w(x)=w(x;s_{1},s_{2};\omega_{1},\omega_{2}),

πm(x)πn(x)w(x)𝑑x=γn2δm,n,m,n.\int_{\mathbb{R}}\pi_{m}(x)\pi_{n}(x)w(x)dx=\gamma_{n}^{-2}\delta_{m,n},\quad m,n\in\mathbb{N}. (1.7)

The orthogonal polynomials satisfy the three term recurrence relation

zπn(z)=πn+1(z)+αnπn(z)+βn2πn1(z),z\pi_{n}(z)=\pi_{n+1}(z)+\alpha_{n}\pi_{n}(z)+\beta_{n}^{2}\pi_{n-1}(z), (1.8)

where αn=αn(s1,s2)\alpha_{n}=\alpha_{n}(s_{1},s_{2}) and βn=βn(s1,s2)\beta_{n}=\beta_{n}(s_{1},s_{2}) are the recurrence coefficients. The polynomials γnπn(x)\gamma_{n}\pi_{n}(x) are the normalized orthogonal polynomials and the leading coefficients γn=γn(s1,s2)\gamma_{n}=\gamma_{n}(s_{1},s_{2}) are connected to the Hankel determinant Dn(s1,s2)=Dn(s1,s2;ω1,ω2)D_{n}(s_{1},s_{2})=D_{n}(s_{1},s_{2};\omega_{1},\omega_{2}) by

Dn(s1,s2)=j=0n1γj2(s1,s2).D_{n}(s_{1},s_{2})=\prod_{j=0}^{n-1}\gamma_{j}^{-2}(s_{1},s_{2}). (1.9)

Consider the gap probability that there is no eigenvalues in the finite interval (s1,s2)(s_{1},s_{2}) for the GUE matrices. On account of (1.1) and (1), the gap probability can be expressed as a ratio of the Hankel determinants

Pro(λj(s1,s2):j=1n)=Dn(s1,s2;0,1)DnGUE,\texttt{Pro}(\lambda_{j}\not\in(s_{1},s_{2}):j=1...n)=\frac{D_{n}(s_{1},s_{2};0,1)}{D_{n}^{\texttt{GUE}}}, (1.10)

where λ1<,,λn\lambda_{1}<,...,\lambda_{n} are the eigenvalues of a matrix in GUE and Dn(s1,s2;ω1,ω2)D_{n}(s_{1},s_{2};\omega_{1},\omega_{2}) is defined in (1). For the gap probability on the infinite interval (s,+)(s,+\infty), we have the distribution of the largest eigenvalue

Pro(λn<s)=Dn(s,s;0,0)DnGUE.\texttt{Pro}(\lambda_{n}<s)=\frac{D_{n}(s,s;0,0)}{D_{n}^{\texttt{GUE}}}. (1.11)

In the large nn limit, the distribution of the largest eigenvalue converges to the celebrated Tracy-Widom distribution

limn+Pro(λn<2n+s2n1/6)=exp(s+(xs)qHM2(x)𝑑x),\lim_{n\to+\infty}\texttt{Pro}(\lambda_{n}<\sqrt{2n}+\frac{s}{\sqrt{2}n^{1/6}})=\exp\left(-\int_{s}^{+\infty}(x-s)q_{\texttt{HM}}^{2}(x)dx\right), (1.12)

where qHM(s)q_{\texttt{HM}}(s) is the Hastings-Mcleod solution of the second Painlevé equation q′′(x)2q3(x)xq(x)=0q^{\prime\prime}(x)-2q^{3}(x)-xq(x)=0 with the asymptotic behavior qHM(x)Ai(x)q_{\texttt{HM}}(x)\sim{\rm Ai}(x) as x+x\to+\infty; see [25].

We proceed to consider the thinned process in GUE by removing each eigenvalues λ1<,,λn\lambda_{1}<,...,\lambda_{n} of the GUE independently with probability p(0,1)p\in(0,1); see [2, 3]. It is observed in [3] that the remaining and removed eigenvalues can be interpreted as observed and unobserved particles, respectively. If we know the information that the largest observed particle λmaxT\lambda^{T}_{\max} is less that yy, then the conditional distribution of the largest eigenvalue λn\lambda_{n} of the original GUE can be expressed by the ratio of Hankel determiants

Pro(λn<x|λmaxT<y)=Dn(y,x;p,0)Dn(y,x;p,p),x>y,\texttt{Pro}(\lambda_{n}<x|\lambda^{T}_{\max}<y)=\frac{D_{n}(y,x;p,0)}{D_{n}(y,x;p,p)},\quad x>y, (1.13)

and

Pro(λn<x|λmaxT<y)=Dn(x,y;0,0)Dn(x,y;p,p),x<y,\texttt{Pro}(\lambda_{n}<x|\lambda^{T}_{\max}<y)=\frac{D_{n}(x,y;0,0)}{D_{n}(x,y;p,p)},\quad x<y, (1.14)

with Dn(s1,s2;ω1,ω2)D_{n}(s_{1},s_{2};\omega_{1},\omega_{2}) defined in (1); see [5, 8] and [1, 7]. It is noted that other thinned random matrices in the situation of circular ensemble is considered in [5] and also in [4] with applications in the studies of Riemann zeros.

Recently, the limits of (1.10) and (1.13) are studied in [6, 8] by considering the Fredholm determinants of the Airy kernel with several discontinuities. More generally, the limits of the gap probabilities on any finite union of intervals near the extreme eigenvalues are considered in [8]. In [27], the second author of the present paper and Dai derive the asymptotics of (1.13) via the Fredholm determinants of the Painlevé XXXIV kernel which is a generalization of the Airy kernel. In both [8] and [27], Tracy-Widom type expressions for the limiting distributions are established by using solutions to the coupled Painlevé II system. The Hankel determinants and orthogonal polynomials associated with the Gaussian weight with one jump discontinuity have also been considered in [1, 16, 17, 22, 26, 28] with applications in random matrices.

The present work is devoted to the studies of the Hankel determinants and the orthogonal polynomials associated with the Gaussian weight with two jump discontinuities both as the degree nn is finite and as nn tends to infinity. When the degree nn is finite, we show that the Hankel determinants and the orthogonal polynomials are described by the coupled Painlevé IV system. As the jump discontinuities tend to the largest eigenvalue of GUE and the degree nn grows to infinity, we establish asymptotic expansions for the Hankel determinants and the orthogonal polynomials. The asymptotics are expressed in terms of solutions to the coupled Painlevé II system. As applications, our results reproduce the asymptotic expansions of the gap probability in a finite interval near the largest eigenvalue of GUE and the conditional distribution of largest eigenvalue of GUE as defined in (1.10) and (1.13), respectively, which are obtained previously in [8, 27].

1.1 Statement of results

The coupled Painlevé IV system

We introduce the Hamiltonian

HIV(a1,a2,b1,b2;x;s)=2(a1b1+a2b2+n)(a1+a2)+2(a1b1(xs)+a2b2(x+s)+nx)(a1b12+a2b22),H_{\texttt{IV}}(a_{1},a_{2},b_{1},b_{2};x;s)=-2(a_{1}b_{1}+a_{2}b_{2}+n)(a_{1}+a_{2})+2(a_{1}b_{1}(x-s)+a_{2}b_{2}(x+s)+nx)-(a_{1}b_{1}^{2}+a_{2}b_{2}^{2}), (1.15)

which is a special Garnier system in two variables in the studies of the classification of 4-dimensional Painlevé-type equations by Kawakami, Nakamura and Sakai [21, Equations (3.12)-(3.13)]. The coupled Painlevé IV system can be written as the following Hamiltonian system

{da1dx=HIVb1(a1,a2,b1,b2;x,s)=2a1(a1+a2+b1x+s),da2dx=HIVb2(a1,a2,b1,b2;x,s)=2a2(a1+a2+b2xs),db1dx=HIVa1(a1,a2,b1,b2;x,s)=b12+2b1(2a1+a2x+s)+2(a2b2+n),db2dx=HIVa2(a1,a2,b1,b2;x,s)=b22+2b2(a1+2a2xs)+2(a1b1+n).\left\{\begin{array}[]{l}\frac{da_{1}}{dx}=\frac{\partial H_{\texttt{IV}}}{\partial b_{1}}(a_{1},a_{2},b_{1},b_{2};x,s)=-2a_{1}(a_{1}+a_{2}+b_{1}-x+s),\\ \frac{da_{2}}{dx}=\frac{\partial H_{\texttt{IV}}}{\partial b_{2}}(a_{1},a_{2},b_{1},b_{2};x,s)=-2a_{2}(a_{1}+a_{2}+b_{2}-x-s),\\ \frac{db_{1}}{dx}=-\frac{\partial H_{\texttt{IV}}}{\partial a_{1}}(a_{1},a_{2},b_{1},b_{2};x,s)=b_{1}^{2}+2b_{1}(2a_{1}+a_{2}-x+s)+2(a_{2}b_{2}+n),\\ \frac{db_{2}}{dx}=-\frac{\partial H_{\texttt{IV}}}{\partial a_{2}}(a_{1},a_{2},b_{1},b_{2};x,s)=b_{2}^{2}+2b_{2}(a_{1}+2a_{2}-x-s)+2(a_{1}b_{1}+n).\end{array}\right. (1.16)

Eliminating b1b_{1} and b2b_{2} from the system, we find that a1a_{1} and a2a_{2} solve a couple of second order nonlinear differential equations

{𝚍2a1𝚍x212a1(𝚍a1𝚍x)26a1(a1+a2)2+8a1(a1+a2)x8a12s+2(2n1)a12a1(xs)2=0,𝚍2a2𝚍x212a2(𝚍a2𝚍x)26a2(a1+a2)2+8a2(a1+a2)x+8a22s+2(2n1)a22a2(x+s)2=0.\left\{\begin{array}[]{l}\frac{\mathtt{d}^{2}a_{1}}{\mathtt{d}x^{2}}-\frac{1}{2a_{1}}\left(\frac{\mathtt{d}a_{1}}{\mathtt{d}x}\right)^{2}-6a_{1}(a_{1}+a_{2})^{2}+8a_{1}(a_{1}+a_{2})x-8a_{1}^{2}s+2(2n-1)a_{1}-2a_{1}(x-s)^{2}=0,\\ \frac{\mathtt{d}^{2}a_{2}}{\mathtt{d}x^{2}}-\frac{1}{2a_{2}}\left(\frac{\mathtt{d}a_{2}}{\mathtt{d}x}\right)^{2}-6a_{2}(a_{1}+a_{2})^{2}+8a_{2}(a_{1}+a_{2})x+8a_{2}^{2}s+2(2n-1)a_{2}-2a_{2}(x+s)^{2}=0.\end{array}\right. (1.17)

If a2=0a_{2}=0, we recover from the above equations the classical Painlevé IV equation (see [13] and [24, Equation (32.2.4)] )

yIV′′=12yIVyIV2+32yIV3+4xyIV2+2(x2+12n)yIV,yIV(x)=2a1(x+s;s).y_{\texttt{IV}}^{\prime\prime}=\frac{1}{2y_{\texttt{IV}}}y_{\texttt{IV}}^{\prime 2}+\frac{3}{2}y_{\texttt{IV}}^{3}+4xy_{\texttt{IV}}^{2}+2(x^{2}+1-2n)y_{\texttt{IV}},\quad y_{\texttt{IV}}(x)=-2a_{1}(x+s;s). (1.18)

Orthogonal polynomials of finite degree: the coupled Painlevé IV system

Our first result shows that, when the degree nn is finite, several quantities of the orthogonal polynomials associated with the weight function (1.5) can be expressed in terms of the coupled Painlevé IV system. These quantities include the the Hankel determiniants, the recurrence coefficients, leading coefficients and the values of the orthogonal polynomials at the jump discontinuities of (1.5). We are interested in the Gaussian weight with two jump discontinuities, thus without loss of generality we assume that the parameters in (1.5) satisfy

s1<s2;ω10,ω20,ω1ω2,ω11.s_{1}<s_{2};\quad\omega_{1}\geqslant 0,\quad\omega_{2}\geqslant 0,\quad\omega_{1}\neq\omega_{2},\quad\omega_{1}\neq 1. (1.19)
Theorem 1.

Let sks_{k} and ωk\omega_{k}, k=1,2k=1,2 be as in (1.19) and Dn(s1,s2)=Dn(s1,s2,ω1,ω2)D_{n}(s_{1},s_{2})=D_{n}(s_{1},s_{2},\omega_{1},\omega_{2}) be the Hankel determinant defined in (1), we denote

F(s1,s2)=s1lnDn(s1,s2)+s2lnDn(s1,s2),F(s_{1},s_{2})=\frac{\partial}{\partial s_{1}}\ln D_{n}(s_{1},s_{2})+\frac{\partial}{\partial s_{2}}\ln D_{n}(s_{1},s_{2}), (1.20)

and

x=s1+s22,s=s2s12.x=\frac{s_{1}+s_{2}}{2},\quad s=\frac{s_{2}-s_{1}}{2}. (1.21)

Then F(s1,s2)F(s_{1},s_{2}) is related to the Hamiltonian for the coupled Painlevé IV system by

F(s1,s2)=HIV(x;s)2nx.F(s_{1},s_{2})=H_{\texttt{IV}}(x;s)-2nx. (1.22)

Moreover, let αn(s1,s2)\alpha_{n}(s_{1},s_{2}), βn(s1,s2)\beta_{n}(s_{1},s_{2}) be the recurrence coefficients defined in (1.8), γn(s1,s2)\gamma_{n}(s_{1},s_{2}) be the leading coefficient of the orthonormal polynomial defined in (1.7) and πn(x)=πn(x;s1,s2)\pi_{n}(x)=\pi_{n}(x;s_{1},s_{2}) be the monic orthogonal polynomial defined in (1.7), we have

αn(s1,s2)=a1(x;s)b12(x;s)+a2(x;s)b22(x;s)2(a1(x;s)b1(x;s)+a2(x;s)b2(x;s)+n),\displaystyle\alpha_{n}(s_{1},s_{2})=\frac{a_{1}(x;s)b_{1}^{2}(x;s)+a_{2}(x;s)b_{2}^{2}(x;s)}{2(a_{1}(x;s)b_{1}(x;s)+a_{2}(x;s)b_{2}(x;s)+n)}, (1.23)
βn2(s1,s2)=12(a1(x;s)b1(x;s)+a2(x;s)b2(x;s)+n),\displaystyle\beta^{2}_{n}(s_{1},s_{2})=\frac{1}{2}\left(a_{1}(x;s)b_{1}(x;s)+a_{2}(x;s)b_{2}(x;s)+n\right), (1.24)
γn12=14πiex2y(x;s)0,\displaystyle\gamma_{n-1}^{2}=\frac{1}{4\pi i}e^{x^{2}}y(x;s)\neq 0, (1.25)
ddxlnγn1(s1,s2)=a1(x;s)+a2(x;s),\displaystyle\frac{d}{dx}\ln\gamma_{n-1}(s_{1},s_{2})=a_{1}(x;s)+a_{2}(x;s), (1.26)
πn(s1)2=2πiω11e2sx+s2a1(x;s)b1(x;s)2y(x;s),\displaystyle\pi_{n}(s_{1})^{2}=\frac{2\pi i}{\omega_{1}-1}e^{-2sx+s^{2}}\frac{a_{1}(x;s)b_{1}(x;s)^{2}}{y(x;s)}, (1.27)
πn(s2)2=2πiω2ω1e2sx+s2a2(x;s)b2(x;s)2y(x;s),\displaystyle\pi_{n}(s_{2})^{2}=\frac{2\pi i}{\omega_{2}-\omega_{1}}e^{2sx+s^{2}}\frac{a_{2}(x;s)b_{2}(x;s)^{2}}{y(x;s)}, (1.28)

where ak(s1,s2)a_{k}(s_{1},s_{2}) and bk(s1,s2)b_{k}(s_{1},s_{2}), k=1,2k=1,2, satisfy the coupled Painlevé IV system (1.16) and y(x;s)y(x;s) is connected to ak(x;s)a_{k}(x;s), k=1,2k=1,2, by dydx=2(a1+a2x)y\frac{dy}{dx}=2(a_{1}+a_{2}-x)y.

Remark 1.

In view of (1.25) and (1.28), we have

a2(x;s)b2(x;s)2=O(ω1ω2),asω1ω2,a_{2}(x;s)b_{2}(x;s)^{2}=O(\omega_{1}-\omega_{2}),\quad\mbox{as}\quad\omega_{1}\to\omega_{2}, (1.29)

where the error bound is uniform for xx and ss in any compact subset of \mathbb{R} and {0}\mathbb{R}\setminus\{0\}, respectively. Using πn1(s2)2=8πiω2ω1e2sx+s2a2(x;s)y(x;s)\pi_{n-1}(s_{2})^{2}=\frac{8\pi i}{\omega_{2}-\omega_{1}}e^{2sx+s^{2}}\frac{a_{2}(x;s)}{y(x;s)} (see (2.53)), we obtain that

a2(x;s)=O(ω1ω2),asω1ω2.a_{2}(x;s)=O(\omega_{1}-\omega_{2}),\quad\mbox{as}\quad\omega_{1}\to\omega_{2}. (1.30)

Since a2(x;s)0a_{2}(x;s)\to 0, as ω1ω2\omega_{1}\to\omega_{2}, the function yIV(x)=2a1(x+s;s)y_{\texttt{IV}}(x)=-2a_{1}(x+s;s) solves the classical Painlevé IV equation as shown before in (1.18). Thus, as ω1ω2\omega_{1}\to\omega_{2}, Theorem 1 implies that the Hankel determinants and the orthogonal polynomials associated with the weight function (1.5) with one discontinuity are related to the classical Painlevé IV equation.

The coupled Painlevé II system

To state our main results on the asymptotics of the orthogonal polynomials, we introduce the following coupled Painlevé II system in dimension four

{dw1dx=HIIv1=2(v1+v2+x2)w12,dv1dx=HIIw1=2v1w1,dw2dx=HIIv2=2(v1+v2+x+s2)w22,dv2dx=HIIw2=2v2w2,\left\{\begin{array}[]{l}\frac{dw_{1}}{dx}=-\frac{\partial H_{\texttt{II}}}{\partial v_{1}}=2(v_{1}+v_{2}+\frac{x}{2})-w_{1}^{2},\\ \frac{dv_{1}}{dx}=\frac{\partial H_{\texttt{II}}}{\partial w_{1}}=2v_{1}w_{1},\\ \frac{dw_{2}}{dx}=-\frac{\partial H_{\texttt{II}}}{\partial v_{2}}=2(v_{1}+v_{2}+\frac{x+s}{2})-w_{2}^{2},\\ \frac{dv_{2}}{dx}=\frac{\partial H_{\texttt{II}}}{\partial w_{2}}=2v_{2}w_{2},\\ \end{array}\right. (1.31)

where vk=vk(x;s)v_{k}=v_{k}(x;s), wk=wk(x;s)w_{k}=w_{k}(x;s), k=1,2k=1,2 and the Hamiltonian HII=HII(v1,v2,w1,w2;x;s)H_{\texttt{II}}=H_{\texttt{II}}(v_{1},v_{2},w_{1},w_{2};x;s) is given by

HII(v1,v2,w1,w2;x;s)=(v1+v2)2(v1+v2)x+v1w12+v2w22sv2.H_{\texttt{II}}(v_{1},v_{2},w_{1},w_{2};x;s)=-(v_{1}+v_{2})^{2}-(v_{1}+v_{2})x+v_{1}w_{1}^{2}+v_{2}w_{2}^{2}-sv_{2}. (1.32)

The coupled Painlevé II system appears in both of the degeneration schemes of the Garnier system in two variables [19, Equations (3.5)-(3.7))] and the Sasano system [20, Equations (3.22)-(3.23)] by Kawakami.

Eliminating w1w_{1} and w2w_{2} from the Hamiltonian system (1.31) gives us the following nonlinear equations for v1v_{1} and v2v_{2}

{v1xxv1x22v14v1(v1+v2+x2)=0,v2xxv2x22v24v2(v1+v2+x+s2)=0.\left\{\begin{array}[]{l}v_{1xx}-\frac{v_{1x}^{2}}{2v_{1}}-4v_{1}(v_{1}+v_{2}+\frac{x}{2})=0,\\ v_{2xx}-\frac{v_{2x}^{2}}{2v_{2}}-4v_{2}(v_{1}+v_{2}+\frac{x+s}{2})=0.\\ \end{array}\right. (1.33)

Let vk(x;s)=uk(x;s)2=uk(x)2v_{k}(x;s)=u_{k}(x;s)^{2}=u_{k}(x)^{2} , k=1,2k=1,2, the above equations are further simplified to

{u1xxxu12u1(u12+u22)=0,u2xx(x+s)u22u2(u12+u22)=0.\left\{\begin{array}[]{l}u_{1xx}-xu_{1}-2u_{1}(u_{1}^{2}+u_{2}^{2})=0,\\ u_{2xx}-(x+s)u_{2}-2u_{2}(u_{1}^{2}+u_{2}^{2})=0.\\ \end{array}\right. (1.34)

If v2(x)=u2(x)2=0v_{2}(x)=u_{2}(x)^{2}=0, then (1.34) is reduced to the classical second Painlevé equation

q′′2q3xq=0.q^{\prime\prime}-2q^{3}-xq=0. (1.35)

The functions vk(x;s)v_{k}(x;s) and uk(x;s)u_{k}(x;s), k=1,2k=1,2, are also connected to HII(x;s)=HII(v1,v2,w1,w2;x;s)H_{\texttt{II}}(x;s)=H_{\texttt{II}}(v_{1},v_{2},w_{1},w_{2};x;s) by

ddxH(x;s)=(v1(x;s)+v2(x;s))=(u1(x;s)2+u2(x;s)2),\frac{d}{dx}H(x;s)=-(v_{1}(x;s)+v_{2}(x;s))=-(u_{1}(x;s)^{2}+u_{2}(x;s)^{2}), (1.36)

which can be obtained be taking derivative on both side of (1.32); see also [27, Equation (7.37)]. The existence of solutions to the coupled Painlevé II system are established in [8, 27].

Proposition 1.

([8, 27]) For the parameters ωk\omega_{k}, k=1,2k=1,2 as given in (1.19) and s>0s>0, there exist real-valued and pole-free solutions vk(x;s)v_{k}(x;s) (or uk(x;s)u_{k}(x;s)), k=1,2k=1,2, to the coupled nonlinear differential equations (1.33) (or (1.34)) subject to the boundary conditions as x+x\to+\infty

v1(x;s)=u1(x;s)2(ω1ω2)Ai(x)2,v2(x;s)=u2(x;s)2(1ω1)Ai(x+s)2,v_{1}(x;s)=u_{1}(x;s)^{2}\sim(\omega_{1}-\omega_{2}){\rm Ai}(x)^{2},\quad v_{2}(x;s)=u_{2}(x;s)^{2}\sim(1-\omega_{1}){\rm Ai}(x+s)^{2}, (1.37)

where Ai{\rm Ai} is the standard Airy function.

Asymptotics of the Hankel determinants and applications in random matrices

Our second result gives the asymptotics of the Hankel determinants expressed in terms of the solutions to the coupled Painlevé II system.

Theorem 2.

Let sks_{k} and ωk\omega_{k}, k=1,2k=1,2 be as in (1.19) and sks_{k} are related to tkt_{k}, k=1,2k=1,2 by

s1=2n+t12n1/6,s2=2n+t22n1/6,s_{1}=\sqrt{2n}+\frac{t_{1}}{\sqrt{2}n^{1/6}},\quad s_{2}=\sqrt{2n}+\frac{t_{2}}{\sqrt{2}n^{1/6}},

with t1<t2t_{1}<t_{2}, then we have the asymptotics of the Hankel determinant Dn(s1,s2)=Dn(s1,s2;ω1,ω2)D_{n}(s_{1},s_{2})=D_{n}(s_{1},s_{2};\omega_{1},\omega_{2}) defined in (1) as nn\to\infty

Dn(s1,s2)=DnGUEexp(t1+(τt1)(u1(τ;t2t1)2+u2(τ;t2t1)2)𝑑τ)(1+O(n1/6)),D_{n}(s_{1},s_{2})=D_{n}^{\texttt{GUE}}\exp\left(-\int_{t_{1}}^{+\infty}(\tau-t_{1})(u_{1}(\tau;t_{2}-t_{1})^{2}+u_{2}(\tau;t_{2}-t_{1})^{2})d\tau\right)\left(1+O(n^{-1/6})\right), (1.38)

where DnGUED_{n}^{\texttt{GUE}} is the Hankel determinant associated with the Gaussian weight with expression given in (1.6), uk(x;s)u_{k}(x;s), k=1,2k=1,2, are solutions to (1.34) subject to the boundary conditions (1.37) and the error bound is uniform for t1t_{1}, t2t_{2} in any compact subset of \mathbb{R}.

Remark 2.

When s0s\to 0, it is shown in [8, Equation (1.28)] that

u1(x;s)2+u2(x;s)2=q2(x;ω2)+O(s),u_{1}(x;s)^{2}+u_{2}(x;s)^{2}=q^{2}(x;\omega_{2})+O(s),

where q(x;ω2)q(x;\omega_{2}) is the Ablowitz-Segur solution to the second Painlevé equation (1.35) with the boundary condition as x+x\to+\infty

q(x;ω2)1ω2Ai(x).q(x;\omega_{2})\sim\sqrt{1-\omega_{2}}~{}{\rm Ai}(x). (1.39)

Therefore, as t2t10t_{2}-t_{1}\to 0, the formula (1.38) is reduced to

Dn(s1)=DnGUEexp(t1+(τt1)q2(τ;ω2))dτ)(1+O(n1/6)),D_{n}(s_{1})=D_{n}^{\texttt{GUE}}\exp\left(-\int_{t_{1}}^{+\infty}(\tau-t_{1})q^{2}(\tau;\omega_{2}))d\tau\right)\left(1+O(n^{-1/6})\right), (1.40)

where Dn(s1)D_{n}(s_{1}) is the Hankel determinant associated with the weigh function (1.5) with one jump discontinuity by taking s1=s2=2n+t12n1/6s_{1}=s_{2}=\sqrt{2n}+\frac{t_{1}}{\sqrt{2}n^{1/6}}. The expansion agrees with the result from [1] where the case with one jump discontinuity is considered.

As an application of the asymptotics of the Hankel determinants. We derive the gap probability of there being no eigenvalues in a finite interval near the extreme eigenvalues of large GUE by using (1.10) and (1.38), which confirms a recent result from [8, Equation (2.6)].

Corollary 1.

([8]) Let s1s_{1} and s2s_{2} be as in Theorem 2, we have the asymptotic approximation of the gap probability of finding no eigenvalues of GUE in the finite interval (s1,s2)(s_{1},s_{2})

Pro(λj(s1,s2):j=1n)\displaystyle\texttt{Pro}(\lambda_{j}\not\in(s_{1},s_{2}):j=1...n) =exp(t1+(τt1)(u1(τ;t2t1)2+u2(τ;t2t1)2)𝑑τ)\displaystyle=\exp\left(-\int_{t_{1}}^{+\infty}(\tau-t_{1})(u_{1}(\tau;t_{2}-t_{1})^{2}+u_{2}(\tau;t_{2}-t_{1})^{2})d\tau\right)
×(1+O(n1/6)),\displaystyle~{}~{}~{}\times\left(1+O(n^{-1/6})\right), (1.41)

where uk(x;s)u_{k}(x;s), k=1,2k=1,2, are solutions to (1.34) subject to the boundary conditions (1.37) with the parameters ω1=0,ω2=1\omega_{1}=0,\omega_{2}=1 and the error bound is uniform for t1t_{1} and t2t_{2} in any compact subset of \mathbb{R}.

In the second application, we derive from (1.38) and (1.40) the large nn limit of the distribution (1.13) in the thinning and conditioning GUE. This reproduces the result in [8, 27].

Corollary 2.

([8, 27]) Let s1s_{1} and s2s_{2} be as in Theorem 2, we have the asymptotics of the conditional gap probability

Pro(λn<s2|λmaxT<s1)\displaystyle\texttt{Pro}(\lambda_{n}<s_{2}|\lambda^{T}_{\max}<s_{1}) =exp(t1+(τt1)(u1(τ;t2t1)2+u2(τ;t2t1)2q2(τ;p))𝑑τ)\displaystyle=\exp\left(-\int_{t_{1}}^{+\infty}(\tau-t_{1})(u_{1}(\tau;t_{2}-t_{1})^{2}+u_{2}(\tau;t_{2}-t_{1})^{2}-q^{2}(\tau;p))d\tau\right)
×(1+O(n1/6)),\displaystyle~{}~{}\times\left(1+O(n^{-1/6})\right), (1.42)

where uk(x;s)u_{k}(x;s), k=1,2k=1,2, are solutions to (1.34) subject to the boundary conditions (1.37) with the parameters ω1=p(0,1),ω2=0\omega_{1}=p\in(0,1),\omega_{2}=0, q(x;p)q(x;p) is the Ablowitz-Segur solution to the second Painlevé equation (1.35) with the asymptotics (1.39) and the error bound is uniform for t1t_{1} and t2t_{2} in any compact subset of \mathbb{R}.

Asymptotics of the coupled Painlevé IV system

Next, we show that the scaling limit of the coupled Painlevé IV system leads to the coupled Painlevé II system.

Theorem 3.

Let sks_{k}, ωk\omega_{k}, k=1,2k=1,2 be as in Theorem 2 and x=(s1+s2)/2x=(s_{1}+s_{2})/2, s=(s2s1)/2s=(s_{2}-s_{1})/2, then we have the asymptotics of the coupled Painlevé IV system as nn\to\infty

a1(x;s)=12n1/6(v1(t1;t2t1)+v1x(t1;t2t1)2n1/3+O(n2/3)),\displaystyle a_{1}(x;s)=-\frac{1}{\sqrt{2}n^{1/6}}\left(v_{1}(t_{1};t_{2}-t_{1})+\frac{v_{1x}(t_{1};t_{2}-t_{1})}{2n^{1/3}}+O(n^{-2/3})\right), (1.43)
a2(x;s)=12n1/6(v2(t1;t2t1)+v2x(t1;t2t1)2n1/3+O(n2/3)),\displaystyle a_{2}(x;s)=-\frac{1}{\sqrt{2}n^{1/6}}\left(v_{2}(t_{1};t_{2}-t_{1})+\frac{v_{2x}(t_{1};t_{2}-t_{1})}{2n^{1/3}}+O(n^{-2/3})\right), (1.44)
b1(x;s)=2n(1v1x(t1;t2t1)2v1(t1;t2t1)n1/3+O(n2/3)),\displaystyle b_{1}(x;s)=\sqrt{2n}\left(1-\frac{v_{1x}(t_{1};t_{2}-t_{1})}{2v_{1}(t_{1};t_{2}-t_{1})n^{1/3}}+O(n^{-2/3})\right), (1.45)
b2(x;s)=2n(1v2x(t1;t2t1)2v2(t1;t2t1)n1/3+O(n2/3)),\displaystyle b_{2}(x;s)=\sqrt{2n}\left(1-\frac{v_{2x}(t_{1};t_{2}-t_{1})}{2v_{2}(t_{1};t_{2}-t_{1})n^{1/3}}+O(n^{-2/3})\right), (1.46)
y(x;s)=2i(2/n)n12en(t1+t2)n1/3(1+HII(t1;t2t1)18(t1+t2)2n1/3+O(n2/3)),\displaystyle y(x;s)=2i(2/n)^{n-\frac{1}{2}}e^{-n-(t_{1}+t_{2})n^{1/3}}\left(1+\frac{H_{\texttt{II}}(t_{1};t_{2}-t_{1})-\frac{1}{8}(t_{1}+t_{2})^{2}}{n^{1/3}}+O(n^{-2/3})\right), (1.47)

where v1(x;s)v_{1}(x;s) and v2(x;s)v_{2}(x;s) are solutions to the coupled Painlevé II system (1.33) subject to the boundary conditions (1.37), HII(x;s)H_{II}(x;s) is the Hamiltonian associated to these solutions and the subscript xx in vkx(x;s)v_{kx}(x;s) denotes the derivative of vk(x;s)v_{k}(x;s) with respect to xx for k=1,2k=1,2.

Asymptotics of the orthogonal polynomials

Applying Theorem 1 and 3, we obtain the asymptotics of the recurrence coefficients, leading coefficients of the orthogonal polynomials and the values of the orthogonal polynomials at s1s_{1} and s2s_{2}.

Theorem 4.

Let sks_{k}, ωk\omega_{k}, k=1,2k=1,2 be as in Theorem 2, we have the asymptotics of the recurrence coefficients and leading coefficients of the orthogonal polynomials as nn\to\infty

αn=12(v1(t1;t2t1)+v2(t1;t2t1))n1/6+O(n1/2),\displaystyle\alpha_{n}=-\frac{1}{\sqrt{2}}\left(v_{1}(t_{1};t_{2}-t_{1})+v_{2}(t_{1};t_{2}-t_{1})\right)n^{-1/6}+O(n^{-1/2}), (1.48)
βn=12n1/223/2(v1(t1;t2t1)+v2(t1;t2t1))n1/6+O(n1/2),\displaystyle\beta_{n}=\frac{1}{\sqrt{2}}n^{1/2}-2^{-3/2}\left(v_{1}(t_{1};t_{2}-t_{1})+v_{2}(t_{1};t_{2}-t_{1})\right)n^{-1/6}+O(n^{-1/2}), (1.49)
γn1=2n234n14n2en2π1/2(1+12HII(t1;t2t1)n1/3+O(n2/3)).\displaystyle\gamma_{n-1}=2^{\frac{n}{2}-\frac{3}{4}}n^{\frac{1}{4}-\frac{n}{2}}e^{\frac{n}{2}}\pi^{-1/2}\left(1+\frac{1}{2}H_{\texttt{II}}(t_{1};t_{2}-t_{1})n^{-1/3}+O(n^{-2/3})\right). (1.50)

Moreover, we derive the asymptotics of the values of the orthogonal polynomials at sks_{k} as nn\to\infty

πn(s1)=(2π1ω1)1/2(ne2)n/2n1/6et1n1/3u1(t1;t2t1)(1+O(n1/3)),\displaystyle\pi_{n}(s_{1})=\left(\frac{2\pi}{1-\omega_{1}}\right)^{1/2}\left(\frac{ne}{2}\right)^{n/2}n^{1/6}e^{t_{1}n^{1/3}}u_{1}(t_{1};t_{2}-t_{1})(1+O(n^{-1/3})), (1.51)
πn(s2)=(2πω1ω2)1/2(ne2)n/2n1/6et2n1/3u2(t1;t2t1)(1+O(n1/3)).\displaystyle\pi_{n}(s_{2})=\left(\frac{2\pi}{\omega_{1}-\omega_{2}}\right)^{1/2}\left(\frac{ne}{2}\right)^{n/2}n^{1/6}e^{t_{2}n^{1/3}}u_{2}(t_{1};t_{2}-t_{1})(1+O(n^{-1/3})). (1.52)

Here vk(x;s)v_{k}(x;s) and uk(x;s)u_{k}(x;s), k=1,2k=1,2 are solutions to (1.33) and (1.34) subject to the boundary conditions (1.37), HII(x;s)H_{II}(x;s) is the Hamiltonian correponding to these solutions.

Remark 3.

When s1s2s_{1}\to s_{2}, the weight function (1.5) is reduced to the Gaussian weight with one jump discontinuity. Then, Theorem 4, together with Remark 2, implies the asymptotics of the recurrence coefficients, leading coefficients and the orthogonal polynomials associated with Gaussian weight with one jump discontinuity. This agrees with a result from [1, Theorem 5].

1.2 Organization of the rest of this paper

The rest of the paper is organized as follows. In Section 2, we consider the Riemann-Hilbert (RH) problem for the orthogonal polynomials associated with the Gaussian weight with two jump discontinuities (1.5). We show that the RH problem is equivalent to the one for the coupled Painlevé IV system. The properties of the Painlevé IV system are studied, including the Lax pair and the Hamiltonian formulation. We then prove Theorem 1 at the end of this section which relates the Hankel determinants and the orthogonal polynomials to the coupled Painlevé IV system. In section 3, we study the asymptotics of the orthogonal polynomials by performing Deift-Zhou steepest descent analysis of the RH problem for the orthogonal polynomials. Finally, the proofs of Theorem 2-4 are given in Section 4.

2 Orthogonal polynomials and the coupled Painlevé IV system

In this section, we will relate the the Hankel determinants and the orthogonal polynomials associated with the weight function (1.5) to the coupled Painlevé IV system. The cennections are collected in Theorem 1. The derivations are based on the RH problem representation of the orthogonal polynomials.

2.1 Riemann-Hilbert problem for the orthogonal polynomials

In this subsection, we first consider the RH problem for the orthogonal polynomials with respect to (1.5), which was introduced by Fokas, Its and Kitaev [14]. We then derive several identities relating the logarithmic derivative of the Hankel determinants to the RH problem. At the end of the subsection, we transform the RH problem to a model RH problem with constant jumps.

Riemann-Hilbert problem for YY

(a)   Y(z;s1,s2)Y(z;s_{1},s_{2}) (Y(z)Y(z) for short) is analytic in \\mathbb{C}\backslash\mathbb{R};

(b)   Y(z)Y(z) satisfies the jump condition

Y+(x)=Y(x)(1w(x)01),x,Y_{+}(x)=Y_{-}(x)\left(\begin{array}[]{cc}1&w(x)\\ 0&1\\ \end{array}\right),\qquad x\in\mathbb{R},

where w(x)=w(x;s1,s2;ω1,ω2)w(x)=w(x;s_{1},s_{2};\omega_{1},\omega_{2}) is defined in (1.5);

(c)   The behavior of Y(z)Y(z) at infinity is

Y(z)=(I+Y1z+O(1z2))(zn00zn),z;Y(z)=\left(I+\frac{Y_{1}}{z}+O\left(\frac{1}{z^{2}}\right)\right)\left(\begin{array}[]{cc}z^{n}&0\\ 0&z^{-n}\\ \end{array}\right),\quad\quad z\rightarrow\infty; (2.1)

(d)   Y(z)=O(ln|zsk|)Y(z)=O(\ln|z-s_{k}|) as zskz\to s_{k} for k=1,2k=1,2.

For ωk0\omega_{k}\geqslant 0, k=1,2k=1,2, it follows from the Sokhotski-Plemelj formula and Liouville’s theorem that the unique solution of the RH problem for YY is given by

Y(z)=(πn(z)12πiπn(x)w(x)xz𝑑x2πiγn12πn1(z)γn12πn1(x)w(x)xz𝑑x),Y(z)=\left(\begin{array}[]{cc}\pi_{n}(z)&\frac{1}{2\pi i}\int_{\mathbb{R}}\frac{\pi_{n}(x)w(x)}{x-z}dx\\ -2\pi i\gamma_{n-1}^{2}\;\pi_{n-1}(z)&-\gamma_{n-1}^{2}\;\int_{\mathbb{R}}\frac{\pi_{n-1}(x)w(x)}{x-z}dx\end{array}\right), (2.2)

where πn(z)\pi_{n}(z) and γn1\gamma_{n-1} are defined in (1.7); see [14].

We establish two differential identities expressing the logarithmic derivative of the Hankel determniant DnD_{n} in terms of the solution YY.

Proposition 2.

Let sks_{k} and ωk\omega_{k}, k=1,2k=1,2 be as in (1.19) and F(s1,s2)F(s_{1},s_{2}) be the logarithmic derivative of the Hankel determinant as defined in (1.20), we have the following relations

F(s1,s2)=1ω12πies12(Y1Y)21(s1)+ω1ω22πies22(Y1Y)21(s2),F(s_{1},s_{2})=\frac{1-\omega_{1}}{2\pi i}e^{-s_{1}^{2}}(Y^{-1}Y^{\prime})_{21}(s_{1})+\frac{\omega_{1}-\omega_{2}}{2\pi i}e^{-s_{2}^{2}}(Y^{-1}Y^{\prime})_{21}(s_{2}), (2.3)

and

F(s1,s2)=2limzz(Y(z)znσ3I)11,F(s_{1},s_{2})=2\lim_{z\to\infty}z(Y(z)z^{-n\sigma_{3}}-I)_{11}, (2.4)

where YY is defined in (2.2).

Proof.

According to (1.20), it follows by taking logarithmic derivative on both sides of the equation (1.9) that

F(s1,s2)\displaystyle F(s_{1},s_{2}) =2j=0n1γj1(γjs1+γjs2)\displaystyle=-2\sum_{j=0}^{n-1}\gamma_{j}^{-1}\left(\frac{\partial\gamma_{j}}{\partial s_{1}}+\frac{\partial\gamma_{j}}{\partial s_{2}}\right) (2.5)
=j=0n1((1ω1)es12γj2πj(s1)2+(ω1ω2)es22γj2πj(s2)2).\displaystyle=\sum_{j=0}^{n-1}((1-\omega_{1})e^{-s_{1}^{2}}\gamma_{j}^{2}\pi_{j}(s_{1})^{2}+(\omega_{1}-\omega_{2})e^{-s_{2}^{2}}\gamma_{j}^{2}\pi_{j}(s_{2})^{2}). (2.6)

Applying the Christoffel-Darboux identity, we obtain

F(s1,s2)=\displaystyle F(s_{1},s_{2})= (1ω1)es12γn12(πn(s1)πn1(s1)πn(s1)πn1(s1))\displaystyle(1-\omega_{1})e^{-s_{1}^{2}}\gamma_{n-1}^{2}(\pi_{n}^{\prime}(s_{1})\pi_{n-1}(s_{1})-\pi_{n}(s_{1})\pi^{\prime}_{n-1}(s_{1}))
+(ω1ω2)es22γn12(πn(s2)πn1(s2)πn(s2)πn1(s2)).\displaystyle~{}+(\omega_{1}-\omega_{2})e^{-s_{2}^{2}}\gamma_{n-1}^{2}(\pi_{n}^{\prime}(s_{2})\pi_{n-1}(s_{2})-\pi_{n}(s_{2})\pi^{\prime}_{n-1}(s_{2})). (2.7)

Then, the differential identity (2.3) follows from the definition of YY and (2).

To prove (2.4), we use a change of variable in (1.7) and obtain

γj2=γj(s1,s2)2=πj(x)2w(x)𝑑x=πj(x+sk)2w(x+sk)𝑑x,k=1,2.\gamma_{j}^{-2}=\gamma_{j}(s_{1},s_{2})^{-2}=\int_{\mathbb{R}}\pi_{j}(x)^{2}w(x)dx=\int_{\mathbb{R}}\pi_{j}(x+s_{k})^{2}w(x+s_{k})dx,\quad k=1,2. (2.8)

Taking derivative with respect to sks_{k} on both sides of (2.8) for k=1,2k=1,2 and using the orthogonality and the definition of the weight function (1.5), we have

2γj1s1γj\displaystyle-2\gamma_{j}^{-1}\frac{\partial}{\partial s_{1}}\gamma_{j} =2γj2xπj(x)2w(x)𝑑x(ω1ω2)es22γj2πj(s2)2,\displaystyle=-2\gamma_{j}^{2}\int_{\mathbb{R}}x\pi_{j}(x)^{2}w(x)dx-(\omega_{1}-\omega_{2})e^{-s_{2}^{2}}\gamma_{j}^{2}\pi_{j}(s_{2})^{2}, (2.9)

and

2γj1s2γj\displaystyle-2\gamma_{j}^{-1}\frac{\partial}{\partial s_{2}}\gamma_{j} =2γj2xπj(x)2w(x)𝑑x(1ω1)es12γj2πj(s1)2.\displaystyle=-2\gamma_{j}^{2}\int_{\mathbb{R}}x\pi_{j}(x)^{2}w(x)dx-(1-\omega_{1})e^{-s_{1}^{2}}\gamma_{j}^{2}\pi_{j}(s_{1})^{2}. (2.10)

Combining the formulas with (2.5)-(2.6) and using the Christoffel-Darboux formula once again, we obtain

F(s1,s2)=2γn12x(ddxπn(x)πn1(x)πn(x)ddxπn1(x))w(x)𝑑x.F(s_{1},s_{2})=-2\gamma_{n-1}^{2}\int_{\mathbb{R}}x\left(\frac{d}{dx}\pi_{n}(x)\pi_{n-1}(x)-\pi_{n}(x)\frac{d}{dx}\pi_{n-1}(x)\right)w(x)dx. (2.11)

From

πn(x)=xn+pnxn1+,\pi_{n}(x)=x^{n}+p_{n}x^{n-1}+\cdots,

we have the decomposition

xddxπn(x)=nπn(x)pnπn1(x)+.x\frac{d}{dx}\pi_{n}(x)=n\pi_{n}(x)-p_{n}\pi_{n-1}(x)+\cdots.

Substituting this into (2.11) and using the orthogonality, we obtain (2.4). This completes Proposition 2. ∎

We define

Φ(z;x,s)=σ1ex22σ3Y(z+x)e12(z+x)2σ3σ1,\Phi(z;x,s)=\sigma_{1}e^{\frac{x^{2}}{2}\sigma_{3}}Y(z+x)e^{-\frac{1}{2}(z+x)^{2}\sigma_{3}}\sigma_{1}, (2.12)

where the variables x,sx,s are related to s1s_{1} and s2s_{2} by (1.21). Then Φ(z)=Φ(z;x,s)\Phi(z)=\Phi(z;x,s) satisfies the following RH problem.

Riemann-Hilbert problem for Φ\Phi

(a)   Φ(z)\Phi(z) is analytic in \\mathbb{C}\backslash\mathbb{R};

(b)   Φ(z)\Phi(z) satisfies the jump condition

Φ+(z)=Φ(z)(1011),z<s,\Phi_{+}(z)=\Phi_{-}(z)\left(\begin{array}[]{cc}1&0\\ 1&1\\ \end{array}\right),\quad z<-s,
Φ+(z)=Φ(z)(10ω11),s<z<s;\Phi_{+}(z)=\Phi_{-}(z)\left(\begin{array}[]{cc}1&0\\ \omega_{1}&1\\ \end{array}\right),\quad-s<z<s;
Φ+(z)=Φ(z)(10ω21),z>s;\Phi_{+}(z)=\Phi_{-}(z)\left(\begin{array}[]{cc}1&0\\ \omega_{2}&1\\ \end{array}\right),\quad z>s;

(c)   The behavior of Φ(z)\Phi(z) at infinity is

Φ(z)=(I+Φ1z+Φ2z2+O(1z3))e(12z2+xz)σ3znσ3;\Phi(z)=\left(I+\frac{\Phi_{1}}{z}+\frac{\Phi_{2}}{z^{2}}+O\left(\frac{1}{z^{3}}\right)\right)e^{(\frac{1}{2}z^{2}+xz)\sigma_{3}}z^{-n\sigma_{3}}; (2.13)

(d)   The behavior of Φ(z)\Phi(z) near s-s is

Φ(z)=Φ(s)(z)(I+1ω12πi(0010)ln(z+s))E(s),\Phi(z)=\Phi^{(-s)}(z)\left(I+\frac{1-\omega_{1}}{2\pi i}\left(\begin{array}[]{cc}0&0\\ 1&0\end{array}\right)\ln(z+s)\right)E^{(-s)}, (2.14)

where arg(z+s)(π,π)\mathop{\rm arg}\nolimits(z+s)\in(-\pi,\pi). Here, Φ(s)(z)\Phi^{(-s)}(z) is analytic near z=sz=-s and has the following expansion

Φ(s)(z)=P0(x,s)(I+P1(x,s)(z+s)+O((z+s)2)).\Phi^{(-s)}(z)=P_{0}(x,s)(I+P_{1}(x,s)(z+s)+O((z+s)^{2})). (2.15)

The piecewise constant matrix E(s)E^{(-s)} is given by

E(s)={(1001),Imz>0,(10ω11),Imz<0.E^{(-s)}=\left\{\begin{array}[]{ll}\left(\begin{array}[]{cc}1&0\\ 0&1\end{array}\right),&\mathop{\rm Im}\nolimits z>0,\\ \left(\begin{array}[]{cc}1&0\\ -\omega_{1}&1\end{array}\right),&\mathop{\rm Im}\nolimits z<0.\end{array}\right.

(e)   The behavior of Φ(z)\Phi(z) near ss is

Φ(z)=Φ(s)(z)(I+ω1ω22πi(0010)ln(zs))E(s),\Phi(z)=\Phi^{(s)}(z)\left(I+\frac{\omega_{1}-\omega_{2}}{2\pi i}\left(\begin{array}[]{cc}0&0\\ 1&0\end{array}\right)\ln(z-s)\right)E^{(s)}, (2.16)

where arg(zs)(π,π)\mathop{\rm arg}\nolimits(z-s)\in(-\pi,\pi). Here, Φ(s)(z)\Phi^{(s)}(z) is analytic near z=sz=s and has the following expansion

Φ(s)(z)=Q0(x,s)(I+Q1(x,s)(zs)+O((zs)2)).\Phi^{(s)}(z)=Q_{0}(x,s)(I+Q_{1}(x,s)(z-s)+O((z-s)^{2})). (2.17)

The piecewise constant matrix E(s)E^{(s)} is defined by

E(s)={(1001),Imz>0,(10ω21),Imz<0.E^{(s)}=\left\{\begin{array}[]{ll}\left(\begin{array}[]{cc}1&0\\ 0&1\end{array}\right),&\mathop{\rm Im}\nolimits z>0,\\ \left(\begin{array}[]{cc}1&0\\ -\omega_{2}&1\end{array}\right),&\mathop{\rm Im}\nolimits z<0.\end{array}\right.

2.2 Lax pair and the coupled Painlevé IV system

In this section, we show that the solution Φ(z;x,s)\Phi(z;x,s) of the RH problem satisfies a system of differential equations in zz and xx when the parameter ss is fixed. The compatibility condition Φzx(z;x,s)=Φxz(z;x,s)\Phi_{zx}(z;x,s)=\Phi_{xz}(z;x,s) gives us the coupled Painlevé IV system. The Hamiltonian for the system is also derived.

Proposition 3.

We have the following Lax pair

Φz(z;x,s)=A(z;x,s)Φ(z;x,s),Φx(z;x,s)=B(z;x,s)Φ(z;x,s),\Phi_{z}(z;x,s)=A(z;x,s)\Phi(z;x,s),\quad\Phi_{x}(z;x,s)=B(z;x,s)\Phi(z;x,s), (2.18)

where

A(z;x,s)=(z+x)σ3+A(x,s)+A1(x,s)z+s+A2(x,s)zs,A(z;x,s)=(z+x)\sigma_{3}+A_{\infty}(x,s)+\frac{A_{1}(x,s)}{z+s}+\frac{A_{2}(x,s)}{z-s}, (2.19)
B(z;x,s)=zσ3+A(x,s),B(z;x,s)=z\sigma_{3}+A_{\infty}(x,s), (2.20)

with the coefficients given below

A(x,s)=(0y(x;s)2(a1(x;s)b1(x;s)+a2(x;s)b2(x;s)+n)/y(x;s)0),A_{\infty}(x,s)=\begin{pmatrix}0&y(x;s)\\ -2\left(a_{1}(x;s)b_{1}(x;s)+a_{2}(x;s)b_{2}(x;s)+n\right)/y(x;s)&0\end{pmatrix}, (2.21)
Ak(x,s)=(ak(x;s)bk(x;s)ak(x;s)y(x;s)ak(x;s)bk2(x;s)/y(x;s)ak(x;s)bk(x;s)),k=1,2.A_{k}(x,s)=\begin{pmatrix}a_{k}(x;s)b_{k}(x;s)&a_{k}(x;s)y(x;s)\\ -a_{k}(x;s)b_{k}^{2}(x;s)/y(x;s)&-a_{k}(x;s)b_{k}(x;s)\end{pmatrix},\quad k=1,2. (2.22)

The compatibility condition of the Lax pair gives us the coupled Painlevé IV system

{dydx=2(a1+a2x)y,da1dx=2a1(a1+a2+b1x+s),da2dx=2a2(a1+a2+b2xs),db1dx=b12+2b1(2a1+a2x+s)+2(a2b2+n),db2dx=b22+2b2(a1+2a2xs)+2(a1b1+n).\left\{\begin{array}[]{l}\frac{dy}{dx}=2(a_{1}+a_{2}-x)y,\\ \frac{da_{1}}{dx}=-2a_{1}(a_{1}+a_{2}+b_{1}-x+s),\\ \frac{da_{2}}{dx}=-2a_{2}(a_{1}+a_{2}+b_{2}-x-s),\\ \frac{db_{1}}{dx}=b_{1}^{2}+2b_{1}(2a_{1}+a_{2}-x+s)+2(a_{2}b_{2}+n),\\ \frac{db_{2}}{dx}=b_{2}^{2}+2b_{2}(a_{1}+2a_{2}-x-s)+2(a_{1}b_{1}+n).\end{array}\right. (2.23)

Eliminating b1b_{1} and b2b_{2} from the system, it is seen that a1a_{1} and a2a_{2} satisfy the following nonlinear differential equations

{𝚍2a1𝚍x212a1(𝚍a1𝚍x)26a1(a1+a2)2+8a1(a1+a2)x8a12s+2(2n1)a12a1(xs)2=0,𝚍2a2𝚍x212a2(𝚍a2𝚍x)26a2(a1+a2)2+8a2(a1+a2)x+8a22s+2(2n1)a22a2(x+s)2=0.\left\{\begin{array}[]{l}\frac{\mathtt{d}^{2}a_{1}}{\mathtt{d}x^{2}}-\frac{1}{2a_{1}}\left(\frac{\mathtt{d}a_{1}}{\mathtt{d}x}\right)^{2}-6a_{1}(a_{1}+a_{2})^{2}+8a_{1}(a_{1}+a_{2})x-8a_{1}^{2}s+2(2n-1)a_{1}-2a_{1}(x-s)^{2}=0,\\ \frac{\mathtt{d}^{2}a_{2}}{\mathtt{d}x^{2}}-\frac{1}{2a_{2}}\left(\frac{\mathtt{d}a_{2}}{\mathtt{d}x}\right)^{2}-6a_{2}(a_{1}+a_{2})^{2}+8a_{2}(a_{1}+a_{2})x+8a_{2}^{2}s+2(2n-1)a_{2}-2a_{2}(x+s)^{2}=0.\end{array}\right. (2.24)
Proof.

Since the jump matrices of the RH problem for Φ(z;x,s)\Phi(z;x,s) are independent of the variables zz and xx, we have that Φz(z;x,s)\Phi_{z}(z;x,s), Φx(z;x,s)\Phi_{x}(z;x,s) and Φ(z;x,s)\Phi(z;x,s) satisfy the same jump condition. Thus, the coefficient A(z;x,s)A(z;x,s) in the differential equations are meromorphic for zz in the complex plane with only possible isolate singularities at z=0z=0, ±s\pm s and the coefficient B(z;x,s)B(z;x,s) is analytic for zz in the complex plane. Then, it follows from the local behavior of Φ(z;x,s)\Phi(z;x,s) as zz\to\infty, z±sz\to\pm s that the coefficients A(z;x,s)A(z;x,s) and B(z;x,s)B(z;x,s) are rational functions in zz with the form given in (2.19)-(2.20). Using the fact that detΦ=1\det\Phi=1, we have trA=trB=0{\rm tr}A={\rm tr}B=0 and thus all the coefficients AkA_{k}, k=0,1,2k=0,1,2 in (2.19) are trace-zero. Using the master equation in (2.18) and the local behavior Φ(z)\Phi(z) at z±sz\pm s, we have

detAk=0,k=1,2.\det A_{k}=0,\quad k=1,2.

