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

One-point asymptotics for half-flat ASEP

Evgeni Dimitrov and Anushka Murthy
Abstract.

We consider the asymmetric simple exclusion process (ASEP) with half-flat initial condition. We show that the one-point marginals of the ASEP height function are described by those of the Airy21\mbox{Airy}_{2\rightarrow 1} process, introduced by Borodin-Ferrari-Sasamoto in (Commun. Pure Appl. Math., 61, 1603-1629, 2008). This result was conjectured by Ortmann-Quastel-Remenik (Ann. Appl. Probab., 26, 507-548), based on an informal asymptotic analysis of exact formulas for generating functions of the half-flat ASEP height function at one spatial point. Our present work provides a fully rigorous derivation and asymptotic analysis of the same generating functions, under certain parameter restrictions of the model.

1. Introduction and main results

1.1. Preface

The asymmetric simple exclusion process (ASEP) is a continuous time Markov process, introduced to the mathematical community by Spitzer [Spi70] in 1970 (and also appearing two years earlier in the biology work of MacDonald, Gibbs and Pipkin [MGP68]).

The state space of the process is {0,1}\{0,1\}^{\mathbb{Z}} and we interpret an element η{0,1}\eta\in\{0,1\}^{\mathbb{Z}} as a particle configuration, where η(x)=1\eta(x)=1 indicates the presence of a particle at location xx, and η(x)=0\eta(x)=0 indicates the presence of a hole at location xx. The dynamics of the model depend on a parameter p[0,1]p\in[0,1], and can be described as follows. Each particle carries an independent exponential clock, which rings at rate 11. When the clock rings the particle attempts to jump one site to the right with probability pp, and with probability q=1pq=1-p it attempts to jump one site to the left. The jump is successful if the site to which the particle attempts to jump is a hole; otherwise, the jump is suppressed and the particle stays put. As mentioned in [Spi70], there is some delicacy in formally constructing a Markov process corresponding to the latter dynamics when the number of particles is infinite; however, a number of papers have been published confirming the existence of such a Markov process – see the works of Harris [Har72], Holley [Hol70], and Liggett [Lig72]. We denote this process by {ηt:t0}\{\eta_{t}:t\geq 0\}. If q=1,p=0q=1,p=0 (or q=0,p=1q=0,p=1) the process {ηt:t0}\{\eta_{t}:t\geq 0\} is called the totally asymmetric simple exclusion process (TASEP), if q=p=1/2q=p=1/2 it is called the symmetric simple exclusion process (SSEP), and if p,q>0p,q>0 and pqp\neq q it is called the (partially) asymmetric simple exclusion process (ASEP). Throughout the paper we will assume that q>p>0q>p>0, so that the particles have a drift to the left.

Given {ηt:t0}\{\eta_{t}:t\geq 0\}, we define {η^t:t0}\{\hat{\eta}_{t}:t\geq 0\} through η^t(x)=2ηt(x)1\hat{\eta}_{t}(x)=2\eta_{t}(x)-1 so that η^t{1,1}\hat{\eta}_{t}\in\{-1,1\}^{\mathbb{Z}}, and then we define the height function of ASEP to be

(1.1) h(t,x)={2N0flux(t)+y=1xη^t(y) if x12N0flux(t), if x=02N0flux(t)y=x+10η^t(y) if x1,h(t,x)=\begin{cases}2N_{0}^{\operatorname{flux}}(t)+\sum_{y=1}^{x}\hat{\eta}_{t}(y)&\mbox{ if }x\geq 1\\ 2N_{0}^{\operatorname{flux}}(t),&\mbox{ if }x=0\\ 2N_{0}^{\operatorname{flux}}(t)-\sum_{y=x+1}^{0}\hat{\eta}_{t}(y)&\mbox{ if }x\leq-1,\end{cases}

where N0flux(t)N_{0}^{\operatorname{flux}}(t) is the net number of particles that crossed from site 11 to 0 up to time tt, i.e. particles that jump from 11 to 0 are counted as +1+1 and those that jumped from 0 to 11 are counted as 1-1.

ASEP is an important member of the one-dimensional Kardar-Parisi-Zhang (KPZ) universality class. For more on the KPZ universality class we refer to the surveys and books [Cor12, HHT15, QS15] and the references therein. In particular, it is expected that the scaled ASEP height function ϵ1/2h(ϵ3/2T,ϵ1x)\epsilon^{1/2}h(\epsilon^{-3/2}T,\epsilon^{-1}x), appropriately shifted, converges as ϵ0+\epsilon\rightarrow 0+ to the fixed time TT distribution of a certain Markov process known as the KPZ fixed point, started from an initial state that depends on the initial condition for ASEP. We mention that the KPZ fixed point was constructed by Matetski-Quastel-Remenik [MQR21], and the convergence of the ASEP height function to the KPZ fixed point was proved for a class of initial conditions by Quastel-Sarkar [QS22].

Despite the remarkable progress in understanding the asymptotic behavior of ASEP, the results in [QS22] only apply when the initial condition asymptotically is either continuous with moderate growth, or consists of multiple narrow wedges. There are still many natural and important initial conditions that are not handled by [QS22], and for which precise statements are not available. One such example is the case of half-flat initial condition, which corresponds to starting ASEP from η0hfl=𝟏{x2}\eta^{\operatorname{h-fl}}_{0}={\bf 1}\{x\in 2\mathbb{N}\} , where ={1,2,3,}\mathbb{N}=\{1,2,3,\dots\}. The half-flat initial condition asymptotically becomes the function that is 0 for positive and -\infty for negative values, so that it is a natural mixture of the two classes of initial conditions handled in [QS22], without belonging to either.

The goal of the present paper is to investigate the large time limit of the ASEP height function h(t,x)h(t,x), started from η0hfl\eta^{\operatorname{h-fl}}_{0}, and show that its one-point marginals are asymptotically described by those of the Airy2→1 process, introduced by Borodin-Ferrari-Sasamoto [BFS08]. This convergence result was conjectured by Ortmann-Quastel-Remenik [OQR16], based on an informal analysis of exact formulas for generating functions of the half-flat ASEP height function at one spatial point. Our present work provides a fully rigorous derivation and asymptotic analysis of the same generating functions, under certain parameter restrictions of the model.

The rest of the introduction is structured as follows. In Section 1.2 we introduce some relevant notation and state a certain moment formula derived in [OQR16] – this is Proposition 1.2. Within the same section we state the exact formulas for the generating functions of the half-flat ASEP height function as Theorem 1.5. We mention that Theorem 1.5 already appeared in [OQR16], but as explained in Remarks 1.8 and 2.4, there are aspects of the derivation of that theorem that are not justified in [OQR16]. In fact, we can only prove the well-posedness of our formulas in Theorem 1.5 when pp is assumed to be sufficiently close to zero. In Section 1.3 we introduce the Airy21\mbox{Airy}_{2\rightarrow 1} process, and in Section 1.4 we present our main asymptotic result.

1.2. Exact formulas for half-flat ASEP

We begin by summarizing some of the notation we require for ASEP in the following definition.

Definition 1.1.

We fix p(0,1/2)p\in(0,1/2) and let q=1pq=1-p, τ=p/q\tau=p/q, γ=qp\gamma=q-p, so that p,q,τ,γ(0,1)p,q,\tau,\gamma\in(0,1). Let {ηt:t0}\{\eta_{t}:t\geq 0\} denote the ASEP started from half-flat initial condition η0hfl=𝟏{x2}\eta^{\operatorname{h-fl}}_{0}={\bf 1}\{x\in 2\mathbb{N}\} for the parameters p,qp,q as in Section 1.1. We denote the distribution of {ηt:t0}\{\eta_{t}:t\geq 0\} from this initial condition by hfl\mathbb{P}^{\operatorname{h-fl}} and write 𝔼hfl\mathbb{E}^{\operatorname{h-fl}} for the expectation with respect to this measure. For xx\in\mathbb{Z} and t>0t>0 we define

(1.2) Nx(t)=y=xηt(y),N_{x}(t)=\sum_{y=-\infty}^{x}\eta_{t}(y),

to be the total number of particles at or to the left of location xx at time tt, and note that from (1.1) we have the following relationship between Nx(t)N_{x}(t) and the height function

(1.3) h(t,x)=2Nx(t)x.h(t,x)=2N_{x}(t)-x.

In the remainder of this section we present the exact formulas for certain observables involving Nx(t)N_{x}(t) that were derived in [OQR16]. We mention that in view of (1.3) all of the observables below can alternatively be expressed in terms of h(t,x)h(t,x); however, we will formulate them with Nx(t)N_{x}(t), following [OQR16]. Both the observables and the formulas for them depend on various special functions, which we present next.

For q[0,1)q\in[0,1) and aa\in\mathbb{C} we let (a;q)(a;q)_{\infty} be the qq-Pochhammer symbol

(1.4) (a;q)=n=0(1aqn).(a;q)_{\infty}=\prod_{n=0}^{\infty}(1-aq^{n}).

We also define the qq-factorial for k0k\in\mathbb{Z}_{\geq 0}

(1.5) kq!=a=1k(1qa)(1q)k,k_{q}!=\frac{\prod_{a=1}^{k}(1-q^{a})}{(1-q)^{k}},

and the qq-exponential function

(1.6) eq(x)=1((1q)x;q)=k=0xkkq!,e_{q}(x)=\frac{1}{((1-q)x;q)_{\infty}}=\sum_{k=0}^{\infty}\frac{x^{k}}{k_{q}!},

where the first identity is the definition of eq(x)e_{q}(x) and is valid for x(1q)1qmx\neq(1-q)^{-1}q^{-m} for m0m\in\mathbb{Z}_{\geq 0} and the second holds when |x|<1|x|<1 – see [AAR99, Corollary 10.2.2a].

For the next several functions we fix p,q,τp,q,\tau as in Definition 1.1, xx\in\mathbb{Z}, and t0t\geq 0. We define

(1.7) 𝔣(w;n)=(1τ)nexp((qp)t1+w(qp)t1+τnw)(1+τnw1+w)x1,𝔤(w;n)=(w;τ)(τ2nw2;τ)(τnw;τ)(τnw2;τ),𝔥(w1,w2;n1,n2)=(w1w2;τ)(τn1+n2w1w2;τ)(τn1w1w2;τ)(τn2w1w2;τ)\begin{split}\mathfrak{f}(w;n)=&\hskip 5.69054pt(1-\tau)^{n}\exp\left(\frac{(q-p)t}{1+w}-\frac{(q-p)t}{1+\tau^{n}w}\right)\cdot\left(\frac{1+\tau^{n}w}{1+w}\right)^{x-1},\\ \mathfrak{g}(w;n)=&\hskip 5.69054pt\frac{(-w;\tau)_{\infty}(\tau^{2n}w^{2};\tau)_{\infty}}{(-\tau^{n}w;\tau)_{\infty}(\tau^{n}w^{2};\tau)_{\infty}},\\ \mathfrak{h}(w_{1},w_{2};n_{1},n_{2})=&\hskip 5.69054pt\frac{(w_{1}w_{2};\tau)_{\infty}(\tau^{n_{1}+n_{2}}w_{1}w_{2};\tau)_{\infty}}{(\tau^{n_{1}}w_{1}w_{2};\tau)_{\infty}(\tau^{n_{2}}w_{1}w_{2};\tau)_{\infty}}\end{split}

Throughout the paper we denote by w\vec{w} the vector w=(w1,,wk)\vec{w}=(w_{1},\dots,w_{k}) and by dwd\vec{w} the complex integration form dw1dwkdw_{1}\cdots dw_{k}. We mention that the dimension kk and the contours over which the integration is performed will be clear from the context.

For complex vectors w,sk\vec{w},\vec{s}\in\mathbb{C}^{k} and ζ[0,)\zeta\in\mathbb{C}\setminus[0,\infty) we write

(1.8) F(s,w)=det[1waτsawb]a,b=1ka=1k𝔣(wa;sa)𝔤(wa;sa)1a<bk𝔥(wa,wb;sa,sb),F(ζ;s,w)=a=1kπ(ζ)sasin(πsa)F(s,w).\begin{split}&F(\vec{s},\vec{w})=\det\left[\frac{-1}{w_{a}\tau^{s_{a}}-w_{b}}\right]_{a,b=1}^{k}\prod_{a=1}^{k}\mathfrak{f}(w_{a};s_{a})\mathfrak{g}(w_{a};s_{a})\cdot\prod_{1\leq a<b\leq k}\mathfrak{h}(w_{a},w_{b};s_{a},s_{b}),\\ &F(\zeta;\vec{s},\vec{w})=\prod_{a=1}^{k}\frac{\pi(-\zeta)^{s_{a}}}{\sin(-\pi s_{a})}\cdot F(\vec{s},\vec{w}).\end{split}

In equation (1.8) and throughout the paper we will always take the principal branch of the logarithm.

With the above notation in place, we can state the exact formulas we use.

Proposition 1.2.

[OQR16, Theorem 1.3]. Assume the same notation as in Definition 1.1. For any m0m\in\mathbb{Z}_{\geq 0} we have

(1.9) 𝔼hfl[τmNx(t)]=mτ!k=0mνk,mhhl(t,x),\mathbb{E}^{\operatorname{h-fl}}\left[\tau^{mN_{x}(t)}\right]=m_{\tau}!\sum_{k=0}^{m}\nu^{\operatorname{h-hl}}_{k,m}(t,x),

where γ1,0\gamma_{-1,0} is the positively oriented circle of radius τ1/8\tau^{-1/8}, centered at the origin and

(1.10) νk,mhhl(t,x)=1k!n1,,nkn1++nk=m1(2π𝗂)kγ1,0k𝑑wF(n,w).\begin{split}\nu^{\operatorname{h-hl}}_{k,m}(t,x)=\hskip 5.69054pt&\frac{1}{k!}\sum_{\begin{subarray}{c}n_{1},\dots,n_{k}\in\mathbb{N}\\ n_{1}+\cdots+n_{k}=m\end{subarray}}\frac{1}{(2\pi\mathsf{i})^{k}}\oint_{\gamma_{-1,0}^{k}}d\vec{w}F(\vec{n},\vec{w}).\end{split}
Remark 1.3.

[OQR16, Theorem 1.3] is formulated for general contours γ1,0\gamma_{-1,0}, and as explained in that theorem γ1,0\gamma_{-1,0} needs to be positively oriented, enclose 0, 1-1, and exclude all other poles of F(n,w)F(\vec{n},\vec{w}). Our choice for γ1,0\gamma_{-1,0} to be the positively oriented circle of radius τ1/8\tau^{-1/8}, centered at the origin, clearly satisfies all of these conditions.

Remark 1.4.

Let us briefly explain how Proposition 1.2 is proved in [OQR16]. If one defines

Q~x(t)=τNx(t)τNx1(t)τ1,\tilde{Q}_{x}(t)=\frac{\tau^{N_{x}(t)}-\tau^{N_{x-1}(t)}}{\tau-1},

it was shown in [BCS14] that 𝔼hfl[Q~x1(t)Q~xk(t)]\mathbb{E}^{\operatorname{h-fl}}\left[\tilde{Q}_{x_{1}}(t)\cdots\tilde{Q}_{x_{k}}(t)\right] is the unique solution to a certain system of differential equations. In [OQR16] the authors discovered a certain kk-fold contour integral that solved this system, and thus provided formulas for the observables 𝔼hfl[Q~x1(t)Q~xk(t)]\mathbb{E}^{\operatorname{h-fl}}\left[\tilde{Q}_{x_{1}}(t)\cdots\tilde{Q}_{x_{k}}(t)\right]. These formulas can be found in [OQR16, Theorem 1.2], and the derivation is done in Section 2 of that paper. From [BCS14, Lemma 4.18] there is a way to express τmNx(t)\tau^{mN_{x}(t)} as a linear combination over Q~x1(t)Q~xk(t)\tilde{Q}_{x_{1}}(t)\cdots\tilde{Q}_{x_{k}}(t) with 0km0\leq k\leq m and x1<<xkxx_{1}<\cdots<x_{k}\leq x, which allows one to express 𝔼hfl[τmNx(t)]\mathbb{E}^{\operatorname{h-fl}}\left[\tau^{mN_{x}(t)}\right] as a certain mm-fold contour integral over different contours. This is done in [OQR16, Proposition 3.2]. Deforming all of these contours to the same one, using a nested contour integral ansatz that is similar to [BC14, Proposition 3.8] and whose proof is inspired by [HO97], one arrives at (1.9).

Using the formulas for the moments of τNx(t)\tau^{N_{x}(t)}, one can obtain a formula for the generating series of the τ\tau-Laplace transform of τNx(t)\tau^{N_{x}(t)}. This formula is given in the following theorem, and forms the basis of our asymptotic analysis.

Theorem 1.5.

Assume the same notation as in Definition 1.1, and let ζ[0,)\zeta\in\mathbb{C}\setminus[0,\infty). We further suppose that τ(0,1)\tau\in(0,1) is sufficiently small, so that

(1.11) (τ1/2+τ1/4)(1τ1/2)2(τ3/4;τ)(τ3/4;τ)(τ1/4;τ)(τ1/4;τ)<1.\frac{(\tau^{1/2}+\tau^{1/4})}{(1-\tau^{1/2})^{2}}\cdot\frac{(-\tau^{3/4};\tau)_{\infty}(-\tau^{3/4};\tau)_{\infty}}{(\tau^{1/4};\tau)_{\infty}(\tau^{1/4};\tau)_{\infty}}<1.

Then, for eτe_{\tau} as in (1.6) we have

(1.12) 𝔼hfl[eτ(ζτNx(t))]=1+k=1Hk(ζ), where \begin{split}\mathbb{E}^{\operatorname{h-fl}}\left[e_{\tau}\left(\zeta\tau^{N_{x}(t)}\right)\right]=\hskip 5.69054pt&1+\sum_{k=1}^{\infty}H_{k}(\zeta),\mbox{ where }\end{split}
(1.13) Hk(ζ)=1k!(2π𝗂)2k(1/2+𝗂)k𝑑sγ1,0k𝑑wF(ζ;s,w),\begin{split}&H_{k}(\zeta)=\frac{1}{k!(2\pi\mathsf{i})^{2k}}\int_{(1/2+\mathsf{i}\mathbb{R})^{k}}d\vec{s}\oint_{\gamma_{-1,0}^{k}}d\vec{w}F(\zeta;\vec{s},\vec{w}),\end{split}

γ1,0\gamma_{-1,0} is as in Proposition 1.2, and F(ζ;s,w)F(\zeta;\vec{s},\vec{w}) is as in (1.8).

Remark 1.6.

Part of the statement of the theorem is that the integral in (1.13) is well-defined and finite, and the sum on the right side of (1.12) is absolutely convergent. We establish these statements in Lemma 2.1.

Remark 1.7.

The way that the formula in (1.12) is derived is by first showing that for |ζ|<1|\zeta|<1

(1.14) 𝔼hfl[eτ(ζτNx(t))]=limNm=0N1mτ!𝔼hfl[τmNx(t)],\mathbb{E}^{\operatorname{h-fl}}\left[e_{\tau}\left(\zeta\tau^{N_{x}(t)}\right)\right]=\lim_{N\rightarrow\infty}\sum_{m=0}^{N}\frac{1}{m_{\tau}!}\cdot\mathbb{E}^{\operatorname{h-fl}}\left[\tau^{mN_{x}(t)}\right],

which is a direct consequence of (1.6). Afterwards, we use the formulas for the moments 𝔼hfl[τmNx(t)]\mathbb{E}^{\operatorname{h-fl}}\left[\tau^{mN_{x}(t)}\right] from (1.9), take the limit NN\rightarrow\infty and rearrange the resulting sum. In this way one obtains

(1.15) 𝔼hfl[eτ(ζτNx(t))]=1+k=1n1=1nk=1ζn1++nkk!(2π𝗂)kγ1,0k𝑑wF(n,w).\mathbb{E}^{\operatorname{h-fl}}\left[e_{\tau}\left(\zeta\tau^{N_{x}(t)}\right)\right]=1+\sum_{k=1}^{\infty}\sum_{n_{1}=1}^{\infty}\cdots\sum_{n_{k}=1}^{\infty}\frac{\zeta^{n_{1}+\cdots+n_{k}}}{k!(2\pi\mathsf{i})^{k}}\oint_{\gamma_{-1,0}^{k}}d\vec{w}F(\vec{n},\vec{w}).

One can express the kk-fold sums over n1,,nkn_{1},\dots,n_{k} as kk-fold contour integrals using a Mellin-Barnes type integral representation from [BC14, Lemma 3.20]. This proves (1.12) for |ζ|<1|\zeta|<1, and then one shows that the equality holds for all ζ[0,)\zeta\in\mathbb{C}\setminus[0,\infty) by analytic continuation. We mention that in taking the NN\rightarrow\infty limit, as well as in applying the analytic continuation argument, we require absolute summability of the series in (1.15), which we can ensure only when τ\tau is sufficiently small. This is the origin of the assumption (1.11). For general τ(0,1)\tau\in(0,1) the integrals in (1.13) are still well-defined; however, we do not know how to show that the series in (1.12) is convergent. The proof of Theorem 1.5 is the content of Section 2.

Remark 1.8.

Theorem 1.5 appears as Theorem 1.4 in [OQR16], although there is a missing term a=1kπsin(πsa)\prod_{a=1}^{k}\frac{\pi}{\sin(-\pi s_{a})} in the integrand in (1.13) and the result is formulated without the assumption (1.11). The authors derive their result following the outline we gave in Remark 1.7 and it is when the authors apply this Mellin-Barnes integral representation that the a=1kπsin(πsa)\prod_{a=1}^{k}\frac{\pi}{\sin(-\pi s_{a})} term is dropped in [OQR16], see [OQR16, Equation (4.13)]. This is a small typo in the paper. A more serious problem, which we discuss in Remark 2.4, is that the authors do not provide sufficient justification to demonstrate that the series in (1.15) is absolutely convergent for |ζ|<1|\zeta|<1, and that one can analytically extend the equality in (1.12) to ζ[0,)\zeta\in\mathbb{C}\setminus[0,\infty). In fact, it is not even clear from the arguments in [OQR16] that the right side of (1.12) is convergent and hence well-defined.

1.3. The Airy21\mbox{Airy}_{2\rightarrow 1} process

In this section we define the Airy21\mbox{Airy}_{2\rightarrow 1} process, which is the object that describes the large time limit of the height function hh from (1.3) under the measure hfl\mathbb{P}^{\operatorname{h-fl}} from Definition 1.1. The Airy21\mbox{Airy}_{2\rightarrow 1} process was introduced by Borodin-Ferrari-Sasamoto [BFS08], where it arises from the limit of the TASEP with half-flat initial condition.

Let us first introduce certain complex contours that arise in our discussion.

Definition 1.9.

For aa\in\mathbb{C} and ϕ(0,π)\phi\in(0,\pi) we define the contour Ca,ϕC_{a,\phi} to be the union of {a+ye𝗂ϕ)}y+\{a+ye^{-\mathsf{i}\phi)}\}_{y\in\mathbb{R}^{+}} and {a+ye𝗂ϕ}y+\{a+ye^{\mathsf{i}\phi}\}_{y\in\mathbb{R}^{+}} oriented to have increasing imaginary part.

The following is Definition 2.1 from [BFS08].

Definition 1.10.

(The Airy21\mbox{Airy}_{2\rightarrow 1} process) Let us set

(1.16) x~=xs2𝟏{s0},y~=yt2𝟏{t0},\tilde{x}=x-s^{2}\cdot{\bf 1}\{s\leq 0\},\hskip 5.69054pt\tilde{y}=y-t^{2}\cdot{\bf 1}\{t\leq 0\},

and define the kernel

(1.17) K(s,x;t,y)=𝟏{t>s}e(y~x~)24(ts)4π(ts)+1(2π𝗂)2C1,π/4𝑑wC0,3π/4𝑑zew3/3+tw2y~wez3/3+sz2x~z2wz2w2,K_{\infty}(s,x;t,y)=-\frac{{\bf 1}\{t>s\}e^{-\frac{(\tilde{y}-\tilde{x})^{2}}{4(t-s)}}}{\sqrt{4\pi(t-s)}}+\frac{1}{(2\pi\mathsf{i})^{2}}\int_{C_{1,\pi/4}}dw\int_{C_{0,3\pi/4}}dz\frac{e^{w^{3}/3+tw^{2}-\tilde{y}w}}{e^{z^{3}/3+sz^{2}-\tilde{x}z}}\cdot\frac{-2w}{z^{2}-w^{2}},

where C1,π/4,C0,3π/4C_{1,\pi/4},C_{0,3\pi/4} are as in Definition 1.9. The Airy2→1 process, denoted by 𝒜21\mathcal{A}_{2\rightarrow 1}, is the process on \mathbb{R} with mm-point joint distribution at t1<t2<<tmt_{1}<t_{2}<\cdots<t_{m} given by the Fredholm determinant

(1.18) (k=1m{𝒜21(tk)yk})=det(IχyKχy)L2({t1,,tm}×).\mathbb{P}\left(\cap_{k=1}^{m}\{\mathcal{A}_{2\rightarrow 1}(t_{k})\leq y_{k}\}\right)=\det\left(I-\chi_{y}K_{\infty}\chi_{y}\right)_{L^{2}(\{t_{1},\dots,t_{m}\}\times\mathbb{R})}.

In (1.18) we have that the measure on {t1,,tm}×\{t_{1},\dots,t_{m}\}\times\mathbb{R} is the product measure of the counting measure on {t1,,tm}\{t_{1},\dots,t_{m}\} and the Lebesgue measure on \mathbb{R}. The function χy\chi_{y} is defined on {t1,,tm}×\{t_{1},\dots,t_{m}\}\times\mathbb{R} through χy(tk,x)=𝟏{x>yk}\chi_{y}(t_{k},x)={\bf 1}\{x>y_{k}\}. We mention that in [BFS08, Appendix B] it was shown that there is a trace class kernel on L2({t1,,tm}×)L^{2}(\{t_{1},\dots,t_{m}\}\times\mathbb{R}) that is conjugate to χyKχy\chi_{y}K_{\infty}\chi_{y}, so that the Fredholm determinant in (1.18) is well-defined.

Remark 1.11.

In [BFS08] the authors chose a different set of contours in the definition of K(s,x;t,y)K_{\infty}(s,x;t,y) in (1.17), denoted by γ+,γ\gamma_{+},\gamma_{-} in that paper. Starting from [BFS08, (2.7)], we can deform γ+,γ\gamma_{+},\gamma_{-} to C1,π/4,C0,3π/4C_{1,\pi/4},C_{0,3\pi/4}, respectively, without crossing any poles of the integrand and thus without affecting the value of K(s,x;t,y)K_{\infty}(s,x;t,y) by Cauchy’s theorem. The decay necessary to deform the contours near infinity comes from the cubic terms in the exponential functions. We also mention that there is an extra minus sign in the integrand in (1.17), compared to [BFS08, (2.7)], which comes from the fact that γ+\gamma_{+} in that paper is oriented in the direction if decreasing imaginary part, whereas C1,π/4C_{1,\pi/4} is oriented in the direction of increasing imaginary part.

Remark 1.12.

As explained in [BFS08], we have that the finite dimensional distributions of 𝒜21(t+m)\mathcal{A}_{2\rightarrow 1}(t+m) converge to those of 21/3𝒜1(22/3t)2^{1/3}\mathcal{A}_{1}(2^{-2/3}t) as mm\rightarrow\infty, where 𝒜1\mathcal{A}_{1} is the Airy1 process, introduced by Sasamoto [Sas05]. In addition, 𝒜21(t+m)\mathcal{A}_{2\rightarrow 1}(t+m) converges to 𝒜2(t)\mathcal{A}_{2}(t) as mm\rightarrow-\infty, where 𝒜2\mathcal{A}_{2} is the Airy2 process, introduced by Prähofer and Spohn [PS02]. We refer the interested reader to [QR13, QR14] for more background on the Airy2→1 and other Airy processes.

The way that the formula (1.18) was derived in [BFS08] was by constructing a certain process XtX_{t}, related to the TASEP with half-flat initial condition, and showing that

(1.19) limt(k=1m{Xt(tk)yk})=det(IχyKχy)L2({t1,,tm}×).\lim_{t\rightarrow\infty}\mathbb{P}\left(\cap_{k=1}^{m}\{X_{t}(t_{k})\leq y_{k}\}\right)=\det\left(I-\chi_{y}K_{\infty}\chi_{y}\right)_{L^{2}(\{t_{1},\dots,t_{m}\}\times\mathbb{R})}.

The fact that the right side of (1.19) is the pointwise limit of distribution functions on m\mathbb{R}^{m} does not a priori imply that it is itself a distribution function. As we could not find a place in the literature where it is shown that the right side of (1.19) is a distribution function, we show it in the following lemma, whose proof is given in Section 6. In the same lemma we also establish some properties of the finite-dimensional distribution functions of 𝒜21\mathcal{A}_{2\rightarrow 1}, which will be required in our arguments.

Lemma 1.13.

The following all hold.

  1. (1)

    There is a probability space (Ω,,)(\Omega,\mathcal{F},\mathbb{P}) and a real-valued process {𝒜21(t):t}\{\mathcal{A}_{2\rightarrow 1}(t):t\in\mathbb{R}\} on that space, whose finite-dimensional distribution is given by (1.18).

  2. (2)

    For each mm-tuple t1<<tmt_{1}<\cdots<t_{m} the right side of (1.18) is continuous in (y1,,ym)m(y_{1},\dots,y_{m})\in\mathbb{R}^{m}.

  3. (3)

    For each y1,t1y_{1},t_{1}\in\mathbb{R} we have the following identity

    (1.20) det(IχyKχy)L2({t1}×)=1+k=1Ik, where Ik=1(2π𝗂)2kk!C1,π/4k𝑑wC0,3π/4k𝑑zdet[2waewa3/3+t1wa2+𝟏{t10}t12way1wa(za2wb2)(waza)eza3/3+t1za2+𝟏{t10}t12zay1za]a,b=1k,\begin{split}&\det\left(I-\chi_{y}K_{\infty}\chi_{y}\right)_{L^{2}(\{t_{1}\}\times\mathbb{R})}=1+\sum_{k=1}^{\infty}I_{k},\mbox{ where }\\ &I_{k}=\frac{1}{(2\pi\mathsf{i})^{2k}k!}\int_{C_{1,\pi/4}^{k}}d\vec{w}\int_{C_{0,3\pi/4}^{k}}d\vec{z}\det\left[\frac{2w_{a}e^{w_{a}^{3}/3+t_{1}w_{a}^{2}+{\bf 1}\{t_{1}\leq 0\}t_{1}^{2}w_{a}-y_{1}w_{a}}}{(z_{a}^{2}-w_{b}^{2})(w_{a}-z_{a})e^{z_{a}^{3}/3+t_{1}z_{a}^{2}+{\bf 1}\{t_{1}\leq 0\}t_{1}^{2}z_{a}-y_{1}z_{a}}}\right]_{a,b=1}^{k},\end{split}

    C1,π/4,C0,3π/4C_{1,\pi/4},C_{0,3\pi/4} are as in Definition 1.9. Part of the statement is that the integrals in (1.20) are all finite, and the series is absolutely convergent.

1.4. Main result

Here we present the main asymptotic result of the paper, which describes the large time limit of the height function hh as in (1.3), under the measure hfl\mathbb{P}^{\operatorname{h-fl}} from Definition 1.1.

Theorem 1.14.

Assume the same notation as in Definition 1.1. There exists τ~(0,1)\tilde{\tau}\in(0,1) sufficiently small, so that for all τ(0,τ~]\tau\in(0,\tilde{\tau}] the following holds. For each α,y\alpha,y\in\mathbb{R} we have

(1.21) limthfl(t1/3(t/2+t1/3(α2/2)𝟏{α0}h(t/γ,t2/3α))y)=(21/3𝒜21(21/3α)y),\begin{split}&\lim_{t\rightarrow\infty}\mathbb{P}^{\operatorname{h-fl}}\left(t^{-1/3}\cdot\left(t/2+t^{1/3}\cdot(\alpha^{2}/2)\cdot{\bf 1}\{\alpha\leq 0\}-h(t/\gamma,\lfloor t^{2/3}\alpha\rfloor)\right)\leq y\right)\\ &=\mathbb{P}\left(2^{-1/3}\mathcal{A}_{2\rightarrow 1}(2^{-1/3}\alpha)\leq y\right),\end{split}

where 𝒜21\mathcal{A}_{2\rightarrow 1} is as in Definition 1.10.

Remark 1.15.

We mention that Theorem 1.14 agrees with the analogous result for TASEP (corresponding to q=1q=1 and p=0p=0) from [BFS08] as we explain here. If we let Yn(t)Y_{n}(t) denote the location of the nn-th leftmost particle of TASEP at time tt, we have from [BFS08, Theorem 2.2] that for each α,y\alpha,{y}\in\mathbb{R}, and n(α,t)=t/4+(1/2)αt2/3n(\alpha,t)=\lfloor t/4+(1/2)\alpha t^{2/3}\rfloor

(1.22) limthfl(Yn(α,t)(t)αt2/3+(α2/2)𝟏{α0}t1/3t1/3y)=(21/3𝒜21(21/3α)y).\lim_{t\rightarrow\infty}\mathbb{P}^{\operatorname{h-fl}}\left(\frac{Y_{n(\alpha,t)}(t)-\alpha t^{2/3}+(\alpha^{2}/2){\bf 1}\{\alpha\leq 0\}t^{1/3}}{t^{1/3}}\leq{y}\right)=\mathbb{P}\left(2^{-1/3}\mathcal{A}_{2\rightarrow 1}(2^{-1/3}\alpha)\leq{y}\right).

We mention that we have set τ=21/3α\tau=2^{-1/3}\alpha in [BFS08, Theorem 2.2] and also Yn(t)=Xn(t)Y_{n}(t)=-X_{n}(t) as in that paper, since we have started from the initial condition η0=𝟏{x2}\eta_{0}={\bf 1}\{x\in 2\mathbb{N}\} and q=1,p=0q=1,p=0, while in [BFS08] the convention η0=𝟏{x2}\eta_{0}={\bf 1}\{-x\in 2\mathbb{N}\} and q=0,p=1q=0,p=1 is used. Let us set

αt=α+t1/3(y(1/2)α2𝟏{α0}),\alpha_{t}=\alpha+t^{-1/3}\cdot\left(y-(1/2)\alpha^{2}{\bf 1}\{\alpha\leq 0\}\right),

and observe that we have the equality of events

{Yn(α,t)(t)αt2/3+(α2/2)𝟏{α0}t1/3t1/3y}={Yn(α,t)(t)t2/3αt}={Nt2/3αt(t)n(α,t)},\left\{\frac{Y_{n(\alpha,t)}(t)-\alpha t^{2/3}+(\alpha^{2}/2){\bf 1}\{\alpha\leq 0\}t^{1/3}}{t^{1/3}}\leq{y}\right\}=\{Y_{n(\alpha,t)}(t)\leq t^{2/3}\alpha_{t}\}=\{N_{\lfloor t^{2/3}\alpha_{t}\rfloor}(t)\geq n(\alpha,t)\},

where we recall that Nx(t)N_{x}(t) is as in (1.2). In particular, from (1.3) and (1.22) we get

(1.23) limthfl(h(t,t2/3αt)+t2/3αt2t/4+t2/3αtt1/3y+t1/3(α2/2)𝟏{α0}2)=limthfl(h(t,t2/3αt)t/2t1/3y+t1/3(α2/2)𝟏{α0})=(21/3𝒜21(21/3α)y).\begin{split}&\lim_{t\rightarrow\infty}\mathbb{P}^{\operatorname{h-fl}}\left(\frac{h(t,\lfloor t^{2/3}\alpha_{t}\rfloor)+\lfloor t^{2/3}\alpha_{t}\rfloor}{2}\geq t/4+t^{2/3}\cdot\frac{\alpha_{t}-t^{-1/3}y+t^{-1/3}\cdot(\alpha^{2}/2)\cdot{\bf 1}\{\alpha\leq 0\}}{2}\right)\\ &=\lim_{t\rightarrow\infty}\mathbb{P}^{\operatorname{h-fl}}\left(h(t,\lfloor t^{2/3}\alpha_{t}\rfloor)\rfloor\geq t/2-t^{1/3}y+t^{1/3}\cdot(\alpha^{2}/2)\cdot{\bf 1}\{\alpha\leq 0\}\right)\\ &=\mathbb{P}\left(2^{-1/3}\mathcal{A}_{2\rightarrow 1}(2^{-1/3}\alpha)\leq{y}\right).\end{split}

Equation (1.23) agrees with (1.21) as γ=qp=1\gamma=q-p=1. We mention that (1.23) does not follow from Theorem 1.14, as p=0p=0 is not allowed, and the point of this computation is to provide a sanity check for the scaling we perform in Theorem 1.14.

Outline

The rest of the paper is organized as follows. In Section 2 we prove Theorem 1.5, which is the starting point of our asymptotic analysis. In Section 3 we present the proof of Theorem 1.14, which relies on taking the limit of the identity (1.12), when ζ\zeta is scaled appropriately with tt and tt\rightarrow\infty. In order to obtain the limit of (1.12), we need to show that each summand converges, which is the statement of Proposition 3.2, and that we can exchange the order of the series and the limit, which is ensured by Proposition 3.3. Proposition 3.2 is proved in Section 4 and relies on the method of steepest descent, as well as various estimates of the functions in (1.7). Proposition 3.3 is proved in Section 5 and is based on estimating the functions in (1.8) along two classes of contours, depending on whether kk is large or small relative to tt. In Section 6 we prove Lemma 1.13.

Acknowledgments

We are grateful to Alexei Borodin, Jeremy Quastel and Daniel Remenik for their useful comments on earlier drafts of the paper. ED is partially supported by NSF grant DMS:2054703.

2. Prelimit formula

In this section we present the proof of Theorem 1.5, which is the starting point of our asymptotic analysis in the next sections. The proof of Theorem 1.5 is given in Section 2.2 and in Section 2.1 below we establish several statements, which will be required. We continue with the same notation as in Section 1.

2.1. Definitions and notation

In this section we present two key lemmas, which will be required in the proof of Theorem 1.5 in the next section, and whose proofs are postponed until Section 2.3. After stating the lemmas we summarize various estimates, which will also be of future use.

The first key result we require is as follows.

Lemma 2.1.

Fix kk\in\mathbb{N} and ζ[0,)\zeta\in\mathbb{C}\setminus[0,\infty). Then, the integral Hk(ζ)H_{k}(\zeta) in (1.13) is well-defined and finite. Moreover, as a function of ζ\zeta we have that Hk(ζ)H_{k}(\zeta) is analytic in [0,)\mathbb{C}\setminus[0,\infty). If τ\tau satisfies (1.11), then the series on the right side of (1.12) is absolutely convergent for each ζ[0,)\zeta\in\mathbb{C}\setminus[0,\infty) and defines an analytic function in [0,)\mathbb{C}\setminus[0,\infty).

The second key lemma we need is the following. It is a special case of a Mellin-Barnes type integral representation from [BC14, Lemma 3.20].

Lemma 2.2.

Let N,kN,k\in\mathbb{N} and ζ[0,)\zeta\in\mathbb{C}\setminus[0,\infty) be such that |ζ|<1|\zeta|<1. Then, the series

(2.1) n1=1nk=1ζn1++nkk!(2π𝗂)kγ1,0k𝑑wF(n,w)\begin{split}&\sum_{n_{1}=1}^{\infty}\cdots\sum_{n_{k}=1}^{\infty}\frac{\zeta^{n_{1}+\cdots+n_{k}}}{k!(2\pi\mathsf{i})^{k}}\oint_{\gamma_{-1,0}^{k}}d\vec{w}F(\vec{n},\vec{w})\end{split}

is absolutely convergent and equals Hk(ζ)H_{k}(\zeta) from (1.13). Here, F(s,w)F(\vec{s},\vec{w}) is as in (1.8) and γ1,0\gamma_{-1,0} is as in Proposition 1.2.

In the remainder of this section we summarize several basic estimates for the various functions that appear in (1.8). We observe that given c>0c>0 we can find c>0c^{\prime}>0 such that if x,yx,y\in\mathbb{R} and d(x+𝗂y,)cd(x+\mathsf{i}y,\mathbb{Z})\geq c, then

(2.2) 1|sin(πx+𝗂πy)|ceπ|y|.\frac{1}{|\sin(\pi x+\mathsf{i}\pi y)|}\leq c^{\prime}e^{-\pi|y|}.

Recall the Cauchy determinant formula, see e.g. [Pra94, 1.3],

(2.3) det[1xiyj]i,j=1N=1i<jN(xixj)(yjyi)i,j=1N(xiyj).\det\left[\frac{1}{x_{i}-y_{j}}\right]_{i,j=1}^{N}=\frac{\prod_{1\leq i<j\leq N}(x_{i}-x_{j})(y_{j}-y_{i})}{\prod_{i,j=1}^{N}(x_{i}-y_{j})}.

The following statement summarizes two different bounds on the above Cauchy determinant.

Lemma 2.3.

[Dim20, Lemma 3.12] Let NN\in\mathbb{N}.

  1. (1)

    Hadamard’s inequality: If AA is an N×NN\times N matrix and v1,,vNv_{1},\dots,v_{N} denote the column vectors of AA, then |detA|i=1Nvi|\det A|\leq\prod_{i=1}^{N}\|v_{i}\| where x=(x12++xN2)1/2\|x\|=(x_{1}^{2}+\cdots+x_{N}^{2})^{1/2} for x=(x1,,xN)x=(x_{1},\dots,x_{N}).

  2. (2)

    Fix r,R(0,)r,R\in(0,\infty) with R>rR>r. Let zi,wiz_{i},w_{i}\in\mathbb{C} be such that |wi|=R|w_{i}|=R and |zi|r|z_{i}|\leq r for i=1,,Ni=1,\dots,N. Then,

    (2.4) |det[1ziwj]i,j=1N|RNNN(r/R)(N2)(1r/R)N2.\left|\det\left[\frac{1}{z_{i}-w_{j}}\right]_{i,j=1}^{N}\right|\leq R^{-N}\cdot\frac{N^{N}\cdot(r/R)^{\binom{N}{2}}}{(1-r/R)^{N^{2}}}.

Let 𝔣,𝔤,𝔥\mathfrak{f},\mathfrak{g},\mathfrak{h} be as in (1.7). Then, we can find A1>0A_{1}>0, depending on τ,t,p,q,x\tau,t,p,q,x, such that for all wγ1,0w\in\gamma_{-1,0} (recall this was the positively oriented circle of radius τ1/8\tau^{-1/8}, centered at the origin) and ss\in\mathbb{C} such that 𝖱𝖾(s)1/2\mathsf{Re}(s)\geq 1/2 we have

(2.5) |𝔣(w;s)𝔤(w;s)|A1.\left|\mathfrak{f}(w;s)\mathfrak{g}(w;s)\right|\leq A_{1}.

Also if w1,w2γ1,0w_{1},w_{2}\in\gamma_{-1,0} and s1,s2s_{1},s_{2}\in\mathbb{C} with 𝖱𝖾(s1),𝖱𝖾(s2)1/2\mathsf{Re}(s_{1}),\mathsf{Re}(s_{2})\geq 1/2 we have

(2.6) |𝔥(w1,w2;s1,s2)|(1+τ1/4)(τ3/4;τ)(τ3/4;τ)(τ1/4;τ)(τ1/4;τ),\left|\mathfrak{h}(w_{1},w_{2};s_{1},s_{2})\right|\leq(1+\tau^{-1/4})\cdot\frac{(-\tau^{3/4};\tau)_{\infty}(-\tau^{3/4};\tau)_{\infty}}{(\tau^{1/4};\tau)_{\infty}(\tau^{1/4};\tau)_{\infty}},

where we used that for |z|=r[0,1)|z|=r\in[0,1) the function |(z;τ)||(z;\tau)_{\infty}| is maximized when z=rz=-r and minimized when z=rz=r.

Finally, from Lemma 2.3(ii) applied to r=τ3/8r=\tau^{3/8} and R=τ1/8R=\tau^{-1/8} we have for all wiγ1,0w_{i}\in\gamma_{-1,0} and sis_{i}\in\mathbb{C} with 𝖱𝖾(si)1/2\mathsf{Re}(s_{i})\geq 1/2 for i=1,,ki=1,\dots,k that

(2.7) |det[1waτsawb]a,b=1k|τk/8kkτ12(k2)(1τ1/2)k2.\left|\det\left[\frac{-1}{w_{a}\tau^{s_{a}}-w_{b}}\right]_{a,b=1}^{k}\right|\leq\tau^{k/8}\frac{k^{k}\cdot\tau^{\frac{1}{2}\binom{k}{2}}}{(1-\tau^{1/2})^{k^{2}}}.

Combining (2.5), (2.6) and (2.7) we conclude that for each kk\in\mathbb{N}, wiγ1,0w_{i}\in\gamma_{-1,0} and sis_{i}\in\mathbb{C} with 𝖱𝖾(si)1/2\mathsf{Re}(s_{i})\geq 1/2 for i=1,,ki=1,\dots,k we have

(2.8) 1k!|det[1waτsawb]a,b=1k𝔣(wa;sa)𝔤(wa;sa)1a<bk𝔥(wa,wb;sa,sb)|Akρ(k2), where A=eA1τ1/8(1τ1/2) and ρ=(τ1/2+τ1/4)(1τ1/2)2(τ3/4;τ)(τ3/4;τ)(τ1/4;τ)(τ1/4;τ).\begin{split}&\frac{1}{k!}\left|\det\left[\frac{-1}{w_{a}\tau^{s_{a}}-w_{b}}\right]_{a,b=1}^{k}\mathfrak{f}(w_{a};s_{a})\mathfrak{g}(w_{a};s_{a})\cdot\prod_{1\leq a<b\leq k}\mathfrak{h}(w_{a},w_{b};s_{a},s_{b})\right|\leq A^{k}\cdot\rho^{\binom{k}{2}},\mbox{ where }\\ &A=\frac{eA_{1}\tau^{1/8}}{(1-\tau^{1/2})}\mbox{ and }\rho=\frac{(\tau^{1/2}+\tau^{1/4})}{(1-\tau^{1/2})^{2}}\cdot\frac{(-\tau^{3/4};\tau)_{\infty}(-\tau^{3/4};\tau)_{\infty}}{(\tau^{1/4};\tau)_{\infty}(\tau^{1/4};\tau)_{\infty}}.\end{split}

In deriving the last inequality we used that kkk!ekk^{k}\leq k!e^{k}, which can be deduced from

(2.9) n!=2πnn+1/2enern for n, where 112n+1<rn<112n,n!=\sqrt{2\pi}n^{n+1/2}e^{-n}e^{r_{n}}\mbox{ for }n\in\mathbb{N},\mbox{ where }\frac{1}{12n+1}<r_{n}<\frac{1}{12n},

see [Rob55, Equation (1)].

2.2. Proof of Theorem 1.5

We continue with the same notation as in the statement of the theorem. For clarity we split the proof into two steps.

Step 1. Note that by Lemma 2.1 each summand on the right side of (1.12) is well-defined and finite, and moreover by our assumption on τ\tau in (1.11) we have that the series on the right side is absolutely convergent and defines an analytic function in ζ\zeta on [0,)\mathbb{C}\setminus[0,\infty).

We claim that (1.12) holds provided that ζ[0,)\zeta\in\mathbb{C}\setminus[0,\infty) is such that |ζ|<1|\zeta|<1. We will prove this statement in the next step. Here we assume its validity and proceed to prove that (1.12) holds for all ζ[0,)\zeta\in\mathbb{C}\setminus[0,\infty).

From (1.6) we have

(2.10) 𝔼hfl[eτ(ζτNx(t))]=n=01((1τ)ζτn;τ)hfl(Nx(t)=n).\mathbb{E}^{\operatorname{h-fl}}\left[e_{\tau}\left(\zeta\tau^{N_{x}(t)}\right)\right]=\sum_{n=0}^{\infty}\frac{1}{((1-\tau)\zeta\tau^{n};\tau)_{\infty}}\cdot\mathbb{P}^{\operatorname{h-fl}}(N_{x}(t)=n).

If K[0,)K\subset\mathbb{C}\setminus[0,\infty) is compact we observe that so is the set K^={0}n=0τnK\hat{K}=\{0\}\cup\cup_{n=0}^{\infty}\tau^{n}\cdot K, and the latter is separated from the zeros of the function ((1τ)z;τ)((1-\tau)z;\tau)_{\infty}. In particular, we conclude that there exists a constant CC, depending on KK and τ\tau, such that for all ζK\zeta\in K and n0n\in\mathbb{Z}_{\geq 0} we have

|1((1τ)ζτn;τ)|C.\left|\frac{1}{((1-\tau)\zeta\tau^{n};\tau)_{\infty}}\right|\leq C.

The latter implies that the right side of (2.10) is the uniform over ζK\zeta\in K limit of

n=0N1((1τ)ζτn;τ)hfl(Nx(t)=n)\sum_{n=0}^{N}\frac{1}{((1-\tau)\zeta\tau^{n};\tau)_{\infty}}\cdot\mathbb{P}^{\operatorname{h-fl}}(N_{x}(t)=n)

as NN\rightarrow\infty. Since each of the above functions are analytic in ζ\zeta on [0,)\mathbb{C}\setminus[0,\infty), we conclude the same is true for the expressions in (2.10), cf. [SS03, Chapter 2, Theorem 5.2].

Our work in the above paragraph shows that the left side of (1.12) is analytic in ζ\zeta on [0,)\mathbb{C}\setminus[0,\infty) and from Lemma 2.1 we know the same is true for the right side. As these two functions agree when ζ[0,)\zeta\in\mathbb{C}\setminus[0,\infty) is such that |ζ|<1|\zeta|<1 by assumption, we conclude that they agree for all ζ[0,)\zeta\in\mathbb{C}\setminus[0,\infty), cf. [SS03, Chapter 2, Theorem 4.8]. This concludes the proof of the theorem.

Step 2. In this step we fix ζ[0,)\zeta\in\mathbb{C}\setminus[0,\infty) such that |ζ|<1|\zeta|<1 and proceed to prove (1.12). From (1.6)

(2.11) 𝔼hfl[eτ(ζτNx(t))]=𝔼hfl[m=0ζmτmNx(t)mτ!]=limNm=0N𝔼hfl[ζmτmNx(t)mτ!],\mathbb{E}^{\operatorname{h-fl}}\left[e_{\tau}\left(\zeta\tau^{N_{x}(t)}\right)\right]=\mathbb{E}^{\operatorname{h-fl}}\left[\sum_{m=0}^{\infty}\frac{\zeta^{m}\tau^{mN_{x}(t)}}{m_{\tau}!}\right]=\lim_{N\rightarrow\infty}\sum_{m=0}^{N}\mathbb{E}^{\operatorname{h-fl}}\left[\frac{\zeta^{m}\tau^{mN_{x}(t)}}{m_{\tau}!}\right],

where we used Fubini’s theorem to exchange the order of the sum and the expectation. Note that the application of Fubini’s theorem is justified since

𝔼hfl[m=0|ζmτmNx(t)mτ!|]𝔼hfl[m=0|ζ|mmτ!]=eτ(|ζ|)<,\mathbb{E}^{\operatorname{h-fl}}\left[\sum_{m=0}^{\infty}\left|\frac{\zeta^{m}\tau^{mN_{x}(t)}}{m_{\tau}!}\right|\right]\leq\mathbb{E}^{\operatorname{h-fl}}\left[\sum_{m=0}^{\infty}\frac{|\zeta|^{m}}{m_{\tau}!}\right]=e_{\tau}\left(|\zeta|\right)<\infty,

where we used that Nx(t)0N_{x}(t)\geq 0 almost surely, τ(0,1)\tau\in(0,1) and |ζ|<1|\zeta|<1 by assumption.

We further have from Proposition 1.2 for each NN\in\mathbb{N} that

(2.12) m=0N𝔼hfl[ζmτmNx(t)mτ!]=1+k=1n1=1nk=1HkN(ζ;n1,,nk), where HkN(ζ;n1,,nk)=𝟏{n1++nkN}ζn1++nkk!(2π𝗂)kγ1,0k𝑑wF(n,s).\begin{split}&\sum_{m=0}^{N}\mathbb{E}^{\operatorname{h-fl}}\left[\frac{\zeta^{m}\tau^{mN_{x}(t)}}{m_{\tau}!}\right]=1+\sum_{k=1}^{\infty}\sum_{n_{1}=1}^{\infty}\cdots\sum_{n_{k}=1}^{\infty}H_{k}^{N}(\zeta;n_{1},\dots,n_{k})\mbox{, where }\\ &H_{k}^{N}(\zeta;n_{1},\dots,n_{k})={\bf 1}\{n_{1}+\cdots+n_{k}\leq N\}\cdot\frac{\zeta^{n_{1}+\cdots+n_{k}}}{k!(2\pi\mathsf{i})^{k}}\oint_{\gamma_{-1,0}^{k}}d\vec{w}F(\vec{n},\vec{s}).\end{split}

From (2.8) and the fact that γ1,0\gamma_{-1,0} has length 2πτ1/82\pi\tau^{-1/8} we have

(2.13) |HkN(ζ;n1,,nk)|Akρ(k2)τk/8|ζ|n1++nk.\left|H_{k}^{N}(\zeta;n_{1},\dots,n_{k})\right|\leq A^{k}\cdot\rho^{\binom{k}{2}}\cdot\tau^{-k/8}|\zeta|^{n_{1}+\cdots+n_{k}}.

In addition, from (2.13) and the dominated convergence theorem we have

limNn1=1nk=1HkN(ζ;n1,,nk)=n1=1nk=1ζn1++nkk!(2π𝗂)kγ1,0k𝑑wF(n,w).\begin{split}&\lim_{N\rightarrow\infty}\sum_{n_{1}=1}^{\infty}\cdots\sum_{n_{k}=1}^{\infty}H_{k}^{N}(\zeta;n_{1},\dots,n_{k})=\sum_{n_{1}=1}^{\infty}\cdots\sum_{n_{k}=1}^{\infty}\frac{\zeta^{n_{1}+\cdots+n_{k}}}{k!(2\pi\mathsf{i})^{k}}\oint_{\gamma_{-1,0}^{k}}d\vec{w}F(\vec{n},\vec{w}).\end{split}

The last equation and Lemma 2.2 together imply

(2.14) limNn1=1nk=1HkN(ζ;n1,,nk)=Hk(ζ).\lim_{N\rightarrow\infty}\sum_{n_{1}=1}^{\infty}\cdots\sum_{n_{k}=1}^{\infty}H_{k}^{N}(\zeta;n_{1},\dots,n_{k})=H_{k}(\zeta).

Combining (2.11) and (2.12) we conclude

(2.15) 𝔼hfl[eτ(ζτNx(t))]=limN1+k=11k!n1=1nk=1HkN(ζ;n1,,nk)=1+k=1limN1k!n1=1nk=1HkN(ζ;n1,,nk)=1+k=1Hk(ζ),\begin{split}&\mathbb{E}^{\operatorname{h-fl}}\left[e_{\tau}\left(\zeta\tau^{N_{x}(t)}\right)\right]=\lim_{N\rightarrow\infty}1+\sum_{k=1}^{\infty}\frac{1}{k!}\sum_{n_{1}=1}^{\infty}\cdots\sum_{n_{k}=1}^{\infty}H_{k}^{N}(\zeta;n_{1},\dots,n_{k})=\\ &1+\sum_{k=1}^{\infty}\lim_{N\rightarrow\infty}\frac{1}{k!}\sum_{n_{1}=1}^{\infty}\cdots\sum_{n_{k}=1}^{\infty}H_{k}^{N}(\zeta;n_{1},\dots,n_{k})=1+\sum_{k=1}^{\infty}H_{k}(\zeta),\end{split}

where in the last equality we used (2.14). We mention that in exchanging the order of the sum and the limit above we used the dominated convergence theorem as from (2.13) we have

n1=1nk=1|HkN(ζ;n1,,nk)|(Aτ1/81|ζ|)kρ(k2),\sum_{n_{1}=1}^{\infty}\cdots\sum_{n_{k}=1}^{\infty}\left|H_{k}^{N}(\zeta;n_{1},\dots,n_{k})\right|\leq\left(\frac{A\tau^{-1/8}}{1-|\zeta|}\right)^{k}\rho^{\binom{k}{2}},

and the latter is summable over kk\in\mathbb{N}. From (2.15) we get (1.12) when |ζ|<1|\zeta|<1, as desired.

Remark 2.4.

Now that we have presented our proof of Theorem 1.5, let us explain how our approach compares with that in [OQR16]. The overall strategy of both proofs is quite similar, in that (1.12) is first established when ζ[0,)\zeta\in\mathbb{C}\setminus[0,\infty) is such that |ζ|<1|\zeta|<1, and then extended to ζ[0,)\zeta\in\mathbb{C}\setminus[0,\infty) by analytic continuation. In the case of |ζ|<1|\zeta|<1 the main difficulty is in taking the NN\rightarrow\infty limit of (2.12) and justifying exchanging the order of the series on the right side of that equation and the limit. To analytically extend (1.12) to ζ[0,)\zeta\in\mathbb{C}\setminus[0,\infty), the main difficulty is in showing that the series on the right side of (1.12) is absolutely convergent and hence analytic in ζ\zeta (as the absolutely convergent series of analytic in ζ\zeta functions).

In order to overcome the two challenges above, one needs to find good bounds on the summands in the two series in (1.12) and (2.12). One runs into trouble, since over γ1,0k\gamma_{-1,0}^{k} the best pointwise estimate for |1a<bk𝔥(wa,wb;sa,sb)||\prod_{1\leq a<b\leq k}\mathfrak{h}(w_{a},w_{b};s_{a},s_{b})| (which appears in both HkN(ζ;n1,,nk)H_{k}^{N}(\zeta;n_{1},\dots,n_{k}) and Hk(ζ)H_{k}(\zeta)) is ec~k2e^{\tilde{c}k^{2}} for some c~>0\tilde{c}>0. This bound behaves very poorly in kk, and cannot be compensated by the k!k! in the denominator of Hk(ζ)H_{k}(\zeta) in (1.13). Faced with this difficulty, the authors [OQR16] suggested to deform γ1,0k\gamma_{-1,0}^{k} to certain kk-dependent contours γ¯kk\overline{\gamma}_{k}^{k}, and for general ζ[0,)\zeta\in\mathbb{C}\setminus[0,\infty) also deform (1/2+𝗂)k(1/2+\mathsf{i}\mathbb{R})^{k} to certain kk-dependent contours D¯kk\overline{D}_{k}^{k}. Over the contours γ¯kk\overline{\gamma}^{k}_{k} and D¯kk\overline{D}_{k}^{k} one can obtain a favorable estimate for |1a<bk𝔥(wa,wb;sa,sb)||\prod_{1\leq a<b\leq k}\mathfrak{h}(w_{a},w_{b};s_{a},s_{b})|. Unfortunately, what was not realized in [OQR16] is that over these deformed contours other parts of the integrands in the definitions of HkN(ζ;n1,,nk)H_{k}^{N}(\zeta;n_{1},\dots,n_{k}) and Hk(ζ)H_{k}(\zeta) behave badly and become difficult to control.

The way we overcome the growth of |1a<bk𝔥(wa,wb;na,nb)||\prod_{1\leq a<b\leq k}\mathfrak{h}(w_{a},w_{b};n_{a},n_{b})| as a function of kk, is to recognize that one has some decay built into the Cauchy determinant in the definitions of HkN(ζ;n1,,nk)H_{k}^{N}(\zeta;n_{1},\dots,n_{k}) and Hk(ζ)H_{k}(\zeta). This decay is sufficient to control the series in (1.12) and (2.12), only when τ\tau is sufficiently small – see (1.11). In fact, it still remains a difficulty for us to prove that the right side of (1.12) is convergent and hence well-defined for general τ(0,1)\tau\in(0,1).

2.3. Proof of Lemmas 2.1 and 2.2

In this section we give the proofs of the two key lemmas from Section 2.1.

Proof of Lemma 2.1.

Let us fix a compact set K[0,)K\subset\mathbb{C}\setminus[0,\infty) and ζK\zeta\in K. It follows from (2.2) that there are constants C1,c1>0C_{1},c_{1}>0, depending on KK alone, such that for all s=1/2+𝗂y1/2+𝗂s=1/2+\mathsf{i}y\in 1/2+\mathsf{i}\mathbb{R}

(2.16) |π(ζ)ssin(πs)|C1ec1|y|.\left|\frac{\pi(-\zeta)^{s}}{\sin(-\pi s)}\right|\leq C_{1}e^{-c_{1}|y|}.

Combining (2.16) with (2.8) we conclude that the integral in (1.13) is well-defined and finite. Moreover, we see that Hk(ζ)H_{k}(\zeta) is the uniform over ζK\zeta\in K limit of

GkN(ζ):=1k!(2π𝗂)2k(1/2+𝗂[N,N])k𝑑sγ1,0k𝑑wF(ζ;s,w) as N.\begin{split}&G^{N}_{k}(\zeta):=\frac{1}{k!(2\pi\mathsf{i})^{2k}}\cdot\int_{(1/2+\mathsf{i}[-N,N])^{k}}d\vec{s}\oint_{\gamma_{-1,0}^{k}}d\vec{w}F(\zeta;\vec{s},\vec{w})\mbox{ as $N\rightarrow\infty$.}\end{split}

The integrand in GkN(ζ)G_{k}^{N}(\zeta) is analytic in ζ\zeta on [0,)\mathbb{C}\setminus[0,\infty) and is jointly continuous in ζ\zeta and w,s\vec{w},\vec{s}, while the contours we are integrating over are compact. The latter implies that GkN(ζ)G_{k}^{N}(\zeta) is analytic in ζ\zeta for each NN, see e.g. [SS03, Theorem 5.4]. We thus conclude that Hk(ζ)H_{k}(\zeta) is analytic as the uniform over compact sets limit of analytic functions, cf. [SS03, Chapter 2, Theorem 5.2].

Using that for λ>0\lambda>0 we have eλ|x|𝑑x=2λ1\int_{\mathbb{R}}e^{-\lambda|x|}dx=2\lambda^{-1}, that the length of γ1,0\gamma_{-1,0} is 2πτ1/82\pi\tau^{-1/8} as well as (2.8) and (2.16) we conclude that

|Hk(ζ)|(C1Aτ1/8πc1)kρ(k2).\left|H_{k}(\zeta)\right|\leq\left(\frac{C_{1}A\tau^{-1/8}}{\pi c_{1}}\right)^{k}\cdot\rho^{\binom{k}{2}}.

The right side above is summable over kk, since ρ(0,1)\rho\in(0,1) in view of (1.11). The latter implies that the series in (1.12) is absolutely convergent. Moreover, since each summand is analytic in ζ\zeta on [0,)\mathbb{C}\setminus[0,\infty), we conclude the same is true for the series by [SS03, Chapter 2, Theorem 5.2]. ∎

Proof of Lemma 2.2.

From (2.8) and the fact that the length of γ1,0\gamma_{-1,0} is 2πτ1/82\pi\tau^{-1/8} we conclude that the (n1,,nk)(n_{1},\dots,n_{k})-th summand in (2.1) is bounded in absolute value by

|ζ|n1++nkAkρ(k2)τk/8(2π)k,|\zeta|^{n_{1}+\cdots+n_{k}}\cdot A^{k}\cdot\rho^{\binom{k}{2}}\cdot\tau^{-k/8}\cdot(2\pi)^{-k},

and the latter is summable over (n1,,nk)k(n_{1},\dots,n_{k})\in\mathbb{N}^{k} (since |ζ|<1|\zeta|<1 by assumption), proving the absolute convergence of the series in (2.1).

To conclude the proof of the lemma, it suffices to show

(2.17) limNn1=1Nnk=1Nζn1++nkk!(2π𝗂)kγ1,0k𝑑wF(n,w)=Hk(ζ).\begin{split}&\lim_{N\rightarrow\infty}\sum_{n_{1}=1}^{N}\cdots\sum_{n_{k}=1}^{N}\frac{\zeta^{n_{1}+\cdots+n_{k}}}{k!(2\pi\mathsf{i})^{k}}\oint_{\gamma_{-1,0}^{k}}d\vec{w}F(\vec{n},\vec{w})=H_{k}(\zeta).\end{split}

Let RN=1/2+NR_{N}=1/2+N and set AN1=1/2𝗂RNA^{1}_{N}=1/2-\mathsf{i}R_{N}, AN2=1/2+𝗂RNA^{2}_{N}=1/2+\mathsf{i}R_{N}, AN3=RN+𝗂RNA^{3}_{N}=R_{N}+\mathsf{i}R_{N}, AN4=RN𝗂RNA^{4}_{N}=R_{N}-\mathsf{i}R_{N}. Denote by γN1\gamma_{N}^{1} the contour, which goes from AN1A_{N}^{1} vertically up to AN2A^{2}_{N}, by γ2N\gamma_{2}^{N} the contour, which goes from AN2A^{2}_{N} horizontally to AN3A_{N}^{3}, by γN3\gamma_{N}^{3} the contour, which goes from AN3A_{N}^{3} vertically down to AN4A_{N}^{4} and by γN4\gamma_{N}^{4} the contour, which goes from AN4A^{4}_{N} horizontally to AN1A_{N}^{1}. Also let γN=i=14γNi\gamma_{N}=\cup_{i=1}^{4}\gamma^{i}_{N} traversed in order, see Figure 1.

Refer to caption
Figure 1. The contours γNi\gamma_{N}^{i} for i=1,,4i=1,\dots,4.

We observe by the Residue Theorem that for each NN\in\mathbb{N}

(2.18) n1=1Nnk=1Nζn1++nkk!(2π𝗂)kγ1,0k𝑑wF(n,w)=1k!(2π𝗂)2kγ1,0k𝑑w(γN)k𝑑sF(ζ;s,w),\begin{split}&\sum_{n_{1}=1}^{N}\cdots\sum_{n_{k}=1}^{N}\frac{\zeta^{n_{1}+\cdots+n_{k}}}{k!(2\pi\mathsf{i})^{k}}\oint_{\gamma_{-1,0}^{k}}d\vec{w}F(\vec{n},\vec{w})=\frac{1}{k!(2\pi\mathsf{i})^{2k}}\oint_{\gamma_{-1,0}^{k}}d\vec{w}\int_{(\gamma_{N})^{k}}d\vec{s}F(\zeta;\vec{s},\vec{w}),\end{split}

where we recall that F(ζ;s,w)F(\zeta;\vec{s},\vec{w}) was defined in (1.8). In deriving the last expression we used that in each variable sas_{a} for a=1,,ka=1,\dots,k the function F(ζ;s,w)F(\zeta;\vec{s},\vec{w}) is analytic in the region enclosed by γN\gamma_{N} except at the points sa{1,,N}s_{a}\in\{1,\dots,N\} where the function has a simple pole coming from πsin(πsa)\frac{\pi}{\sin(-\pi s_{a})}, the fact that

𝖱𝖾𝗌z=mπsin(πsa)=(1)m+1,\mathsf{Res}_{z=m}\frac{\pi}{\sin(-\pi s_{a})}=(-1)^{m+1},

and also that γN\gamma_{N} is negatively oriented.

We next note by the dominated convergence theorem that

(2.19) limN1k!(2π𝗂)2kγ1,0k𝑑w(γN1)k𝑑sF(ζ;s,w)=Hk(ζ).\begin{split}&\lim_{N\rightarrow\infty}\frac{1}{k!(2\pi\mathsf{i})^{2k}}\oint_{\gamma_{-1,0}^{k}}d\vec{w}\int_{(\gamma^{1}_{N})^{k}}d\vec{s}F(\zeta;\vec{s},\vec{w})=H_{k}(\zeta).\end{split}

In deriving the last statement we used (2.8) and (2.16), which justify the dominated convergence theorem with dominating function (C1A)kρ(k2)a=1kec1|ya|(C_{1}A)^{k}\rho^{\binom{k}{2}}\prod_{a=1}^{k}e^{-c_{1}|y_{a}|} (here C1,c1C_{1},c_{1} are as in (2.16) and we have written sa=xa+𝗂yas_{a}=x_{a}+\mathsf{i}y_{a} for a=1,,ka=1,\dots,k).

In addition, we have from (2.2) and (2.8), the fact that the length of γ0,1\gamma_{0,-1} is 2πτ1/82\pi\tau^{-1/8}, the length of γNr\gamma_{N}^{r} is NN for r=2,4r=2,4 and |ζ|<1|\zeta|<1 that for r=2,4r=2,4

(2.20) |1k!(2π𝗂)2kγ1,0k𝑑w(γNr)k𝑑sF(ζ;s,w)|Akρ(k2)Nkτk/8(2π)kckekπN.\begin{split}&\left|\frac{1}{k!(2\pi\mathsf{i})^{2k}}\oint_{\gamma_{-1,0}^{k}}d\vec{w}\int_{(\gamma^{r}_{N})^{k}}d\vec{s}F(\zeta;\vec{s},\vec{w})\right|\leq\frac{A^{k}\rho^{\binom{k}{2}}N^{k}\tau^{-k/8}\cdot}{(2\pi)^{k}}\cdot c^{\prime k}e^{-k\pi N}.\end{split}

Using also that the length of γN3\gamma_{N}^{3} is 2N+12N+1, and the same statements as above we get

(2.21) |1k!(2π𝗂)2kγ1,0k𝑑w(γN3)k𝑑sF(ζ;s,w)||ζ|k(N+1/2)Akρ(k2)(2N+1)kτk/8(2π)kck.\begin{split}&\left|\frac{1}{k!(2\pi\mathsf{i})^{2k}}\oint_{\gamma_{-1,0}^{k}}d\vec{w}\int_{(\gamma^{3}_{N})^{k}}d\vec{s}F(\zeta;\vec{s},\vec{w})\right|\leq|\zeta|^{k(N+1/2)}\frac{A^{k}\rho^{\binom{k}{2}}(2N+1)^{k}\tau^{-k/8}}{(2\pi)^{k}}\cdot c^{\prime k}.\end{split}

We mention that in (2.20) and (2.21) the constant cc^{\prime} is as in (2.2) for c=1/2c=1/2, and we used that γNi\gamma_{N}^{i} are at least distance 1/21/2 from \mathbb{Z} by construction.

Since |ζ|<1|\zeta|<1 by assumption, we have that the right sides of (2.20) and (2.21) both converge to zero as NN\rightarrow\infty. Combining (2.18), (2.19), (2.20) and (2.21) we conclude (2.17), which concludes the proof of the lemma.

3. Weak convergence

The goal of this section is to prove Theorem 1.14, for which we need to study equation (1.12) in Theorem 1.5 as tt\rightarrow\infty. In Seciton 3.1 we explain how we need to scale the parameters in (1.12), and formulate two key asymptotic statements about the summands Hk(ζ)H_{k}(\zeta) – see Propositions 3.2 and 3.3. In Section 3.2 we use these two propositions to complete the proof of Theorem 1.14. In Section 3.3 we present a useful way to rewrite Hk(ζ)H_{k}(\zeta), which will help us establish Propositions 3.2 and 3.3, whose proofs are given in Sections 4 and 5, respectively. Throughout this section we continue with the same notation as in Sections 1 and 2.

3.1. Two key propositions

In this section we formulate the key results we require in the proof of Theorem 1.14 in Section 3.2. We begin by stating our assumptions on parameters and their scaling.

Definition 3.1.

We assume the same notation as in Definition 1.1. We further fix r~,α\tilde{r},\alpha\in\mathbb{R}, and for t>0t>0 define

(3.1) ζ=(1τ)1τ(1/4)t(1/2)(t2/3α1)+t1/3r~.\zeta=-(1-\tau)^{-1}\tau^{-(1/4)t-(1/2)(\lfloor t^{2/3}\alpha\rfloor-1)+t^{1/3}\tilde{r}}.

With this data we let I(k,t)=Hk(ζ)I(k,t)=H_{k}(\zeta), where Hk(ζ)H_{k}(\zeta) is as in (1.13) with p,q,τ,ζp,q,\tau,\zeta as just specified, x=t2/3αx=\lfloor t^{2/3}\alpha\rfloor and tt replaced with t/γt/\gamma. We mention that I(k,t)I(k,t) is well-defined and finite in view of Lemma 2.1 and the fact that ζ\zeta from (3.1) lies in (,0)(-\infty,0).

The first key proposition we require is as follows.

Proposition 3.2.

Assume the same notation as in Definition 3.1. For kk\in\mathbb{N}

(3.2) limtI(k,t)=1k!(2π𝗂)2kC0,π/4k𝑑uC1,3π/4k𝑑vdet[2vaeua3/48ua2α/8uar~(uava)(vb2ua2)eva3/48va2α/8var~]a,b=1k,\lim_{t\rightarrow\infty}I(k,t)=\frac{1}{k!(2\pi\mathsf{i})^{2k}}\int_{C_{0,\pi/4}^{k}}d\vec{u}\int_{C_{-1,3\pi/4}^{k}}d\vec{v}\det\left[\frac{2v_{a}e^{u_{a}^{3}/48-u_{a}^{2}\alpha/8-u_{a}\tilde{r}}}{(u_{a}-v_{a})(v_{b}^{2}-u_{a}^{2})e^{v_{a}^{3}/48-v_{a}^{2}\alpha/8-v_{a}\tilde{r}}}\right]_{a,b=1}^{k},

where Ca,ϕC_{a,\phi} is as in Definition 1.9.

The second key proposition we require is as follows.

Proposition 3.3.

Assume the same notation as in Definition 3.1. There exists τ0(0,1)\tau_{0}\in(0,1) sufficiently small, so that the following holds. If τ(0,τ0]\tau\in(0,\tau_{0}], we can find constants A,T>0A,T>0, depending on τ,r~,α\tau,\tilde{r},\alpha, such that for tTt\geq T and kk\in\mathbb{N}

(3.3) |I(k,t)|Akkk/2.\left|I(k,t)\right|\leq A^{k}\cdot k^{-k/2}.

We end this section with the following elementary probability lemma from [BC14], which will also be required in our arguments. We mention that analogues of the below lemma have been known for a while in the physics literature, see e.g. [CLDR10, Equation (14)].

Lemma 3.4.

[BC14, Lemma 4.39]. Suppose that fnf_{n} is a sequence of functions fn:[0,1]f_{n}:\mathbb{R}\rightarrow[0,1], such that for each nn, fn(y)f_{n}(y) is strictly decreasing in yy with a limit of 11 at y=y=-\infty and 0 at y=y=\infty. Assume that for each δ>0\delta>0 one has on \[δ,δ]\mathbb{R}\backslash[-\delta,\delta], fn𝟏{y<0}f_{n}\rightarrow{\bf 1}_{\{y<0\}} uniformly. Let XnX_{n} be a sequence of random variables such that for each xx\in\mathbb{R}

𝔼[fn(Xnx)]p(x),\mathbb{E}[f_{n}(X_{n}-x)]\rightarrow p(x),

and assume that p(x)p(x) is a continuous probability distribution function. Then XnX_{n} converges in distribution to a random variable XX, such that (Xx)=p(x)\mathbb{P}(X\leq x)=p(x).

3.2. Proof of Theorem 1.14

In this section we present the proof of Theorem 1.14. For clarity, we split the proof into two steps.

Step 1. Let τ0\tau_{0} be as in Proposition 3.3, and let τ~(0,1)\tilde{\tau}\in(0,1) be sufficiently small so that τ~τ0\tilde{\tau}\leq\tau_{0} and for τ(0,τ~]\tau\in(0,\tilde{\tau}] we have that (1.11) holds. This specifies τ~\tilde{\tau} in the statement of the theorem. In the remainder we fix τ(0,τ~]\tau\in(0,\tilde{\tau}] and a sequence tn>0t_{n}>0, such that tnt_{n}\uparrow\infty as nn\rightarrow\infty.

For yy\in\mathbb{R} and nn\in\mathbb{N} we define

fn(y)=1(τtn1/3y;τ).f_{n}(y)=\frac{1}{(-\tau^{-t_{n}^{1/3}y};\tau)_{\infty}}.

We further introduce the random variables

(3.4) Xn=tn1/3(tn/4+(1/2)(tn2/3α1)Ntn2/3α(tn/γ)),X_{n}=t_{n}^{-1/3}\cdot\left(t_{n}/4+(1/2)(\lfloor t_{n}^{2/3}\alpha\rfloor-1)-N_{\lfloor t_{n}^{2/3}\alpha\rfloor}(t_{n}/\gamma)\right),

where Nx(t)N_{x}(t) is as in (1.2). We claim that for each r~\tilde{r}\in\mathbb{R} we have

(3.5) limn𝔼hfl[fn(Xnr~)]=(𝒜21(21/3α)24/3r~+𝟏{α0}22/3α2),\lim_{n\rightarrow\infty}\mathbb{E}^{\operatorname{h-fl}}\left[f_{n}(X_{n}-\tilde{r})\right]=\mathbb{P}\left(\mathcal{A}_{2\rightarrow 1}(2^{-1/3}\alpha)\leq 2^{4/3}\cdot\tilde{r}+{\bf 1}\{\alpha\leq 0\}\cdot 2^{-2/3}\alpha^{2}\right),

where 𝒜21\mathcal{A}_{2\rightarrow 1} is as in Definition 1.10. We prove (3.5) in Step 2. Here, we assume its validity and conclude the proof of the theorem.

From [FV15, Lemma 5.1] we have that

(3.6) hq(y):=1(τqy;τ)=k=111+τqy+kh_{q}(y):=\frac{1}{(-\tau^{-qy};\tau)_{\infty}}=\prod_{k=1}^{\infty}\frac{1}{1+\tau^{-qy+k}}

is strictly decreasing for all q>0q>0. Moreover, for each δ>0\delta>0 one has hq(y)1{y<0}h_{q}(y)\rightarrow{1}_{\{y<0\}} uniformly on \[δ,δ]\mathbb{R}\backslash[-\delta,\delta] as qq\rightarrow\infty. The latter implies that fnf_{n} satisfy the conditions of Lemma 3.4. We further note that the right side of (3.5) is continuous in r~\tilde{r} from the second part of Lemma 1.13. In particular, we see that the conditions of Lemma 3.4 are all satisfied, and so

(3.7) limnhfl(tn1/3(tn/4+(1/2)(tn2/3α1)Ntn2/3α(tn/γ))r~)=(𝒜21(21/3α)24/3r~+𝟏{α0}22/3α2).\begin{split}&\lim_{n\rightarrow\infty}\mathbb{P}^{\operatorname{h-fl}}\left(t_{n}^{-1/3}\cdot\left(t_{n}/4+(1/2)(\lfloor t_{n}^{2/3}\alpha\rfloor-1)-N_{\lfloor t_{n}^{2/3}\alpha\rfloor}(t_{n}/\gamma)\right)\leq\tilde{r}\right)\\ &=\mathbb{P}\left(\mathcal{A}_{2\rightarrow 1}(2^{-1/3}\alpha)\leq 2^{4/3}\cdot\tilde{r}+{\bf 1}\{\alpha\leq 0\}\cdot 2^{-2/3}\alpha^{2}\right).\end{split}

Combining (1.3) and (3.7), we readily deduce (1.21) with y=2r~+(1/2)𝟏{α0}α2y=2\tilde{r}+(1/2){\bf 1}\{\alpha\leq 0\}\cdot\alpha^{2}.

Step 2. In this step we prove (3.5). From Theorem 1.5 (here we use that τ(0,τ~]\tau\in(0,\tilde{\tau}]) and Definition 3.1 we know that

(3.8) 𝔼hfl[fn(Xnr~)]=1+k=1I(k,tn).\mathbb{E}^{\operatorname{h-fl}}\left[f_{n}(X_{n}-\tilde{r})\right]=1+\sum_{k=1}^{\infty}I(k,t_{n}).

We observe that each summand in (3.8) converges by Proposition 3.2, and also from Proposition 3.3 we may exchange the order of the series and the limit by the dominated convergence theorem with dominating series Akkk/2A^{k}\cdot k^{-k/2}. Consequently, we conclude from Propositions 3.2 and 3.3 that

limn𝔼hfl[fn(Xnr~)]=1+k=11k!(2π𝗂)2kC0,π/4k𝑑uC1,3π/4k𝑑vdet[2vaeua3/48ua2α/8uar~(uava)(vb2ua2)eva3/48va2α/8var~]a,b=1k.\begin{split}\lim_{n\rightarrow\infty}\mathbb{E}^{\operatorname{h-fl}}\left[f_{n}(X_{n}-\tilde{r})\right]=1+\sum_{k=1}^{\infty}&\frac{1}{k!(2\pi\mathsf{i})^{2k}}\int_{C_{0,\pi/4}^{k}}d\vec{u}\int_{C_{-1,3\pi/4}^{k}}d\vec{v}\\ &\det\left[\frac{2v_{a}e^{u_{a}^{3}/48-u_{a}^{2}\alpha/8-u_{a}\tilde{r}}}{(u_{a}-v_{a})(v_{b}^{2}-u_{a}^{2})e^{v_{a}^{3}/48-v_{a}^{2}\alpha/8-v_{a}\tilde{r}}}\right]_{a,b=1}^{k}.\end{split}

We may now apply the change of variables wa=24/3vaw_{a}=-2^{-4/3}v_{a}, za=24/3uaz_{a}=-2^{-4/3}u_{a} for a=1,,ka=1,\dots,k to get

(3.9) limn𝔼hfl[fn(Xnr~)]=1+k=11k!(2π𝗂)2kC0,3π/4k𝑑zC24/3,π/4k𝑑wdet[2waeza3/3za221/3α+za24/3r~(waza)(za2wb2)ewa3/3wa221/3α+wa24/3r~]a,b=1k.\begin{split}\lim_{n\rightarrow\infty}\mathbb{E}^{\operatorname{h-fl}}\left[f_{n}(X_{n}-\tilde{r})\right]=1+\sum_{k=1}^{\infty}&\frac{1}{k!(2\pi\mathsf{i})^{2k}}\int_{C_{0,3\pi/4}^{k}}d\vec{z}\int_{C_{2^{-4/3},\pi/4}^{k}}d\vec{w}\\ &\det\left[\frac{2w_{a}e^{-z_{a}^{3}/3-z_{a}^{2}2^{-1/3}\alpha+z_{a}2^{4/3}\tilde{r}}}{(w_{a}-z_{a})(z_{a}^{2}-w_{b}^{2})e^{-w_{a}^{3}/3-w_{a}^{2}2^{-1/3}\alpha+w_{a}2^{4/3}\tilde{r}}}\right]_{a,b=1}^{k}.\end{split}

In the last integral we may deform the C24/3,π/4C_{2^{-4/3},\pi/4} contours to C1,π/4C_{1,\pi/4} without crossing any poles of the integrands, and thus without affecting the value of the integrals. The decay necessary to deform the contours near infinity comes from the cubic terms in the exponential functions. At this point, we see that the right side of (3.9) agrees with the right side of the first line of (1.20) with t1=21/3αt_{1}=2^{-1/3}\alpha and y1=24/3r~+𝟏{α0}22/3α2y_{1}=2^{4/3}\cdot\tilde{r}+{\bf 1}\{\alpha\leq 0\}\cdot 2^{-2/3}\alpha^{2}. From (3.9) and Lemma 1.13 we conclude (3.5).

3.3. Change of variables

In this section we rewrite the function Hk(ζ)H_{k}(\zeta) from (1.13) in a way that is suitable for our asymptotic analysis in the next two sections, see Lemma 3.8. In order to state our new formula we require a bit of notation, and our exposition here follows [Dim20, Section 3].

Definition 3.5.

Fix τ(0,1)\tau\in(0,1). Let w,z,uw,z,u\in\mathbb{C} be such that zw0zw\neq 0, |z|τn|w||z|\neq\tau^{n}|w| for any nn\in\mathbb{Z} and u[0,)u\not\in[0,\infty). For such a set of parameters we define the function

(3.10) S(w,z;u,τ)=mπ[u][logτ]1[logzlogw2mπ𝗂]sin(π[logτ]1[logzlogw2mπ𝗂]),S(w,z;u,\tau)=\sum_{m\in\mathbb{Z}}\frac{\pi\cdot[-u]^{[\log\tau]^{-1}[\log z-\log w-2m\pi\mathsf{i}]}}{\sin(-\pi[\log\tau]^{-1}[\log z-\log w-2m\pi\mathsf{i}])},

where everywhere we take the principal branch of the logarthm, i.e. if v=re𝗂θv=re^{\mathsf{i}\theta} with r>0r>0 and θ(π,π]\theta\in(-\pi,\pi] we set logv=logr+𝗂θ\log v=\log r+\mathsf{i}\theta. