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

Wronskian structures of planar symplectic ensembles

Sung-Soo Byun Center for Mathematical Challenges, Korea Institute for Advanced Study, 85 Hoegiro, Dongdaemun-gu, Seoul 02455, Republic of Korea [email protected] Markus Ebke Department of Mathematics, Friedrich-Alexander-Universität Erlangen-Nürnberg, Cauerstrasse 11, 91058 Erlangen, Germany [email protected]  and  Seong-Mi Seo Department of Mathematics, Chungnam National University, 99 Daehak-ro, Yuseong-gu, Daejeon 34134, Republic of Korea. [email protected]
Abstract.

We consider the eigenvalues of non-Hermitian random matrices in the symmetry class of the symplectic Ginibre ensemble, which are known to form a Pfaffian point process in the plane. It was recently discovered that the limiting correlation kernel of the symplectic Ginibre ensemble in the vicinity of the real line can be expressed in a unified form of a Wronskian. We derive scaling limits for variations of the symplectic Ginibre ensemble and obtain such Wronskian structures for the associated universality classes. These include almost-Hermitian bulk/edge scaling limits of the elliptic symplectic Ginibre ensemble and edge scaling limits of the symplectic Ginibre ensemble with boundary confinement. Our proofs follow from the generalised Christoffel-Darboux formula for the former and from the Laplace method for the latter. Based on such a unified integrable structure of Wronskian form, we also provide an intimate relation between the function in the argument of the Wronskian in the symplectic symmetry class and the kernel in the complex symmetry class which form determinantal point processes in the plane.

2020 Mathematics Subject Classification:
Primary 60B20; Secondary 33C45
Sung-Soo Byun was partially supported by Samsung Science and Technology Foundation (SSTF-BA1401-51), by the National Research Foundation of Korea (NRF-2019R1A5A1028324) and by a KIAS Individual Grant (SP083201) via the Center for Mathematical Challenges at Korea Institute for Advanced Study. Markus Ebke was partially supported by Deutsche Forschungsgemeinschaft (IRTG 2235). Seong-Mi Seo was partially supported by the KIAS Individual Grant (MG063103) at Korea Institute for Advanced Study and by the National Research Foundation of Korea (2019R1A5A1028324, 2019R1F1A1058006).

1. Introduction

In the study of integrable models in random matrix theory, the underlying structure is one of the most important features that makes asymptotic analysis possible. We shall study such a structure for the planar symplectic ensemble 𝜻={ζj}j=1N\bm{\zeta}=\{\zeta_{j}\}_{j=1}^{N} whose joint probability distribution follows

(1.1) d𝐏N(𝜻)=1ZNj>k=1N|ζjζk|2|ζjζ¯k|2j=1N|ζjζ¯j|2eNQ(ζj)dA(ζj),(dA(ζ):=1πd2ζ),d\mathbf{P}_{N}(\bm{\zeta})=\frac{1}{Z_{N}}\prod_{j>k=1}^{N}\lvert\zeta_{j}-\zeta_{k}\rvert^{2}\lvert\zeta_{j}-\overline{\zeta}_{k}\rvert^{2}\prod_{j=1}^{N}\lvert\zeta_{j}-\overline{\zeta}_{j}\rvert^{2}e^{-NQ(\zeta_{j})}\,dA(\zeta_{j}),\qquad(dA(\zeta):=\tfrac{1}{\pi}d^{2}\zeta),

where ZNZ_{N} is the partition function. Here Q:Q:\mathbb{C}\to\mathbb{R} is a suitable function satisfying the complex conjugation symmetry Q(ζ)=Q(ζ¯)Q(\zeta)=Q(\bar{\zeta}), called the external potential. For the special case when Q(ζ)=|ζ|2Q(\zeta)=|\zeta|^{2}, the measure (1.1) describes the distribution of the eigenvalues of quaternionic Gaussian random matrices, also known as the symplectic Ginibre ensemble [37].

It is well known [18] that as N,N\to\infty, the system 𝜻\bm{\zeta} tends to occupy a certain compact subset SS\subset\mathbb{C} called the droplet. We shall study the local statistics of the ensemble (1.1) in the vicinity of the real line. For this purpose, we define the rescaled point process 𝒛={z1,,zN}\bm{z}=\{z_{1},\dots,z_{N}\} as

(1.2) zj:=eiθγN1(ζjp),pS,z_{j}:=e^{-i\theta}\gamma_{N}^{-1}\cdot(\zeta_{j}-p),\qquad p\in S\cap\mathbb{R},

where γN\gamma_{N} is an appropriate NN-dependent factor called the microscopic scale, cf. (2.1) and (2.20). Here θ\theta\in\mathbb{R} is the angle of the outward normal direction at the boundary if pSp\in\partial S and otherwise θ=0\theta=0.

By definition, the kk-point correlation function RN,kR_{N,k} is the normalised probability that kk of the eigenvalues lie in infinitesimal neighbourhoods of z1,,zkz_{1},\dots,z_{k}, i.e.

(1.3) RN,k(z1,,zk):=limε0( at least one particle in 𝔻(zj,ε),j=1,,k)ε2k,R_{N,k}(z_{1},\dots,z_{k}):=\lim_{\varepsilon\downarrow 0}\frac{\mathbb{P}(\,\exists\text{ at least one particle in }\mathbb{D}(z_{j},\varepsilon),j=1,\cdots,k)}{\varepsilon^{2k}},

see also (3.1) for a more standard definition. Recall that for a 2n×2n2n\times 2n skew-symmetric matrix A=(aj,k)A=(a_{j,k}), its Pfaffian Pf(A){\textup{Pf}}(A) is given by

Pf(A)=12nn!σS2nsgn(σ)j=1naσ(2j1),σ(2j),{\textup{Pf}}(A)=\frac{1}{2^{n}n!}\sum_{\sigma\in S_{2n}}\textup{sgn}(\sigma)\prod_{j=1}^{n}a_{\sigma(2j-1),\sigma(2j)},

where S2nS_{2n} is the symmetric group of order (2n)!(2n)!. Recently, it was shown in [5] that for the symplectic Ginibre ensemble, when Q(ζ)=|ζ|2Q(\zeta)=|\zeta|^{2} and thus S=𝔻(0,2)S=\mathbb{D}(0,\sqrt{2}), the kk-point correlation function RN,k(z1,,zk)R_{N,k}(z_{1},\dots,z_{k}) converges, as NN\to\infty, locally uniformly to the limit

(1.4) Rk(z1,,zk)=j=1k(z¯jzj)Pf[e|zj|2|zl|2(κ(zj,zl)κ(zj,z¯l)κ(z¯j,zl)κ(z¯j,z¯l))]j,l=1k,R_{k}(z_{1},\dots,z_{k})=\prod_{j=1}^{k}(\bar{z}_{j}-z_{j}){\textup{Pf}}\Big{[}e^{-|z_{j}|^{2}-|z_{l}|^{2}}\begin{pmatrix}\kappa(z_{j},z_{l})&\kappa(z_{j},\bar{z}_{l})\vskip 3.0pt plus 1.0pt minus 1.0pt\\ \kappa(\bar{z}_{j},z_{l})&\kappa(\bar{z}_{j},\bar{z}_{l})\end{pmatrix}\Big{]}_{j,l=1}^{k},

where the pre-kernel κ\kappa is of the form

(1.5) κ(z,w):=πez2+w2EW(fw,fz)(u)𝑑u.\kappa(z,w):=\sqrt{\pi}e^{z^{2}+w^{2}}\int_{E}W(f_{w},f_{z})(u)\,du.

Here, W(f,g):=fggfW(f,g):=fg^{\prime}-gf^{\prime} is the Wronskian, and

(1.6) fz(u):=12erfc(2(zu)),E:={(,)for the bulk case,(,0)for the edge case.f_{z}(u):=\tfrac{1}{2}\operatorname{erfc}(\sqrt{2}(z-u)),\qquad E:=\begin{cases}(-\infty,\infty)&\textup{for the bulk case},\vskip 3.0pt plus 1.0pt minus 1.0pt\\ (-\infty,0)&\textup{for the edge case}.\end{cases}

(See [42] for a related result. We also refer to [49, 30] for the spectral radius of the symplectic Ginibre ensemble.)

We emphasise that the function fzf_{z} is not necessarily unique. For instance, the variation f~z(u):=afz(u)+b(z)\tilde{f}_{z}(u):=af_{z}(u)+b(z) leads to the same pre-kernel in (1.5) if a=±1a=\pm 1 and b(z)=fz(u)|Eb(z)=f_{z}(u)|_{\partial E}. Thus, instead of fzf_{z} in (1.6), one may take

(1.7) {fz(u):=12erf(2(zu))for the bulk case,fz(u):=12(erfc(2(zu))erfc(2z))for the edge case.\begin{cases}f_{z}(u):=\tfrac{1}{2}\operatorname{erf}(\sqrt{2}(z-u))&\text{for the bulk case},\\ f_{z}(u):=\tfrac{1}{2}\Big{(}\operatorname{erfc}(\sqrt{2}(z-u))-\operatorname{erfc}(\sqrt{2}z)\Big{)}&\text{for the edge case}.\end{cases}

Note in particular that the 11-point function RR1R\equiv R_{1} is given by

(1.8) R(z)=(z¯z)e2|z|2κ(z,z¯)=4πImze4(Imz)2EIm[fz(u)fz¯(u)]𝑑u.R(z)=(\bar{z}-z)e^{-2|z|^{2}}\kappa(z,\bar{z})=4\sqrt{\pi}\,\operatorname{Im}z\,e^{-4(\operatorname{Im}z)^{2}}\int_{E}\operatorname{Im}[f_{z}^{\prime}(u)f_{\bar{z}}(u)]\,du.

For the symplectic Ginibre ensemble, it follows from S=𝔻(0,2)S=\mathbb{D}(0,\sqrt{2}) that the real bulk case corresponds to the regime p(2,2)p\in(-\sqrt{2},\sqrt{2}), whereas the real edge case corresponds to the regime p=±2p=\pm\sqrt{2}. We also remark that for the real bulk case, one can evaluate the integral in (1.5) in terms of the error function, which corresponds to the representation of the pre-kernel in [41, 9], see [5, Remark 2.3.(ii)] for further details.

Let us now define

(1.9) 𝒦(z,w):=2πez2+w¯2Efz(u)fw¯(u)𝑑u.\mathcal{K}(z,w):=2\sqrt{\pi}\,e^{z^{2}+\bar{w}^{2}}\int_{E}f^{\prime}_{z}(u)f^{\prime}_{\bar{w}}(u)\,du.

With the choice of (1.6), it is easy to see that the function 𝒦\mathcal{K} in (1.9) evaluates to

(1.10) e|z|2|w|2𝒦(z,w)={2e|z|2|w|2+2zw¯for the bulk case,2e|z|2|w|2+2zw¯erfc(z+w¯)for the edge case.e^{-|z|^{2}-|w|^{2}}\mathcal{K}(z,w)=\begin{cases}2\,e^{-|z|^{2}-|w|^{2}+2z\bar{w}}&\textup{for the bulk case},\vskip 3.0pt plus 1.0pt minus 1.0pt\\ 2\,e^{-|z|^{2}-|w|^{2}+2z\bar{w}}\operatorname{erfc}(z+\bar{w})&\textup{for the edge case}.\end{cases}

Here, one may notice that (1.10) corresponds to the limiting local correlation kernel of the complex Ginibre ensemble whose joint probability distribution is given by

(1.11) d𝓟N(𝝀)=1𝒵Nj>k=1N|λjλk|2j=1NeNQ(λj)dA(λj)\displaystyle\begin{split}d\bm{\mathcal{P}}_{N}(\bm{\lambda})&=\frac{1}{\mathcal{Z}_{N}}\prod_{j>k=1}^{N}\lvert\lambda_{j}-\lambda_{k}\rvert^{2}\prod_{j=1}^{N}e^{-NQ(\lambda_{j})}\,dA(\lambda_{j})\end{split}

with Q(ζ)=|ζ|2Q(\zeta)=\lvert\zeta\rvert^{2}, which forms a determinantal point process, see e.g. [32]. In other words, a properly defined limiting local correlation function R^k(z^1,,z^k)\widehat{R}_{k}(\widehat{z}_{1},\dots,\widehat{z}_{k}) of the ensemble (1.11) is given by

(1.12) R^k(z^1,,z^k)=det[e|z^j|2|z^l|2𝒦(z^j,z^l)]j,l=1k.\widehat{R}_{k}(\widehat{z}_{1},\dots,\widehat{z}_{k})=\det\Big{[}e^{-|\widehat{z}_{j}|^{2}-|\widehat{z}_{l}|^{2}}\mathcal{K}(\widehat{z}_{j},\widehat{z}_{l})\Big{]}_{j,l=1}^{k}.

In what follows, we call the ensemble (1.11) the complex counterpart to the symplectic ensemble (1.1).

The above discussions already lead us to the following natural questions.

  • Question 1. For a general potential QQ in the symplectic ensemble (1.1), does the limiting pre-kernel κ\kappa at pp\in\mathbb{R} have the Wronskian structure (1.5) for some function fzf_{z} and EE\subset\mathbb{R}?

  • Question 2. If the answer to the first question is affirmative, is the function defined in (1.9) the kernel of the corresponding complex ensemble? cf. (1.12).

The main purpose of this study is to examine these questions for the elliptic Ginibre ensemble in the almost-Hermitian regime and the Ginibre ensemble with boundary confinements.

2. Discussions of main results

Beyond the standard scaling limits (1.6) arising from the symplectic Ginibre ensemble, we aim to derive further universality classes whose complex counterparts have been well studied. In this section, we introduce our models and state the main results.

Until further notice, the microscopic scale in (1.2) is given by

(2.1) γN=1Nδ,δ:=ΔQ(p)2,(Δ:=¯).\gamma_{N}=\frac{1}{\sqrt{N\delta}},\qquad\delta:=\dfrac{\Delta Q(p)}{2},\qquad(\Delta:=\partial{\bar{\partial}}).

This specific choice of the rescaling comes from the fact that the macroscopic density of 𝜻\bm{\zeta} with respect to the area measure is 12ΔQ,\frac{1}{2}\Delta Q, see [18, 50].

2.1. Symplectic elliptic Ginibre ensemble in the almost-Hermitian regime

We first study the symplectic elliptic Ginibre ensemble, a one-parameter family of random matrices indexed by a non-Hermiticity parameter τ[0,1)\tau\in[0,1). This model describes an interpolation between the Gaussian symplectic ensemble (τ=1)(\tau=1) and the symplectic Ginibre ensemble (τ=0)(\tau=0). Its eigenvalue statistics correspond to (1.1) with the potential

(2.2) Q(ζ):=11τ2(|ζ|2τReζ2).Q(\zeta):=\tfrac{1}{1-\tau^{2}}(|\zeta|^{2}-\tau\operatorname{Re}\zeta^{2}).

As the terminology “elliptic” indicates, the associated droplet SS is given by the ellipse

(2.3) S:={x+iy|(x2(1+τ))2+(y2(1τ))21}.S:=\{x+iy\in\mathbb{C}\,|\,(\tfrac{x}{\sqrt{2}(1+\tau)})^{2}+(\tfrac{y}{\sqrt{2}(1-\tau)})^{2}\leq 1\}.

From the general universality principle of random matrix theory, it can be expected that for any fixed τ[0,1)\tau\in[0,1), the local statistics of the ensemble in the large system coincide with the one (1.6) obtained from the case τ=0\tau=0. Based on the skew-orthogonal Hermite polynomials introduced by Kanzieper [41], such a statement was recently proved in [7] for p=0p=0 and in [22] for general pp\in\mathbb{R}. (We mention that the local statistic at p=0p=0 can be special for certain random matrix models [21, 23, 40, 8, 31].)

On the other hand, when τ1\tau\to 1 as NN\to\infty with an appropriate rate, the limiting correlation functions obtained from the double scaling limit are no longer described in terms of (1.6), and it is natural to expect a non-trivial transition between the scaling limits of non-Hermitian and Hermitian random matrices. In general, such a transition appears in the so-called almost-Hermitian regime (or weak non-Hermiticity) introduced in the series of works [36, 34, 35] by Fyodorov, Khoruzhenko, and Sommers.

Refer to caption
Figure 1. Eigenvalues of the elliptic Ginibre ensemble in the almost-Hermitian regime

In Theorems 2.1 and  2.2 below, we state the scaling limits of the symplectic elliptic Ginibre ensembles in the almost-Hermitian regime, see Figure 1. We emphasise that for the special cases p=0p=0 and p=±2(1+τ)p=\pm\sqrt{2}(1+\tau), (cf. (2.3)) the scaling limits were studied in [41] and [10] respectively, based on the idea of Riemann sum approximation of skew-orthogonal Hermite polynomial kernels. This method is very useful in finding an explicit formula of the limiting pre-kernel, but it is not easy to perform the asymptotic analysis for general pp\in\mathbb{R} or to precisely control the error term.

Instead, we exploit the generalised Christoffel-Darboux formula introduced in [22], which allows us to obtain unified proofs for any points on the real line and to perform precise asymptotic analysis. (Indeed, this method can also be used to derive the subleading correction terms as well, see [22].) Furthermore, we describe the scaling limits in terms of the unified Wronskian form (1.5) and show that the associated kernels 𝒦\mathcal{K} of the form (1.9) correspond to those obtained from their complex counterparts. These provide affirmative answers to the two questions in the previous section for the almost-Hermitian symplectic ensembles.

Our first main result is on the bulk scaling limit of the symplectic elliptic Ginibre ensemble in the almost-Hermitian regime.

Theorem 2.1.

(Almost-Hermitian bulk scaling limit) Let QQ be the elliptic potential (2.2) with

(2.4) ττN=1c22N,c>0.\tau\equiv\tau_{N}=1-\frac{c^{2}}{2N},\qquad c>0.

Then for p(2(1+τ),2(1+τ))p\in(-\sqrt{2}(1+\tau),\sqrt{2}(1+\tau)), the kk-point correlation function RN,kR_{N,k} converges, as NN\to\infty, locally uniformly to the limit RkR_{k} of the form (1.4) with

(2.5) fz(u):=12πIet2/2sin(2t(zu))dtt,E:=.f_{z}(u):=\frac{1}{2\pi}\int_{I}e^{-t^{2}/2}\sin(2t(z-u))\,\frac{dt}{t},\qquad E:=\mathbb{R}.

Here

(2.6) I:=(c~,c~),c~:=c1p28.I:=(-\tilde{c},\tilde{{c}}),\qquad\tilde{c}:=c\sqrt{1-\tfrac{p^{2}}{8}}.

With the choice (2.5), the pre-kernel κ\kappa has an alternative representation

(2.7) κ(z,w)=1πez2+w2Ieu2sin(2u(zw))duu,\kappa(z,w)=\frac{1}{\sqrt{\pi}}e^{z^{2}+w^{2}}\int_{I}e^{-u^{2}}\sin(2u(z-w))\frac{du}{u},

see Subsection 4.2. For p=0p=0, this form of the limiting pre-kernel was investigated in [41]. We emphasise that by means of Ward’s equation, it was shown in [5, Theorem 2.10] that the pre-kernel of the form (2.7) is a unique translation invariant scaling limit of general planar symplectic ensembles.

Note in particular that with (2.5), the kernel 𝒦\mathcal{K} in (1.9) is given by

(2.8) 𝒦(z,w)=2πez2+w¯2Iet2cos(2t(zw¯))𝑑t.\mathcal{K}(z,w)=\frac{2}{\sqrt{\pi}}\,e^{z^{2}+\bar{w}^{2}}\int_{I}e^{-t^{2}}\cos(2t(z-\bar{w}))\,dt.

The kernel (2.8) corresponds to the one obtained from the complex elliptic Ginibre, see e.g. [34, 6]. As one may notice, the factor 1p2/8\sqrt{1-p^{2}/8} in (2.6) corresponds to the density of the semi-circle law. We refer to [11] for the geometric interpretation of such a density term, which follows from the limiting shape of the droplet (also called cross-section convergence).

It is also easy to show that the function fzf_{z} in (2.5) satisfies

(2.9) limcfz(u)=12erf(2(zu)).\lim_{c\to\infty}f_{z}(u)=\tfrac{1}{2}\operatorname{erf}(\sqrt{2}(z-u)).

Thus one can recover (1.6) (cf. (1.7)) for the bulk case in the non-Hermitian limit when cc\to\infty.

Refer to caption
(a) c~=1\tilde{c}=1
Refer to caption
(b) c~=2\tilde{c}=2
Refer to caption
(c) c~=3\tilde{c}=3
Figure 2. The graphs of the one-point function RR in (1.8) with (2.5).

We now turn to the edge scaling limit.

Theorem 2.2.

(Almost-Hermitian edge scaling limit) Let QQ be the elliptic potential (2.2) with

(2.10) ττN=1c2(2N)1/3,c>0.\tau\equiv\tau_{N}=1-\frac{c^{2}}{(2N)^{1/3}},\qquad c>0.

Then for p=±2(1+τ)p=\pm\sqrt{2}(1+\tau), the kk-point correlation function RN,kR_{N,k} converges, as NN\to\infty, locally uniformly to the limit RkR_{k} of the form (1.4) with

(2.11) fz(u):=2c0uec3(zt)+c612Ai(2c(zt)+c44)𝑑t,E:=(,0).f_{z}(u):=2c\int_{0}^{u}e^{c^{3}(z-t)+\frac{c^{6}}{12}}\operatorname{Ai}\Big{(}2c(z-t)+\frac{c^{4}}{4}\Big{)}\,dt,\qquad E:=(-\infty,0).

Notice that with (2.11), the kernel 𝒦\mathcal{K} in (1.9) is given by

(2.12) 𝒦(z,w)=8πc2ez2+w¯2+c660ec3(z+w¯2u)Ai(2c(zu)+c44)Ai(2c(w¯u)+c44)𝑑u.\mathcal{K}(z,w)=8\sqrt{\pi}\,c^{2}\,e^{z^{2}+\bar{w}^{2}+\frac{c^{6}}{6}}\int_{-\infty}^{0}e^{c^{3}(z+\bar{w}-2u)}\operatorname{Ai}\Big{(}2c(z-u)+\frac{c^{4}}{4}\Big{)}\operatorname{Ai}\Big{(}2c(\bar{w}-u)+\frac{c^{4}}{4}\Big{)}\,du.

Again, the kernel (2.12) corresponds to the complex counterpart obtained first in [19, 3]. (We also refer the reader to [11, 10] for alternative derivations of (2.12).)

We now briefly discuss the non-Hermitian limit when cc\to\infty. It follows from the asymptotic behaviour of the Airy function (see e.g. [47, Eq.(9.7.5)])

(2.13) Ai(z)=exp(23z3/2)2πz1/4(1+O(z3/2)),(z)\operatorname{Ai}(z)=\frac{\exp(-\frac{2}{3}z^{3/2})}{2\sqrt{\pi}z^{1/4}}\cdot(1+O(z^{-3/2})),\qquad(z\to\infty)

that the function fzf_{z} in (2.11) satisfies

(2.14) limcfz(u)=12(erfc(2(zu))erfc(2z)).\lim_{c\to\infty}f_{z}(u)=\tfrac{1}{2}\Big{(}\operatorname{erfc}(\sqrt{2}(z-u))-\operatorname{erfc}(\sqrt{2}z)\Big{)}.

Thus one can recover (1.6) (cf. (1.7)) for the edge case as well.

Refer to caption
(a) c=1c=1
Refer to caption
(b) c=2c=2
Refer to caption
(c) c=3c=3
Figure 3. The graphs of the one-point function RR in (1.8) with (2.11).

2.2. Symplectic Ginibre ensemble with boundary confinements

In the previous subsection we have discussed the ensembles (1.1) whose potential does not have any constraints near the boundary. In what follows, we call such a situation as free boundary (or soft edge) condition. On the other hand different universality classes naturally arise at the edge, if appropriate boundary constraints are imposed. Typical examples of such boundary conditions are the so-called soft/hard edge and hard edge constraints.

In the soft/hard edge setting, we completely confine the particles inside of the droplet SS by redefining the potential Q(ζ)=+Q(\zeta)=+\infty outside SS. This type of boundary confinement does not change the limiting spectral distribution. The term soft/hard comes from this situation being called when “the soft edge meets the hard edge” [27]. For the complex ensembles (1.11), such a boundary condition has been investigated in [15, 16, 20, 51] for example.

In the hard edge setting, we confine the particles further inside of the droplet SS. This leads to a modified associated equilibrium measure, in particular giving rise to some non-trivial measure on a certain one-dimensional subset. From a statistical physics point of view, this confinement has the effect of condensing a non-trivial portion of the particles onto the hard edge. We refer to [52, 38, 46] and references therein for the studies of complex ensembles (1.11) with such type of boundary confinement. (See also [57] for a similar situation in the context of truncated unitary ensembles.)

To our knowledge, the edge scaling limits of the symplectic ensembles associated with the above boundary conditions have not been investigated, and we aim to contribute to these problems. In particular, in Theorems 2.3 and 2.4, we derive the scaling limits of the symplectic Ginibre ensembles with soft/hard edge and hard edge constraints, which provide new universality classes. See Figure 4 for the graphs of the corresponding one-point functions. Furthermore, we shall show that the limiting pre-kernels are again of Wronskian form and that the relation to their complex counterparts again holds.

First, we consider the soft/hard edge Ginibre ensemble, which corresponds to the configuration (1.1) with the potential

(2.15) Q(ζ)={|ζ|2if |ζ|2,otherwise.Q(\zeta)=\begin{cases}|\zeta|^{2}&\text{if }|\zeta|\leq\sqrt{2},\\ \infty&\text{otherwise}.\end{cases}

By construction, all the eigenvalues are completely confined inside the droplet S=𝔻(0,2)S=\mathbb{D}(0,\sqrt{2}). As a result, the rescaled point process (1.2) at the real edge of the spectrum lies only in the left-half plane \mathbb{H}_{-}. Since the limiting spectral distribution (the circular law) is the same as the usual symplectic Ginibre ensemble, we rescale the process with the choice of microscopic scale (2.1). We then obtain the following.

Theorem 2.3.

(Non-Hermitian soft/hard edge scaling limit) Let QQ be the soft/hard edge Ginibre potential (2.15). Then for p=±2p=\pm\sqrt{2}, the kk-point correlation function RN,kR_{N,k} converges, as NN\to\infty, locally uniformly to the limit RkR_{k} of the form (1.4) with

(2.16) fz(u):=2πue2(zt)2erfc(2t)𝑑t,E:=(,0).f_{z}(u):=\frac{2}{\sqrt{\pi}}\int_{-\infty}^{u}\frac{e^{-2(z-t)^{2}}}{\sqrt{\operatorname{erfc}(2t)}}\,dt,\qquad E:=(-\infty,0).

Note that with (2.16), the kernel 𝒦\mathcal{K} in (1.9) is given by

(2.17) 𝒦(z,w)=4πe2zw¯0e(z+w¯2u)2erfc(2u)𝑑u.\mathcal{K}(z,w)=\frac{4}{\sqrt{\pi}}\,e^{2z\bar{w}}\int_{-\infty}^{0}\frac{e^{-(z+\bar{w}-2u)^{2}}}{\operatorname{erfc}(2u)}\,du.

Again, (2.17) corresponds to the kernel of the complex Ginibre point process with soft/hard edge condition, see e.g. [14, 15, 32].

We now turn to the hard edge Ginibre ensemble. This corresponds to the ensemble (1.1) with the potential

(2.18) Q(ζ)={|ζ|2if |ζ|2ρ,otherwise,ρ(0,1).Q(\zeta)=\begin{cases}|\zeta|^{2}&\text{if }|\zeta|\leq\sqrt{2}\rho,\\ \infty&\text{otherwise},\end{cases}\qquad\rho\in(0,1).

By definition, the eigenvalues are confined in a disk Dρ:=𝔻(0,2ρ)D_{\rho}:=\mathbb{D}(0,\sqrt{2}\rho). Thus the rescaled processes are again confined in \mathbb{H}_{-}. In this case, the associated equilibrium measure is no longer absolutely continuous with respect to the area measure dAdA and rather it is of the form

(2.19) 12ΔQ𝟏DρdA+1ρ22ρds,\frac{1}{2}\Delta Q\mathbf{1}_{D_{\rho}}\,dA+\frac{1-\rho^{2}}{\sqrt{2}\rho}\,ds,

where dsds is the normalized arc-length measure on Dρ\partial D_{\rho}. Therefore we choose the micro-scale at the edge-point p=2ρp=\sqrt{2}\rho as

(2.20) γN=2ρN(1ρ2).\gamma_{N}=\frac{\sqrt{2}\rho}{N(1-\rho^{2})}.

We then obtain the following.

Theorem 2.4.

(Non-Hermitian hard edge scaling limit) Let QQ be the hard edge Ginibre potential (2.18). Then for p=±2ρp=\pm\sqrt{2}\rho, the kk-point correlation function RN,kR_{N,k} converges, as NN\to\infty, locally uniformly to the limit

(2.21) Rk(z1,,zk)=j=1k(z¯jzj)Pf[(κ(zj,zl)κ(zj,z¯l)κ(z¯j,zl)κ(z¯j,z¯l))]j,l=1k,R_{k}(z_{1},\dots,z_{k})=\prod_{j=1}^{k}(\bar{z}_{j}-z_{j}){\textup{Pf}}\Big{[}\begin{pmatrix}\kappa(z_{j},z_{l})&\kappa(z_{j},\bar{z}_{l})\vskip 3.0pt plus 1.0pt minus 1.0pt\\ \kappa(\bar{z}_{j},z_{l})&\kappa(\bar{z}_{j},\bar{z}_{l})\end{pmatrix}\Big{]}_{j,l=1}^{k},

where the pre-kernel κ\kappa is of Wronskian form

(2.22) κ(z,w):=EW(fw,fz)(u)𝑑u,fz(u):=0ut12e2zt𝑑t,E:=(0,1).\kappa(z,w):=\int_{E}W(f_{w},f_{z})(u)\,du,\qquad f_{z}(u):=\int_{0}^{u}t^{\frac{1}{2}}e^{2zt}\,dt,\qquad E:=(0,1).

We mention that the function fzf_{z} can be written in terms of the incomplete gamma function γ(a,z)\gamma(a,z) as

(2.23) fz(u)=1(2z)32γ(32,2uz).f_{z}(u)=\frac{1}{(-2z)^{\frac{3}{2}}}\gamma(\tfrac{3}{2},-2uz).

Compared to the limiting correlation functions of the form (1.4) with (1.5), in the hard edge scaling limit (2.21), there are no Gaussian factors. This comes from the micro-scale (2.20) used only for the hard edge ensemble among the models under consideration in the present work. By a similar reason, with the choice of (2.22), it is natural to consider

(2.24) 𝒦(z,w):=4Efz(u)fw¯(u)𝑑u=4Eue2u(z+w¯)𝑑u\mathcal{K}(z,w):=4\int_{E}f_{z}^{\prime}(u)f_{\bar{w}}^{\prime}(u)\,du=4\int_{E}u\,e^{2u(z+\bar{w})}\,du

rather than the one of the form (1.9). Then this kernel (2.24) corresponds to the one obtained from the complex Ginibre ensemble with hard edge constraint, see [52].

Concerning the equivalence of the truncated unitary and hard edge Ginibre ensemble [52, 57], we expect that the scaling limit (2.22) coincides with the one obtained in the context of the truncated symplectic ensemble, see [45, 43] for the scaling limit away from the real edge.

Refer to caption
(a) free boundary
Refer to caption
(b) soft/hard edge
Refer to caption
(c) hard edge
Figure 4. The graphs of the one-point function RR in (1.8) with (1.6), (2.16), and (2.22) respectively.
Remark (Numerics on subleading terms).

Beyond the limiting correlation functions, a natural question arising in the study of scaling limits is their rates of convergence as NN\to\infty. For instance, these were obtained in [22] for the symplectic elliptic Ginibre ensemble with fixed τ[0,1)\tau\in[0,1). (See also [33] for the Hermitian counterparts.) In particular, it was shown that in the edge scaling limit, the convergence rate is of order O(N1/2)O(N^{-1/2}).

The precise asymptotic expansions in the situations of Theorems 2.22.3, and  2.4 exceed the scope of this paper. Nevertheless, we present some relevant numerical simulations. In particular, from the numerics below, we observe that the rates of convergences are of order O(N1/3)O(N^{-1/3}) in Theorem 2.2 (cf. [33]), of order O(N1/2)O(N^{-1/2}) in Theorem 2.3 and of order O(N3/4)O(N^{-3/4}) in Theorem 2.4.

Refer to caption
(a) almost-Hermitian edge, y=0.7y=0.7
Refer to caption
(b) soft/hard edge, y=1y=1
Refer to caption
(c) hard edge, y=1y=1
Refer to caption
(d) almost-Hermitian edge, x=2x=2
Refer to caption
(e) soft/hard edge, x=1x=1
Refer to caption
(f) hard edge, x=1x=1
Figure 5. (A)–(C) display the graphs of xRN(r)(x+iy)x\mapsto R_{N}^{(r)}(x+iy) for a given value of yy\in\mathbb{R}, where RN(r)(z):=Nr(RN(z)R(z))R_{N}^{(r)}(z):=N^{r}(R_{N}(z)-R(z)). Here, r=1/3r=1/3 for the almost-Hermitian edge (with c=1c=1), r=1/2r=1/2 for the soft/hard edge and r=3/4r=3/4 for the hard edge cases respectively. (D)–(F) are the same figures for yRN(r)(x+iy)y\mapsto R_{N}^{(r)}(x+iy).

We end this section by giving a remark on universality.

Remark (Towards local universality).

Let υ(z,w):=ez2w2κ(z,w).\upsilon(z,w):=e^{-z^{2}-w^{2}}\kappa(z,w). Then integration by parts gives rise to

(2.25) zυ(z,w)=2Ezfz(u)fw(u)𝑑u+fw(u)zfz(u)|E.\displaystyle\begin{split}\partial_{z}\upsilon(z,w)=-2\int_{E}\frac{\partial}{\partial z}f_{z}(u)\cdot f_{w}^{\prime}(u)\,du+f_{w}(u)\frac{\partial}{\partial z}f_{z}(u)\Big{|}_{\partial E}.\end{split}

For the free boundary cases, the first term on the right-hand side corresponds to the kernel of the complex ensembles, cf. (1.9).

The finite-NN version of this equation for the (elliptic) Ginibre potential was introduced in [5, 22] as a version of the Christoffel-Darboux formula. Such a relation for some singular potentials has been investigated as well, which requires higher (or fractional) order differential operators, see [2, 48, 4] for the Laguerre ensembles and [5, 17, 26] for the Mittag-Leffler ensembles.

We also refer to [1, 56] for similar equations in the Hermitian random matrix theory. Together with the local universality of the complex ensembles, this relation plays an important role in the study of local universality for symplectic ensembles [28, 29]. In a similar spirit, we expect that the finite-NN version of the identity (2.25) together with the bulk/edge universality of the determinantal Coulomb gas [13, 39] (cf.[12]) provides key ingredients in universality problems for planar symplectic ensembles.

Organisation of the paper.

In Section 3, we compile and summarise the relevant materials on the planar symplectic ensembles such as the skew-orthogonal polynomial representation of the pre-kernel.

Section 4 is devoted to the proofs of Theorems 2.1 and  2.2. In Subsection 4.1, we recall the Christoffel-Darboux formula in [22] and provide the general strategy of the proofs.

In Section 5, we prove Theorems 2.3 and 2.4. The strategy of the proof of these theorems using the Laplace method is outlined in Subsection 5.1.

3. Preliminaries

By definition, the kk-point correlation function 𝐑N,k\mathbf{R}_{N,k} of the system (1.1) is given by

(3.1) 𝐑N,k(ζ1,,ζk):=N!(Nk)!Nk𝐏N(ζ1,,ζN)j=k+1NdA(ζj).\mathbf{R}_{N,k}(\zeta_{1},\cdots,\zeta_{k}):=\frac{N!}{(N-k)!}\int_{\mathbb{C}^{N-k}}\mathbf{P}_{N}(\zeta_{1},\dots,\zeta_{N})\prod_{j=k+1}^{N}\,dA(\zeta_{j}).

It is well known [41] that the ensemble (1.1) forms a Pfaffian point process. In other words, there is a two-variable function ϰN{\bm{\varkappa}}_{N}, called the (skew) pre-kernel, such that

(3.2) 𝐑N,k(ζ1,,ζk)=j=1k(ζ¯jζj)Pf[eNQ(ζj)/2NQ(ζl)/2(ϰN(ζj,ζl)ϰN(ζj,ζ¯l)ϰN(ζ¯j,ζl)ϰN(ζ¯j,ζ¯l))]j,l=1,,k.\mathbf{R}_{N,k}(\zeta_{1},\cdots,\zeta_{k})=\prod_{j=1}^{k}(\overline{\zeta}_{j}-\zeta_{j}){\textup{Pf}}\Big{[}e^{-NQ(\zeta_{j})/2-NQ(\zeta_{l})/2}\begin{pmatrix}{\bm{\varkappa}}_{N}(\zeta_{j},\zeta_{l})&{\bm{\varkappa}}_{N}(\zeta_{j},\bar{\zeta}_{l})\vskip 3.0pt plus 1.0pt minus 1.0pt\\ {\bm{\varkappa}}_{N}(\bar{\zeta}_{j},\zeta_{l})&{\bm{\varkappa}}_{N}(\bar{\zeta}_{j},\bar{\zeta}_{l})\end{pmatrix}\Big{]}_{j,l=1,\cdots,k}.

By a change of measures, it is easy to see that the correlation function RN,kR_{N,k} of the rescaled process 𝒛\bm{z} in (1.2) is given by

(3.3) RN,k(z1,,zk)=γN2k𝐑N,k(ζ1,,ζk).R_{N,k}(z_{1},\cdots,z_{k})=\gamma_{N}^{2k}\,\mathbf{R}_{N,k}(\zeta_{1},\cdots,\zeta_{k}).

In particular, with the rescaled pre-kernel

(3.4) κN(z,w):=γN3ϰN(ζ,η),\displaystyle\begin{split}\kappa_{N}(z,w):=\gamma_{N}^{3}{\bm{\varkappa}}_{N}(\zeta,\eta),\end{split}

we have

(3.5) RN,k(z1,,zk)=j=1k(z¯jzj)Pf[eN2(Q(p+zjNδ)+Q(p+zlNδ))(κN(zj,zl)κN(zj,z¯l)κN(z¯j,zl)κN(z¯j,z¯l))]j,l=1,k.R_{N,k}(z_{1},\cdots,z_{k})=\prod_{j=1}^{k}(\bar{z}_{j}-z_{j}){\textup{Pf}}\Big{[}e^{-\frac{N}{2}(Q(p+\frac{z_{j}}{\sqrt{N\delta}})+Q(p+\frac{z_{l}}{\sqrt{N\delta}}))}\begin{pmatrix}\kappa_{N}(z_{j},z_{l})&\kappa_{N}(z_{j},\bar{z}_{l})\\ \kappa_{N}(\bar{z}_{j},z_{l})&\kappa_{N}(\bar{z}_{j},\bar{z}_{l})\end{pmatrix}\Big{]}_{j,l=1,\cdots k}.

We remark that different pre-kernels may give rise to the same correlation functions. In particular, we call two pre-kernels κN\kappa_{N} and κ~N\widetilde{\kappa}_{N} equivalent if there exists a sequence of unimodular functions gN:g_{N}:\mathbb{C}\to\mathbb{C} with gN(ζ¯)=1/gN(ζ)g_{N}(\overline{\zeta})=1/g_{N}(\zeta) such that κ~N(z,w)=gN(z)gN(w)κN(z,w)\widetilde{\kappa}_{N}(z,w)=g_{N}(z)g_{N}(w)\kappa_{N}(z,w). In what follows, we also call cN(z,w):=gN(z)gN(w)c_{N}(z,w):=g_{N}(z)g_{N}(w) a cocycle.

The skew-symmetric form ,s\langle\cdot,\cdot\rangle_{s} is given by

f,gs:=(f(ζ)g(ζ¯)g(ζ)f(ζ¯))(ζζ¯)eNQ(ζ)𝑑A(ζ).\langle f,g\rangle_{s}:=\int_{\mathbb{C}}\Big{(}f(\zeta)g(\bar{\zeta})-g(\zeta)f(\bar{\zeta})\Big{)}(\zeta-\bar{\zeta})e^{-NQ(\zeta)}\,dA(\zeta).

Let qmq_{m} be a family of monic polynomials of degree mm that satisfy the following skew-orthogonality conditions with skew-norms rk>0r_{k}>0: for all k,lk,l\in\mathbb{N}

(3.6) q2k,q2ls=q2k+1,q2l+1s=0,q2k,q2l+1s=q2l+1,q2ks=rkδk,l,\langle q_{2k},q_{2l}\rangle_{s}=\langle q_{2k+1},q_{2l+1}\rangle_{s}=0,\qquad\langle q_{2k},q_{2l+1}\rangle_{s}=-\langle q_{2l+1},q_{2k}\rangle_{s}=r_{k}\,\delta_{k,l},

where δk,l\delta_{k,l} is the Kronecker delta. Then the pre-kernel ϰN{\bm{\varkappa}}_{N} has a canonical representation in terms of the skew-orthogonal polynomials

(3.7) ϰN(ζ,η)=k=0N1q2k+1(ζ)q2k(η)q2k(ζ)q2k+1(η)rk.{\bm{\varkappa}}_{N}(\zeta,\eta)=\sum_{k=0}^{N-1}\frac{q_{2k+1}(\zeta)q_{2k}(\eta)-q_{2k}(\zeta)q_{2k+1}(\eta)}{r_{k}}.

We present some examples of skew-orthogonal polynomials that will be used in the following sections.

  • Example 1. (Elliptic potential) For the elliptic potential QQ in (2.2), it was obtained by Kanzieper [41] that the associated skew-orthogonal polynomials qkq_{k} can be expressed in terms of the Hermite polynomial

    (3.8) Hk(z):=(1)kez2dkdzkez2=k!m=0k/2(1)mm!(k2m)!(2z)k2m.H_{k}(z):=(-1)^{k}e^{z^{2}}\frac{d^{k}}{dz^{k}}e^{-z^{2}}=k!\sum_{m=0}^{\lfloor k/2\rfloor}\frac{(-1)^{m}}{m!(k-2m)!}(2z)^{k-2m}.

    To be precise, we have

    (3.9) q2k+1(ζ)=(τ2N)k+12H2k+1(N2τζ),q2k(ζ)=(2N)kk!l=0k(τ/2)l(2l)!!H2l(N2τζ)q_{2k+1}(\zeta)=\Big{(}\frac{\tau}{2N}\Big{)}^{k+\frac{1}{2}}H_{2k+1}\Big{(}\sqrt{\tfrac{N}{2\tau}}\zeta\Big{)},\qquad q_{2k}(\zeta)=\Big{(}\frac{2}{N}\Big{)}^{k}k!\sum_{l=0}^{k}\frac{(\tau/2)^{l}}{(2l)!!}H_{2l}\Big{(}\sqrt{\tfrac{N}{2\tau}}\zeta\Big{)}

    and their skew-norms rkr_{k} are given by

    (3.10) rk=2(1τ)3/2(1+τ)1/2(2k+1)!N2k+2.r_{k}=2(1-\tau)^{3/2}(1+\tau)^{1/2}\frac{(2k+1)!}{N^{2k+2}}.

    This also follows from a more general method of constructing skew-orthogonal polynomials [7, Theorem 3.1].

  • Example 2. (Radially symmetric potential) Let us consider a general radially symmetric potential Q(ζ)=Q(|ζ|)Q(\zeta)=Q(|\zeta|) with Q(ζ)4log|ζ|Q(\zeta)\gg 4\log|\zeta| as ζ\zeta\to\infty. We write

    (3.11) hk:=|ζ|2keNQ(ζ)𝑑A(ζ).h_{k}:=\int_{\mathbb{C}}|\zeta|^{2k}e^{-NQ(\zeta)}\,dA(\zeta).

    for the orthogonal norm. Then it is easy to show that

    (3.12) q2k+1(ζ)=ζ2k+1,q2k(ζ)=ζ2k+l=0k1ζ2lj=0kl1h2l+2j+2h2l+2j+1,rk=2h2k+1q_{2k+1}(\zeta)=\zeta^{2k+1},\qquad q_{2k}(\zeta)=\zeta^{2k}+\sum_{l=0}^{k-1}\zeta^{2l}\prod_{j=0}^{k-l-1}\frac{h_{2l+2j+2}}{h_{2l+2j+1}},\qquad r_{k}=2h_{2k+1}

    forms a family of skew-orthogonal polynomials, see e.g. [40, p.7] and [7, Theorem 3.1].

4. Scaling limits of the almost-Hermitian ensembles

In this section, we study the elliptic Ginibre ensemble in the almost-Hermitian regime and prove Theorems 2.1 and  2.2. In Subsection 4.1, we outline the strategy of our proofs based on the Christoffel-Darboux formula. Subsections 4.2 and  4.3 are devoted to the study of the bulk scaling limit (Theorem 2.1) and the edge scaling limit (Theorem 2.2) respectively.

4.1. Strategy of the proof: the Christoffel-Darboux formula

Combining (3.4), (3.7), (3.9), (3.10), we have the canonical representation of the pre-kernel

κN(z,w)=2(1+τ)k=0N1(τ/2)k+12(2k+1)!!H2k+1(N2τp+1τ2τz)l=0k(τ/2)l(2l)!!H2l(N2τp+1τ2τw)2(1+τ)k=0N1(τ/2)k+12(2k+1)!!H2k+1(N2τp+1τ2τw)l=0k(τ/2)l(2l)!!H2l(N2τp+1τ2τz).\displaystyle\begin{split}\kappa_{N}(z,w)&=\sqrt{2}(1+\tau)\sum_{k=0}^{N-1}\frac{(\tau/2)^{k+\frac{1}{2}}}{(2k+1)!!}H_{2k+1}\Big{(}\sqrt{\tfrac{N}{2\tau}}p+\sqrt{\tfrac{1-\tau^{2}}{\tau}}z\Big{)}\sum_{l=0}^{k}\frac{(\tau/2)^{l}}{(2l)!!}H_{2l}\Big{(}\sqrt{\tfrac{N}{2\tau}}p+\sqrt{\tfrac{1-\tau^{2}}{\tau}}w\Big{)}\\ &\quad-\sqrt{2}(1+\tau)\sum_{k=0}^{N-1}\frac{(\tau/2)^{k+\frac{1}{2}}}{(2k+1)!!}H_{2k+1}\Big{(}\sqrt{\tfrac{N}{2\tau}}p+\sqrt{\tfrac{1-\tau^{2}}{\tau}}w\Big{)}\sum_{l=0}^{k}\frac{(\tau/2)^{l}}{(2l)!!}H_{2l}\Big{(}\sqrt{\tfrac{N}{2\tau}}p+\sqrt{\tfrac{1-\tau^{2}}{\tau}}z\Big{)}.\end{split}

Following [22], let us introduce

(4.1) κ^N(z,w):=ωN(z,w)κN(z,w),\widehat{\kappa}_{N}(z,w):=\omega_{N}(z,w)\kappa_{N}(z,w),

where

(4.2) ωN(z,w)=exp[τ(pN2(1τ2)+z)2+τ(pN2(1τ2)+w)2(pN1τ2+2z)(pN1τ2+2w)].\omega_{N}(z,w)=\exp\Big{[}\tau\,\Big{(}p\sqrt{\tfrac{N}{2(1-\tau^{2})}}+z\Big{)}^{2}+\tau\,\Big{(}p\sqrt{\tfrac{N}{2(1-\tau^{2})}}+w\Big{)}^{2}-\Big{(}p\sqrt{\tfrac{N}{1-\tau^{2}}}+\sqrt{2}z\Big{)}\Big{(}p\sqrt{\tfrac{N}{1-\tau^{2}}}+\sqrt{2}w\Big{)}\Big{]}.

Note that by (2.2) and (2.1), we have

(4.3) eN2(Q(p+zNδ)+Q(p+wNδ))1ωN(z,w)=e|z|2|w|2+2zw1cN(z,w),e^{-\frac{N}{2}(Q(p+\frac{z}{\sqrt{N\delta}})+Q(p+\frac{w}{\sqrt{N\delta}}))}\frac{1}{\omega_{N}(z,w)}=e^{-\lvert z\rvert^{2}-\lvert w\rvert^{2}+2zw}\,\frac{1}{c_{N}(z,w)},

where the cocycle cN(z,w)c_{N}(z,w) is given by

(4.4) cN(z,w)=exp(i2N1τ1+τpImzi2N1τ1+τpImw+iτImz2+iτImw2).c_{N}(z,w)=\exp\Big{(}-i\sqrt{2N\tfrac{1-\tau}{1+\tau}}\,p\,\operatorname{Im}z-i\sqrt{2N\tfrac{1-\tau}{1+\tau}}\,p\,\operatorname{Im}w+i\tau\operatorname{Im}z^{2}+i\tau\operatorname{Im}w^{2}\Big{)}.

Therefore as NN\to\infty, we have the uniform limit

(4.5) limNcN(z,w)eN2(Q(p+zNδ)+Q(p+wNδ))κN(z,w)=e|z|2|w|2+2zwκ^(z,w),\lim_{N\to\infty}c_{N}(z,w)e^{-\frac{N}{2}(Q(p+\frac{z}{\sqrt{N\delta}})+Q(p+\frac{w}{\sqrt{N\delta}}))}\kappa_{N}(z,w)=e^{-\lvert z\rvert^{2}-\lvert w\rvert^{2}+2zw}\,\widehat{\kappa}(z,w),

where κ^:=limNκ^N\widehat{\kappa}:=\lim_{N\to\infty}\widehat{\kappa}_{N}. Here the convergence is uniform on compact subsets of \mathbb{C}.

The key idea which allows us to perform the asymptotic analysis is the following version of the Christoffel-Darboux formula [22, Proposition 1.1].

Lemma 4.1.

(Christoffel-Darboux formula for the skew-orthogonal Hermite polynomial kernel) We have

(4.6) zκ^N(z,w)=2(zw)κ^N(z,w)+IN(z,w)IIN(z,w),\partial_{z}\widehat{\kappa}_{N}(z,w)=2(z-w)\widehat{\kappa}_{N}(z,w)+\textup{I}_{N}(z,w)-\textup{II}_{N}(z,w),

where

(4.7) IN(z,w)=21τ2ωN(z,w)k=02N1(τ/2)kk!Hk(N2τp+1τ2τz)Hk(N2τp+1τ2τw)\textup{I}_{N}(z,w)=2\sqrt{1-\tau^{2}}\,\omega_{N}(z,w)\sum_{k=0}^{2N-1}\frac{(\tau/2)^{k}}{k!}H_{k}\Big{(}\sqrt{\tfrac{N}{2\tau}}p+\sqrt{\tfrac{1-\tau^{2}}{\tau}}z\Big{)}H_{k}\Big{(}\sqrt{\tfrac{N}{2\tau}}p+\sqrt{\tfrac{1-\tau^{2}}{\tau}}w\Big{)}

and

(4.8) IIN(z,w)=21τ2ωN(z,w)(τ/2)N(2N1)!!H2N(N2τp+1τ2τz)l=0N1(τ/2)l(2l)!!H2l(N2τp+1τ2τw).\textup{II}_{N}(z,w)=2\sqrt{1-\tau^{2}}\,\omega_{N}(z,w)\frac{(\tau/2)^{N}}{(2N-1)!!}H_{2N}\Big{(}\sqrt{\tfrac{N}{2\tau}}p+\sqrt{\tfrac{1-\tau^{2}}{\tau}}z\Big{)}\sum_{l=0}^{N-1}\frac{(\tau/2)^{l}}{(2l)!!}H_{2l}\Big{(}\sqrt{\tfrac{N}{2\tau}}p+\sqrt{\tfrac{1-\tau^{2}}{\tau}}w\Big{)}.

We now write

(4.9) υ(z,w):=ez2w2κ(z,w)=e(zw)2κ^(z,w)\upsilon(z,w):=e^{-z^{2}-w^{2}}\kappa(z,w)=e^{-(z-w)^{2}}\widehat{\kappa}(z,w)

and

(4.10) I(z,w):=limNIN(z,w),II(z,w):=limNIIN(z,w),\textup{I}(z,w):=\lim_{N\to\infty}\textup{I}_{N}(z,w),\qquad\textup{II}(z,w):=\lim_{N\to\infty}\textup{II}_{N}(z,w),

where the convergence is uniform on compact subsets of \mathbb{C} (we will show this later in the proof of Proposition 4.4). Then by Lemma 4.1, we have

(4.11) zυ(z,w)=e(zw)2(I(z,w)II(z,w)).\frac{\partial}{\partial z}\upsilon(z,w)=e^{-(z-w)^{2}}\Big{(}\textup{I}(z,w)-\textup{II}(z,w)\Big{)}.

We analyse the large-NN limit of IN\textup{I}_{N} and IIN\textup{II}_{N} using the expressions

(4.12) IN(z,w)=IN(z0,w)+z0z(zIN)(t,w)𝑑t,IIN(z,w)=IIN(z,w0)+w0w(wIIN)(z,t)𝑑t.\textup{I}_{N}(z,w)=\textup{I}_{N}(z_{0},w)+\int_{z_{0}}^{z}\big{(}\partial_{z}\textup{I}_{N}\big{)}(t,w)\,dt,\qquad\textup{II}_{N}(z,w)=\textup{II}_{N}(z,w_{0})+\int_{w_{0}}^{w}\big{(}\partial_{w}\textup{II}_{N}\big{)}(z,t)\,dt.

The following version of the Christoffel-Darboux formula for the kernel of the complex elliptic Ginibre ensemble was obtained by Lee and Riser [44, Proposition 2.3], see also [25, Section 3] for more general identities of such kind.

Lemma 4.2.

(Christoffel-Darboux formula for the orthogonal Hermite polynomial kernel) The function

(4.13) SN(ζ,η):=k=0N1(τ/2)kk!Hk(ζ)Hk(η)S_{N}(\zeta,\eta):=\sum_{k=0}^{N-1}\frac{(\tau/2)^{k}}{k!}H_{k}(\zeta)H_{k}(\eta)

satisfies

(4.14) ζSN(ζ,η)=2τ1τ2(ητζ)SN(ζ,η)+21τ2(τ2)NτHN(ζ)HN1(η)HN1(ζ)HN(η)(N1)!.\begin{split}\partial_{\zeta}S_{N}(\zeta,\eta)&=\frac{2\tau}{1-\tau^{2}}(\eta-\tau\zeta)S_{N}(\zeta,\eta)+\frac{2}{1-\tau^{2}}\Big{(}\frac{\tau}{2}\Big{)}^{N}\frac{\tau H_{N}(\zeta)H_{N-1}(\eta)-H_{N-1}(\zeta)H_{N}(\eta)}{(N-1)!}.\end{split}

Using this lemma, we have the following expressions.

Lemma 4.3.

We have

(4.15) zIN(z,w)\displaystyle\partial_{z}\textup{I}_{N}(z,w) =4τ(τ/2)2N(2N1)!ωN(z,w)(τH2N(ζ)H2N1(η)H2N1(ζ)H2N(η))\displaystyle=\frac{4}{\sqrt{\tau}}\frac{(\tau/2)^{2N}}{(2N-1)!}\omega_{N}(z,w)\Big{(}\tau H_{2N}(\zeta)H_{2N-1}(\eta)-H_{2N-1}(\zeta)H_{2N}(\eta)\Big{)}

and

(4.16) w[e(zw)2IIN(z,w)]=(τ1)4τ(τ/2)2N(2N1)!e(zw)2ωN(z,w)H2N(ζ)H2N1(η),\partial_{w}\Big{[}e^{-(z-w)^{2}}\textup{II}_{N}(z,w)\Big{]}=(\tau-1)\frac{4}{\sqrt{\tau}}\frac{(\tau/2)^{2N}}{(2N-1)!}e^{-(z-w)^{2}}\omega_{N}(z,w)H_{2N}(\zeta)H_{2N-1}(\eta),

where

(4.17) ζ=N2τp+1τ2τz,η=N2τp+1τ2τw.\zeta=\sqrt{\tfrac{N}{2\tau}}p+\sqrt{\tfrac{1-\tau^{2}}{\tau}}z,\qquad\eta=\sqrt{\tfrac{N}{2\tau}}p+\sqrt{\tfrac{1-\tau^{2}}{\tau}}w.

In the following subsections, we derive the asymptotics of the inhomogeneous terms in Lemma 4.1, see Propositions 4.4 and  4.8 below. Then, by solving the resulting limiting differential equations for the pre-kernel, we complete the proof of Theorems 2.1 and  2.2.

4.2. Almost-Hermitian bulk scaling limit

In this subsection, we consider the almost-Hermitian bulk scaling limit where τ\tau is given by (2.4) and p(2(1+τ),2(1+τ))p\in(-\sqrt{2}(1+\tau),\sqrt{2}(1+\tau)).

We aim to show the following proposition.

Proposition 4.4.

We have

(4.18) I(z,w)=2πe(zw)2Iet2cos(2t(zw))𝑑t,II(z,w)=0,\textup{I}(z,w)=\frac{2}{\sqrt{\pi}}\,e^{(z-w)^{2}}\int_{I}e^{-t^{2}}\cos(2t(z-w))\,dt,\qquad\textup{II}(z,w)=0,

where II is given by (2.6).

We prove Theorem 2.1 using Proposition 4.4.

Proof of Theorem 2.1.

Let us first show the alternative representation (2.7). Note that

(4.19) fz(u)=1πIet2/2cos(2t(uz))𝑑t,I=(c~,c~).\displaystyle f_{z}^{\prime}(u)=-\frac{1}{\pi}\int_{I}e^{-t^{2}/2}\cos(2t(u-z))\,dt,\quad I=(-\tilde{c},\tilde{c}).

Thus we have

W(fw,fz)(u)=fw(u)fz(u)fz(u)fw(u)\displaystyle\quad W(f_{w},f_{z})(u)=f_{w}(u)f_{z}^{\prime}(u)-f_{z}(u)f_{w}^{\prime}(u)
=12π2I2et2/2s2/2(sin(2t(uw))cos(2s(uz))sin(2t(uz))cos(2s(uw)))1t𝑑t𝑑s\displaystyle=\frac{1}{2\pi^{2}}\int_{I^{2}}e^{-t^{2}/2-s^{2}/2}\Big{(}\sin(2t(u-w))\cos(2s(u-z))-\sin(2t(u-z))\cos(2s(u-w))\Big{)}\frac{1}{t}\,dt\,ds

Notice here that

sin(2t(uw))cos(2s(uz))=14i(e2it(uw)e2it(uw))(e2is(uz)+e2is(uz))\displaystyle\quad\sin(2t(u-w))\cos(2s(u-z))=\frac{1}{4i}(e^{2it(u-w)}-e^{-2it(u-w)})(e^{2is(u-z)}+e^{-2is(u-z)})
=14i(e2i(t+s)ue2itw2isz+e2i(ts)ue2itw+2isze2i(st)ue2itw2isze2i(t+s)ue2itw+2isz).\displaystyle=\frac{1}{4i}\Big{(}e^{2i(t+s)u}e^{-2itw-2isz}+e^{2i(t-s)u}e^{-2itw+2isz}-e^{2i(s-t)u}e^{2itw-2isz}-e^{-2i(t+s)u}e^{2itw+2isz}\Big{)}.

Since

δ(x)=12πeiux𝑑u=1πe2iux𝑑u\delta(x)=\frac{1}{2\pi}\int_{\mathbb{R}}e^{iux}\,du=\frac{1}{\pi}\int_{\mathbb{R}}e^{2iux}\,du

in the sense of distribution, we have

sin(2t(uw))cos(2s(uz))𝑑u\displaystyle\int_{\mathbb{R}}\sin(2t(u-w))\cos(2s(u-z))\,du =π4i(δ(t+s)+δ(ts))(e2it(zw)e2it(zw))\displaystyle=\frac{\pi}{4i}\Big{(}\delta(t+s)+\delta(t-s)\Big{)}\Big{(}e^{2it(z-w)}-e^{-2it(z-w)}\Big{)}
=π2(δ(t+s)+δ(ts))sin(2t(zw)).\displaystyle=\frac{\pi}{2}\Big{(}\delta(t+s)+\delta(t-s)\Big{)}\sin(2t(z-w)).

Therefore we have

Iet2/2s2/2(sin(2t(uw))cos(2s(uz))sin(2t(uz))cos(2s(uw)))𝑑u𝑑s\displaystyle\quad\int_{I}\int_{\mathbb{R}}e^{-t^{2}/2-s^{2}/2}\Big{(}\sin(2t(u-w))\cos(2s(u-z))-\sin(2t(u-z))\cos(2s(u-w))\Big{)}\,du\,ds
=πI(δ(t+s)+δ(ts))et2/2s2/2sin(2t(zw))𝑑s=2πet2sin(2t(zw)).\displaystyle=\pi\int_{I}\Big{(}\delta(t+s)+\delta(t-s)\Big{)}e^{-t^{2}/2-s^{2}/2}\sin(2t(z-w))\,ds=2\pi\,e^{-t^{2}}\sin(2t(z-w)).

By interchanging the integrals, this gives rise to

EW(fw,fz)(u)𝑑u\displaystyle\int_{E}W(f_{w},f_{z})(u)\,du =1πIet2sin(2t(zw))dtt,E=,\displaystyle=\frac{1}{\pi}\int_{I}e^{-t^{2}}\sin(2t(z-w))\frac{dt}{t},\qquad E=\mathbb{R},

which leads to (2.7).

Now it suffices to show that the function

(4.20) υ(z,w):=ez2w2κ(z,w)=1πIeu2sin(2u(zw))duu\upsilon(z,w):=e^{-z^{2}-w^{2}}\kappa(z,w)=\frac{1}{\sqrt{\pi}}\int_{I}e^{-u^{2}}\sin(2u(z-w))\frac{du}{u}

is a unique anti-symmetric solution, which satisfies (4.11) with (4.18). This is immediate from the definition since

zυ(z,w)\displaystyle\frac{\partial}{\partial z}\upsilon(z,w) =z1πIeu2sin(2u(zw))duu=2πIeu2cos(2u(zw))𝑑u.\displaystyle=\frac{\partial}{\partial z}\frac{1}{\sqrt{\pi}}\int_{I}e^{-u^{2}}\sin(2u(z-w))\frac{du}{u}=\frac{2}{\sqrt{\pi}}\int_{I}e^{-u^{2}}\cos(2u(z-w))\,du.

Now the proof is complete. ∎

The rest of this subsection is devoted to the proof of Proposition 4.4. In the sequel, we shall consider the case p[0,2(1+τ))p\in[0,\sqrt{2}(1+\tau)) as the other case follows along the same lines with slight modifications. For the asymptotic analysis, we shall use the following strong asymptotics of the Hermite polynomials.

Lemma 4.5.

As NN\to\infty, we have for p=0p=0 (and ζ\zeta given in (4.17))

(4.21) H2N(ζ)(1)N(2N)!N!cos(2cz),H2N1(ζ)(1)N+1(2N)!4N1N!sin(2cz).H_{2N}(\zeta)\sim(-1)^{N}\frac{(2N)!}{N!}\cos(2cz),\qquad H_{2N-1}(\zeta)\sim(-1)^{N+1}\frac{(2N)!}{\sqrt{4N-1}N!}\sin(2cz).

For p(0,2(1+τ))p\in(0,\sqrt{2}(1+\tau)), we have

H2N(ζ)254(4Ne)N(8p2)14eN4p2+c2p28+pc2z×cos[(2arccos(p22)p8p24)N+12arccos(p22)(c2p8+cz2)8p2π4]\displaystyle\begin{split}H_{2N}(\zeta)&\sim 2^{\frac{5}{4}}\Big{(}\frac{4N}{e}\Big{)}^{N}(8-p^{2})^{-\frac{1}{4}}e^{\frac{N}{4}p^{2}+\frac{c^{2}p^{2}}{8}+\frac{pc}{\sqrt{2}}z}\\ &\times\cos\Big{[}\Big{(}2\arccos(\tfrac{p}{2\sqrt{2}})-\tfrac{p\sqrt{8-p^{2}}}{4}\Big{)}N+\tfrac{1}{2}\arccos(\tfrac{p}{2\sqrt{2}})-(\tfrac{c^{2}p}{8}+\tfrac{cz}{\sqrt{2}})\sqrt{8-p^{2}}-\tfrac{\pi}{4}\Big{]}\end{split}

and

H2N1(ζ)254e12(4Ne)N12(8p2)14eN4p2+c2p28+pc2z×cos[(2arccos(p22)p8p24)N12arccos(p22)(c2p8+cz2)8p2π4].\displaystyle\begin{split}H_{2N-1}(\zeta)&\sim 2^{\frac{5}{4}}e^{-\frac{1}{2}}\Big{(}\frac{4N}{e}\Big{)}^{N-\frac{1}{2}}(8-p^{2})^{-\frac{1}{4}}e^{\frac{N}{4}p^{2}+\frac{c^{2}p^{2}}{8}+\frac{pc}{\sqrt{2}}z}\\ &\times\cos\Big{[}\Big{(}2\arccos(\tfrac{p}{2\sqrt{2}})-\tfrac{p\sqrt{8-p^{2}}}{4}\Big{)}N-\tfrac{1}{2}\arccos(\tfrac{p}{2\sqrt{2}})-(\tfrac{c^{2}p}{8}+\tfrac{cz}{\sqrt{2}})\sqrt{8-p^{2}}-\tfrac{\pi}{4}\Big{]}.\end{split}
Proof.

This immediately follows from the Mehler-Heine formula and Plancherel-Rotach asymptotics, see e.g. [47, Section 18.11(ii)] for p=0p=0 and [55, Corollary 4.1], cf. [11, Appendix A], for p0p\neq 0. ∎

Lemma 4.6.

As NN\to\infty, we have

(4.22) zIN(z,w)4πec~2e(zw)2sin(2c~(zw))\displaystyle\begin{split}\partial_{z}\textup{I}_{N}(z,w)\sim\frac{4}{\sqrt{\pi}}e^{-\tilde{c}^{2}}e^{(z-w)^{2}}\sin(2\tilde{c}(z-w))\end{split}

and

(4.23) w[e(zw)2IIN(z,w)]c2N82π(8p2)12ec2+c2p28×(sin[2c~(z+w)+p8p24(c2+2N)4Narccos(p22)]cos[2c~(wz)+arccos(p22)]).\displaystyle\begin{split}&\quad\partial_{w}\Big{[}e^{-(z-w)^{2}}\textup{II}_{N}(z,w)\Big{]}\sim\frac{c^{2}}{N}\frac{8\sqrt{2}}{\sqrt{\pi}}(8-p^{2})^{-\frac{1}{2}}e^{-c^{2}+\frac{c^{2}p^{2}}{8}}\\ &\times\Big{(}\sin\Big{[}2\tilde{c}(z+w)+\tfrac{p\sqrt{8-p^{2}}}{4}(c^{2}+2N)-4N\arccos(\tfrac{p}{2\sqrt{2}})\Big{]}-\cos\Big{[}2\tilde{c}(w-z)+\arccos(\tfrac{p}{2\sqrt{2}})\Big{]}\Big{)}.\end{split}
Proof.

By (4.17), we have

ζ=p2N+(c2p42+cz)1N+O(N32),ζ22=p24N+(p2c28+pc2z)+O(N1),\zeta=\frac{p}{\sqrt{2}}\sqrt{N}+\Big{(}\frac{c^{2}p}{4\sqrt{2}}+cz\Big{)}\frac{1}{\sqrt{N}}+O(N^{-\frac{3}{2}}),\qquad\frac{\zeta^{2}}{2}=\frac{p^{2}}{4}N+\Big{(}\frac{p^{2}c^{2}}{8}+\frac{pc}{\sqrt{2}}z\Big{)}+O(N^{-1}),

and similarly for η\eta in (4.17). Furthermore we have

logωN(z,w)=p22N+(zw)2cp2(z+w)c2p28+O(N1),2Nlogτ=c2+O(N1).\displaystyle\begin{split}\log\omega_{N}(z,w)&=-\frac{p^{2}}{2}N+(z-w)^{2}-\frac{cp}{\sqrt{2}}(z+w)-\frac{c^{2}p^{2}}{8}+O(N^{-1}),\qquad 2N\log\tau=-c^{2}+O(N^{-1}).\end{split}

We first consider the case p=0p=0 (note that c~=c\tilde{c}=c). In this case, it follows from Lemma 4.5 that

(4.24) H2N(ζ)H2N1(η)24N1N!(N1)!πNcos(2cz)sin(2cw).\displaystyle\begin{split}H_{2N}(\zeta)H_{2N-1}(\eta)&\sim-\frac{2^{4N-1}N!(N-1)!}{\pi\sqrt{N}}\cos(2cz)\sin(2cw).\end{split}

Also notice that

(4.25) 4τ(τ/2)2N(2N1)!24N1N!(N1)!πN4πec2.\frac{4}{\sqrt{\tau}}\frac{(\tau/2)^{2N}}{(2N-1)!}\frac{2^{4N-1}N!(N-1)!}{\pi\sqrt{N}}\sim\frac{4}{\sqrt{\pi}}e^{-c^{2}}.

Then by Lemma 4.3, we obtain

(4.26) zIN(z,w)4πe(zw)2c2sin(2c(zw))\partial_{z}\textup{I}_{N}(z,w)\sim\frac{4}{\sqrt{\pi}}e^{(z-w)^{2}-c^{2}}\sin(2c(z-w))

and

(4.27) w[e(zw)2IIN(z,w)]c2N8πec2cos(2cz)sin(2cw).\partial_{w}\Big{[}e^{-(z-w)^{2}}\textup{II}_{N}(z,w)\Big{]}\sim\frac{c^{2}}{N}\frac{8}{\sqrt{\pi}}\,e^{-c^{2}}\cos(2cz)\sin(2cw).

Next, we prove the lemma for the case p(0,2(1+τ))p\in(0,\sqrt{2}(1+\tau)). By Lemma 4.5 and the elementary identity of trigonometric functions, we have

H2N(ζ)H2N1(η)232e12(4Ne)2N12(8p2)12eN2p2+c2p24+cp2(z+w)×(cos[2c~(wz)+arccos(p22)]sin[8p24(22c(z+w)+p(c2+2N))4Narccos(p22)]).\displaystyle\begin{split}&\quad H_{2N}(\zeta)H_{2N-1}(\eta)\sim 2^{\frac{3}{2}}e^{-\frac{1}{2}}\Big{(}\frac{4N}{e}\Big{)}^{2N-\frac{1}{2}}(8-p^{2})^{-\frac{1}{2}}e^{\frac{N}{2}p^{2}+\frac{c^{2}p^{2}}{4}+\frac{cp}{\sqrt{2}}(z+w)}\\ &\times\Big{(}\cos\Big{[}2\tilde{c}(w-z)+\arccos(\tfrac{p}{2\sqrt{2}})\Big{]}-\sin\Big{[}\tfrac{\sqrt{8-p^{2}}}{4}\Big{(}2\sqrt{2}c(z+w)+p(c^{2}+2N)\Big{)}-4N\arccos(\tfrac{p}{2\sqrt{2}})\Big{]}\Big{)}.\end{split}

Then it follows from

cos[2c~(wz)+arccos(p22)]cos[2c~(zw)+arccos(p22)]=212(8p2)12sin(2c~(zw))\displaystyle\quad\cos\Big{[}2\tilde{c}(w-z)+\arccos(\tfrac{p}{2\sqrt{2}})\Big{]}-\cos\Big{[}2\tilde{c}(z-w)+\arccos(\tfrac{p}{2\sqrt{2}})\Big{]}=2^{-\frac{1}{2}}(8-p^{2})^{\frac{1}{2}}\sin(2\tilde{c}(z-w))

that

H2N(ζ)H2N1(η)H2N(η)H2N1(ζ)2e12(4Ne)2N12eN2p2+c2p24+cp2(z+w)sin(2c~(zw)).\displaystyle\begin{split}H_{2N}(\zeta)H_{2N-1}(\eta)-H_{2N}(\eta)H_{2N-1}(\zeta)\sim 2\,e^{-\frac{1}{2}}\Big{(}\frac{4N}{e}\Big{)}^{2N-\frac{1}{2}}e^{\frac{N}{2}p^{2}+\frac{c^{2}p^{2}}{4}+\frac{cp}{\sqrt{2}}(z+w)}\sin(2\tilde{c}(z-w)).\end{split}

Notice here that we have

4τ(τ/2)2N(2N1)!ωN(z,w)2e12(4Ne)2N12eN2p2+c2p24+cp2(z+w)4πec2+c2p28+(zw)2.\displaystyle\quad\frac{4}{\sqrt{\tau}}\frac{(\tau/2)^{2N}}{(2N-1)!}\omega_{N}(z,w)\cdot 2\,e^{-\frac{1}{2}}\Big{(}\frac{4N}{e}\Big{)}^{2N-\frac{1}{2}}e^{\frac{N}{2}p^{2}+\frac{c^{2}p^{2}}{4}+\frac{cp}{\sqrt{2}}(z+w)}\sim\frac{4}{\sqrt{\pi}}\,e^{-c^{2}+\frac{c^{2}p^{2}}{8}+(z-w)^{2}}.

Now Lemma 4.3 completes the proof.

Lemma 4.7.

The function I given by (4.18) satisfies

(4.28) zI(z,w)=4πe(zw)2c2sin(2c(zw)).\partial_{z}\textup{I}(z,w)=\frac{4}{\sqrt{\pi}}e^{(z-w)^{2}-c^{2}}\sin(2c(z-w)).
Proof.

By definition, we have

zI(z,w)\displaystyle\partial_{z}\textup{I}(z,w) =2πz[e(zw)2Iet2cos(2t(zw))𝑑t]\displaystyle=\frac{2}{\sqrt{\pi}}\partial_{z}\Big{[}e^{(z-w)^{2}}\int_{I}e^{-t^{2}}\cos(2t(z-w))\,dt\Big{]}
=4πe(zw)2Iet2((zw)cos(2t(zw))tsin(2t(zw)))𝑑t.\displaystyle=\frac{4}{\sqrt{\pi}}\,e^{(z-w)^{2}}\int_{I}e^{-t^{2}}\Big{(}(z-w)\cos(2t(z-w))-t\sin(2t(z-w))\Big{)}\,dt.

Now the lemma follows from integration by parts. ∎

We are now ready to prove Proposition 4.4.

Proof of Proposition 4.4.

By Lemmas 4.6 and  4.7, all we need to show is the proposition for some initial values.

For the function I, we choose the initial value I(w¯,w).\textup{I}(\bar{w},w). As discussed in [22], this corresponds to the microscopic density of the complex elliptic Ginibre ensemble. Thus the initial value follows from for instance [34, 6, 11].

Now it remains to show that IIN(z,w)II(z,w)=0\textup{II}_{N}(z,w)\to\textup{II}(z,w)=0 for some choice of ww. Here we choose w=0w=0.

Let us write

(4.29) ez2IIN(z,0)=21τ2ez2ωN(z,0)(τ/2)N(2N1)!!H2N(N2τp+1τ2τz)eN2(1+τ)p2TN(p),e^{-z^{2}}\textup{II}_{N}(z,0)=2\sqrt{1-\tau^{2}}\,e^{-z^{2}}\omega_{N}(z,0)\frac{(\tau/2)^{N}}{(2N-1)!!}H_{2N}\Big{(}\sqrt{\tfrac{N}{2\tau}}p+\sqrt{\tfrac{1-\tau^{2}}{\tau}}z\Big{)}e^{\frac{N}{2(1+\tau)}p^{2}}T_{N}(p),

where

(4.30) TN(p):=eN2(1+τ)p2l=0N1(τ/2)l(2l)!!H2l(N2τp).T_{N}(p):=e^{-\frac{N}{2(1+\tau)}p^{2}}\sum_{l=0}^{N-1}\frac{(\tau/2)^{l}}{(2l)!!}H_{2l}\Big{(}\sqrt{\tfrac{N}{2\tau}}p\Big{)}.

We claim that TN(p)1/2T_{N}(p)\sim 1/\sqrt{2}. Then it follows from straightforward computations using Lemma 4.5 that IIN(z,0)=O(1/N).\textup{II}_{N}(z,0)=O(1/\sqrt{N}).

To analyse the term TN(p)T_{N}(p), let us write

(4.31) TN(p)=TN(0)+0pTN(t)𝑑t.T_{N}(p)=T_{N}(0)+\int_{0}^{p}T^{\prime}_{N}(t)\,dt.

Due to the Hermite numbers H2l(0)=(1)l(2l)!/l!H_{2l}(0)=(-1)^{l}(2l)!/l!, we have by Taylor expansion

(4.32) TN(0)=l=0N1τll!(1)l(2l)!2l(2l)!!=11+τ+O(τNN)=12(1+O(1N)).T_{N}(0)=\sum_{l=0}^{N-1}\frac{\tau^{l}}{l!}\frac{(-1)^{l}(2l)!}{2^{l}(2l)!!}=\frac{1}{\sqrt{1+\tau}}+O\Big{(}\frac{\tau^{N}}{\sqrt{N}}\Big{)}=\frac{1}{\sqrt{2}}\Big{(}1+O\Big{(}\frac{1}{\sqrt{N}}\Big{)}\Big{)}.

By applying Lemmas 4.2 and 4.5,

(4.33) TN(t)=N2ττN1+τ14N1(N1)!eN2(1+τ)t2H2N1(N2τt)N2π(1t28)14ec22+c216t2cos[(2arccos(t22)t8t24)N12arccos(t22)c2t88t2π4].\begin{split}&\quad T^{\prime}_{N}(t)=-\sqrt{\frac{N}{2\tau}}\frac{\tau^{N}}{1+\tau}\frac{1}{4^{N-1}(N-1)!}e^{-\frac{N}{2(1+\tau)}t^{2}}H_{2N-1}\Big{(}\sqrt{\tfrac{N}{2\tau}}t\Big{)}\\ &\sim-\frac{\sqrt{N}}{\sqrt{2\pi}}(1-\tfrac{t^{2}}{8})^{-\frac{1}{4}}e^{-\frac{c^{2}}{2}+\frac{c^{2}}{16}t^{2}}\cos\Big{[}\Big{(}2\arccos(\tfrac{t}{2\sqrt{2}})-\tfrac{t\sqrt{8-t^{2}}}{4}\Big{)}N-\tfrac{1}{2}\arccos(\tfrac{t}{2\sqrt{2}})-\tfrac{c^{2}t}{8}\sqrt{8-t^{2}}-\tfrac{\pi}{4}\Big{]}.\end{split}

Invoking this asymptotics, the resulting oscillatory integral in (4.31) can be analysed in the same way as [24, Lemma 4.4]. To be more precise, it has the asymptotic form NRe0pa(t)eiNϕ(t)𝑑t\sqrt{N}\operatorname{Re}\int_{0}^{p}a(t)e^{iN\phi(t)}\,dt, where

(4.34) a(t)=12π(1t28)14exp(c22+c216t2i(12arccos(t22)+c2t88t2+π4))a(t)=-\tfrac{1}{\sqrt{2\pi}}\Big{(}1-\tfrac{t^{2}}{8}\Big{)}^{-\frac{1}{4}}\exp\Big{(}-\tfrac{c^{2}}{2}+\tfrac{c^{2}}{16}t^{2}-i\big{(}\tfrac{1}{2}\arccos(\tfrac{t}{2\sqrt{2}})+\tfrac{c^{2}t}{8}\sqrt{8-t^{2}}+\tfrac{\pi}{4}\big{)}\Big{)}

and

(4.35) ϕ(t)=2arccos(t22)t48t2.\phi(t)=2\arccos(\tfrac{t}{2\sqrt{2}})-\tfrac{t}{4}\sqrt{8-t^{2}}.

Since ϕ(t)=128t20\phi^{\prime}(t)=-\frac{1}{2}\sqrt{8-t^{2}}\neq 0 as p<22p<2\sqrt{2}, the Riemann–Lebesgue lemma yields

Re0pa(t)eiNϕ(t)𝑑t=O(1N),\operatorname{Re}\int_{0}^{p}a(t)e^{iN\phi(t)}\,dt=O\Big{(}\frac{1}{N}\Big{)},

see [24, Lemma 4.4] for further details. This completes the proof.

4.3. Almost-Hermitian edge scaling limit

In this subsection, we consider the almost-Hermitian edge scaling limit where τ\tau is given by (2.10) and p=±2(1+τ)p=\pm\sqrt{2}(1+\tau). As in the previous subsection, we need to show the following.

Proposition 4.8.

We have

(4.36) I(z,w)=8πc2e(zw)20ec3(z+w2u)+c66Ai(2c(zu)+c44)Ai(2c(wu)+c44)𝑑u\textup{I}(z,w)=8\sqrt{\pi}\,c^{2}\,e^{(z-w)^{2}}\int_{-\infty}^{0}e^{c^{3}(z+w-2u)+\frac{c^{6}}{6}}\operatorname{Ai}\Big{(}2c(z-u)+\frac{c^{4}}{4}\Big{)}\operatorname{Ai}\Big{(}2c(w-u)+\frac{c^{4}}{4}\Big{)}\,du

and

(4.37) II(z,w)=4πc2e(zw)2ec3zAi(2cz+c44)0ec3(wu)+c66Ai(2c(wu)+c44)𝑑u.\textup{II}(z,w)=4\sqrt{\pi}\,c^{2}\,e^{(z-w)^{2}}e^{c^{3}z}\operatorname{Ai}\Big{(}2cz+\frac{c^{4}}{4}\Big{)}\int_{-\infty}^{0}e^{c^{3}(w-u)+\frac{c^{6}}{6}}\operatorname{Ai}\Big{(}2c(w-u)+\frac{c^{4}}{4}\Big{)}\,du.

We remark that by (2.13), it is easy to see that

(4.38) limcI(z,w)=erfc(z+w),limcII(z,w)=12e(zw)22z2erfc(2w).\lim_{c\to\infty}\textup{I}(z,w)=\operatorname{erfc}(z+w),\qquad\lim_{c\to\infty}\textup{II}(z,w)=\tfrac{1}{\sqrt{2}}e^{(z-w)^{2}-2z^{2}}\operatorname{erfc}(\sqrt{2}w).

These terms appear in the limiting differential equations for the non-Hermitian regime, see [5, 22] .

We prove Theorem 2.2 using Proposition 4.8.

Proof of Theorem 2.2.

Let us write

(4.39) υ(z,w):=π0W(fw,fz)(u)𝑑u=π0(fw(u)fz(u)fw(u)fz(u))𝑑u,\upsilon(z,w):=\sqrt{\pi}\int_{-\infty}^{0}W(f_{w},f_{z})(u)\,du=\sqrt{\pi}\int_{-\infty}^{0}\Big{(}f_{w}(u)f^{\prime}_{z}(u)-f^{\prime}_{w}(u)f_{z}(u)\Big{)}\,du,

where fzf_{z} and EE are given by (2.11). Then similarly as above, it suffices to show that the function υ\upsilon is a unique anti-symmetric solution, which satisfies (4.11) with (4.36) and (4.37).

To see this, note that

(4.40) fz(u)=2cexp(c3(zu)+c612)Ai(2c(zu)+c44).f^{\prime}_{z}(u)=2c\,\exp\Big{(}c^{3}(z-u)+\frac{c^{6}}{12}\Big{)}\operatorname{Ai}\Big{(}2c(z-u)+\frac{c^{4}}{4}\Big{)}.

This gives

(4.41) zfz(u)=fz′′(u).\frac{\partial}{\partial z}f^{\prime}_{z}(u)=-f^{\prime\prime}_{z}(u).

Notice that since

zfz(u)\displaystyle\frac{\partial}{\partial z}f_{z}(u) =2c0uz[exp(c3(zt)+c612)Ai(2c(zt)+c44)]𝑑t\displaystyle=2c\int_{0}^{u}\frac{\partial}{\partial z}\Big{[}\exp\Big{(}c^{3}(z-t)+\frac{c^{6}}{12}\Big{)}\operatorname{Ai}\Big{(}2c(z-t)+\frac{c^{4}}{4}\Big{)}\Big{]}\,dt
=2c0ut[exp(c3(zt)+c612)Ai(2c(zt)+c44)]𝑑t,\displaystyle=-2c\int_{0}^{u}\frac{\partial}{\partial t}\Big{[}\exp\Big{(}c^{3}(z-t)+\frac{c^{6}}{12}\Big{)}\operatorname{Ai}\Big{(}2c(z-t)+\frac{c^{4}}{4}\Big{)}\Big{]}\,dt,

we have

(4.42) zfz(u)=fz(u)+fz(0).\frac{\partial}{\partial z}f_{z}(u)=-f^{\prime}_{z}(u)+f^{\prime}_{z}(0).

Therefore we obtain

zυ(z,w)\displaystyle\frac{\partial}{\partial z}\upsilon(z,w) =π0z(fw(u)fz(u)fw(u)fz(u))𝑑u\displaystyle=\sqrt{\pi}\int_{-\infty}^{0}\frac{\partial}{\partial z}\Big{(}f_{w}(u)f^{\prime}_{z}(u)-f^{\prime}_{w}(u)f_{z}(u)\Big{)}\,du
=π0(fw(u)fz′′(u)+fw(u)fz(u)fw(u)fz(0))𝑑u\displaystyle=\sqrt{\pi}\int_{-\infty}^{0}\Big{(}-f_{w}(u)f^{\prime\prime}_{z}(u)+f^{\prime}_{w}(u)f^{\prime}_{z}(u)-f^{\prime}_{w}(u)f_{z}^{\prime}(0)\Big{)}\,du
=2π0fz(u)fw(u)𝑑uπ(fz(u)+fz(0))fw(u)|0,\displaystyle=2\sqrt{\pi}\int_{-\infty}^{0}f^{\prime}_{z}(u)f^{\prime}_{w}(u)\,du-\sqrt{\pi}\,\Big{(}f_{z}^{\prime}(u)+f_{z}^{\prime}(0)\Big{)}f_{w}(u)\Big{|}_{-\infty}^{0},

which completes the proof. ∎

The rest of this subsection is devoted to the proof of Proposition 4.8.

For the asymptotic analysis, we shall use the following (critical) strong asymptotics of the Hermite polynomials.

Lemma 4.9.

As NN\to\infty, we have (with ζ\zeta again given by (4.17))

(4.43) H2N(ζ)(2π)142N(2N)!(2N)112e12ζ2Ai(2cz+c44)H_{2N}(\zeta)\sim(2\pi)^{\frac{1}{4}}2^{N}\sqrt{(2N)!}(2N)^{-\frac{1}{12}}e^{\frac{1}{2}\zeta^{2}}\operatorname{Ai}\Big{(}2cz+\frac{c^{4}}{4}\Big{)}

and

(4.44) H2N1(ζ)(2π)142N12(2N1)!(2N)112e12ζ2Ai(2cz+c44).H_{2N-1}(\zeta)\sim(2\pi)^{\frac{1}{4}}2^{N-\frac{1}{2}}\sqrt{(2N-1)!}(2N)^{-\frac{1}{12}}e^{\frac{1}{2}\zeta^{2}}\operatorname{Ai}\Big{(}2cz+\frac{c^{4}}{4}\Big{)}.
Proof.

This immediately follows from the critical Plancherel-Rotach estimate

(4.45) HN(2N+z2N1/6)(2π)142N2N!N112e12(2N+z2N1/6)2Ai(z),H_{N}(\sqrt{2N}+\tfrac{z}{\sqrt{2}N^{1/6}})\sim(2\pi)^{\frac{1}{4}}2^{\frac{N}{2}}\sqrt{N!}N^{-\frac{1}{12}}e^{\frac{1}{2}(\sqrt{2N}+\frac{z}{\sqrt{2}N^{1/6}})^{2}}\operatorname{Ai}(z),

see e.g. [54]. ∎

Lemma 4.10.

As NN\to\infty, we have

(4.46) zIN(z,w)=4πc2ec3(z+w)+c66+(zw)2Ai(2cz+c44)Ai(2cw+c44)+O(N16)\partial_{z}\textup{I}_{N}(z,w)=-4\sqrt{\pi}c^{2}e^{c^{3}(z+w)+\frac{c^{6}}{6}+(z-w)^{2}}\operatorname{Ai}\Big{(}2cz+\frac{c^{4}}{4}\Big{)}\operatorname{Ai}\Big{(}2cw+\frac{c^{4}}{4}\Big{)}+O(N^{-\frac{1}{6}})

and

(4.47) w[e(zw)2IIN(z,w)]=4πc2ec3(z+w)+c66Ai(2cz+c44)Ai(2cw+c44)+O(N16).\partial_{w}\Big{[}e^{-(z-w)^{2}}\textup{II}_{N}(z,w)\Big{]}=-4\sqrt{\pi}c^{2}e^{c^{3}(z+w)+\frac{c^{6}}{6}}\operatorname{Ai}\Big{(}2cz+\frac{c^{4}}{4}\Big{)}\operatorname{Ai}\Big{(}2cw+\frac{c^{4}}{4}\Big{)}+O(N^{-\frac{1}{6}}).
Proof.

First note that for ζ\zeta given in (4.17) we have

(4.48) ζ=2N+c4/4+2cz2(2N)1/6+O(N12),ζ22=2N+(253c4+243cz)N13+c64+c32z+O(N16).\zeta=2\sqrt{N}+\frac{c^{4}/4+2cz}{\sqrt{2}\,(2N)^{1/6}}+O(N^{-\frac{1}{2}}),\qquad\frac{\zeta^{2}}{2}=2N+(2^{-\frac{5}{3}}c^{4}+2^{\frac{4}{3}}cz)N^{\frac{1}{3}}+\frac{c^{6}}{4}+\frac{c^{3}}{2}z+O(N^{-\frac{1}{6}}).

We also have

(4.49) logωN(z,w)=4N+223c2N23243c(z+w)N13+(zw)2+c32(z+w)+O(N13)\displaystyle\begin{split}\log\omega_{N}(z,w)&=-4N+2^{\frac{2}{3}}c^{2}N^{\frac{2}{3}}-2^{\frac{4}{3}}\,c(z+w)N^{\frac{1}{3}}+(z-w)^{2}+\frac{c^{3}}{2}(z+w)+O(N^{-\frac{1}{3}})\end{split}

and

(4.50) 2Nlogτ=223c2N23223c4N13c63+O(N13).2N\log\tau=-2^{\frac{2}{3}}c^{2}N^{\frac{2}{3}}-2^{-\frac{2}{3}}c^{4}N^{\frac{1}{3}}-\frac{c^{6}}{3}+O(N^{-\frac{1}{3}}).

These give rise to

(4.51) logωN(z,w)+ζ2+η22+2Nlogτ=c3(z+w)+c66+(zw)2+O(N16).\displaystyle\begin{split}\log\omega_{N}(z,w)+\frac{\zeta^{2}+\eta^{2}}{2}+2N\log\tau&=c^{3}(z+w)+\frac{c^{6}}{6}+(z-w)^{2}+O(N^{-\frac{1}{6}}).\end{split}

Note that by Lemma 4.9, we have

(4.52) H2N(ζ)H2N1(η)(2π)1222N12(2N)!(2N1)!(2N)16e12(ζ2+η2)Ai(2cz+c44)Ai(2cw+c44).H_{2N}(\zeta)H_{2N-1}(\eta)\sim(2\pi)^{\frac{1}{2}}2^{2N-\frac{1}{2}}\sqrt{(2N)!(2N-1)!}(2N)^{-\frac{1}{6}}e^{\frac{1}{2}(\zeta^{2}+\eta^{2})}\operatorname{Ai}\Big{(}2cz+\frac{c^{4}}{4}\Big{)}\operatorname{Ai}\Big{(}2cw+\frac{c^{4}}{4}\Big{)}.

Also notice that

(4.53) (2π)1222N12(2N)!(2N1)!(2N)164τ22N(2N1)!(τ1)=4πc2+O(N13).(2\pi)^{\frac{1}{2}}2^{2N-\frac{1}{2}}\sqrt{(2N)!(2N-1)!}(2N)^{-\frac{1}{6}}\frac{4}{\sqrt{\tau}}\frac{2^{-2N}}{(2N-1)!}(\tau-1)=-4\sqrt{\pi}c^{2}+O(N^{-\frac{1}{3}}).

Combining all of the above with Lemma 4.3, we obtain

(4.54) zIN(z,w)=4πc2ec3(z+w)+c66+(zw)2Ai(2cz+c44)Ai(2cw+c44)+O(N16)\partial_{z}\textup{I}_{N}(z,w)=-4\sqrt{\pi}c^{2}e^{c^{3}(z+w)+\frac{c^{6}}{6}+(z-w)^{2}}\operatorname{Ai}\Big{(}2cz+\frac{c^{4}}{4}\Big{)}\operatorname{Ai}\Big{(}2cw+\frac{c^{4}}{4}\Big{)}+O(N^{-\frac{1}{6}})

and

(4.55) w[e(zw)2IIN(z,w)]=4πc2ec3(z+w)+c66Ai(2cz+c44)Ai(2cw+c44)+O(N16).\partial_{w}\Big{[}e^{-(z-w)^{2}}\textup{II}_{N}(z,w)\Big{]}=-4\sqrt{\pi}c^{2}e^{c^{3}(z+w)+\frac{c^{6}}{6}}\operatorname{Ai}\Big{(}2cz+\frac{c^{4}}{4}\Big{)}\operatorname{Ai}\Big{(}2cw+\frac{c^{4}}{4}\Big{)}+O(N^{-\frac{1}{6}}).

This completes the proof. ∎

Lemma 4.11.

The functions I and II given by (4.36) and (4.37) satisfy

(4.56) zI(z,w)=πe(zw)2fz(0)fw(0),w[e(zw)2II(z,w)]=πfz(0)fw(0).\frac{\partial}{\partial z}\textup{I}(z,w)=-\sqrt{\pi}\,e^{(z-w)^{2}}\,f_{z}^{\prime}(0)f_{w}^{\prime}(0),\qquad\frac{\partial}{\partial w}\Big{[}e^{-(z-w)^{2}}\textup{II}(z,w)\Big{]}=-\sqrt{\pi}\,f_{z}^{\prime}(0)f_{w}^{\prime}(0).
Proof.

The second identity is obvious since

w[e(zw)2II(z,w)]\displaystyle\partial_{w}\Big{[}e^{-(z-w)^{2}}\textup{II}(z,w)\Big{]} =πfz(0)wfw()\displaystyle=-\sqrt{\pi}\,f_{z}^{\prime}(0)\frac{\partial}{\partial w}f_{w}(-\infty)
=πfz(0)(fw(0)fw())=πfz(0)fw(0).\displaystyle=-\sqrt{\pi}\,f_{z}^{\prime}(0)\Big{(}f^{\prime}_{w}(0)-f^{\prime}_{w}(-\infty)\Big{)}=-\sqrt{\pi}\,f_{z}^{\prime}(0)f_{w}^{\prime}(0).

For the first one, note that

zI(z,w)\displaystyle\frac{\partial}{\partial z}\textup{I}(z,w) =2πz[e(zw)20fz(u)fw(u)𝑑u]\displaystyle=2\sqrt{\pi}\,\frac{\partial}{\partial z}\Big{[}e^{(z-w)^{2}}\int_{-\infty}^{0}f^{\prime}_{z}(u)f^{\prime}_{w}(u)\,du\Big{]}
=2πe(zw)2[2(zw)0fz(u)fw(u)𝑑u+0zfz(u)fw(u)𝑑u]\displaystyle=2\sqrt{\pi}\,e^{(z-w)^{2}}\Big{[}2(z-w)\int_{-\infty}^{0}f^{\prime}_{z}(u)f^{\prime}_{w}(u)\,du+\int_{-\infty}^{0}\frac{\partial}{\partial z}f^{\prime}_{z}(u)f^{\prime}_{w}(u)\,du\Big{]}
=2πe(zw)2[2(zw)0fz(u)fw(u)𝑑u0fz′′(u)fw(u)𝑑u].\displaystyle=2\sqrt{\pi}\,e^{(z-w)^{2}}\Big{[}2(z-w)\int_{-\infty}^{0}f^{\prime}_{z}(u)f^{\prime}_{w}(u)\,du-\int_{-\infty}^{0}f^{\prime\prime}_{z}(u)f^{\prime}_{w}(u)\,du\Big{]}.

Thus we have

zI(z,w)|w=z\displaystyle\frac{\partial}{\partial z}\textup{I}(z,w)\Big{|}_{w=z} =2π0fz′′(u)fz(u)𝑑u=2πfz(0)2zI(z,w)|w=z.\displaystyle=-2\sqrt{\pi}\,\int_{-\infty}^{0}f^{\prime\prime}_{z}(u)f^{\prime}_{z}(u)\,du=-2\sqrt{\pi}\,f^{\prime}_{z}(0)^{2}-\frac{\partial}{\partial z}\textup{I}(z,w)\Big{|}_{w=z}.

Therefore all we need to show is

(4.57) 2zwI(z,w)=2π(zw)e(zw)2fz(0)fw(0)+πe(zw)2fz(0)fw′′(0)=2π(zw)e(zw)2fz(0)fw(0)+πe(zw)2fz′′(0)fw(0)=π2e(zw)2(fz(0)fw′′(0)+fz′′(0)fw(0)).\displaystyle\begin{split}\frac{\partial^{2}}{\partial z\partial w}\textup{I}(z,w)&=2\sqrt{\pi}(z-w)e^{(z-w)^{2}}f_{z}^{\prime}(0)f_{w}^{\prime}(0)+\sqrt{\pi}\,e^{(z-w)^{2}}\,f_{z}^{\prime}(0)f_{w}^{\prime\prime}(0)\\ &=-2\sqrt{\pi}(z-w)e^{(z-w)^{2}}f_{z}^{\prime}(0)f_{w}^{\prime}(0)+\sqrt{\pi}\,e^{(z-w)^{2}}\,f_{z}^{\prime\prime}(0)f_{w}^{\prime}(0)\\ &=\frac{\sqrt{\pi}}{2}e^{(z-w)^{2}}\Big{(}f_{z}^{\prime}(0)f_{w}^{\prime\prime}(0)+f_{z}^{\prime\prime}(0)f_{w}^{\prime}(0)\Big{)}.\end{split}

This follows from straightforward computations using integration by parts. ∎

We now prove Proposition 4.8.

Proof of Proposition 4.8.

By Lemmas 4.10 and  4.11, it suffices to show the proposition for some initial values.

As above, for the function I, we choose the initial value z=w¯z=\bar{w}. Then the value of I(w¯,w)\textup{I}(\bar{w},w) follows from [19, 3, 11].

Therefore again, it remains to show IIN(z,w)II(z,w)\textup{II}_{N}(z,w)\to\textup{II}(z,w) for some ww. For w=0w=0, we shall show that

(4.58) ez2IIN(z,0)4πc2Ai(2cz+c44)0ec3(zu)+c66Ai(2cu+c44)𝑑u.e^{-z^{2}}\textup{II}_{N}(z,0)\to 4\sqrt{\pi}c^{2}\operatorname{Ai}\Big{(}2cz+\frac{c^{4}}{4}\Big{)}\int_{-\infty}^{0}e^{c^{3}(z-u)+\frac{c^{6}}{6}}\operatorname{Ai}\Big{(}-2cu+\frac{c^{4}}{4}\Big{)}\,du.

For this purpose, we again use the expression (4.29). After inserting (2.10), the limit of the prefactor is easy to compute and for the Hermite polynomials in zz we use the asymptotics from Lemma 4.9. Only the sum TN(pN)T_{N}(p_{N}) with pN=2(1+τ)p_{N}=\sqrt{2}(1+\tau) remains for which we use the following ansatz:

(4.59) TN(pN)=TN(0)+0rNTN(t)𝑑t+rNpNTN(t)𝑑t,rN:=pN(2N)23+α,T_{N}(p_{N})=T_{N}(0)+\int_{0}^{r_{N}}T^{\prime}_{N}(t)\,dt+\int_{r_{N}}^{p_{N}}T^{\prime}_{N}(t)\,dt,\qquad r_{N}:=p_{N}-(2N)^{-\frac{2}{3}+\alpha},

where 0<α<4150<\alpha<\frac{4}{15}. Note that by Lemma 4.2, we have

(4.60) TN(t)=N2ττN1+τ14N1(N1)!eN2(1+τ)t2H2N1(N2τt).T^{\prime}_{N}(t)=-\sqrt{\frac{N}{2\tau}}\frac{\tau^{N}}{1+\tau}\frac{1}{4^{N-1}(N-1)!}e^{-\frac{N}{2(1+\tau)}t^{2}}H_{2N-1}\Big{(}\sqrt{\tfrac{N}{2\tau}}t\Big{)}.

Recall also that by (4.32), we have limNTN(0)=12\lim_{N\to\infty}T_{N}(0)=\frac{1}{\sqrt{2}}.

In the bulk t[0,rN]t\in[0,r_{N}] we use the Hermite asymptotics from the oscillatory regime

(4.61) H2N1(N2τt)=(4t22)1422NNN12exp(N+N4t2+c28t2(2N)23+c48t2(2N)13+c68t2)O(1)H_{2N-1}\Big{(}\sqrt{\tfrac{N}{2\tau}}t\Big{)}=\Big{(}4-\tfrac{t^{2}}{2}\Big{)}^{-\frac{1}{4}}2^{2N}N^{N-\frac{1}{2}}\exp\Big{(}-N+\tfrac{N}{4}t^{2}+\tfrac{c^{2}}{8}t^{2}(2N)^{\frac{2}{3}}+\tfrac{c^{4}}{8}t^{2}(2N)^{\frac{1}{3}}+\tfrac{c^{6}}{8}t^{2}\Big{)}O(1)

where the O(1)O(1)-term is uniform because α>0\alpha>0, cf. [53, Eq. (3.3), (3.5)]. Hence the derivative has the asymptotic form

(4.62) TN(t)=N(1t28)14exp(c22(1t28)(2N)23c44(13t28)(2N)13c68(437t28))O(1).T^{\prime}_{N}(t)=-\sqrt{N}\Big{(}1-\tfrac{t^{2}}{8}\Big{)}^{-\frac{1}{4}}\exp\Big{(}-\tfrac{c^{2}}{2}(1-\tfrac{t^{2}}{8})(2N)^{\frac{2}{3}}-\tfrac{c^{4}}{4}(1-\tfrac{3t^{2}}{8})(2N)^{\frac{1}{3}}-\tfrac{c^{6}}{8}(\tfrac{4}{3}-\tfrac{7t^{2}}{8})\Big{)}O(1).

Thus we conclude that there exists a constant C>0C>0 such that for all sufficiently large NN and all t[0,rN]t\in[0,r_{N}] the following inequality holds

(4.63) |TN(t)|CNc(2N)112exp(c222Nα+c612).\lvert T^{\prime}_{N}(t)\rvert\leq C\frac{\sqrt{N}}{\sqrt{c}}(2N)^{\frac{1}{12}}\exp\Big{(}-\frac{c^{2}}{2\sqrt{2}}N^{\alpha}+\frac{c^{6}}{12}\Big{)}.

Since α>0\alpha>0 we obtain 0rNTN(t)𝑑t0\int_{0}^{r_{N}}T^{\prime}_{N}(t)dt\to 0 for NN\to\infty.

Near the edge we set t=2(1+τ)+(2N)23st=\sqrt{2}(1+\tau)+(2N)^{-\frac{2}{3}}s. Then by Lemma 4.9 we have

(4.64) TN(t)=213N23exp(c612+c222s)[Ai(s2+c44)+O(N23+52α)].T^{\prime}_{N}(t)=-2^{-\frac{1}{3}}N^{\frac{2}{3}}\exp\Big{(}\tfrac{c^{6}}{12}+\tfrac{c^{2}}{2\sqrt{2}}s\Big{)}\Big{[}\operatorname{Ai}\big{(}\tfrac{s}{\sqrt{2}}+\tfrac{c^{4}}{4}\big{)}+O(N^{-\frac{2}{3}+\frac{5}{2}\alpha})\Big{]}.

and the error term is again uniform because α<13\alpha<\frac{1}{3}, cf. [53, Eq. (3.11), (3.13)]. Thus we obtain that

(4.65) rNpNTN(t)𝑑t=(2N)23(2N)α0TN(pN+(2N)23s)𝑑s12ec6120ec222sAi(s2+c44)𝑑s=2cec6120ec3uAi(2cu+c44)𝑑u.\begin{split}\int_{r_{N}}^{p_{N}}T^{\prime}_{N}(t)\,dt&=(2N)^{-\frac{2}{3}}\int_{-(2N)^{\alpha}}^{0}T^{\prime}_{N}\big{(}p_{N}+(2N)^{-\frac{2}{3}}s\big{)}\,ds\\ &\sim-\frac{1}{2}e^{\frac{c^{6}}{12}}\int_{-\infty}^{0}e^{\frac{c^{2}}{2\sqrt{2}}s}\operatorname{Ai}\Big{(}\frac{s}{\sqrt{2}}+\frac{c^{4}}{4}\Big{)}\,ds=-\sqrt{2}\,c\,e^{\frac{c^{6}}{12}}\int_{0}^{\infty}e^{-c^{3}u}\operatorname{Ai}\Big{(}-2cu+\frac{c^{4}}{4}\Big{)}\,du.\end{split}

Also note that the Airy-function has the following Laplace transform [47, Eq. (9.10.13)]

(4.66) eavAi(v)𝑑v=ea33,Rea>0.\int_{-\infty}^{\infty}e^{av}\operatorname{Ai}(v)dv=e^{\frac{a^{3}}{3}},\qquad\operatorname{Re}a>0.

Therefore we can rewrite the initial value 1/21/\sqrt{2} as an integral over ec3uAi(2cu+c44)e^{-c^{3}u}\operatorname{Ai}(2cu+\frac{c^{4}}{4}) with u(,)u\in(-\infty,\infty). Combining all of the above with (4.59), we arrive at

(4.67) TN(pN)2ce2Nc22(2N)23+c6120ec3uAi(2cu+c44)𝑑u.T_{N}(p_{N})\sim\sqrt{2}c\,e^{2N-\frac{c^{2}}{2}(2N)^{\frac{2}{3}}+\frac{c^{6}}{12}}\int_{-\infty}^{0}e^{-c^{3}u}\operatorname{Ai}\Big{(}-2cu+\frac{c^{4}}{4}\Big{)}\,du.

Now, it follows from the straightforward computations of the pre-factors that the claimed formula (4.58) holds.

5. Scaling limits of the soft/hard and hard edge ensembles

In this section, we study the Ginibre ensemble with boundary confinements and prove Theorems 2.3 and  2.4. In the first subsection, we summarise the strategy of our proofs using the Laplace method. In Subsections 5.2 and  5.3, we derive the boundary scaling limits of the Ginibre ensemble with soft/hard edge (Theorem 2.3) and hard edge (Theorem 2.4) constraint respectively.

5.1. Strategy of the proof: the Laplace method

We consider the potential QQ of the form (2.15) and (2.18). Let us write

(5.1) wm(ζ):=(hm)12ζmeN|ζ|2/2w_{m}(\zeta):=(h_{m})^{-\frac{1}{2}}\zeta^{m}e^{-N|\zeta|^{2}/2}

for the weighted orthonormal polynomial, where hmh_{m} is the orthogonal norm given by (3.11). Since QQ is radially symmetric, it follows from (3.12), (3.4) and (3.7) that the rescaled and weighted pre-kernel κN\kappa_{N} can be written as

(5.2) κN(z,w)eN(|ζ|2+|η|2)/2=GN(z,w)GN(w,z),\kappa_{N}(z,w)e^{-N(|\zeta|^{2}+|\eta|^{2})/2}=G_{N}(z,w)-G_{N}(w,z),

where

(5.3) GN(z,w):=γN32k=0N1w2k+1(ζ)[(h2kh2k+1)12w2k(η)+l=0k1(h2lh2k+1)12j=0kl1h2l+2j+2h2l+2j+1w2l(η)].G_{N}(z,w):=\frac{\gamma_{N}^{3}}{2}\sum_{k=0}^{N-1}w_{2k+1}(\zeta)\Big{[}\Big{(}\frac{h_{2k}}{h_{2k+1}}\Big{)}^{\frac{1}{2}}w_{2k}(\eta)+\sum_{l=0}^{k-1}\Big{(}\frac{h_{2l}}{h_{2k+1}}\Big{)}^{\frac{1}{2}}\prod_{j=0}^{k-l-1}\frac{h_{2l+2j+2}}{h_{2l+2j+1}}w_{2l}(\eta)\Big{]}.

Here and in the sequel, we denote ζ=p+γNz\zeta=p+\gamma_{N}z and η=p+γNw\eta=p+\gamma_{N}w, where γN\gamma_{N} is the microscopic scale given by (2.1) for the soft/hard edge case and by (2.20) for the hard edge case. Throughout this section, we write

(5.4) δN=N12logN,τm=mN.\delta_{N}=N^{-\frac{1}{2}}\log N,\qquad\tau_{m}=\frac{m}{N}.

Recall also that Dρ:={z:|z|2ρ}D_{\rho}:=\{z\in\mathbb{C}:|z|\leq\sqrt{2}\rho\}.

The general strategy of deriving the large-NN behaviour of GNG_{N} is as follows:

  • for m{1,2,,2N1}m\in\{1,2,\dots,2N-1\}, we compute the asymptotic behaviours of wmw_{m} by means of Laplace’s method;

  • we show that in the case of soft/hard edge ensemble, the dominant terms of GNG_{N} near D1\partial D_{1} consist of wmw_{m} with 2(1δN)τm22(1-\delta_{N})\leq\tau_{m}\leq 2, whereas in the case of hard edge ensemble, those near Dρ\partial D_{\rho} consist of wmw_{m} with 2(ρ2+δN)τm22(\rho^{2}+\delta_{N})\leq\tau_{m}\leq 2;

  • by discarding the lower degree terms, we compile the contributions of dominant terms using the Riemann sum approximation.

We now present some asymptotic behaviours of hmh_{m}. For this purpose, we define

(5.5) Vτ(r):=r22τlogr,r>0.V_{\tau}(r):=r^{2}-2\tau\log r,\qquad r>0.

Observe that the function VτV_{\tau} has a unique critical point at rτ:=τr_{\tau}:=\sqrt{\tau}. It is convenient to write

(5.6) Φ(x):=xet2𝑑t=π2erfc(x).\Phi(x):=\int_{x}^{\infty}e^{-t^{2}}\,dt=\frac{\sqrt{\pi}}{2}\operatorname{erfc}(x).

We first obtain the following. (See [52, Lemmas 3.1, 3.3]) for a related statement.)

Lemma 5.1.

For a given ρ(0,1]\rho\in(0,1], let Q(ζ)=|ζ|2+1DρQ(\zeta)=|\zeta|^{2}+\infty\cdot 1_{\mathbb{C}\setminus D_{\rho}}. Then we have the following.

  • (i)

    For each mm with |τm2ρ2|<2δN|\tau_{m}-2\rho^{2}|<2\delta_{N}, we have

    (5.7) hm=em(mN)m2ρNΦ(ξm)(1+o(1)),ξm:=N2ρ(τm2ρ2),h_{m}=e^{-m}\Big{(}\frac{m}{N}\Big{)}^{m}\frac{2\rho}{\sqrt{N}}\,\Phi(\xi_{m})(1+o(1)),\qquad\xi_{m}:=\tfrac{\sqrt{N}}{2\rho}(\tau_{m}-2\rho^{2}),

    where o(1)0o(1)\to 0 uniformly for mm as NN\to\infty.

  • (ii)

    For each mm with 2(ρ2+δN)τm22(\rho^{2}+\delta_{N})\leq\tau_{m}\leq 2, we have

    (5.8) hm=(2ρ2)m+1N(mN2ρ2)e2ρ2N(1+o(1)),h_{m}=\frac{(2\rho^{2})^{m+1}}{N(\frac{m}{N}-2\rho^{2})}e^{-2\rho^{2}N}(1+o(1)),

    where o(1)0o(1)\to 0 uniformly for mm as NN\to\infty.

Proof.

For mm with |τm2ρ2|<2δN|\tau_{m}-2\rho^{2}|<2\delta_{N}, the orthogonal norm hmh_{m} can be written as

(5.9) hm:=Dρ|ζ|2meN|ζ|2𝑑A(ζ)=2rτmδN2ρeNVτm(r)r𝑑r+20rτmδNeNVτm(r)r𝑑r,h_{m}:=\int_{D_{\rho}}|\zeta|^{2m}e^{-N|\zeta|^{2}}\,dA(\zeta)=2\int_{r_{\tau_{m}}-\delta_{N}^{\prime}}^{\sqrt{2}\rho}e^{-NV_{\tau_{m}}(r)}r\,dr+2\int_{0}^{r_{\tau_{m}}-\delta_{N}^{\prime}}e^{-NV_{\tau_{m}}(r)}r\,dr,

where δN:=N12(logN)2\delta_{N}^{\prime}:=N^{-\frac{1}{2}}(\log N)^{2}. By the Taylor expansion

(5.10) Vτ(r)=Vτ(rτ)+2(rrτ)2+O(rrτ)3,rrτV_{\tau}(r)=V_{\tau}(r_{\tau})+2(r-r_{\tau})^{2}+O(r-r_{\tau})^{3},\qquad r\to r_{\tau}

at the critical point r=rτr=r_{\tau}, we obtain that as NN\to\infty,

(5.11) 2rτmδN2ρreNVτm(r)𝑑r=2eNVτm(rτm)(2ρ+O(δN))rτmδN2ρe2N(rrτm)2+O(NδN3)𝑑r.2\int_{r_{\tau_{m}}-\delta_{N}^{\prime}}^{\sqrt{2}\rho}re^{-NV_{\tau_{m}}(r)}\,dr=2e^{-NV_{\tau_{m}}(r_{\tau_{m}})}(\sqrt{2}\rho+O(\delta_{N}^{\prime}))\int_{r_{\tau_{m}}-\delta_{N}^{\prime}}^{\sqrt{2}\rho}e^{-2N(r-r_{\tau_{m}})^{2}+O(N\delta_{N}^{\prime 3})}\,dr.

Here, OO-constant can be taken independent of mm. Note that by (5.4),

rτm2ρ=τm2ρ=122ρ(τm2ρ2)+O(δN2),N.r_{\tau_{m}}-\sqrt{2}\rho=\sqrt{\tau_{m}}-\sqrt{2}\rho=\frac{1}{2\sqrt{2}\rho}(\tau_{m}-2\rho^{2})+O(\delta_{N}^{2}),\qquad N\to\infty.

Therefore, by the change of variable t=2N(rrτm)t=\sqrt{2N}(r-r_{\tau_{m}}), we have

(5.12) rτmδN2ρe2N(rrτm)2𝑑r=12Nξmet2𝑑t(1+o(1))\int_{r_{\tau_{m}}-\delta_{N}^{\prime}}^{\sqrt{2}\rho}e^{-2N(r-r_{\tau_{m}})^{2}}\,dr=\frac{1}{\sqrt{2N}}\int_{\xi_{m}}^{\infty}e^{-t^{2}}\,dt\,(1+o(1))

as NN\to\infty. Since Vτm(rτm)=mNmNlogmNV_{\tau_{m}}(r_{\tau_{m}})=\frac{m}{N}-\frac{m}{N}\log\frac{m}{N}, it follows from (5.11) and (5.12) that

(5.13) 2rτmδN2ρeNVτm(r)r𝑑r=em(mN)m2ρNΦ(ξm)(1+o(1)),2\int_{r_{\tau_{m}}-\delta_{N}^{\prime}}^{\sqrt{2}\rho}e^{-NV_{\tau_{m}}(r)}r\,dr=e^{-m}\Big{(}\frac{m}{N}\Big{)}^{m}\,\frac{2\rho}{\sqrt{N}}\,\Phi(\xi_{m})\,(1+o(1)),

where o(1)0o(1)\to 0 uniformly for mm as NN\to\infty.

On the other hand, the second integral on the right-hand side of (5.9) is negligible. To see this, observe first that Vτm(rτm)=0V^{\prime}_{\tau_{m}}(r_{\tau_{m}})=0 and Vτm(r)<0V^{\prime}_{\tau_{m}}(r)<0 for r<rτmr<r_{\tau_{m}}. Thus for r<rτmδNr<r_{\tau_{m}}-\delta_{N}^{\prime}

Vτm(r)Vτm(rτmδN)=Vτm(rτm)+CδN2V_{\tau_{m}}(r)\geq V_{\tau_{m}}(r_{\tau_{m}}-\delta_{N}^{\prime})=V_{\tau_{m}}(r_{\tau_{m}})+C\delta_{N}^{\prime 2}

for some positive constant CC. This gives that as N,N\to\infty,

(5.14) 20rτmδNeNVτm(r)r𝑑rCeNVτm(rτm)eCNδN2=em(mN)mO(eC(logN)4).2\int_{0}^{r_{\tau_{m}}-\delta_{N}^{\prime}}e^{-NV_{\tau_{m}}(r)}r\,dr\leq C^{\prime}e^{-NV_{\tau_{m}}(r_{\tau_{m}})}e^{-CN\delta_{N}^{\prime 2}}=e^{-m}\Big{(}\frac{m}{N}\Big{)}^{m}O(e^{-C(\log N)^{4}}).

We have shown the first assertion of the lemma.

Next, we show the second assertion. The proof is similar to that of Lemma 5.1 (i). A key observation is that for mm with 2N(ρ2+δN)m2N2N(\rho^{2}+\delta_{N})\leq m\leq 2N

hm\displaystyle h_{m} =2ρδN2ρ2rm+1eNr2𝑑r(1+o(1))\displaystyle=\int_{\sqrt{2}\rho-\delta_{N}}^{\sqrt{2}\rho}2r^{m+1}e^{-Nr^{2}}\,dr\cdot(1+o(1))
=22ρeNVτm(2ρ)0δNe2Nρ(mN2ρ2)t𝑑t(1+o(1))=(2ρ2)m+1N(mN2ρ2)e2ρ2N(1+o(1)),\displaystyle=2\sqrt{2}\rho\,e^{-NV_{\tau_{m}}(\sqrt{2}\rho)}\int_{0}^{\delta_{N}}e^{-\frac{\sqrt{2}N}{\rho}(\frac{m}{N}-2\rho^{2})t}\,dt\cdot(1+o(1))=\frac{(2\rho^{2})^{m+1}}{N(\frac{m}{N}-2\rho^{2})}e^{-2\rho^{2}N}\cdot(1+o(1)),

which completes the proof. ∎

5.2. Non-Hermitian soft/hard edge scaling limit

In this subsection, we prove Theorem 2.3. We first show the following proposition (here ={z:Rez<0}\mathbb{H}_{-}=\{z\in\mathbb{C}:\mathrm{Re}z<0\} denotes the left half plane).

Proposition 5.2.

There exists a sequence of cocycles (cN)N1(c_{N})_{N\geq 1} such that

limNcN(z,w)GN(z,w)=2ez2|z|2+w2|w|20e2(zs)2Φ(2s)se2(wt)2Φ(2t)𝑑t𝑑s,\lim_{N\to\infty}c_{N}(z,w)G_{N}(z,w)=2e^{z^{2}-|z|^{2}+w^{2}-|w|^{2}}\int_{-\infty}^{0}\frac{e^{-2(z-s)^{2}}}{\sqrt{\Phi(2s)}}\int_{-\infty}^{s}\frac{e^{-2(w-t)^{2}}}{\sqrt{\Phi(2t)}}dt\,ds,

where the convergence is uniform for z,wz,w in compact subsets of \mathbb{H}_{-}. Here, the function Φ\Phi is given by (5.6).

Using Proposition 5.2, we first complete the proof of Theorem 2.3.

Proof of Theorem 2.3. .

Recall that the potential QQ is given by (2.15). Then by combining Proposition 5.2 with (5.2) and (3.5), the theorem follows. Here, we again use the fact that a sequence of cocycles cancels out when forming a Pfaffian. ∎

We first discard the negligible terms in the sum (5.3). Recall that δN\delta_{N} and τm\tau_{m} are given by (5.4).

Lemma 5.3.

For mm with |τm2|2δN|\tau_{m}-2|\leq 2\delta_{N} we have

(5.15) wm(ζ)=(N4)14(Φ(ξm))12e2iNImzez2|z|2e2(zξm/2)2(1+o(1)),w_{m}(\zeta)=\Big{(}\frac{N}{4}\Big{)}^{\frac{1}{4}}(\Phi(\xi_{m}))^{-\frac{1}{2}}e^{2i\sqrt{N}\operatorname{Im}z}e^{z^{2}-|z|^{2}}e^{-2(z-\xi_{m}/2)^{2}}(1+o(1)),

where ξm=N(m2N1)\xi_{m}=\sqrt{N}(\frac{m}{2N}-1). Here, o(1)0o(1)\to 0 uniformly for all zz in any compact set as NN\to\infty.

Proof.

By Lemma 5.1 (i), we have

wm(ζ)=(N4)14(Φ(ξm))12eNVτm(rτm)/2eN(|ζ|22mNlogζ)/2(1+o(1)),w_{m}(\zeta)=\Big{(}\frac{N}{4}\Big{)}^{\frac{1}{4}}(\Phi(\xi_{m}))^{-\frac{1}{2}}e^{NV_{\tau_{m}}(r_{\tau_{m}})/2}e^{-N(|\zeta|^{2}-\frac{2m}{N}\log\zeta)/2}(1+o(1)),

where VτV_{\tau} is given in (5.5) and rτm=mNr_{\tau_{m}}=\sqrt{\frac{m}{N}}. Since

|ζ|2=|p+γNz|2=(p+γNz)2+γN2(|z|2z2)2ipγNImz=ζ2+2N(|z|2z2)4iNImz,|\zeta|^{2}=|p+\gamma_{N}z|^{2}=(p+\gamma_{N}z)^{2}+\gamma_{N}^{2}(|z|^{2}-z^{2})-2ip\gamma_{N}\operatorname{Im}z=\zeta^{2}+\frac{2}{N}(|z|^{2}-z^{2})-\frac{4i}{\sqrt{N}}\operatorname{Im}z,

we obtain that

eN(|ζ|22mNlogζ)/2=ez2|z|2e2iNImzeN(ζ22mNlogζ)/2.\displaystyle e^{-N(|\zeta|^{2}-\frac{2m}{N}\log\zeta)/2}=e^{z^{2}-|z|^{2}}e^{2i\sqrt{N}\operatorname{Im}z}e^{-N(\zeta^{2}-\frac{2m}{N}\log\zeta)/2}.

Now the lemma follows from straightforward computations using the Taylor series expansion at r=rτmr=r_{\tau_{m}}

ζ22mNlogζ=Vτm(rτm)+2(ζrτm)2+O(|ζrτm|3)\zeta^{2}-\frac{2m}{N}\log\zeta=V_{\tau_{m}}(r_{\tau_{m}})+2(\zeta-r_{\tau_{m}})^{2}+O(|\zeta-r_{\tau_{m}}|^{3})

and the following asymptotic expansion

ζrτm=p+γNzrτm=2Nz+2(1m2N)=2Nz+12(1m2N)+O(δN2).\zeta-r_{\tau_{m}}=p+\gamma_{N}z-r_{\tau_{m}}=\sqrt{\frac{2}{N}}z+\sqrt{2}\Big{(}1-\sqrt{\frac{m}{2N}}\Big{)}=\sqrt{\frac{2}{N}}z+\frac{1}{\sqrt{2}}\Big{(}1-{\frac{m}{2N}}\Big{)}+O(\delta_{N}^{2}).

Lemma 5.4.

Let KK be a compact subset of \mathbb{H}_{-}. Then there exist positive constants cc and CC such that for all mm with m2N(1δN)m\leq 2N(1-\delta_{N}) and for all zKz\in K,

(5.16) |wm(2+γNz)|2Cec(logN)2.|w_{m}(\sqrt{2}+\gamma_{N}z)|^{2}\leq Ce^{-c(\log N)^{2}}.
Proof.

For m2N(1δN)m\leq 2N(1-\delta_{N}), the critical point rτmr_{\tau_{m}} of VτmV_{\tau_{m}} satisfies

(5.17) rτm=τm2(1δN)=2(112δN)+O(δN2),N.r_{\tau_{m}}=\sqrt{\tau_{m}}\leq\sqrt{2(1-\delta_{N})}=\sqrt{2}\Big{(}1-\frac{1}{2}\delta_{N}\Big{)}+O(\delta_{N}^{2}),\qquad N\to\infty.

Choose a constant c1>0c_{1}>0 such that rτm22c1δNr_{\tau_{m}}\leq\sqrt{2}-2c_{1}\delta_{N}. Since Vτm(r)0V^{\prime}_{\tau_{m}}(r)\geq 0 for rrτmr\geq r_{\tau_{m}}, we obtain

(5.18) hm22c1δN2c1δN2reNVτm(r)𝑑rC1δNeNVτm(2c1δN)h_{m}\geq\int_{\sqrt{2}-2c_{1}\delta_{N}}^{\sqrt{2}-c_{1}\delta_{N}}2re^{-NV_{\tau_{m}}(r)}dr\geq C_{1}\delta_{N}e^{-NV_{\tau_{m}}(\sqrt{2}-c_{1}\delta_{N})}

for some constant C1C_{1}. By (5.1), this gives that for zKz\in K,

|wm(2+γNz)|2C2δN1eN(Vτm(|2+γNz|)Vτm(2c1δN))Cec(logN)2|w_{m}(\sqrt{2}+\gamma_{N}z)|^{2}\leq C_{2}\delta_{N}^{-1}e^{-N(V_{\tau_{m}}(|\sqrt{2}+\gamma_{N}z|)-V_{\tau_{m}}(\sqrt{2}-c_{1}\delta_{N}))}\leq Ce^{-c(\log N)^{2}}

for some constants cc and CC. ∎

Lemma 5.4 asserts that for m2N(1δN)m\leq 2N(1-\delta_{N}), wmw_{m} is negligible in the sum (5.3). Indeed, we obtain the following lemma.

Lemma 5.5.

Let KK be a compact subset of \mathbb{H}_{-}. Then for z,wKz,w\in K, we have

(5.19) GN(z,w)γN32k=N(1δN)N1w2k+1(ζ)[(h2kh2k+1)12w2k(η)+l=N(1δN)k1(h2lh2k+1)12j=0kl1h2l+2j+2h2l+2j+1w2l(η)].G_{N}(z,w)\sim\frac{\gamma_{N}^{3}}{2}\sum_{k=N(1-\delta_{N})}^{N-1}w_{2k+1}(\zeta)\Big{[}\Big{(}\frac{h_{2k}}{h_{2k+1}}\Big{)}^{\frac{1}{2}}w_{2k}(\eta)+\sum_{l=N(1-\delta_{N})}^{k-1}\Big{(}\frac{h_{2l}}{h_{2k+1}}\Big{)}^{\frac{1}{2}}\prod_{j=0}^{k-l-1}\frac{h_{2l+2j+2}}{h_{2l+2j+1}}w_{2l}(\eta)\Big{]}.
Proof.

We first mention that the number of summands NδN=N1/2logNN\delta_{N}=N^{1/2}\log N\to\infty as NN\to\infty. For each mm, the norm hmh_{m} can be expressed in terms of the lower incomplete function γ\gamma as

hm=022r2m+1eNr2𝑑r=N(m+1)γ(m+1,2N).\displaystyle h_{m}=\int_{0}^{\sqrt{2}}2r^{2m+1}e^{-Nr^{2}}\,dr=N^{-(m+1)}\gamma(m+1,2N).

It is well known that if XNX_{N} is a Poisson(2N)\mathrm{Poisson}(2N) random variable, then

Γ(m+1)γ(m+1,2N)Γ(m+1)=𝐏[XN<m+1].\frac{\Gamma(m+1)-\gamma(m+1,2N)}{\Gamma(m+1)}=\mathbf{P}[X_{N}<m+1].

For all m<2N(1δN)m<2N(1-\delta_{N}), the normal approximation of Poisson random variables gives

𝐏[XNm]𝐏[XN2N2N2logN]=𝐏[Z2logN](1+o(1)),\mathbf{P}[X_{N}\leq m]\leq\mathbf{P}\Big{[}\frac{X_{N}-2N}{\sqrt{2N}}\leq-\sqrt{2}\log N\Big{]}=\mathbf{P}[Z\leq-\sqrt{2}\log N](1+o(1)),

where ZZ is a standard normal random variable. Note that 𝐏[Z2logN]=O(ec(logN)2)\mathbf{P}[Z\leq-2\log N]=O(e^{-c(\log N)^{2}}) for some constant cc. This implies that there exists a constant cc^{\prime} such that

(5.20) hmhm+1=Nm+1(1+O(ec(logN)2)),hmhm+1hm+2hm+1=m+2m+1(1+O(ec(logN)2)),\frac{h_{m}}{h_{m+1}}=\frac{N}{m+1}(1+O(e^{-c^{\prime}(\log N)^{2}})),\qquad\frac{h_{m}}{h_{m+1}}\frac{h_{m+2}}{h_{m+1}}=\frac{m+2}{m+1}(1+O(e^{-c^{\prime}(\log N)^{2}})),

where O(ec(logN)2)0O(e^{-c^{\prime}(\log N)^{2}})\to 0 uniformly for mm as NN\to\infty. This implies that hm/hm+1=O(N){h_{m}}/{h_{m+1}}=O(N) and for k<N(1δN)k<N(1-\delta_{N})

(5.21) h2lh2k+1j=0kl1h2l+2j+2h2l+2j+1=h2kh2k+1j=0kl1h2l+2j+2h2l+2j+1h2l+2jh2l+2j+1=O(N).\sqrt{\frac{{h_{2l}}}{{h_{2k+1}}}}\prod_{j=0}^{k-l-1}\frac{h_{2l+2j+2}}{h_{2l+2j+1}}=\sqrt{\frac{h_{2k}}{h_{2k+1}}}\prod_{j=0}^{k-l-1}\sqrt{\frac{h_{2l+2j+2}}{h_{2l+2j+1}}}\sqrt{\frac{h_{2l+2j}}{h_{2l+2j+1}}}=O(N).

Combining all of the above with Lemma 5.4, we obtain

(5.22) k=0N(1δN)w2k+1(ζ)[(h2kh2k+1)12w2k(η)+l=0k1(h2lh2k+1)12j=0kl1h2l+2j+2h2l+2j+1w2l(η)]=o(1)\sum_{k=0}^{N(1-\delta_{N})}w_{2k+1}(\zeta)\Big{[}\Big{(}\frac{h_{2k}}{h_{2k+1}}\Big{)}^{\frac{1}{2}}w_{2k}(\eta)+\sum_{l=0}^{k-1}\Big{(}\frac{h_{2l}}{h_{2k+1}}\Big{)}^{\frac{1}{2}}\prod_{j=0}^{k-l-1}\frac{h_{2l+2j+2}}{h_{2l+2j+1}}w_{2l}(\eta)\Big{]}=o(1)

uniformly for z,wKz,w\in K. Similarly, we obtain

(5.23) k=N(1δN)N1w2k+1(ζ)l=0N(1δN)(h2lh2k+1)12j=0kl1h2l+2j+2h2l+2j+1w2l(η)=o(1),\sum_{k=N(1-\delta_{N})}^{N-1}w_{2k+1}(\zeta)\sum_{l=0}^{N(1-\delta_{N})}\Big{(}\frac{h_{2l}}{h_{2k+1}}\Big{)}^{\frac{1}{2}}\prod_{j=0}^{k-l-1}\frac{h_{2l+2j+2}}{h_{2l+2j+1}}w_{2l}(\eta)=o(1),

which completes the proof. ∎

We now show the following.

Lemma 5.6.

For k,lk,l with N(1δN)k,lN1N(1-\delta_{N})\leq k,l\leq N-1, we have

(5.24) 1h2k+1=12h2k(1+o(1)),1h2k+1j=0kl1h2l+2j+2h2l+2j+1=12h2l(1+o(1)).\frac{1}{\sqrt{h_{2k+1}}}=\frac{1}{\sqrt{2h_{2k}}}(1+o(1)),\qquad\frac{1}{\sqrt{h_{2k+1}}}\prod_{j=0}^{k-l-1}\frac{h_{2l+2j+2}}{h_{2l+2j+1}}=\frac{1}{\sqrt{2h_{2l}}}(1+o(1)).
Proof.

Similar to the proof of Lemma 5.5, the ratio of the norms can be estimated by

(5.25) hm+1hm=m+1N1𝐏[XNm+1]1𝐏[XNm]=m+1N(1Φ(ξm)Φ(ξm+1)Φ(ξm)(1+o(1)))\frac{h_{m+1}}{h_{m}}=\frac{m+1}{N}\cdot\frac{1-\mathbf{P}[X_{N}\leq m+1]}{1-\mathbf{P}[X_{N}\leq m]}=\frac{m+1}{N}\Big{(}1-\frac{\Phi(\xi_{m})-\Phi(\xi_{m+1})}{\Phi(\xi_{m})}(1+o(1))\Big{)}

as m,Nm,N\to\infty with |m2N|2NδN|m-2N|\leq 2N\delta_{N}. Here, XNPoisson(2N)X_{N}\sim\mathrm{Poisson}(2N) and Φ\Phi, ξm\xi_{m} are defined in (5.6) and Lemma 5.1 (i). Since there exists cm(ξm,ξm+1)c_{m}\in(\xi_{m},\xi_{m+1}) such that

(5.26) Φ(ξm)Φ(ξm+1)=12Necm2\Phi(\xi_{m})-\Phi(\xi_{m+1})=\frac{1}{2\sqrt{N}}e^{-c_{m}^{2}}

by the mean value theorem, we have for kk with N(1δN)kN1N(1-\delta_{N})\leq k\leq N-1,

(5.27) h2k+1h2k=2(1+o(1)),\frac{h_{2k+1}}{h_{2k}}=2\cdot(1+o(1)),

where o(1)0o(1)\to 0 uniformly for mm as NN\to\infty. This gives the first assertion.

For the second asymptotic equation, we obtain from (5.25) and (5.26) that

hmhm+1hm+2hm+1=1+O(N1)\frac{h_{m}}{h_{m+1}}\frac{h_{m+2}}{h_{m+1}}=1+O(N^{-1})

for mm with 2N(1δN)m2N2N(1-\delta_{N})\leq m\leq 2N. This gives

1h2k+1j=0kl1h2l+2j+2h2l+2j+1=h2kh2k+1h2lj=0kl1h2l+2j+2h2l+2j+1h2l+2jh2l+2j+1=12h2l(1+o(1)),\frac{1}{\sqrt{h_{2k+1}}}\prod_{j=0}^{k-l-1}\frac{h_{2l+2j+2}}{h_{2l+2j+1}}=\sqrt{\frac{h_{2k}}{h_{2k+1}h_{2l}}}\prod_{j=0}^{k-l-1}\sqrt{\frac{h_{2l+2j+2}}{h_{2l+2j+1}}}\sqrt{\frac{h_{2l+2j}}{h_{2l+2j+1}}}=\frac{1}{\sqrt{2h_{2l}}}(1+o(1)),

which yields the desired equation. ∎

We are now ready to show Proposition 5.2.

Proof of Proposition 5.2.

By Lemmas 5.3, 5.5 and  5.6, we have

(5.28) GN(z,w)e2iNIm(z+w)2Nez2|z|2+w2|w|2k=N(1δN)N1e2(zξ2k+1/2)2Φ(ξ2k+1)l=N(1δN)ke2(wξ2l/2)2Φ(ξ2l).G_{N}(z,w)\sim\frac{e^{2i\sqrt{N}\operatorname{Im}(z+w)}}{2N}e^{z^{2}-|z|^{2}+w^{2}-|w|^{2}}\sum_{k=N(1-\delta_{N})}^{N-1}\frac{e^{-2(z-\xi_{2k+1}/2)^{2}}}{\sqrt{\Phi(\xi_{2k+1})}}\sum_{l=N(1-\delta_{N})}^{k}\frac{e^{-2(w-\xi_{2l}/2)^{2}}}{\sqrt{\Phi(\xi_{2l})}}.

Then the proposition follows from the Riemann sum approximation. ∎

5.3. Non-Hermitian hard edge scaling limit

This subsection is devoted to the proof of Theorem 2.4. As in the previous subsection, by (5.2), it is enough to show the following proposition.

Proposition 5.7.

There exists a sequence of cocycles (cN)N1(c_{N})_{N\geq 1} such that

limNcN(z,w)GN(z,w)=01s12e2sz0st12e2tw𝑑t𝑑s,\lim_{N\to\infty}c_{N}(z,w)\,G_{N}(z,w)=\int_{0}^{1}s^{\frac{1}{2}}e^{2sz}\int_{0}^{s}t^{\frac{1}{2}}e^{2tw}\,dt\,ds,

where the convergence is uniform for z,wz,w in any compact subset of \mathbb{H}_{-}.

We first show the following lemma.

Lemma 5.8.

For each mm with (2ρ2+2δN)N<m<2N(2\rho^{2}+2\delta_{N})N<m<2N and for all zz in a compact set, we have

wm(ζ)=gN(z)ρN(m2Nρ2)e21ρ2(m2Nρ2)z(1+o(1)),w_{m}(\zeta)=\frac{g_{N}(z)}{\rho}\sqrt{N\Big{(}\frac{m}{2N}-\rho^{2}\Big{)}}e^{\frac{2}{1-\rho^{2}}(\frac{m}{2N}-\rho^{2})z}(1+o(1)),

where gN(z)g_{N}(z) is a unimodular function.

Proof.

By Lemma 5.1 (ii), we obtain

wm(ζ)=1ρN(m2Nρ2)eN(|ζ|22τmNlogζVτm(2ρ))/2(1+o(1)),w_{m}(\zeta)=\frac{1}{\rho}\sqrt{N\Big{(}\frac{m}{2N}-\rho^{2}\Big{)}}e^{-N(|\zeta|^{2}-2\tau_{m}{N}\log\zeta-V_{\tau_{m}}(\sqrt{2}\rho))/2}(1+o(1)),

where τm=m/N\tau_{m}=m/N and Vτm(r)=r22τmlogrV_{\tau_{m}}(r)=r^{2}-2\tau_{m}\log r defined in (5.5). Using the expansion

|2ρ+γNz|22τmlog(2ρ+γNz)Vτm(2ρ)\displaystyle\quad|\sqrt{2}\rho+\gamma_{N}z|^{2}-2\tau_{m}\log(\sqrt{2}\rho+\gamma_{N}z)-V_{\tau_{m}}(\sqrt{2}\rho)
=22ργNRez2m2ρNγNz+O(N2)=4N(1ρ2)(ρ2m2N)zi4ρ2N(1ρ2)Imz+O(N2),\displaystyle=2\sqrt{2}\rho\gamma_{N}\operatorname{Re}z-\frac{2m}{\sqrt{2}\rho N}\gamma_{N}z+O(N^{-2})=\frac{4}{N(1-\rho^{2})}\Big{(}\rho^{2}-\frac{m}{2N}\Big{)}z-i\frac{4\rho^{2}}{N(1-\rho^{2})}\operatorname{Im}z+O(N^{-2}),

we have the following asymptotic expansion:

wm(2ρ+γNz)=1ρN(m2Nρ2)ei2ρ21ρ2Imze21ρ2(m2Nρ2)z(1+o(1)).w_{m}(\sqrt{2}\rho+\gamma_{N}z)=\frac{1}{\rho}\sqrt{N\Big{(}\frac{m}{2N}-\rho^{2}\Big{)}}\,e^{i\frac{2\rho^{2}}{1-\rho^{2}}\operatorname{Im}z}e^{\frac{2}{1-\rho^{2}}(\frac{m}{2N}-\rho^{2})z}(1+o(1)).

Lemma 5.9.

Let KK be a compact subset of \mathbb{H}_{-}. Then there exists a positive constant CC such that for all mm with |τm2ρ2|<2δN|\tau_{m}-2\rho^{2}|<2\delta_{N} and for all zKz\in K,

|wm(p+γNz)|2CNlogN.|w_{m}(p+\gamma_{N}z)|^{2}\leq C\sqrt{N}\log N.
Proof.

By Lemma 5.1 (i), we have

(5.29) |wm(p+γNz)|2=(hm)1eNVτm(|p+γNz|)=N2ρΦ(ξm)eN(Vτm(|p+γNz|)Vτm(rτm))(1+o(1)),|w_{m}(p+\gamma_{N}z)|^{2}=(h_{m})^{-1}e^{-NV_{\tau_{m}}(|p+\gamma_{N}z|)}=\frac{\sqrt{N}}{2\rho\,\Phi(\xi_{m})}e^{-N(V_{\tau_{m}}(|p+\gamma_{N}z|)-V_{\tau_{m}}(r_{\tau_{m}}))}(1+o(1)),

where ξm=N2ρ(τm2ρ2)\xi_{m}=\frac{\sqrt{N}}{2\rho}(\tau_{m}-2\rho^{2}) and rτmr_{\tau_{m}} is the critical point of VτmV_{\tau_{m}}. The Taylor expansion of VτmV_{\tau_{m}} at rτmr_{\tau_{m}} gives

(5.30) |wm(p+γNz)|2=N2ρΦ(ξm)e2N(p+γNRezrτm)2(1+o(1))=N2ρΦ(ξm)eξm2(1+o(1))|w_{m}(p+\gamma_{N}z)|^{2}=\frac{\sqrt{N}}{2\rho\,\Phi(\xi_{m})}e^{-2N(p+\gamma_{N}\operatorname{Re}z-r_{\tau_{m}})^{2}}(1+o(1))=\frac{\sqrt{N}}{2\rho\,\Phi(\xi_{m})}e^{-\xi_{m}^{2}}(1+o(1))

since prτm=2ρmN=ξm2N+O(N1|ξm|2)p-r_{\tau_{m}}=\sqrt{2}\rho-\sqrt{\frac{m}{N}}=\frac{\xi_{m}}{\sqrt{2N}}+O(N^{-1}|\xi_{m}|^{2}) as NN\to\infty. Here o(1)0o(1)\to 0 uniformly for zKz\in K. Note that |ξm|=O(logN)|\xi_{m}|=O(\log N). By (5.6), it follows from the well-known asymptotics of the complementary error function (see e.g. [47, Section 7.12]) that

eξm2Φ(ξm)C(logN)1.e^{\xi_{m}^{2}}\,\Phi(\xi_{m})\geq C^{\prime}(\log N)^{-1}.

This completes the proof. ∎

As in Section 5.2, the following lemma asserts that the weighted polynomials wmw_{m} of lower degree m<2N(ρ2δN)m<2N(\rho^{2}-\delta_{N}) tend to exponentially decay in a neighbourhood of the hard edge. For the proof of the lemma, we refer to Lemma 3.5 in [52].

Lemma 5.10.

Let KK be a compact subset of \mathbb{H}_{-}. Then for mm with 0m<N(2ρ22δN)0\leq m<N(2\rho^{2}-2\delta_{N}) and for zKz\in K, we have

(5.31) |wm(p+γNz)|2Cec(logN)2),|w_{m}(p+\gamma_{N}z)|^{2}\leq Ce^{-c(\log N)^{2}}),

where cc and CC are positive constants.

As a counterpart of Lemma 5.5, we obtain the following lemma which allows us to discard the lower degree polynomials in the sum (5.3).

Lemma 5.11.

Let KK be a compact subset of \mathbb{H}_{-}. Then for z,wKz,w\in K, we have

(5.32) GN(z,w)γN32k=(ρ2+δN)NN1w2k+1(ζ)((h2kh2k+1)12w2k(η)+l=(ρ2+δN)Nk1(h2lh2k+1)12j=0kl1h2l+2j+2h2l+2j+1w2l(η)).G_{N}(z,w)\sim\frac{\gamma_{N}^{3}}{2}\!\!\sum_{k=(\rho^{2}+\delta_{N})N}^{N-1}\!w_{2k+1}(\zeta)\Big{(}\Big{(}\frac{h_{2k}}{h_{2k+1}}\Big{)}^{\frac{1}{2}}w_{2k}(\eta)+\!\!\sum_{l=(\rho^{2}+\delta_{N})N}^{k-1}\!\Big{(}\frac{h_{2l}}{h_{2k+1}}\Big{)}^{\frac{1}{2}}\prod_{j=0}^{k-l-1}\frac{h_{2l+2j+2}}{h_{2l+2j+1}}w_{2l}(\eta)\Big{)}.
Proof.

Similarly as in the proof of Lemma 5.5, we obtain for all mm with m<2N(ρ2δN)m<2N(\rho^{2}-\delta_{N})

(5.33) hmhm+1=Nγ(m+1,2ρ2N)γ(m+2,2ρ2N)=Nm+1(1+O(ec(logN)2)),hmhm+1hm+2hm+1=m+2m+1(1+O(ec(logN)2))\frac{h_{m}}{h_{m+1}}=\frac{N\,\gamma(m+1,2\rho^{2}N)}{\gamma(m+2,2\rho^{2}N)}=\frac{N}{m+1}(1+O(e^{-c(\log N)^{2}})),\quad\frac{h_{m}}{h_{m+1}}\frac{h_{m+2}}{h_{m+1}}=\frac{m+2}{m+1}(1+O(e^{-c(\log N)^{2}}))

for some constant cc. Together with Lemma 5.10, this implies for all z,wKz,w\in K

(5.34) k=0N(ρ2δN)w2k+1(ζ)[(h2kh2k+1)12w2k(η)+l=0k1(h2lh2k+1)12j=0kl1h2l+2j+2h2l+2j+1w2l(η)]=o(1)\sum_{k=0}^{N(\rho^{2}-\delta_{N})}w_{2k+1}(\zeta)\Big{[}\Big{(}\frac{h_{2k}}{h_{2k+1}}\Big{)}^{\frac{1}{2}}w_{2k}(\eta)+\sum_{l=0}^{k-1}\Big{(}\frac{h_{2l}}{h_{2k+1}}\Big{)}^{\frac{1}{2}}\prod_{j=0}^{k-l-1}\frac{h_{2l+2j+2}}{h_{2l+2j+1}}w_{2l}(\eta)\Big{]}=o(1)

since all the weighted polynomials w2k+1w_{2k+1} and w2lw_{2l} in the sum are exponentially small as NN\to\infty.

Now for mm with |τm2ρ2|<2δN|\tau_{m}-2\rho^{2}|<2\delta_{N}, it follows from Lemma 5.1 (i) (see also Lemma 5.6) that

(5.35) hmhm+1=12ρ2(1+O(N12)),hmhm+1hm+2hm+1=1+O(N1).\frac{h_{m}}{h_{m+1}}=\frac{1}{2\rho^{2}}(1+O(N^{-\frac{1}{2}})),\qquad\frac{h_{m}}{h_{m+1}}\frac{h_{m+2}}{h_{m+1}}=1+O(N^{-1}).

This gives the estimate for k,lk,l with |kρ2N|<NδN|k-\rho^{2}N|<N\delta_{N} and |lρ2N|<NδN|l-\rho^{2}N|<N\delta_{N}

h2lh2k+1j=0kl1h2l+2j+2h2l+2j+1=h2kh2k+1j=0kl1h2l+2j+2h2l+2j+1h2l+2jh2l+2j+1=12ρ2(1+o(1)).\sqrt{\frac{h_{2l}}{{h_{2k+1}}}}\prod_{j=0}^{k-l-1}\frac{h_{2l+2j+2}}{h_{2l+2j+1}}=\sqrt{\frac{h_{2k}}{h_{2k+1}}}\prod_{j=0}^{k-l-1}\sqrt{\frac{h_{2l+2j+2}}{h_{2l+2j+1}}}\sqrt{\frac{h_{2l+2j}}{h_{2l+2j+1}}}=\frac{1}{\sqrt{2\rho^{2}}}(1+o(1)).

Note that Lemma 5.9 gives a bound |wm(η)|2<CNlogN|w_{m}(\eta)|^{2}<C\sqrt{N}\log N for all mm with |τm2ρ2|<2δN|\tau_{m}-2\rho^{2}|<2\delta_{N} and all zKz\in K. Since the sum

(5.36) γN3k=N(ρ2δN)N(ρ2+δN)w2k+1(η)[(h2kh2k+1)12w2k(η)+l=0k1(h2lh2k+1)12j=0kl1h2l+2j+2h2l+2j+1w2l(η)]\gamma_{N}^{3}\sum_{k=N(\rho^{2}-\delta_{N})}^{N(\rho^{2}+\delta_{N})}w_{2k+1}(\eta)\Big{[}\Big{(}\frac{h_{2k}}{h_{2k+1}}\Big{)}^{\frac{1}{2}}w_{2k}(\eta)+\sum_{l=0}^{k-1}\Big{(}\frac{h_{2l}}{h_{2k+1}}\Big{)}^{\frac{1}{2}}\prod_{j=0}^{k-l-1}\frac{h_{2l+2j+2}}{h_{2l+2j+1}}w_{2l}(\eta)\Big{]}

contains O(N2δN)O(N^{2}\delta_{N}) terms and γN=O(N1)\gamma_{N}=O(N^{-1}), the sum (5.36) has a bound of O(N1log2N)O(N^{-1}\log^{2}N).

Finally, we obtain that

(5.37) γN3k=N(ρ2+δN)N1w2k+1(η)l=0N(ρ2+δN)(h2lh2k+1)12j=0kl1h2l+2j+2h2l+2j+1w2l(η)=O(N34log2N)\gamma_{N}^{3}\sum_{k=N(\rho^{2}+\delta_{N})}^{N-1}w_{2k+1}(\eta)\sum_{l=0}^{N(\rho^{2}+\delta_{N})}\Big{(}\frac{h_{2l}}{h_{2k+1}}\Big{)}^{\frac{1}{2}}\prod_{j=0}^{k-l-1}\frac{h_{2l+2j+2}}{h_{2l+2j+1}}w_{2l}(\eta)=O(N^{-\frac{3}{4}}\log^{2}N)

since |w2k+1|2|w_{2k+1}|^{2} is at most O(N)O(N) for kN(ρ2+δN)k\geq N(\rho^{2}+\delta_{N}), |w2l|2=O(NlogN)|w_{2l}|^{2}=O(\sqrt{N}\log N) for |lρ2N|NδN|l-\rho^{2}N|\leq N\delta_{N}, and w2lw_{2l} with lN(ρ2δN)l\leq N(\rho^{2}-\delta_{N}) is exponentially small. Hence, the proof is complete. ∎

Lemma 5.12.

For kk, ll with lk1l\leq k-1 and (ρ2+δN)Nk,l<N(\rho^{2}+\delta_{N})N\leq k,l<N, we have

(5.38) h2k+1h2k=2ρ(1+o(1)),h2lh2k+1j=0kl1h2l+2j+2h2l+2j+1=12ρ(1+o(1)),\sqrt{\frac{h_{2k+1}}{h_{2k}}}=\sqrt{2}\rho\,(1+o(1)),\qquad\sqrt{\frac{h_{2l}}{h_{2k+1}}}\prod_{j=0}^{k-l-1}\frac{h_{2l+2j+2}}{h_{2l+2j+1}}=\frac{1}{\sqrt{2}\rho}\,(1+o(1)),

where o(1)0o(1)\to 0 as NN\to\infty.

Proof.

By Lemma 5.1 (ii), the ratio of the norm in (5.38) is obtained directly from the following observation: for (2ρ2+2δN)Nm<2N(2\rho^{2}+2\delta_{N})N\leq m<2N

hm+1hm=2ρ2m2ρ2Nm+12ρ2N(1+o(1))=2ρ2(1+o(1)).\frac{h_{m+1}}{h_{m}}=2\rho^{2}\frac{m-2\rho^{2}N}{m+1-2\rho^{2}N}(1+o(1))=2\rho^{2}(1+o(1)).

For the latter asymptotic expansion in (5.38), we see that

h2lh2k+1j=0kl1h2l+2j+2h2l+2j+1\displaystyle\sqrt{\frac{h_{2l}}{h_{2k+1}}}\prod_{j=0}^{k-l-1}\frac{h_{2l+2j+2}}{h_{2l+2j+1}} =h2kh2k+1j=0kl1h2l+2j+2h2l+2j+1h2l+2jh2l+2j+1\displaystyle=\sqrt{\frac{h_{2k}}{h_{2k+1}}}\prod_{j=0}^{k-l-1}\sqrt{\frac{h_{2l+2j+2}}{h_{2l+2j+1}}}\sqrt{\frac{h_{2l+2j}}{h_{2l+2j+1}}}
=12ρm=lk12m+12ρ2N2m+22ρ2N2m+12ρ2N2m2ρ2N(1+o(1)).\displaystyle=\frac{1}{\sqrt{2}\rho}\prod_{m=l}^{k-1}\sqrt{\frac{2m+1-2\rho^{2}N}{2m+2-2\rho^{2}N}\cdot\frac{2m+1-2\rho^{2}N}{2m-2\rho^{2}N}}(1+o(1)).

Here, the product can be approximated by 11 since for l,kl,k with (ρ2+δN)Nl,kN1(\rho^{2}+\delta_{N})N\leq l,k\leq N-1

m=lk1(2m+22ρ2N)(2m2ρ2N)(2m+12ρ2N)2=m=lk1(1(12m+12ρ2N)2)=1+o(1),\displaystyle\prod_{m=l}^{k-1}\frac{(2m+2-2\rho^{2}N)(2m-2\rho^{2}N)}{(2m+1-2\rho^{2}N)^{2}}=\prod_{m=l}^{k-1}\Big{(}1-\Big{(}\frac{1}{2m+1-2\rho^{2}N}\Big{)}^{2}\Big{)}=1+o(1),

which completes the proof. ∎

We now prove Proposition  5.7.

Proof of Proposition 5.7.

Combining Lemmas  5.8, 5.11, and 5.12, we obtain

GN(z,w)γN322ρk=N(ρ2+δN)N1w2k+1(ζ)l=N(ρ2+δN)k1w2l(η)\displaystyle\quad G_{N}(z,w)\sim\frac{\gamma_{N}^{3}}{2\sqrt{2}\rho}\sum_{k=N(\rho^{2}+\delta_{N})}^{N-1}w_{2k+1}(\zeta)\sum_{l=N(\rho^{2}+\delta_{N})}^{k-1}w_{2l}(\eta)
gN(z)gN(w)N2(1ρ2)3k=N(ρ2+δN)N1(2k+12Nρ2)12e21ρ2(2k+12Nρ2)zl=N(ρ2+δN)k1(lNρ2)12e21ρ2(lNρ2)w.\displaystyle\sim\frac{g_{N}(z)g_{N}(w)}{N^{2}(1-\rho^{2})^{3}}\sum_{k=N(\rho^{2}+\delta_{N})}^{N-1}\Big{(}\frac{2k+1}{2N}-\rho^{2}\Big{)}^{\frac{1}{2}}e^{\frac{2}{1-\rho^{2}}(\frac{2k+1}{2N}-\rho^{2})z}\!\!\!\sum_{l=N(\rho^{2}+\delta_{N})}^{k-1}\!\!\Big{(}\frac{l}{N}-\rho^{2}\Big{)}^{\frac{1}{2}}e^{\frac{2}{1-\rho^{2}}(\frac{l}{N}-\rho^{2})w}.

Now the proposition follows from the Riemann sum approximation. ∎

Acknowledgement.

We wish to express our gratitude to Gernot Akemann, Nam-Gyu Kang, and Iván Parra for helpful discussions.

References

  • [1] M. Adler, P. J. Forrester, T. Nagao, and P. van Moerbeke. Classical skew orthogonal polynomials and random matrices. J. Statist. Phys., 99(1-2):141–170, 2000.
  • [2] G. Akemann. The complex Laguerre symplectic ensemble of non-Hermitian matrices. Nuclear Phys. B, 730(3):253–299, 2005.
  • [3] G. Akemann and M. Bender. Interpolation between Airy and Poisson statistics for unitary chiral non-Hermitian random matrix ensembles. J. Math. Phys., 51(10):103524, 2010.
  • [4] G. Akemann, S.-S. Byun, and N.-G. Kang. A non-Hermitian generalisation of the Marchenko-Pastur distribution: from the circular law to multi-criticality. Ann. Henri Poincaré, 22(4):1035–1068, 2021.
  • [5] G. Akemann, S.-S. Byun, and N.-G. Kang. Scaling limits of planar symplectic ensembles. SIGMA Symmetry Integrability Geom. Methods Appl., 18:Paper No. 007, 40, 2022.
  • [6] G. Akemann, M. Cikovic, and M. Venker. Universality at weak and strong non-Hermiticity beyond the elliptic Ginibre ensemble. Comm. Math. Phys., 10.1007/s00220-018-3201-1, 2018.
  • [7] G. Akemann, M. Ebke, and I. Parra. Skew-orthogonal polynomials in the complex plane and their Bergman-like kernels. Comm. Math. Phys., 389:621–659, 2022.
  • [8] G. Akemann, J. R. Ipsen, and E. Strahov. Permanental processes from products of complex and quaternionic induced Ginibre ensembles. Random Matrices Theory Appl., 3(4):1450014, 54, 2014.
  • [9] G. Akemann, M. Kieburg, A. Mielke, and T. Prosen. Universal signature from integrability to chaos in dissipative open quantum systems. Phys. Rev. Lett., 123(25):254101, 2019.
  • [10] G. Akemann and M. J. Phillips. The interpolating Airy kernels for the β=1\beta=1 and β=4\beta=4 elliptic Ginibre ensembles. J. Stat. Phys., 155(3):421–465, 2014.
  • [11] Y. Ameur and S.-S. Byun. Almost-Hermitian random matrices and bandlimited point processes. preprint arXiv:2101.03832, 2021.
  • [12] Y. Ameur and J. Cronvall. Szegö type asymptotics for the reproducing kernel in spaces of full-plane weighted polynomials. preprint arXiv:2107.11148, 2021.
  • [13] Y. Ameur, H. Hedenmalm, and N. Makarov. Fluctuations of eigenvalues of random normal matrices. Duke Math. J., 159(1):31–81, 2011.
  • [14] Y. Ameur, N.-G. Kang, and N. Makarov. Rescaling Ward identities in the random normal matrix model. Constr. Approx., 50(1):63–127, 2019.
  • [15] Y. Ameur, N.-G. Kang, N. Makarov, and A. Wennman. Scaling limits of random normal matrix processes at singular boundary points. J. Funct. Anal., 278(3):108340, 2020.
  • [16] Y. Ameur, N.-G. Kang, and S.-M. Seo. On boundary confinements for the Coulomb gas. Anal. Math. Phys., 10(4):Paper No. 68, 42, 2020.
  • [17] Y. Ameur, N.-G. Kang, and S.-M. Seo. The random normal matrix model: insertion of a point charge. Potential Anal. (online), 2021.
  • [18] F. Benaych-Georges and F. Chapon. Random right eigenvalues of Gaussian quaternionic matrices. Random Matrices Theory Appl., 1(2):1150009, 18, 2012.
  • [19] M. Bender. Edge scaling limits for a family of non-Hermitian random matrix ensembles. Probab. Theory Relat. Fields, 147(1):241–271, 2010.
  • [20] P. M. Bleher and A. B. J. Kuijlaars. Orthogonal polynomials in the normal matrix model with a cubic potential. Adv. Math., 230(3):1272–1321, 2012.
  • [21] S.-S. Byun and C. Charlier. On the almost-circular symplectic induced Ginibre ensemble. preprint arXiv:2206.06021, 2022.
  • [22] S.-S. Byun and M. Ebke. Universal scaling limits of the symplectic elliptic Ginibre ensembles. Random Matrices Theory Appl. (online), 2022.
  • [23] S.-S. Byun and P. J. Forrester. Spherical induced ensembles with symplectic symmetry. preprint arXiv:2209.01934, 2022.
  • [24] S.-S. Byun, N.-G. Kang, J. O. Lee, and J. Lee. Real eigenvalues of elliptic random matrices. Int. Math. Res. Not. (online), 2021.
  • [25] S.-S. Byun, S.-Y. Lee, and M. Yang. Lemniscate ensembles with spectral singularity. preprint arXiv:2107.07221, 2021.
  • [26] L.-L. Chau and O. Zaboronsky. On the structure of correlation functions in the normal matrix model. Commun. Math. Phys., 196(1):203–247, 1998.
  • [27] T. Claeys and A. B. J. Kuijlaars. Universality in unitary random matrix ensembles when the soft edge meets the hard edge. In Integrable systems and random matrices, volume 458 of Contemp. Math., pages 265–279. Amer. Math. Soc., Providence, RI, 2008.
  • [28] P. Deift and D. Gioev. Universality at the edge of the spectrum for unitary, orthogonal, and symplectic ensembles of random matrices. Comm. Pure Appl. Math., 60(6):867–910, 2007.
  • [29] P. Deift and D. Gioev. Universality in random matrix theory for orthogonal and symplectic ensembles. Int. Math. Res. Pap. IMRP, 2007(2):Art. ID rpm004, 116, 2007.
  • [30] G. Dubach. Symmetries of the quaternionic ginibre ensemble. Random Matrices Theory Appl., 10(01):2150013, 2021.
  • [31] J. Fischmann, W. Bruzda, B. A. Khoruzhenko, H.-J. Sommers, and K. Życzkowski. Induced Ginibre ensemble of random matrices and quantum operations. J. Phys. A, 45(7):075203, 31, 2012.
  • [32] P. J. Forrester. Log-gases and Random Matrices (LMS-34). Princeton University Press, Princeton, 2010.
  • [33] P. J. Forrester, N. E. Frankel, and T. M. Garoni. Asymptotic form of the density profile for Gaussian and Laguerre random matrix ensembles with orthogonal and symplectic symmetry. J. Math. Phys., 47(2):023301, 26, 2006.
  • [34] Y. V. Fyodorov, B. A. Khoruzhenko, and H.-J. Sommers. Almost Hermitian random matrices: crossover from Wigner-Dyson to Ginibre eigenvalue statistics. Phys. Rev. Lett., 79(4):557–560, 1997.
  • [35] Y. V. Fyodorov, B. A. Khoruzhenko, and H.-J. Sommers. Almost-Hermitian random matrices: eigenvalue density in the complex plane. Phys. Lett. A, 226(1-2):46–52, 1997.
  • [36] Y. V. Fyodorov, H.-J. Sommers, and B. A. Khoruzhenko. Universality in the random matrix spectra in the regime of weak non-Hermiticity. Ann. Inst. H. Poincaré Phys. Théor., 68(4):449–489, 1998.
  • [37] J. Ginibre. Statistical ensembles of complex, quaternion, and real matrices. J. Math. Phys., 6(3):440–449, 1965.
  • [38] H. Hedenmalm and A. Wennman. Riemann-Hilbert hierarchies for hard edge planar orthogonal polynomials. preprint arXiv:2008.02682, 2020.
  • [39] H. Hedenmalm and A. Wennman. Planar orthogonal polynomials and boundary universality in the random normal matrix model. Acta Math., 227(2):309–406, 2021.
  • [40] J. R. Ipsen. Products of independent quaternion Ginibre matrices and their correlation functions. J. Phys. A, 46(26):265201, 16, 2013.
  • [41] E. Kanzieper. Eigenvalue correlations in non-Hermitean symplectic random matrices. J. Phys. A, 35(31):6631–6644, 2002.
  • [42] B. A. Khoruzhenko and S. Lysychkin. Scaling limits of eigenvalue statistics in the quaternion-real Ginibre ensemble. in preparation.
  • [43] B. A. Khoruzhenko and S. Lysychkin. Truncations of random symplectic unitary matrices. preprint arXiv:2111.02381.
  • [44] S.-Y. Lee and R. Riser. Fine asymptotic behavior for eigenvalues of random normal matrices: ellipse case. J. Math. Phys., 57(2):023302, 29, 2016.
  • [45] S. Lysychkin. Complex eigenvalues of high dimensional quaternion random matrices. PhD Thesis, Queen Mary University of London United Kingdom, 2021.
  • [46] T. Nagao, G. Akemann, M. Kieburg, and I. Parra. Families of two-dimensional Coulomb gases on an ellipse: correlation functions and universality. J. Phys. A, 53(7):075201, 36, 2020.
  • [47] F. W. Olver, D. W. Lozier, R. F. Boisvert, and C. W. Clark (Editors). NIST Handbook of Mathematical Functions. Cambridge University Press, Cambridge, 2010.
  • [48] J. C. Osborn. Universal results from an alternate random-matrix model for QCD with a baryon chemical potential. Phys. Rev. Lett., 93(22):222001, 2004.
  • [49] B. Rider. A limit theorem at the edge of a non-Hermitian random matrix ensemble. J. Phys. A, 36(12):3401–3409, 2003.
  • [50] E. B. Saff and V. Totik. Logarithmic potentials with external fields, volume 316 of Grundlehren der Mathematischen Wissenschaften [Fundamental Principles of Mathematical Sciences]. Springer-Verlag, Berlin, 1997. Appendix B by Thomas Bloom.
  • [51] S.-M. Seo. Edge scaling limit of the spectral radius for random normal matrix ensembles at hard edge. J. Stat. Phys., 181(5):1473–1489, 2020.
  • [52] S.-M. Seo. Edge behavior of two-dimensional Coulomb gases near a hard wall. Ann. Henri Poincaré, 23:2247–2275, 2022.
  • [53] H. Skovgaard. Asymptotic forms of Hermite polynomials. Technical Report, 18, Department of Mathematics, California Institute of Technology, 1959.
  • [54] G. Szegö. Orthogonal Polynomials. American Mathematical Society Colloquium Publications, Vol. 23. American Mathematical Society, New York, 1939.
  • [55] X.-S. Wang. Plancherel-Rotach asymptotics of second-order difference equations with linear coefficients. J. Approx. Theory, 188:1–18, 2014.
  • [56] H. Widom. On the relation between orthogonal, symplectic and unitary matrix ensembles. J. Statist. Phys., 94(3-4):347–363, 1999.
  • [57] K. Życzkowski and H.-J. Sommers. Truncations of random unitary matrices. J. Phys. A, 33(10):2045–2057, 2000.