We denote (Ak)11=akbk(A_{k})_{11}=a_{k}b_{k} for k=1,2k=1,2.

Substituting the behavior of Φ\Phi at infinity into the master equation of the Lax pair (2.18), we find after comparing the coefficients of z0z^{0} and z1z^{-1} on both sides of the equation that

A=[Φ1,σ3],A_{\infty}=[\Phi_{1},\sigma_{3}], (2.25)

and

A1+A2=nσ3+[Φ2+xΦ1,σ3]+[σ3,Φ1]Φ1,A_{1}+A_{2}=-n\sigma_{3}+[\Phi_{2}+x\Phi_{1},\sigma_{3}]+[\sigma_{3},\Phi_{1}]\Phi_{1}, (2.26)

where Φk\Phi_{k} is the coefficient of zkz^{-k} in the large zz asymptotic expansion of Φ(z)\Phi(z). In view of (2.25), we get

A=(02(Φ1)122(Φ1)210).A_{\infty}=\begin{pmatrix}0&-2(\Phi_{1})_{12}\\ 2(\Phi_{1})_{21}&0\end{pmatrix}. (2.27)

From the diagonal entries of the equation (2.26), we find the relation

2(Φ1)12(Φ1)21=n+(A1+A2)11=a1b1+a2b2+n.2(\Phi_{1})_{12}(\Phi_{1})_{21}=n+(A_{1}+A_{2})_{11}=a_{1}b_{1}+a_{2}b_{2}+n. (2.28)

We define

y=2(Φ1)12,y=-2(\Phi_{1})_{12}, (2.29)

then the above relations imply that

(A)12=y,(A)21=2y(a1b1+a2b2+n).(A_{\infty})_{12}=y,\quad(A_{\infty})_{21}=-\frac{2}{y}(a_{1}b_{1}+a_{2}b_{2}+n). (2.30)

We define (Ak)12=aky(A_{k})_{12}=a_{k}y for k=1,2k=1,2. Then, the other entries of AkA_{k} can be expressed in terms of aka_{k} and bkb_{k} for k=1,2k=1,2, as given in (2.22).

Similarly, the coefficient B(z)=B(z;x,s)B(z)=B(z;x,s) can be determined by using the behavior of Φ\Phi at infinity

B(z)=Φx(z)Φ(z)1=zσ3+(02(Φ1)122(Φ1)210)=zσ3+A.B(z)=\Phi_{x}(z)\Phi(z)^{-1}=z\sigma_{3}+\begin{pmatrix}0&-2(\Phi_{1})_{12}\\ 2(\Phi_{1})_{21}&0\end{pmatrix}=z\sigma_{3}+A_{\infty}. (2.31)

The compatibility condition Φzx=Φxz\Phi_{zx}=\Phi_{xz} gives us the zero-curve equation

AxBz+[A,B]=0.A_{x}-B_{z}+[A,B]=0. (2.32)

Substituting (2.19) and (2.20) into the above equation, the compatibility condition is equivalent to

{dAdx=x[A,σ3][A1,σ3][A2,σ3],dA1dx=[A1,A]+s[A1,σ3],dA2dx=[A2,A]s[A2,σ3].\left\{\begin{array}[]{l}\frac{dA_{\infty}}{dx}=x[A_{\infty},\sigma_{3}]-[A_{1},\sigma_{3}]-[A_{2},\sigma_{3}],\\ \frac{dA_{1}}{dx}=-[A_{1},A_{\infty}]+s[A_{1},\sigma_{3}],\\ \frac{dA_{2}}{dx}=-[A_{2},A_{\infty}]-s[A_{2},\sigma_{3}].\end{array}\right. (2.33)

We then obtain the system of differential equations (2.23). Deleting b1b_{1} and b2b_{2} from the system, we obtain the differential equations for a1a_{1} and a2a_{2} as given in (2.24). This completes the proof of Proposition 3.

Proposition 4.

The Hamiltonian for the coupled Painlevé IV system is

HIV(a1,a2,b1,b2;x,s)=2(a1b1+a2b2+n)(a1+a2)+2(a1b1(xs)+a2b2(x+s)+nx)(a1b12+a2b22).H_{\texttt{IV}}(a_{1},a_{2},b_{1},b_{2};x,s)=-2(a_{1}b_{1}+a_{2}b_{2}+n)(a_{1}+a_{2})+2(a_{1}b_{1}(x-s)+a_{2}b_{2}(x+s)+nx)-(a_{1}b_{1}^{2}+a_{2}b_{2}^{2}). (2.34)

And the coupled Painlevé IV system (2.23) can be written as the Hamiltonian system:

{da1dx=HIV,b1(a1,a2,b1,b2;x,s),da2dx=HIV,b2(a1,a2,b1,b2;x,s),db1dx=HIV,a1(a1,a2,b1,b2;x,s),db2dx=HIV,a2(a1,a2,b1,b2;x,s),\left\{\begin{array}[]{l}\frac{da_{1}}{dx}=H_{\texttt{IV},b_{1}}(a_{1},a_{2},b_{1},b_{2};x,s),\\ \frac{da_{2}}{dx}=H_{\texttt{IV},b_{2}}(a_{1},a_{2},b_{1},b_{2};x,s),\\ \frac{db_{1}}{dx}=-H_{\texttt{IV},a_{1}}(a_{1},a_{2},b_{1},b_{2};x,s),\\ \frac{db_{2}}{dx}=-H_{\texttt{IV},a_{2}}(a_{1},a_{2},b_{1},b_{2};x,s),\end{array}\right. (2.35)

where HIV,aH_{\texttt{IV},a} denotes the partial derivative of HIVH_{\texttt{IV}} with respect to aa.

Proof.

The Hamiltonian introduced by Jimbo, Miwa and Ueno [18] is given by

HIV(x;s)=𝚁𝚎𝚜z=Φ()(z)1ddzΦ()(z)ddxΘ(z;x)σ3=2(Φ1)11,H_{\texttt{IV}}(x;s)=-\mathtt{Res}_{z=\infty}\Phi^{(\infty)}(z)^{-1}\frac{d}{dz}\Phi^{(\infty)}(z)\frac{d}{dx}\Theta(z;x)\sigma_{3}=-2(\Phi_{1})_{11}, (2.36)

where Θ(z;x)=(12z2+xz)σ3\Theta(z;x)=(\frac{1}{2}z^{2}+xz)\sigma_{3} and Φ1\Phi_{1} is the coefficient of z1z^{-1} in the large zz expansion of Φ\Phi in (2.13). Using (2.3), (2.4) and (2.12), we have

HIV(x;s)\displaystyle H_{\texttt{IV}}(x;s) =F(s1,s2)+2nx\displaystyle=F(s_{1},s_{2})+2nx (2.37)
=2nx+1ω12πi(Φ1Φz)12(s)+ω1ω22πi(Φ1Φz)12(s).\displaystyle=2nx+\frac{1-\omega_{1}}{2\pi i}\left(\Phi^{-1}\Phi_{z}\right)_{12}(-s)+\frac{\omega_{1}-\omega_{2}}{2\pi i}\left(\Phi^{-1}\Phi_{z}\right)_{12}(s).

From (2.14) and (2.16), we get

HIV(x;s)=2nx+1ω12πi(P1)12+ω1ω22πi(Q1)12,H_{\texttt{IV}}(x;s)=2nx+\frac{1-\omega_{1}}{2\pi i}\left(P_{1}\right)_{12}+\frac{\omega_{1}-\omega_{2}}{2\pi i}\left(Q_{1}\right)_{12}, (2.38)

where P1=P1(x,s)P_{1}=P_{1}(x,s) and Q1=Q1(x,s)Q_{1}=Q_{1}(x,s) are defined in (2.15) and (2.17), respectively. Substituting the expansions (2.14) and (2.16) into the master equation of (2.18), we obtain

1ω12πiP0(0010)P01=A1,\displaystyle\frac{1-\omega_{1}}{2\pi i}P_{0}\left(\begin{array}[]{cc}0&0\\ 1&0\end{array}\right)P_{0}^{-1}=A_{1}, (2.41)
P1+1ω12πi[P1,(0010)]=P01((xs)σ3+AA22s)P0,\displaystyle P_{1}+\frac{1-\omega_{1}}{2\pi i}\left[P_{1},\left(\begin{array}[]{cc}0&0\\ 1&0\end{array}\right)\right]=P_{0}^{-1}((x-s)\sigma_{3}+A_{\infty}-\frac{A_{2}}{2s})P_{0}, (2.44)
ω1ω22πiQ0(0010)Q01=A2,\displaystyle\frac{\omega_{1}-\omega_{2}}{2\pi i}Q_{0}\left(\begin{array}[]{cc}0&0\\ 1&0\end{array}\right)Q_{0}^{-1}=A_{2}, (2.47)
Q1+ω1ω22πi[Q1,(0010)]=Q01((x+s)σ3+A+A12s)Q0.\displaystyle Q_{1}+\frac{\omega_{1}-\omega_{2}}{2\pi i}\left[Q_{1},\left(\begin{array}[]{cc}0&0\\ 1&0\end{array}\right)\right]=Q_{0}^{-1}((x+s)\sigma_{3}+A_{\infty}+\frac{A_{1}}{2s})Q_{0}. (2.50)

Now

P0(0010)P01=((P0)12(P0)22(P0)122(P0)222(P0)12(P0)22).P_{0}\left(\begin{array}[]{cc}0&0\\ 1&0\end{array}\right)P_{0}^{-1}=\left(\begin{array}[]{cc}(P_{0})_{12}(P_{0})_{22}&-(P_{0})_{12}^{2}\\ (P_{0})_{22}^{2}&-(P_{0})_{12}(P_{0})_{22}\\ \end{array}\right).

Then, a substitution of the above equation into (2.41) gives

1ω12πi(P0)222=a1b12y,1ω12πi(P0)122=a1y,1ω12πi(P0)12(P0)22=a1b1.\frac{1-\omega_{1}}{2\pi i}(P_{0})_{22}^{2}=-\frac{a_{1}b_{1}^{2}}{y},\quad\frac{1-\omega_{1}}{2\pi i}(P_{0})_{12}^{2}=-a_{1}y,\quad\frac{1-\omega_{1}}{2\pi i}(P_{0})_{12}(P_{0})_{22}=a_{1}b_{1}. (2.51)

Let

𝒜=(xs)σ3+AA22s,\mathcal{A}=(x-s)\sigma_{3}+A_{\infty}-\frac{A_{2}}{2s},

we obtain from (2.44) that

1ω12πi(P1)12\displaystyle\frac{1-\omega_{1}}{2\pi i}(P_{1})_{12} =1ω12πi(2(P0)12(P0)22𝒜11(P0)122𝒜21+(P0)222𝒜12)\displaystyle=\frac{1-\omega_{1}}{2\pi i}(2(P_{0})_{12}(P_{0})_{22}\mathcal{A}_{11}-(P_{0})_{12}^{2}\mathcal{A}_{21}+(P_{0})_{22}^{2}\mathcal{A}_{12})
=2a1b1𝒜11+a1y𝒜21a1b12y𝒜12\displaystyle=2a_{1}b_{1}\mathcal{A}_{11}+a_{1}y\mathcal{A}_{21}-\frac{a_{1}b_{1}^{2}}{y}\mathcal{A}_{12}
=2a1b1(xs)2a1(a1b1+a2b2+n)a1b12+12sa1a2(b1b2)2.\displaystyle=2a_{1}b_{1}(x-s)-2a_{1}(a_{1}b_{1}+a_{2}b_{2}+n)-a_{1}b_{1}^{2}+\frac{1}{2s}a_{1}a_{2}(b_{1}-b_{2})^{2}. (2.52)

Similarly, we get after some straightforward calculations

ω1ω22πi(Q0)222=a2b22y,ω1ω22πi(Q0)122=a2y,ω1ω22πi(Q0)12(Q0)22=a2b2,\frac{\omega_{1}-\omega_{2}}{2\pi i}(Q_{0})_{22}^{2}=-\frac{a_{2}b_{2}^{2}}{y},\quad\frac{\omega_{1}-\omega_{2}}{2\pi i}(Q_{0})_{12}^{2}=-a_{2}y,\quad\frac{\omega_{1}-\omega_{2}}{2\pi i}(Q_{0})_{12}(Q_{0})_{22}=a_{2}b_{2}, (2.53)

and

ω1ω22πi(Q1)12=2a2b2(x+s)2a1(a1b1+a2b2+n)a2b2212sa1a2(b1b2)2.\frac{\omega_{1}-\omega_{2}}{2\pi i}(Q_{1})_{12}=2a_{2}b_{2}(x+s)-2a_{1}(a_{1}b_{1}+a_{2}b_{2}+n)-a_{2}b_{2}^{2}-\frac{1}{2s}a_{1}a_{2}(b_{1}-b_{2})^{2}. (2.54)

Then, the expression of the Hamiltonian (2.34) follows directly by substituting (2.2) and (2.54) into (2.38). In view of the Hamiltonian (2.34), it is seen that the coupled Painlevé IV system (2.23) is equivalent to the Hamiltonian system (2.35). This completes the proof of Proposition 4. ∎

2.3 Proof of Theorem 1

The relation (1.22) follows from (2.37). Let Y1Y_{1} and Y2Y_{2} be the coefficients of 1/z1/z and 1/z21/z^{2} in the expansion of YY near infinity (2.1), we have the following relations for the recurrence coefficients αn=αn(s1,s2)\alpha_{n}=\alpha_{n}(s_{1},s_{2}), βn1=βn1(s1,s2)\beta_{n-1}=\beta_{n-1}(s_{1},s_{2}) and the leading coefficient γn=γn(s1,s2)\gamma_{n}=\gamma_{n}(s_{1},s_{2}) of the monic orthogonal polynomial of degree n1n-1:

αn=(Y1)11+(Y2)12(Y1)12,βn2=(Y1)12(Y1)21andγn12=12πi(Y1)21;\alpha_{n}=(Y_{1})_{11}+\frac{(Y_{2})_{12}}{(Y_{1})_{12}},\quad\quad\beta^{2}_{n}=(Y_{1})_{12}(Y_{1})_{21}\quad\mbox{and}\quad\gamma_{n-1}^{2}=-\frac{1}{2\pi i}(Y_{1})_{21}; (2.55)

see [11]. In view of (2.12), it is then seen that

(Y1)11=(Φ1)11nx,(Y1)12=ex2(Φ1)21,(Y1)21=ex2(Φ1)12,(Y_{1})_{11}=-(\Phi_{1})_{11}-nx,\quad\quad(Y_{1})_{12}=e^{-x^{2}}(\Phi_{1})_{21},\quad\quad(Y_{1})_{21}=e^{x^{2}}(\Phi_{1})_{12}, (2.56)

and

(Y2)12=ex2((n+1)x(Φ1)21+(Φ2)21),(Y_{2})_{12}=e^{-x^{2}}((n+1)x(\Phi_{1})_{21}+(\Phi_{2})_{21}), (2.57)

where Φ1\Phi_{1} and Φ2\Phi_{2} are defined in (2.13). From the relation (2.26), we have

x(Φ1)21+(Φ2)21=12(A1+A2)21+(Φ1)11(Φ1)21.x(\Phi_{1})_{21}+(\Phi_{2})_{21}=\frac{1}{2}(A_{1}+A_{2})_{21}+(\Phi_{1})_{11}(\Phi_{1})_{21}. (2.58)

Substituting (2.56) into (2.55) and recalling (2.28)-(2.30), we obtain that

βn2=(Φ1)12(Φ1)21=12(a1(x;s)b1(x;s)+a2(x;s)b2(x;s)+n),\beta^{2}_{n}=(\Phi_{1})_{12}(\Phi_{1})_{21}=\frac{1}{2}\left(a_{1}(x;s)b_{1}(x;s)+a_{2}(x;s)b_{2}(x;s)+n\right), (2.59)

and

γn12=12πiex2(Φ1)12=14πiex2y(x;s)0.\gamma_{n-1}^{2}=-\frac{1}{2\pi i}e^{x^{2}}(\Phi_{1})_{12}=\frac{1}{4\pi i}e^{x^{2}}y(x;s)\neq 0. (2.60)

On account of (2.23), we have

ddxlnγn1=a1(x;s)+a2(x;s).\frac{d}{dx}\ln\gamma_{n-1}=a_{1}(x;s)+a_{2}(x;s). (2.61)

Inserting (2.56)-(2.58) into (2.55) yields

αn\displaystyle\alpha_{n} =(Φ1)11+(Φ2)21+x(Φ1)21(Φ1)21\displaystyle=-(\Phi_{1})_{11}+\frac{(\Phi_{2})_{21}+x(\Phi_{1})_{21}}{(\Phi_{1})_{21}}
=12(A1+A2)21(Φ1)21\displaystyle=\frac{1}{2}\frac{(A_{1}+A_{2})_{21}}{(\Phi_{1})_{21}}
=a1(x;s)b12(x;s)+a2(x;s)b22(x;s)2(a1(x;s)b1(x;s)+a2(x;s)b2(x;s)+n).\displaystyle=\frac{a_{1}(x;s)b_{1}^{2}(x;s)+a_{2}(x;s)b_{2}^{2}(x;s)}{2(a_{1}(x;s)b_{1}(x;s)+a_{2}(x;s)b_{2}(x;s)+n)}. (2.62)

In summary, we obtain (1.23)-(1.26) by collecting (2.59)-(2.3).

From (2.2), (2.12) and (2.51), it is seen that

πn(s1)=(Y)11(s1)=e12s2sx(Φ)22(s)=e12s2sx(P0)22,\pi_{n}(s_{1})=(Y)_{11}(s_{1})=e^{\frac{1}{2}s^{2}-sx}(\Phi)_{22}(-s)=e^{\frac{1}{2}s^{2}-sx}(P_{0})_{22}, (2.63)

and

πn(s2)=(Y)11(s2)=e12s2+sx(Φ)22(s)=e12s2+sx(Q0)22.\pi_{n}(s_{2})=(Y)_{11}(s_{2})=e^{\frac{1}{2}s^{2}+sx}(\Phi)_{22}(s)=e^{\frac{1}{2}s^{2}+sx}(Q_{0})_{22}. (2.64)

Therefore, we obtain (1.27) and (1.28) after replacing the expressions of (P0)22(P_{0})_{22} and (Q0)22(Q_{0})_{22} by (2.51) and (2.53). This completes the proof of Theorem 1.

3 Nonlinear steepest descent analysis of the Riemann-Hilbert problem for YY

In this section, we take s1=2n+t12n1/6s_{1}=\sqrt{2n}+\frac{t_{1}}{\sqrt{2}n^{1/6}} and s2=2n+t22n1/6s_{2}=\sqrt{2n}+\frac{t_{2}}{\sqrt{2}n^{1/6}} in the weight function (1.5). Then, we perform Deift-Zhou nonlinear steepest descent analysis [10, 11, 12] for the Riemann-Hilbert problem for Y(z;s1,s2)Y(z;s_{1},s_{2}) as nn\to\infty. The analysis will allow us to find the asymptotics of Hankel determinants and the orthogonal polynomials associated with (1.5). The analysis of a Riemann-Hilbert problem with one jump singularity in the weight function (1.5) is considered in [28] by the second author and Zhao.

3.1 The first transformation: YTY\rightarrow T

The first transformation is defined by

T(z)=(2n)12nσ3e12nlσ3Y(2nz)en(12lg(z))σ3,z\,T(z)=(2n)^{-\frac{1}{2}n\sigma_{3}}e^{-\frac{1}{2}nl\sigma_{3}}Y(\sqrt{2n}z)e^{n\left(\frac{1}{2}l-g(z)\right)\sigma_{3}},\quad z\in\mathbb{C}\backslash\mathbb{R}, (3.1)

where the constant l=12ln2l=-1-2\ln 2. The gg-function therein is defined by

g(z)=2π11ln(zx)1x2𝑑x,g(z)=\frac{2}{\pi}\int_{-1}^{1}\ln(z-x)\sqrt{1-x^{2}}dx, (3.2)

where the logarithm takes the principle branch arg(zx)(π,π)\mathop{\rm arg}\nolimits(z-x)\in(-\pi,\pi). We then introduce the ϕ\phi-function

ϕ(z)=zz21ln(z+z21),\phi(z)=z\sqrt{z^{2}-1}-\ln\left(z+\sqrt{z^{2}-1}\right), (3.3)

where the principle branches are chosen. The ϕ\phi-function and gg-function are related by

2[g(z)+ϕ(z)]2z2l=0,z\(,1].2\left[g(z)+\phi(z)\right]-2z^{2}-l=0,\quad z\in\mathbb{C}\backslash(-\infty,1]. (3.4)

As a consequence, TT is normalized at infinity

T(z)=I+O(1/z),T(z)=I+O(1/z),

and satisfies the jump condition

T+(x)=T(x){(1θ(x)e2nϕ(x)01),x(1,+);(e2nϕ+(x)θ(x)0e2nϕ(x)),x(1,1);(1e2nϕ+(x)01),x(,1),T_{+}(x)=T_{-}(x)\left\{\begin{array}[]{ll}\left(\begin{array}[]{cc}1&\theta(x)e^{-2n\phi(x)}\\ 0&1\\ \end{array}\right),&x\in(1,+\infty);\\ &\\ \left(\begin{array}[]{cc}e^{2n\phi_{+}(x)}&\theta(x)\\ 0&e^{2n\phi_{-}(x)}\\ \end{array}\right),&x\in(-1,1);\\ &\\ \left(\begin{array}[]{cc}1&e^{-2n\phi_{+}(x)}\\ 0&1\\ \end{array}\right),&x\in(-\infty,-1),\end{array}\right. (3.5)

where θ(x)={1x<λ1ω1λ1<x<λ2,ω2x>λ2,\theta(x)=\left\{\begin{array}[]{cc}1&x<\lambda_{1}\\ \omega_{1}&\lambda_{1}<x<\lambda_{2},\\ \omega_{2}&x>\lambda_{2},\end{array}\right. with λ1=1+t12n2/3\lambda_{1}=1+\frac{t_{1}}{2n^{2/3}} and λ2=1+t22n2/3\lambda_{2}=1+\frac{t_{2}}{2n^{2/3}}.

3.2 The second transformation: TST\rightarrow S

In the second transformation, we define

S(z)={T(z),for z outside the lens,T(z)(10e2nϕ(z)1),for z in the upper lens,T(z)(10e2nϕ(z)1),for z in the lower lens,S(z)=\left\{\begin{array}[]{ll}T(z),&\mbox{for $z$ outside the lens,}\\ &\\ T(z)\left(\begin{array}[]{cc}1&0\\ -e^{2n\phi(z)}&1\\ \end{array}\right),&\mbox{for $z$ in the upper lens,}\\ &\\ T(z)\left(\begin{array}[]{cc}1&0\\ e^{2n\phi(z)}&1\\ \end{array}\right),&\mbox{for $z$ in the lower lens,}\end{array}\right. (3.6)

where the regions are illustrated in Fig.1. Then SS satisfies the jump condition

S+(z)=S(z)JS(z).S_{+}(z)=S_{-}(z)J_{S}(z). (3.7)
Refer to caption
Figure 1: The jump contours and regions for the RH problem for SS when λ1>1\lambda_{1}>1.

For λ1>1\lambda_{1}>1, we have

JS(z)={(1ω2e2nϕ(z)01),z(λ2,+),(1ω1e2nϕ(z)01),z(λ1,λ2),(0e2nϕ(z)e2nϕ(z)0),z(1,λ1),(0110),z(1,1)(10e2nϕ(z)1),zon lens,(1e2nϕ+(z)01),z(,1),J_{S}(z)=\left\{\begin{array}[]{ll}\left(\begin{array}[]{cc}1&\omega_{2}e^{-2n\phi(z)}\\ 0&1\\ \end{array}\right),&z\in(\lambda_{2},+\infty),\\[11.38092pt] \left(\begin{array}[]{cc}1&\omega_{1}e^{-2n\phi(z)}\\ 0&1\\ \end{array}\right),&z\in(\lambda_{1},\lambda_{2}),\\[11.38092pt] \left(\begin{array}[]{cc}0&e^{-2n\phi(z)}\\ -e^{2n\phi(z)}&0\\ \end{array}\right),&z\in(1,\lambda_{1}),\\[11.38092pt] \left(\begin{array}[]{cc}0&1\\ -1&0\\ \end{array}\right),&z\in(-1,1)\\ \left(\begin{array}[]{cc}1&0\\ e^{2n\phi(z)}&1\\ \end{array}\right),&z~{}\mbox{on lens},\\[11.38092pt] \left(\begin{array}[]{cc}1&e^{-2n\phi_{+}(z)}\\ 0&1\\ \end{array}\right),&z\in(-\infty,-1),\end{array}\right. (3.8)

where the contours are indicated in Fig. 1.
For λ1<1<λ2\lambda_{1}<1<\lambda_{2}, we have

JS(z)={(1ω2e2nϕ(z)01),z(λ2,+),(1ω1e2nϕ(z)01),z(1,λ2),(e2nϕ+(z)ω10e2nϕ(z)),z(λ1,1),(0110),z(1,λ1),(10e2nϕ(z)1),zon lens,(1e2nϕ+(z)01),z(,1).J_{S}(z)=\left\{\begin{array}[]{ll}\left(\begin{array}[]{cc}1&\omega_{2}e^{-2n\phi(z)}\\ 0&1\\ \end{array}\right),&z\in(\lambda_{2},+\infty),\\[11.38092pt] \left(\begin{array}[]{cc}1&\omega_{1}e^{-2n\phi(z)}\\ 0&1\\ \end{array}\right),&z\in(1,\lambda_{2}),\\[11.38092pt] \left(\begin{array}[]{cc}e^{2n\phi_{+}(z)}&\omega_{1}\\ 0&e^{2n\phi_{-}(z)}\\ \end{array}\right),&z\in(\lambda_{1},1),\\[11.38092pt] \left(\begin{array}[]{cc}0&1\\ -1&0\\ \end{array}\right),&z\in(-1,\lambda_{1}),\\[11.38092pt] \left(\begin{array}[]{cc}1&0\\ e^{2n\phi(z)}&1\\ \end{array}\right),&z~{}\mbox{on lens},\\[11.38092pt] \left(\begin{array}[]{cc}1&e^{-2n\phi_{+}(z)}\\ 0&1\\ \end{array}\right),&z\in(-\infty,-1).\end{array}\right.

For λ1<λ2<1\lambda_{1}<\lambda_{2}<1, we have

JS(z)={(1ω2e2nϕ(z)01),z(1,+),(e2nϕ+(z)ω20e2nϕ(z)),z(λ2,1),(e2nϕ+(z)ω10e2nϕ(z)),z(λ1,λ2),(0110),z(1,λ1),(10e2nϕ(z)1),zon lens,(1e2nϕ+(z)01),z(,1).J_{S}(z)=\left\{\begin{array}[]{ll}\left(\begin{array}[]{cc}1&\omega_{2}e^{-2n\phi(z)}\\ 0&1\\ \end{array}\right),&z\in(1,+\infty),\\[11.38092pt] \left(\begin{array}[]{cc}e^{2n\phi_{+}(z)}&\omega_{2}\\ 0&e^{2n\phi_{-}(z)}\\ \end{array}\right),&z\in(\lambda_{2},1),\\[11.38092pt] \left(\begin{array}[]{cc}e^{2n\phi_{+}(z)}&\omega_{1}\\ 0&e^{2n\phi_{-}(z)}\\ \end{array}\right),&z\in(\lambda_{1},\lambda_{2}),\\[11.38092pt] \left(\begin{array}[]{cc}0&1\\ -1&0\\ \end{array}\right),&z\in(-1,\lambda_{1}),\\[11.38092pt] \left(\begin{array}[]{cc}1&0\\ e^{2n\phi(z)}&1\\ \end{array}\right),&z~{}\mbox{on lens},\\[11.38092pt] \left(\begin{array}[]{cc}1&e^{-2n\phi_{+}(z)}\\ 0&1\\ \end{array}\right),&z\in(-\infty,-1).\end{array}\right.

3.3 Global Parametrix

The global parametrix solves the following approximating RH problem, with the jump along (1,λ1)(-1,\lambda_{1}):

(a)   N(z)N(z) is analytic in \[1,λ1]\mathbb{C}\backslash[-1,\lambda_{1}];

(b)

N+(x)=N(x)(0110),x(1,λ1);N_{+}(x)=N_{-}(x)\left(\begin{array}[]{cc}0&1\\ -1&0\\ \end{array}\right),\quad x\in(-1,\lambda_{1}); (3.9)

(c)

N(z)=I+O(z1),z.N(z)=I+O(z^{-1}),\quad z\rightarrow\infty. (3.10)

The solution of the RH problem is constructed explicitly ( see [26] ):

N(z)=(η(z)+η1(z)2η(z)η1(z)2iη(z)η1(z)2iη(z)+η1(z)2),η(z)=(zλ1z+1)1/4,N(z)=\left(\begin{array}[]{cc}\frac{\eta(z)+\eta^{-1}(z)}{2}&\frac{\eta(z)-\eta^{-1}(z)}{2i}\\ -\frac{\eta(z)-\eta^{-1}(z)}{2i}&\frac{\eta(z)+\eta^{-1}(z)}{2}\\ \end{array}\right),\quad\eta(z)=\left(\frac{z-\lambda_{1}}{z+1}\right)^{1/4}, (3.11)

where the branch is chosen such that η(z)\eta(z) is analytic in [1,λ1]\mathbb{C}\setminus[-1,\lambda_{1}], and η(z)1\eta(z)\sim 1 as zz\to\infty.

3.4 Local parametrix near z=1z=1

The jump matrices for S(z)S(z) are not close to the identity matrix near the node points z=±1z=\pm 1. Thus, local parametrices have to be constructed in the neighborhoods of z=±1z=\pm 1. Near z=1z=-1, the parametrix P(1)(z)P^{(-1)}(z) can be constructed in terms of the Airy function [9, 11]. We proceed to find a local parametrix P(1)(z)P^{(1)}(z) in U(1,r)U(1,r), which is an open disc centered at z=1z=1 with radius r>0r>0. The parametrix solves the following RH problem:

Riemann-Hilbert problem for P(1)P^{(1)}

(a)   P(1)(z)P^{(1)}(z) is analytic in U(1,r)\ΣSU(1,r)\backslash\Sigma_{S};

(b)   On ΣSU(1,r)\Sigma_{S}\cap U(1,r), P(1)(z)P^{(1)}(z) satisfies the same jump condition as S(z)S(z),

P+(1)(z)=P(1)(z)JS,zΣSU(1,r);P^{(1)}_{+}(z)=P^{(1)}_{-}(z)J_{S},~{}~{}z\in\Sigma_{S}\cap U(1,r); (3.12)

(c)   P(1)(z)P^{(1)}(z) satisfies the following matching condition on U(1,r)\partial U(1,r):

P(1)(z)N1(z)=I+O(n1/3);P^{(1)}(z)N^{-1}(z)=I+O\left(n^{-1/3}\right); (3.13)

(d)   The behavior of P(1)(z)=O(ln(zλk)P^{(1)}(z)=O(\ln(z-\lambda_{k}) as zλkz\to\lambda_{k} for k=1,2k=1,2.

To construct the local parametrix, we introduce the following model RH problem, which shares the same jump condition as P(1)(z)enϕ(z)σ3P^{(1)}(z)e^{-n\phi(z)\sigma_{3}}.

The Riemann-Hilbert problem for Ψ\Psi

0ssΣ4\Sigma_{4}Σ2\Sigma_{2}Σ3\Sigma_{3}Σ1\Sigma_{1}Σ0\Sigma_{0}IIVIIIII
Figure 2: The jump contours and regions for the RH problem for Ψ\Psi for s>0s>0.

(a)   Ψ(ζ;x,s)\Psi(\zeta;x,s) (Ψ(ζ)\Psi(\zeta), for short) is analytic in \j=04Σj\mathbb{C}\backslash\bigcup_{j=0}^{4}\Sigma_{j}, where the jump contours are indicated in Fig. 2;

(b)   Ψ(ζ)\Psi(\zeta) satisfies the jump condition for s>0s>0

Ψ+(ζ)=Ψ(ζ){(1ω201),ζ(s,+),(1ω101),ζ(0,s),(1011),ζΣ2,(0110),ζΣ3,(1011),ζΣ4;\displaystyle\Psi_{+}(\zeta)=\Psi_{-}(\zeta)\left\{\begin{array}[]{ll}\left(\begin{array}[]{cc}1&\omega_{2}\\ 0&1\\ \end{array}\right),&\zeta\in(s,+\infty),\\[11.38092pt] \left(\begin{array}[]{cc}1&\omega_{1}\\ 0&1\\ \end{array}\right),&\zeta\in(0,s),\\[11.38092pt] \left(\begin{array}[]{cc}1&0\\ 1&1\\ \end{array}\right),&\zeta\in{\Sigma}_{2},\\[11.38092pt] \left(\begin{array}[]{cc}0&1\\ -1&0\\ \end{array}\right),&\zeta\in{\Sigma}_{3},\\[11.38092pt] \left(\begin{array}[]{cc}1&0\\ 1&1\\ \end{array}\right),&\zeta\in\Sigma_{4};\end{array}\right. (3.29)

(c)   As ζ\zeta\rightarrow\infty,

Ψ(ζ)=(10ir(x,s)1)[I+Ψ1(x,s)ζ+O(ζ2)]ζ14σ3I+iσ12e(23ζ3/2+xζ1/2)σ3,\Psi(\zeta)=\left(\begin{array}[]{cc}1&0\\ ir(x,s)&1\end{array}\right)\left[I+\frac{\Psi_{1}(x,s)}{\zeta}+O\left(\zeta^{-2}\right)\right]\zeta^{-\frac{1}{4}\sigma_{3}}\frac{I+i\sigma_{1}}{\sqrt{2}}e^{-(\frac{2}{3}\zeta^{3/2}+x\zeta^{1/2})\sigma_{3}}, (3.30)

where r(x,s)=i(Ψ1(x,s))12r(x,s)=i(\Psi_{1}(x,s))_{12}.

(d)   As ζ0\zeta\rightarrow 0,

Ψ(ζ)=Ψ(0)(ζ)(I+1ω12πi(0100)lnζ)E,\Psi(\zeta)=\Psi^{(0)}(\zeta)\left(I+\frac{1-\omega_{1}}{2\pi i}\left(\begin{array}[]{cc}0&1\\ 0&0\end{array}\right)\ln\zeta\right)E, (3.31)

where Ψ(0)(ζ)\Psi^{(0)}(\zeta) is analytic at ζ=0\zeta=0 with the expansion

Ψ(0)(ζ)=P0^(x,s)(I+P^1(x,s)ζ+O(ζ2)).\Psi^{(0)}(\zeta)=\hat{P_{0}}(x,s)(I+\hat{P}_{1}(x,s)\zeta+O(\zeta^{2})). (3.32)

And the piecewise constant matrix

E={(1001),ζΩ1,(1011),ζΩ2,(1ω1ω111),ζΩ3,(1ω101),ζΩ4.{E}=\left\{\begin{array}[]{ll}\left(\begin{array}[]{cc}1&0\\ 0&1\end{array}\right),&\zeta\in\Omega_{1},\\ \left(\begin{array}[]{cc}1&0\\ -1&1\end{array}\right),&\zeta\in\Omega_{2},\\ \left(\begin{array}[]{cc}1-\omega_{1}&-\omega_{1}\\ 1&1\end{array}\right),&\zeta\in\Omega_{3},\\ \left(\begin{array}[]{cc}1&-\omega_{1}\\ 0&1\end{array}\right),&\zeta\in\Omega_{4}.\end{array}\right.

(e)   As ζs\zeta\rightarrow s,

Ψ(ζ)=Ψ(1)(ζ)(I+ω1ω22πi(0100)ln(ζs))E^,\Psi(\zeta)=\Psi^{(1)}(\zeta)\left(I+\frac{\omega_{1}-\omega_{2}}{2\pi i}\left(\begin{array}[]{cc}0&1\\ 0&0\end{array}\right)\ln(\zeta-s)\right)\widehat{E}, (3.33)

where Ψ(1)(ζ)\Psi^{(1)}(\zeta) is analytic at ζ=s\zeta=s with the following expansion

Ψ(1)(ζ)=Q^0(x,s)(I+Q^1(x,s)(ζs)+O((ζs)2)).\Psi^{(1)}(\zeta)=\hat{Q}_{0}(x,s)(I+\hat{Q}_{1}(x,s)(\zeta-s)+O((\zeta-s)^{2})). (3.34)

Here, the piecewise constant matrix

E^={(1001),Imζ>0,(1ω201),Imζ<0.\widehat{E}=\left\{\begin{array}[]{ll}\left(\begin{array}[]{cc}1&0\\ 0&1\end{array}\right),&\mathop{\rm Im}\nolimits\zeta>0,\\ \left(\begin{array}[]{cc}1&-\omega_{2}\\ 0&1\end{array}\right),&\mathop{\rm Im}\nolimits\zeta<0.\end{array}\right.

The RH problem for Ψ\Psi appears recently in the studies of the Fredholm determinants of Painlevé II kernel and Painlevé XXXIV kernel in [27] by the second author of the present work and Dai. It also arises in the studies of the determinants of the Airy kernel with several discontinuities in [8] by Claeys and Doeraene, when the number of discontinuities therein equals to two. The existence of solution to the RH problem for Ψ\Psi is proved. It is also shown that Ψ(ζ;x,s)\Psi(\zeta;x,s) satisfies the following Lax pair

Ψζ(ζ;x,s)=(v1x2ζ+v2x2(ζs)iiv1ζiv2ζsi(ζ+x+v1+v2+v1x24v1ζ+v2x24v2(ζs))v1x2ζv2x2(ζs))Ψ(ζ;x,s),\displaystyle\Psi_{\zeta}(\zeta;x,s)=\left(\begin{array}[]{cc}\frac{v_{1x}}{2\zeta}+\frac{v_{2x}}{2(\zeta-s)}&i-\frac{iv_{1}}{\zeta}-\frac{iv_{2}}{\zeta-s}\\ -i\left(\zeta+x+v_{1}+v_{2}+\frac{v_{1x}^{2}}{4v_{1}\zeta}+\frac{v_{2x}^{2}}{4v_{2}(\zeta-s)}\right)&-\frac{v_{1x}}{2\zeta}-\frac{v_{2x}}{2(\zeta-s)}\end{array}\right)\Psi(\zeta;x,s), (3.37)
Ψx(ζ;x,s)=(0iiζ2i(v1+v2+x2)0)Ψ(ζ;x,s).\displaystyle\Psi_{x}(\zeta;x,s)=\left(\begin{array}[]{cc}0&i\\ -i\zeta-2i(v_{1}+v_{2}+\tfrac{x}{2})&0\end{array}\right)\Psi(\zeta;x,s). (3.40)

The compatibility condition of the Lax pair is described by the coupled Painlevé II system (1.31). Moreover, the Hamiltonian (1.32) is related to the coefficient of 1/ζ1/\zeta in the large-ζ\zeta expansion of Ψ(ζ)\Psi(\zeta) in (3.30) by

HII(x;s)=x24+r(x,s);H_{\texttt{II}}(x;s)=\frac{x^{2}}{4}+r(x,s); (3.41)

see [27, Equation (4.21)].

We introduce the conformal mapping

f(z)=(32ϕ(z))2/3=2(z1)+15(z1)2+O((z1)3),f(z)=\left(\frac{3}{2}\phi(z)\right)^{2/3}=2(z-1)+\frac{1}{5}(z-1)^{2}+O\left((z-1)^{3}\right), (3.42)

from a neighborhood of z=1z=1 to that of the origin. Then the local parametrix P(1)(z){P}^{(1)}(z) can be constructed for zU(1,r)z\in U(1,r) as follows

P(1)(z)=E(z)Ψ(n2/3(f(z)f(λ1));n2/3f(λ1),n2/3(f(λ2)f(λ1)))enϕ(z)σ3,P^{(1)}(z)=E(z)\Psi\left(n^{2/3}(f(z)-f(\lambda_{1}));n^{2/3}f(\lambda_{1}),n^{2/3}(f(\lambda_{2})-f(\lambda_{1}))\right)e^{n\phi(z)\sigma_{3}}, (3.43)

and the pre-factor

E(z)=N(z)12(Iiσ1)[n2/3(f(z)f(λ1))]σ3/4(10iHII(n2/3f(λ1);n2/3(f(λ2)f(λ1)))1),E(z)=N(z)\frac{1}{\sqrt{2}}(I-i\sigma_{1})\left[n^{2/3}(f(z)-f(\lambda_{1}))\right]^{\sigma_{3}/4}\left(\begin{array}[]{cc}1&0\\ -iH_{\texttt{II}}(n^{2/3}f(\lambda_{1});n^{2/3}(f(\lambda_{2})-f(\lambda_{1})))&1\end{array}\right), (3.44)

where HII(x;t)H_{\texttt{II}}(x;t) is the Hamiltonian given in (1.32) and related to the coefficient of 1/ζ1/\zeta in the large-ζ\zeta expansion of Ψ(ζ)\Psi(\zeta) by (3.41).

Proposition 5.

For

λk=sk2n=1+tk2n2/3\lambda_{k}=\frac{s_{k}}{\sqrt{2n}}=1+\frac{t_{k}}{2n^{2/3}}

with bounded real parameters tkt_{k}, k=1,2k=1,2, the local parametrix defined in (3.43) and (3.44) solves the RH problem for P(1)P^{(1)}. Moreover, we have the expansion for zU(1,r)z\in\partial U(1,r):

P(1)(z)N(z)1=I+Δ(z)n1/3+O(n2/3),P^{(1)}(z)N(z)^{-1}=I+\frac{\Delta(z)}{n^{1/3}}+O(n^{-2/3}), (3.45)

where

Δ(z)=HII(n2/3f(λ1);n2/3(f(λ2)f(λ1)))2(f(z)f(λ1))1/2N(z)(σ3iσ1)N1(z)=O(1),\Delta(z)=\frac{H_{\texttt{II}}(n^{2/3}f(\lambda_{1});n^{2/3}(f(\lambda_{2})-f(\lambda_{1})))}{2(f(z)-f(\lambda_{1}))^{1/2}}N(z)(\sigma_{3}-i\sigma_{1})N^{-1}(z)=O(1), (3.46)

and HII(x,t)H_{\texttt{II}}(x,t) is the Hamiltonian defined in (1.32).

Proof.

Taking the principle branch for the fractional power, it follows from (3.42) that

(f(x)f(λ1))+14=(f(x)f(λ1))14eπ2i,x<λ1.(f(x)-f(\lambda_{1}))^{\frac{1}{4}}_{+}=(f(x)-f(\lambda_{1}))^{\frac{1}{4}}_{-}~{}e^{\frac{\pi}{2}i},\quad x<\lambda_{1}.

This, together with the expression of N(z)N(z) in (3.11), implies that E(z)E(z) is analytic for z in U(1,r)U(1,r). Recalling the properties of Ψ(ζ)\Psi(\zeta) in (3.29)-(3.33), it is then seen that the jump condition and the local behaviors near λk,k=1,2\lambda_{k},k=1,2, in the Riemann-Hilbert problem for P(1)P^{(1)}, are fulfilled.

We then proceed to check the matching condition (3.13). Substituting the large-ζ\zeta behavior of Ψ(ζ)\Psi(\zeta) (3.30) into (3.43) leads us to the expansion for zz on U(1,r)\partial U(1,r) as nn\to\infty:

P(1)(z)N(z)1\displaystyle P^{(1)}(z)N(z)^{-1} =N(z)12(Iiσ1)[n2/3(f(z)f(λ1))]σ3/4(10i4(n2/3f(λ1))21)\displaystyle=N(z)\frac{1}{\sqrt{2}}(I-i\sigma_{1})\left[n^{2/3}(f(z)-f(\lambda_{1}))\right]^{\sigma_{3}/4}\left(\begin{array}[]{cc}1&0\\ -\frac{i}{4}(n^{2/3}f(\lambda_{1}))^{2}&1\end{array}\right) (3.49)
(I+Ψ1(n2/3f(λ1),n2/3(f(λ2)f(λ1)))n2/3(f(z)f(λ1))+O(n4/3))\displaystyle~{}~{}\left(I+\frac{\Psi_{1}(n^{2/3}f(\lambda_{1}),n^{2/3}(f(\lambda_{2})-f(\lambda_{1})))}{n^{2/3}(f(z)-f(\lambda_{1}))}+O(n^{-4/3})\right)
[n2/3(f(z)f(λ1))]σ3/412(I+iσ1)enρ(z;λ1)σ3N(z)1,\displaystyle~{}~{}\left[n^{2/3}(f(z)-f(\lambda_{1}))\right]^{-\sigma_{3}/4}\frac{1}{\sqrt{2}}(I+i\sigma_{1})e^{n\rho(z;\lambda_{1})\sigma_{3}}N(z)^{-1}, (3.50)

where

ρ(z;λ1)\displaystyle\rho(z;\lambda_{1}) =23f(z)3/223(f(z)f(λ1))3/2f(λ1)(f(z)f(λ1))1/2\displaystyle=\frac{2}{3}f(z)^{3/2}-\frac{2}{3}(f(z)-f(\lambda_{1}))^{3/2}-f(\lambda_{1})(f(z)-f(\lambda_{1}))^{1/2}
=1(f(z)f(λ1))1/2(23f(z)2(1f(λ1)f(z))1/223f(z)2(1f(λ1)f(z))2\displaystyle=\frac{1}{\left(f(z)-f(\lambda_{1})\right)^{1/2}}\left(\frac{2}{3}f(z)^{2}\left(1-\frac{f(\lambda_{1})}{f(z)}\right)^{1/2}-\frac{2}{3}f(z)^{2}\left(1-\frac{f(\lambda_{1})}{f(z)}\right)^{2}\right.
f(z)f(λ1)(1f(λ1)f(z)))\displaystyle~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}\left.-f(z)f(\lambda_{1})\left(1-\frac{f(\lambda_{1})}{f(z)}\right)\right)
=f(λ1)24(f(z)f(λ1))1/2(1+O(f(λ1)).\displaystyle=\frac{f(\lambda_{1})^{2}}{4\left(f(z)-f(\lambda_{1})\right)^{1/2}}(1+O(f(\lambda_{1})). (3.51)

Here f(λ1)2(λ11)f(\lambda_{1})\sim 2(\lambda_{1}-1) as λ11\lambda_{1}\to 1. Inserting the definition of Ψ1\Psi_{1} in (3.30) and (3) into (3.49), we obtain (3.45). For zU(1,r)z\in\partial U(1,r), the denominator in (3.46), namely f(z)f(λ1)f(z)-f(\lambda_{1}), is bounded away from zero. It follows from Proposition 1 and (1.36), H(x;s)H(x;s) is analytic for real variables xx and ss and thus bounded. Therefore, the factor Δ\Delta defined in (3.46) is bounded for zU(1,r)z\in\partial U(1,r). Thus, we obtain the matching condition (3.13) and complete the proof of Proposition 5. ∎

Remark 4.

The estimate in (3.46) and thus the matching condition (3.13) can be established for more general parameters:

λk=sk2n=1+tk2n2/3\lambda_{k}=\frac{s_{k}}{\sqrt{2n}}=1+\frac{t_{k}}{2n^{2/3}}

where

c1t1<t2c2n1/6,t2t1c3,-c_{1}\leqslant t_{1}<t_{2}\leqslant c_{2}n^{1/6},\qquad t_{2}-t_{1}\leqslant c_{3}, (3.52)

for any given positive constants ckc_{k} , k=1,2,3k=1,2,3. Actually, for such parameters, we have

n2/3(f(λ2)f(λ1))=t2t1+O(n1/3).n^{2/3}(f(\lambda_{2})-f(\lambda_{1}))=t_{2}-t_{1}+O(n^{-1/3}).

In view of the asymptotic behavior (1.37) and the relation (1.36), we know that HII(x;s)H_{\texttt{II}}(x;s) is exponentially small for bounded ss and large positive xx. Therefore, we have the estimate (3.46) for dt1c2n1/6d\leqslant t_{1}\leqslant c_{2}n^{1/6} with a certain big enough constant dd. This, together with the estimate (3.46) derived before for c1t1d-c_{1}\leqslant t_{1}\leqslant d, leads us to the claim.

3.5 The final transformation: SRS\rightarrow R

The final transformation is defined by

R(z)={S(z)N1(z),z\{U(1,r)U(1,r)ΣS},S(z){P(1)(z)}1,zU(1,r)\ΣS,S(z){P(1)(z)}1,zU(1,r)\ΣS.R(z)=\left\{\begin{array}[]{ll}S(z)N^{-1}(z),&z\in\mathbb{C}\backslash\left\{U(-1,r)\cup U(1,r)\cup\Sigma_{S}\right\},\\ S(z)\left\{P^{(-1)}(z)\right\}^{-1},&z\in U(-1,r)\backslash\Sigma_{S},\\ S(z)\left\{P^{(1)}(z)\right\}^{-1},&z\in U(1,r)\backslash\Sigma_{S}.\end{array}\right. (3.53)

From the matching condition (3.13), we have

JR(z)IL2L(ΣR)=O(n1/3),\|J_{R}(z)-I\|_{L^{2}\cap L^{\infty}(\Sigma_{R})}=O(n^{-1/3}), (3.54)

where the error bound is uniform for the parameters t1t_{1} and t2t_{2} specified by (3.52). Thus, by a standard argument as given in [9, 10, 11], we have the estimate

R(z)=I+O(n1/3),R(z)=I+O(n^{-1/3}), (3.55)

where the error bound is uniform for zz in whole complex plane.

4 Proofs of Theorem 2-4

In this section, we will prove the main results on the asymptotics of the Hankel determinants and several quantities related to the orthogonal polynomials, including the recurrence coefficients and the leading coefficients. Moreover, we will derive the asymptotics of the coupled Painlevé IV system.

4.1 Proof of Theorem 2: asymptotic of the Hankel determinants

Lemma 1.

Let

sk=2n+tk2n1/6,k=1,2,s_{k}=\sqrt{2n}+\frac{t_{k}}{2n^{1/6}},\quad k=1,2,

and F(s1,s2)F(s_{1},s_{2}) be the logarithmic derivative of the Hankel determinant defined in (1.20), we have

F(s1,s2)=2n1/6HII(t1;t2t1)+O(n1/6),F(s_{1},s_{2})=\sqrt{2}n^{1/6}H_{\texttt{II}}(t_{1};t_{2}-t_{1})+O(n^{-1/6}), (4.1)

where HII(x;s)H_{\texttt{II}}(x;s) is the Hamiltonian for the coupled Painlevé II system as defined in (1.32). The error bound is uniform for c1t1<t2c2n1/6-c_{1}\leqslant t_{1}<t_{2}\leqslant c_{2}n^{1/6} and t2t1c3t_{2}-t_{1}\leqslant c_{3} for any given positive constants ckc_{k} , k=1,2,3k=1,2,3; see also (3.52).

Proof.

Tracing back the series of invertible transformations (3.1), (3.6) and (3.53)

YTSR,Y\to T\to S\to R,

we have

Y+(2nz)=(2n)12nσ3e12nlσ3R(z)E(z)Ψ+(n2/3(f(z)f(λ1));xn,sn)enz2σ3,λ1<z<1+r,Y_{+}(\sqrt{2n}z)=(2n)^{\frac{1}{2}n\sigma_{3}}e^{\frac{1}{2}nl\sigma_{3}}R(z)E(z)\Psi_{+}\left(n^{2/3}(f(z)-f(\lambda_{1}));x_{n},s_{n}\right)e^{nz^{2}\sigma_{3}},\quad\lambda_{1}<z<1+r, (4.2)

where E(z)E(z) as defined in (3.44) is analytic for |z1|<r|z-1|<r. With the parameters specified by (3.52), we have

xn=n2/3f(λ1)=t1+O(n1/3),sn=n2/3(f(λ2)f(λ1))=t2t1+O(n1/3).x_{n}=n^{2/3}f(\lambda_{1})=t_{1}+O(n^{-1/3}),\quad s_{n}=n^{2/3}(f(\lambda_{2})-f(\lambda_{1}))=t_{2}-t_{1}+O(n^{-1/3}). (4.3)

Thus, substituting (4.2) and the estimate (3.55) into the differential identity (2.3), we obtain

F(s1,s2)=1ω12πin1/6(Ψ1Ψζ)21(0)+ω1ω22πin1/6(Ψ1Ψζ)21(sn)+O(n1/6),\displaystyle F(s_{1},s_{2})=\frac{1-\omega_{1}}{\sqrt{2}\pi i}n^{1/6}\left(\Psi^{-1}\Psi_{\zeta}\right)_{21}(0)+\frac{\omega_{1}-\omega_{2}}{\sqrt{2}\pi i}n^{1/6}\left(\Psi^{-1}\Psi_{\zeta}\right)_{21}(s_{n})+O(n^{-1/6}), (4.4)

where the error bound is uniform for s1s_{1} and s2s_{2} specified by (3.52). Using the expansions of Ψ(z)\Psi(z) near z=0z=0 and z=sz=s in (3.31) and (3.33), we have

F(s1,s2)=1ω12πin1/6(P^1)21(xn,sn)+ω1ω22πin1/6(Q^1)21(xn,sn)+O(n1/6),\displaystyle F(s_{1},s_{2})=\frac{1-\omega_{1}}{\sqrt{2}\pi i}n^{1/6}(\hat{P}_{1})_{21}(x_{n},s_{n})+\frac{\omega_{1}-\omega_{2}}{\sqrt{2}\pi i}n^{1/6}(\hat{Q}_{1})_{21}(x_{n},s_{n})+O(n^{-1/6}), (4.5)

where P^1\hat{P}_{1} and Q^1\hat{Q}_{1} are defined in (3.32) and (3.34), respectively.

Next, we express P^1\hat{P}_{1} and Q^1\hat{Q}_{1} in terms of the coupled Painlevé II system. Applying the differential equation (3.37), we obtain

1ω12πiP^0(0100)P^01=(v1x2iv1iv1x24v1v1x2),\displaystyle\frac{1-\omega_{1}}{2\pi i}\hat{P}_{0}\left(\begin{array}[]{cc}0&1\\ 0&0\end{array}\right)\hat{P}_{0}^{-1}=\left(\begin{array}[]{cc}\frac{v_{1x}}{2}&-iv_{1}\\ -i\frac{v_{1x}^{2}}{4v_{1}}&-\frac{v_{1x}}{2}\end{array}\right), (4.10)
P^1+1ω12πi[P^1,(0100)]=P^01(v2x2si+iv2si(x+v1+v2v2x24v2s)v2x2s)P^0.\displaystyle\hat{P}_{1}+\frac{1-\omega_{1}}{2\pi i}\left[\hat{P}_{1},\left(\begin{array}[]{cc}0&1\\ 0&0\end{array}\right)\right]=\hat{P}_{0}^{-1}\left(\begin{array}[]{cc}-\frac{v_{2x}}{2s}&i+\frac{iv_{2}}{s}\\ -i\left(x+v_{1}+v_{2}-\frac{v_{2x}^{2}}{4v_{2}s}\right)&\frac{v_{2x}}{2s}\end{array}\right)\hat{P}_{0}. (4.15)

Now

P^0(0100)P^01=((P^0)11(P^0)21(P^0)112(P^0)212(P^0)11(P^0)21).\hat{P}_{0}\left(\begin{array}[]{cc}0&1\\ 0&0\end{array}\right)\hat{P}_{0}^{-1}=\left(\begin{array}[]{cc}-(\hat{P}_{0})_{11}(\hat{P}_{0})_{21}&(\hat{P}_{0})_{11}^{2}\\ -(\hat{P}_{0})_{21}^{2}&(\hat{P}_{0})_{11}(\hat{P}_{0})_{21}\\ \end{array}\right).

Then, this, together with (4.10), leads us to

1ω12πi(P^0)112=iv1,1ω12πi(P^0)212=iv1x24v1,1ω12πi(P^0)11(P^0)21=v1x2.\frac{1-\omega_{1}}{2\pi i}(\hat{P}_{0})_{11}^{2}=-iv_{1},\quad\frac{1-\omega_{1}}{2\pi i}(\hat{P}_{0})_{21}^{2}=i\frac{v_{1x}^{2}}{4v_{1}},\quad\frac{1-\omega_{1}}{2\pi i}(\hat{P}_{0})_{11}(\hat{P}_{0})_{21}=-\frac{v_{1x}}{2}. (4.16)

From the (21)(21) entry of the matrix equation (4.15), it is seen that

(P^1)21=i(x+v1+v2v2x24v2s)(P0^)112+v2xs(P^0)11(P^0)21+i(1+v2s)(P0^)212.(\hat{P}_{1})_{21}=-i\left(x+v_{1}+v_{2}-\frac{v_{2x}^{2}}{4v_{2}s}\right)(\hat{P_{0}})_{11}^{2}+\frac{v_{2x}}{s}(\hat{P}_{0})_{11}(\hat{P}_{0})_{21}+i\left(1+\frac{v_{2}}{s}\right)(\hat{P_{0}})_{21}^{2}. (4.17)

Substituting (4.16) into (4.17), we obtain

1ω1πi(P^1)21=2v1(x+v1+v2v2x24v2s)v1xv2xs+(1+v2s)v1x22v1.\frac{1-\omega_{1}}{\pi i}(\hat{P}_{1})_{21}=-2v_{1}\left(x+v_{1}+v_{2}-\frac{v_{2x}^{2}}{4v_{2}s}\right)-\frac{v_{1x}v_{2x}}{s}+\left(1+\frac{v_{2}}{s}\right)\frac{v_{1x}^{2}}{2v_{1}}. (4.18)

Similarly, we have that

ω1ω22πi(Q^0)112=iv2,ω1ω22πi(Q^0)212=iv2x24v2,\frac{\omega_{1}-\omega_{2}}{2\pi i}(\hat{Q}_{0})_{11}^{2}=-iv_{2},\quad\frac{\omega_{1}-\omega_{2}}{2\pi i}(\hat{Q}_{0})_{21}^{2}=i\frac{v_{2x}^{2}}{4v_{2}}, (4.19)

and

ω1ω2πi(Q^1)21=2v2(xs+v1+v2+v1x24v1s)+v1xv2xs(1v1s)v2x22v2.\frac{\omega_{1}-\omega_{2}}{\pi i}(\hat{Q}_{1})_{21}=-2v_{2}\left(x-s+v_{1}+v_{2}+\frac{v_{1x}^{2}}{4v_{1}s}\right)+\frac{v_{1x}v_{2x}}{s}-\left(1-\frac{v_{1}}{s}\right)\frac{v_{2x}^{2}}{2v_{2}}. (4.20)

Therefore, we obtain from (1.31), (4.3) (4.5), (4.18) and (4.20) that

F(s1,s2)\displaystyle F(s_{1},s_{2}) =12n1/6(2v2sn2(v1+v2)(xn+v1+v2)+v1x22v1+v2x22v2)+O(n1/6)\displaystyle=\frac{1}{\sqrt{2}}n^{1/6}\left(-2v_{2}s_{n}-2(v_{1}+v_{2})(x_{n}+v_{1}+v_{2})+\frac{v_{1x}^{2}}{2v_{1}}+\frac{v_{2x}^{2}}{2v_{2}}\right)+O(n^{-1/6}) (4.21)
=2n1/6HII(xn;sn)+O(n1/6)\displaystyle=\sqrt{2}n^{1/6}H_{\texttt{II}}(x_{n};s_{n})+O(n^{-1/6}) (4.22)
=2n1/6HII(t1;t2t1)+O(n1/6),\displaystyle=\sqrt{2}n^{1/6}H_{\texttt{II}}(t_{1};t_{2}-t_{1})+O(n^{-1/6}), (4.23)

where HII(x;s)H_{\texttt{II}}(x;s) is the Hamiltonian for the coupled Painlevé II system as defined in (1.32). This completes the proof Lemma 1. ∎

Next we derive the asymptotic expansion for the Hankel determinant DnD_{n} defined by (1) when the jump discontinuities of the weight function (1.5) are large enough.

Lemma 2.

For s2>s12n+c0s_{2}>s_{1}\geqslant\sqrt{2n}+c_{0} and any given positive constant c0c_{0}, we have the asymptotic approximation for the Hankel determinant Dn(s1,s2)=Dn(s1,s2;ω1,ω2)D_{n}(s_{1},s_{2})=D_{n}(s_{1},s_{2};\omega_{1},\omega_{2}) defined by (1)

Dn(s1,s2)=DnGUE(1+O(ecn1/4)),D_{n}(s_{1},s_{2})=D_{n}^{\texttt{GUE}}\left(1+O\left(e^{-cn^{1/4}}\right)\right), (4.24)

where cc is some positive constant and DnGUED_{n}^{\texttt{GUE}}, given explicitly in (1.6), is the Hankel determinant associated with the pure Gaussian weight.

Proof.

For s2>s12n+c0s_{2}>s_{1}\geqslant\sqrt{2n}+c_{0}, we have λ2>λ11+c02n>1\lambda_{2}>\lambda_{1}\geqslant 1+\frac{c_{0}}{\sqrt{2n}}>1. On account of (3.42), there is some constant c>0c>0 such that

nϕ(λ)>cn1/4n\phi(\lambda)>cn^{1/4}

for λ>λ1\lambda>\lambda_{1}. Thus, the jump matrices JS(z)J_{S}(z) defined in (3.8) tend to the identity matrix exponentially fast for z(λ1,λ2)(λ2,+)z\in(\lambda_{1},\lambda_{2})\cup(\lambda_{2},+\infty). Therefore, we have

S(z)=(I+O(e2nϕ(λ1))S0(z),S(z)=(I+O(e^{-2n\phi(\lambda_{1})})S_{0}(z), (4.25)

where S0(z)S_{0}(z) is solution to the RH problem for SS when the parameters ω1=ω2=1\omega_{1}=\omega_{2}=1 and λ1=1\lambda_{1}=1 therein. Tracing back the sequence of transformations YTSY\to T\to S, given in (3.1), (3.6), we have

Y(2nz)=(2n)12nσ3e12nlσ3(I+O(e2nϕ(λ1))S0(z)eng(z)σ312nlσ3,Y(\sqrt{2n}z)=(2n)^{\frac{1}{2}n\sigma_{3}}e^{\frac{1}{2}nl\sigma_{3}}(I+O(e^{-2n\phi(\lambda_{1})})S_{0}(z)e^{ng(z)\sigma_{3}-\frac{1}{2}nl\sigma_{3}}, (4.26)

where eng(z)σ3=(I+O(1z2))znσ3e^{ng(z)\sigma_{3}}=(I+O\left(\frac{1}{z^{2}}\right))z^{n\sigma_{3}}. In view of (2.2) and the differential identity (2.4), we obtain

F(s1,s2)=2pn+O(e2nϕ(λ1))=O(e2nϕ(λ1)),F(s_{1},s_{2})=2p_{n}+O\left(e^{-2n\phi(\lambda_{1})}\right)=O\left(e^{-2n\phi(\lambda_{1})}\right), (4.27)

where pn=0p_{n}=0 is the sub-leading coefficient of the monic Hermite polynomial of degree nn. We integrate on both sides of the above equation and obtain

lnDn(s1,s2)lnDn(s1+L,s2+L)=s1L+s1O(e2nϕ(λ1))𝑑τ.\ln D_{n}(s_{1},s_{2})-\ln D_{n}(s_{1}+L,s_{2}+L)=\int_{s_{1}}^{L+s_{1}}O\left(e^{-2n\phi(\lambda_{1})}\right)d\tau. (4.28)

Let L+L\to+\infty, we get (4.24) and complete the proof of Lemma 2. ∎

Now, we are ready to prove Theorem 2. Integrating on both sides of the equation (4.1), we obtain for some positive constant c0c_{0} that

lnDn(s1,s2)lnDn(s1+c0,s2+c0)=t1t0HII(τ;t2t1)𝑑τ+O(n1/6),\ln D_{n}(s_{1},s_{2})-\ln D_{n}(s_{1}+c_{0},s_{2}+c_{0})=-\int_{t_{1}}^{t_{0}}H_{\texttt{II}}(\tau;t_{2}-t_{1})d\tau+O(n^{-1/6}), (4.29)

where

s1=2n+t12n1/6,s2=2n+t22n1/6s_{1}=\sqrt{2n}+\frac{t_{1}}{\sqrt{2}n^{1/6}},\quad s_{2}=\sqrt{2n}+\frac{t_{2}}{\sqrt{2}n^{1/6}}

for tkt_{k}, k=1,2k=1,2, in any compact subset of \mathbb{R} and

t0=n2/3f(s1+c02n)2c0n1/6.t_{0}=n^{2/3}f(\frac{s_{1}+c_{0}}{\sqrt{2n}})\sim\sqrt{2}c_{0}n^{1/6}. (4.30)

In view of (4.24), we have

lnDn(s1+c0,s2+c0)=DnGUE(1+O(ecn1/4)),\ln D_{n}(s_{1}+c_{0},s_{2}+c_{0})=D_{n}^{\texttt{GUE}}\left(1+O\left(e^{-cn^{1/4}}\right)\right), (4.31)

where cc is some positive constant and DnGUED_{n}^{\texttt{GUE}} is the Hankel determinant associated with the Gaussian weight; see (1.6). Recalling (1.36), we obtain from an integration by parts that

t1t0HII(τ;t2t1)𝑑τ=t1t0(τt1)(u1(τ;t2t1)2+u2(τ;t2t1)2)𝑑τ,\int_{t_{1}}^{t_{0}}H_{\texttt{II}}(\tau;t_{2}-t_{1})d\tau=\int_{t_{1}}^{t_{0}}(\tau-t_{1})\left(u_{1}(\tau;t_{2}-t_{1})^{2}+u_{2}(\tau;t_{2}-t_{1})^{2}\right)d\tau, (4.32)

where u1(x)u_{1}(x) and u2(x)u_{2}(x) are solutions to the coupled nonlinear differential equations (1.34) subject to the boundary conditions (1.37) as x+x\to+\infty. On account of (1.37) and (4.30), we have

t0+(τt1)(u1(τ;t2t1)2+u22(τ;t2t1))𝑑τ=O(ecn1/4),\int_{t_{0}}^{+\infty}(\tau-t_{1})\left(u_{1}(\tau;t_{2}-t_{1})^{2}+u_{2}^{2}(\tau;t_{2}-t_{1})\right)d\tau=O\left(e^{-cn^{1/4}}\right), (4.33)

for some constant c>0c>0. Inserting (4.31) and (4.33) into (4.29), we obtain (1.38). This completes the proof of Theorem 2.

4.2 Proof of theorem 3: asymptotics of the coupled Painlevé IV

From this section, the parameters in (1.5) are defined by s1=2nλ1=2n+t12n1/6s_{1}=\sqrt{2n}\lambda_{1}=\sqrt{2n}+\frac{t_{1}}{\sqrt{2}n^{1/6}} and s2=2nλ2=2n+t22n1/6,s_{2}=\sqrt{2n}\lambda_{2}=\sqrt{2n}+\frac{t_{2}}{\sqrt{2}n^{1/6}}, with t1t_{1} and t2t_{2} in any compact subset of \mathbb{R}. Tracing back the sequence of transformations YTSRY\to T\to S\to R, given in (3.1), (3.6) and (3.53), we have the expression for large zz:

Y(2nz)=(2n)12nσ3e12nlσ3R(z)N(z)eng(z)σ312nlσ3,Y(\sqrt{2n}z)=(2n)^{\frac{1}{2}n\sigma_{3}}e^{\frac{1}{2}nl\sigma_{3}}R(z)N(z)e^{ng(z)\sigma_{3}-\frac{1}{2}nl\sigma_{3}}, (4.34)

where l=12ln2l=-1-2\ln 2. From the definition of g(z)g(z) in (3.2), it is seen that

eng(z)σ3znσ3=I+O(1z2),z.e^{ng(z)\sigma_{3}}z^{-n\sigma_{3}}=I+O\left(\frac{1}{z^{2}}\right),\quad z\to\infty.

By the expression of N(z)N(z) in (3.11), we have the expansion

N(z)=I+N1z+O(1z2),z,N(z)=I+\frac{N_{1}}{z}+O\left(\frac{1}{z^{2}}\right),\quad z\to\infty, (4.35)

where N1=14(1+λ1)σ2N_{1}=-\frac{1}{4}(1+\lambda_{1})\sigma_{2}. It follows from the expansion of the jump for R(z)R(z) in (3.45) and (3.46) that

R(z)=I+R(1)(z)n1/3+O(1n2/3),n.R(z)=I+\frac{R^{(1)}(z)}{n^{1/3}}+O\left(\frac{1}{n^{2/3}}\right),\quad n\rightarrow\infty. (4.36)

Here, R(1)R^{(1)} satisfy the jump relation

R+(1)(z)R(1)(z)=Δ(z),R^{(1)}_{+}(z)-R^{(1)}_{-}(z)=\Delta(z), (4.37)

where Δ(z)\Delta(z) is given in (3.46) and R(1)(z)=O(1/z)R^{(1)}(z)=O(1/z) for zz large. Applying Cauchy’s theorem, it is seen that

R(1)(z)={1+λ1HII(n2/3f(λ1);n2/3(f(λ2)f(λ1)))2f(λ1)(zλ1)(σ3iσ1)Δ(z),zU(1,r),1+λ1HII(n2/3f(λ1);n2/3(f(λ2)f(λ1)))2f(λ1)(zλ1)(σ3iσ1),zU(1,r)¯.R^{(1)}(z)=\left\{\begin{array}[]{ll}\frac{\sqrt{1+\lambda_{1}}H_{\texttt{II}}(n^{2/3}f(\lambda_{1});n^{2/3}(f(\lambda_{2})-f(\lambda_{1})))}{2\sqrt{f^{\prime}(\lambda_{1})}(z-\lambda_{1})}(\sigma_{3}-i\sigma_{1})-\Delta(z),&z\in U(1,r),\\[8.5359pt] \frac{\sqrt{1+\lambda_{1}}H_{\texttt{II}}(n^{2/3}f(\lambda_{1});n^{2/3}(f(\lambda_{2})-f(\lambda_{1})))}{2\sqrt{f^{\prime}(\lambda_{1})}(z-\lambda_{1})}(\sigma_{3}-i\sigma_{1}),&z\not\in\overline{U(1,r)}.\end{array}\right. (4.38)

Thus, we get from the expression the following expansion as zz\to\infty

R(z)=I+R1z+O(1z2),R(z)=I+\frac{R_{1}}{z}+O\left(\frac{1}{z^{2}}\right), (4.39)

and

R1\displaystyle R_{1} =1+λ12f(λ1)n1/3HII(n2/3f(λ1);n2/3(f(λ2)f(λ1)))(σ3iσ1)\displaystyle=\frac{\sqrt{1+\lambda_{1}}}{2\sqrt{f^{\prime}(\lambda_{1})}n^{1/3}}H_{\texttt{II}}(n^{2/3}f(\lambda_{1});n^{2/3}(f(\lambda_{2})-f(\lambda_{1})))(\sigma_{3}-i\sigma_{1})
=12n1/3HII(t1;t2t1)(σ3iσ1)+O(n2/3),\displaystyle=\frac{1}{2n^{1/3}}H_{\texttt{II}}(t_{1};t_{2}-t_{1})(\sigma_{3}-i\sigma_{1})+O(n^{-2/3}), (4.40)

where use is also made of (3.42). In view of (3.11), (3.42) and (3.46), we have the following expansion as zλ1z\to\lambda_{1}

R(1)(z)=110HII(t1;t2t1)(σ3iσ1)+O(n2/3)+O(zλ1).R^{(1)}(z)=-\frac{1}{10}H_{\texttt{II}}(t_{1};t_{2}-t_{1})(\sigma_{3}-i\sigma_{1})+O(n^{-2/3})+O(z-\lambda_{1}). (4.41)

Substituting (4.35), (4.39) and (4.2) into (4.34) yields

Y1\displaystyle Y_{1} =2n(2n)12nσ3e12nlσ3(R1+N1)e12nlσ3(2n)12nσ3\displaystyle=\sqrt{2n}(2n)^{\frac{1}{2}n\sigma_{3}}e^{\frac{1}{2}nl\sigma_{3}}(R_{1}+N_{1})e^{-\frac{1}{2}nl\sigma_{3}}(2n)^{-\frac{1}{2}n\sigma_{3}} (4.42)
=2n(2n)12nσ3e12nlσ3(12σ2+HII(t1;t2t1)2n1/3(σ3iσ1)+O(n2/3))e12nlσ3(2n)12nσ3.\displaystyle=\sqrt{2n}(2n)^{\frac{1}{2}n\sigma_{3}}e^{\frac{1}{2}nl\sigma_{3}}\left(-\frac{1}{2}\sigma_{2}+\frac{H_{\texttt{II}}(t_{1};t_{2}-t_{1})}{2n^{1/3}}(\sigma_{3}-i\sigma_{1})+O(n^{-2/3})\right)e^{-\frac{1}{2}nl\sigma_{3}}(2n)^{-\frac{1}{2}n\sigma_{3}}. (4.43)

This, together with the relation (2.29), implies that

y(x;s)\displaystyle y(x;s) =2(Y1)21ex2\displaystyle=-2(Y_{1})_{21}e^{-x^{2}}
=i(2n)n+12ex2nl(1+HII(t1;t2t1)n1/3+O(n2/3)),\displaystyle=i(2n)^{-n+\frac{1}{2}}e^{-x^{2}-nl}\left(1+\frac{H_{II}(t_{1};t_{2}-t_{1})}{n^{1/3}}+O(n^{-2/3})\right), (4.44)

where x=s1+s22=2n+t1+t222n1/6x=\frac{s_{1}+s_{2}}{2}=\sqrt{2n}+\frac{t_{1}+t_{2}}{2\sqrt{2}n^{1/6}} and s=s2s12=t2t122n1/6s=\frac{s_{2}-s_{1}}{2}=\frac{t_{2}-t_{1}}{2\sqrt{2}n^{1/6}}. Recalling l=12ln2l=-1-2\ln 2, we obtain the asymptotic expansion of y(x;s)y(x;s) as given in (1.47).

Next, we consider the asymptotics of ak(x;s)a_{k}(x;s) and ak(x;s)a_{k}(x;s) for k=1,2k=1,2. It follows from the mastar equation of the Lax pair (2.18) that

a1(x;s)=1y(x;s)limzs(z+s)(ΦzΦ1)12,\displaystyle a_{1}(x;s)=\frac{1}{y(x;s)}\lim_{z\rightarrow-s}(z+s)(\Phi_{z}\Phi^{-1})_{12}, (4.45)
a2(x;s)=1y(x;s)limzs(zs)(ΦzΦ1)12,\displaystyle a_{2}(x;s)=\frac{1}{y(x;s)}\lim_{z\rightarrow s}(z-s)(\Phi_{z}\Phi^{-1})_{12}, (4.46)
b1(x;s)=1a1(x;s)limzs(z+s)(ΦzΦ1)11,\displaystyle b_{1}(x;s)=\frac{1}{a_{1}(x;s)}\lim_{z\rightarrow-s}(z+s)(\Phi_{z}\Phi^{-1})_{11}, (4.47)
b2(x;s)=1a2(x;s)limzs(zs)(ΦzΦ1)11,\displaystyle b_{2}(x;s)=\frac{1}{a_{2}(x;s)}\lim_{z\rightarrow s}(z-s)(\Phi_{z}\Phi^{-1})_{11}, (4.48)

where the subscript zz denotes the derivative with respect to zz. It is seen from (2.12) that

ΦzΦ1=σ1ex22σ3[Yz(z+x)Y1(z+x)+Y(z+x)((z+x)σ3)Y1(z+x)]ex22σ3σ1.\Phi_{z}\Phi^{-1}=\sigma_{1}e^{\frac{x^{2}}{2}\sigma_{3}}[Y_{z}(z+x)Y^{-1}(z+x)+Y(z+x)(-(z+x)\sigma_{3})Y^{-1}(z+x)]e^{-\frac{x^{2}}{2}\sigma_{3}}\sigma_{1}. (4.49)

According to (4.45) and (4.49) and in view of the fact that Y(z)Y(z) has at most logarithm singularity at s1s_{1}, we have

a1(x;s)\displaystyle a_{1}(x;s) =1y(x;s)ex2limzs(z+s)(Yz(z+x)Y1(z+x))21,\displaystyle=\frac{1}{y(x;s)}e^{-x^{2}}\lim_{z\rightarrow-s}(z+s)(Y_{z}(z+x)Y^{-1}(z+x))_{21},
=1y(x;s)ex2limzs1(zs1)(Yz(z)Y1(z))21,\displaystyle=\frac{1}{y(x;s)}e^{-x^{2}}\lim_{z\rightarrow s_{1}}(z-s_{1})(Y_{z}(z)Y^{-1}(z))_{21},
=1y(x;s)ex2limzλ12n(zλ1)(Yz(2nz)Y1(2nz))21.\displaystyle=\frac{1}{y(x;s)}e^{-x^{2}}\lim_{z\rightarrow\lambda_{1}}\sqrt{2n}(z-\lambda_{1})(Y_{z}(\sqrt{2n}z)Y^{-1}(\sqrt{2n}z))_{21}. (4.50)

Inserting (4.2) into (4.2), it is seen that

a1(x;s)=(2n)ny(x;s)ex2nl(R(λ1)E(λ1)(limζ0ζΨζ(ζ)Ψ1(ζ))E1(λ1)R1(λ1))21,a_{1}(x;s)=\frac{(2n)^{-n}}{y(x;s)}e^{-x^{2}-nl}\left(R(\lambda_{1})E(\lambda_{1})\left(\lim_{\zeta\rightarrow 0}\zeta\Psi_{\zeta}(\zeta)\Psi^{-1}(\zeta)\right)E^{-1}(\lambda_{1})R^{-1}(\lambda_{1})\right)_{21}, (4.51)

where use is made of the fact that E(z)E(z) and R(z)R(z) are analytic at z=λ1z=\lambda_{1}. It follows from the behavior of Ψ(z)\Psi(z) near z=λ1z=\lambda_{1} as given in (3.31) that

limζ0ζΨζ(ζ)Ψ1(ζ)=1ω12πiP^0(0100)P^01.\lim_{\zeta\rightarrow 0}\zeta\Psi_{\zeta}(\zeta)\Psi^{-1}(\zeta)=\frac{1-\omega_{1}}{2\pi i}\hat{P}_{0}\left(\begin{array}[]{cc}0&1\\ 0&0\end{array}\right)\hat{P}_{0}^{-1}. (4.52)

From the expression (3.44), we get for k=1,2k=1,2,

E(λk)=12(Iiσ1)nσ3/62σ3/2(10iHII(t1;t2t1)1)(I+O(n2/3)).E(\lambda_{k})=\frac{1}{\sqrt{2}}(I-i\sigma_{1})n^{\sigma_{3}/6}2^{\sigma_{3}/2}\left(\begin{array}[]{cc}1&0\\ -iH_{\texttt{II}}(t_{1};t_{2}-t_{1})&1\end{array}\right)(I+O(n^{-2/3})). (4.53)

Thus, we obtain from (3.55), (4.51)-(4.53) that

a1(x;s)=(1ω1)(2n)n2πiy(x;s)ex2nl((P^0)112n1/3+i(P^0)11(P^0)21+HII(t1;t2t1)(P^0)112+O(n1/3)).a_{1}(x;s)=\frac{(1-\omega_{1})(2n)^{-n}}{2\pi iy(x;s)}e^{-x^{2}-nl}\\ \left((\hat{P}_{0})_{11}^{2}n^{1/3}+i(\hat{P}_{0})_{11}(\hat{P}_{0})_{21}+H_{\texttt{II}}(t_{1};t_{2}-t_{1})(\hat{P}_{0})_{11}^{2}+O(n^{-1/3})\right).

Using (4.16) and (4.2), we obtain the asymptotic approximation of a1(x;s)a_{1}(x;s) as stated in (1.43).

Similarly, from (3.33), (3.43) and (4.46), we have

a2(x;s)\displaystyle a_{2}(x;s) =(2n)ny(x;s)ex2nl(R(λ2)E(λ2)ω1ω22πiQ^0(0100)Q^01E1(λ2)R1(λ2))21.\displaystyle=\frac{(2n)^{-n}}{y(x;s)}e^{-x^{2}-nl}\left(R(\lambda_{2})E(\lambda_{2})\frac{\omega_{1}-\omega_{2}}{2\pi i}\hat{Q}_{0}\left(\begin{array}[]{cc}0&1\\ 0&0\end{array}\right)\hat{Q}_{0}^{-1}E^{-1}(\lambda_{2})R^{-1}(\lambda_{2})\right)_{21}. (4.56)

Substituting (3.55), (4.19), (4.2) and (4.53) into (4.56), we obtain the asymptotic expansion of a2(x;s)a_{2}(x;s) as given in (1.44). In view of (4.47) and (4.48), we obtain the asymptotics of b1(x;s)b_{1}(x;s) and b2(x;s)b_{2}(x;s) by considering the (2,2)(2,2) entry of the matrices in (4.51) and (4.56):

b1(x;s)\displaystyle b_{1}(x;s) =(1ω1)2πia1(x;s)(i(P^0)112n1/3+O(n1/3))\displaystyle=\frac{(1-\omega_{1})}{2\pi ia_{1}(x;s)}\left(-i(\hat{P}_{0})_{11}^{2}n^{1/3}+O(n^{-1/3})\right)
=2n(1v1x(t1;t2t1)2v1(t1;t2t1)n1/3+O(n2/3)),n,\displaystyle=\sqrt{2n}\left(1-\frac{v_{1x}(t_{1};t_{2}-t_{1})}{2v_{1}(t_{1};t_{2}-t_{1})n^{1/3}}+O(n^{-2/3})\right),\quad n\rightarrow\infty, (4.57)

and

b2(x;s)\displaystyle b_{2}(x;s) =(ω1ω2)2πia2(x;s)(i(Q^0)112n1/3+O(n1/3))\displaystyle=\frac{(\omega_{1}-\omega_{2})}{2\pi ia_{2}(x;s)}\left(-i(\hat{Q}_{0})_{11}^{2}n^{1/3}+O(n^{-1/3})\right)
=2n(1v2x(t1;t2t1)2v2(t1;t2t1)n1/3+O(n2/3)),n,\displaystyle=\sqrt{2n}\left(1-\frac{v_{2x}(t_{1};t_{2}-t_{1})}{2v_{2}(t_{1};t_{2}-t_{1})n^{1/3}}+O(n^{-2/3})\right),\quad n\rightarrow\infty, (4.58)

where x=s1+s22=2n+t1+t222n1/6x=\frac{s_{1}+s_{2}}{2}=\sqrt{2n}+\frac{t_{1}+t_{2}}{2\sqrt{2}n^{1/6}} and s=s2s12=t2t122n1/6s=\frac{s_{2}-s_{1}}{2}=\frac{t_{2}-t_{1}}{2\sqrt{2}n^{1/6}}. This completes the proof of Theorem 3.

4.3 Proof of Theorem 4: asymptotics of the orthogonal polynomials

From (1.43)-(1.46), we have

a1(x;s)b1(x;s)2=2n5/6(v1(t1;t2t1)12v1x(t1;t2t1)n1/3+O(n2/3)),a_{1}(x;s)b_{1}(x;s)^{2}=-\sqrt{2}n^{5/6}\left(v_{1}(t_{1};t_{2}-t_{1})-\frac{1}{2}v_{1x}(t_{1};t_{2}-t_{1})n^{-1/3}+O(n^{-2/3})\right), (4.59)
a2(x;s)b2(x;s)2=2n5/6(v2(t1;t2t1)12v2x(t1;t2t1)n1/3+O(n2/3)),a_{2}(x;s)b_{2}(x;s)^{2}=-\sqrt{2}n^{5/6}\left(v_{2}(t_{1};t_{2}-t_{1})-\frac{1}{2}v_{2x}(t_{1};t_{2}-t_{1})n^{-1/3}+O(n^{-2/3})\right), (4.60)
a1(x;s)b1(x;s)+a2(x;s)b2(x;s)=n1/3((v1(t1;t2t1)+v2(t1;t2t1))+O(n2/3)).a_{1}(x;s)b_{1}(x;s)+a_{2}(x;s)b_{2}(x;s)=-n^{1/3}\left((v_{1}(t_{1};t_{2}-t_{1})+v_{2}(t_{1};t_{2}-t_{1}))+O(n^{-2/3})\right). (4.61)

Substituting (4.59)-(4.61) into (1.23), (1.24), (1.27) and (1.28), we obtain the asymptotics of the recurrence coefficients and πn(sk)\pi_{n}(s_{k}), k=1,2k=1,2, as given in (1.48), (1.49), (1.51) and (1.52), respectively. In view of (2.55), (4.35), (4.2) and (4.43), we derive the asymptotic expansion for the leading coefficient γn1\gamma_{n-1} of the orthonormal polynomial of degree n1n-1:

γn1\displaystyle\gamma_{n-1} =(12πi(Y1)21)1/2\displaystyle=\left(-\frac{1}{2\pi i}(Y_{1})_{21}\right)^{1/2}
=(2n(2n)nenl(R1+N1)212πi)1/2\displaystyle=\left(\frac{\sqrt{2n}(2n)^{-n}e^{-nl}(R_{1}+N_{1})_{21}}{-2\pi i}\right)^{1/2}
=2n234n14n2en2π(1+HII(t1;t2t1))2n1/3+O(n2/3)),n.\displaystyle=\frac{2^{\frac{n}{2}-\frac{3}{4}}n^{\frac{1}{4}-\frac{n}{2}}e^{\frac{n}{2}}}{\sqrt{\pi}}\left(1+\frac{H_{\texttt{II}}(t_{1};t_{2}-t_{1}))}{2n^{1/3}}+O(n^{-2/3})\right),\quad n\rightarrow\infty. (4.62)

Thus, we complete the proof of Theorem 4.

Acknowledgements

The authors are grateful to Dan Dai for useful comments. Xiao-Bo Wu was partially supported by National Natural Science Foundation of China under grant number 11801376 and Science Foundation of Education Department of Jiangxi Province under grant number GJJ170937. Shuai-Xia Xu was partially supported by National Natural Science Foundation of China under grant numbers 11971492, 11571376 and 11201493.

References

  • [1] A. Bogatskiy, T. Claeys and A. Its, Hankel determinant and orthogonal polynomials for a Gaussian weight with a discontinuity at the edge, Comm. Math. Phys., 347, 127–162 (2016).
  • [2] O. Bohigas and M.P. Pato, Missing levels in correlated spectra, Phys. Lett. B , 595 , 171–176 (2004).
  • [3] O. Bohigas and M.P. Pato, Randomly incomplete spectra and intermediate statistics, Phys. Rev. E (3), 74, 036212 (2006).
  • [4] F. Bornemann, P. J. Forrester and A. Mays, Finite size effects for spacing distributions in random matrix theory: circular ensembles and Riemann zeros, Stud. Appl. Math. , 138, 401–437 (2017).
  • [5] C. Charlier and T. Claeys, Thinning and conditioning of the Circular Unitary Ensemble, Random Matrices Theory Appl., 6, no. 2, 1750007 (2017).
  • [6] C. Charlier and T. Claeys, Large Gap Asymptotics for Airy Kernel Determinants with Discontinuities, Commun. Math. Phys., https://doi.org/10.1007/s00220-019-03538-w (2019).
  • [7] C. Charlier and A. Deaño, Asymptotics for Hankel determinants associated to a Hermite weight with a varying discontinuity, SIGMA Symmetry Integrability Geom. Methods Appl., 14, Paper No. 018, 43 pp (2018).
  • [8] T. Claeys and A. Doeraene, The generating function for the Airy point process and a system of coupled Painlevé II equations, Stud. Appl. Math., 140, 403–437 (2018).
  • [9] P. Deift, Orthogonal polynomials and random matrices: a Riemann-Hilbert approach, Courant Lecture Notes 3, New York University, Amer. Math. Soc., Providence, RI, 1999.
  • [10] P. Deift, T. Kriecherbauer, K. T.-R. McLaughlin, S. Venakides and X. Zhou, Uniform asymptotics for polynomials orthogonal with respect to varying exponential weights and applications to universality questions in random matrix theory, Comm. Pure Appl. Math., 52, 1335–1425 (1999).
  • [11] P. Deift, T. Kriecherbauer, K. T.-R. McLaughlin, S. Venakides and X. Zhou, Strong asymptotics of orthogonal polynomials with respect to exponential weights, Comm. Pure Appl. Math., 52, 1491–1552 (1999).
  • [12] P. Deift and X. Zhou, A steepest descent method for oscillatory Riemann-Hilbert problems, Asymptotics for the MKdV equation, Ann. Math., 137, 295–368 (1993).
  • [13] A.S. Fokas, A.R. Its, A.A. Kapaev and V.Y. Novokshenov, Painlevé Transcendents: The Riemann–Hilbert Approach, AMS Mathematical Surveys and Monographs, 128, Amer. Math. Soc., Providence, RI, 2006.
  • [14] A.S. Fokas, A.R. Its and A.V. Kitaev, The isomonodromy approach to matrix models in 2D2D quantum gravity, Comm. Math. Phys., 147, 395–430 (1992).
  • [15] P.J. Forrester, Log-gases and Random Matrices, London Mathematical Society Monographs Series, 34., Princeton University Press, Princeton, NJ, 2010.
  • [16] P.J. Forrester and N.S. Witte, Application of the τ\tau-function theory of Painlevé equations to random matrices: PIV, PII and the GUE, Comm. Math. Phys., 219, 357–398 (2001).
  • [17] A. Its and I. Krasovsky, Hankel determinant and orthogonal polynomials for the Gaussian weight with a jump, Integrable Systems and Random Matrices, J. Baik et al., eds., Contemp. Math., 458, Amer. Math. Soc., Providence, RI, 2008, 215–247.
  • [18] M. Jimbo, T. Miwa and K. Ueno, Monodromy preserving deformation of linear ordinary differential equations with rational coefficients II, Phys. D, 2, 407–448 (1981).
  • [19] H. Kawakami, Four-dimensional Painlevé-type equations associated with ramified linear equations III: Garnier systems and Fuji-Suzuki systerms, SIGMA , 13, 096, 50 pp (2017).
  • [20] H. Kawakami, Four-dimensional Painlevé-type equations associated with ramified linear equations II: Sasano systems, J. Integrable Syst., 3, no.1, xyy013, 36 pp (2018).
  • [21] H. Kawakami, A. Nakamura, and H. Sakai, Degeneration scheme of 4-dimensional Painlevé-type equations, arXiv:1209.3836.
  • [22] C. Min and Y. Chen, Painlevé transcendents and the Hankel determinants generated by a discontinuous Gaussian weight, Math. Methods Appl. Sci., 42, no.1, 301–321 (2019).
  • [23] M.L. Mehta, Random Matrices, 3rd ed., Elsevier/Academic Press, Amsterdam, 2004.
  • [24] F.W.J. Olver, A.B. Olde Daalhuis, D.W. Lozier, B.I. Schneider, R.F. Boisvert, C.W. Clark, B.R. Miller and B.V. Saunders, eds, NIST Digital Library of Mathematical Functions, http://dlmf.nist.gov/, Release 1.0.21 of 2018-12-15.
  • [25] C. Tracy and H. Widom, Level-spacing distributions and the Airy kernel, Comm. Math. Phys., 159, 151–174 (1994).
  • [26] X.-B. Wu, S.-X. Xu and Y.-Q. Zhao, Gaussian Unitary Ensemble with Boundary Spectrum Singularity and σ\sigma-Form of the Painlevé II Equation, Stud. Appl. Math., 140, 221–251(2018).
  • [27] S.-X. Xu and D. Dai, Tracy-Widom distributions in critical unitary random matrix ensembles and the coupled Painlevé II system, Comm. Math. Phys., 365 no. 2, 515–567 (2019).
  • [28] S.-X. Xu and Y.-Q. Zhao, Painlevé XXXIV asymptotics of orthogonal polynomials for the Gaussian weight with a jump at the edge, Stud. Appl. Math., 127, 67–105 (2011).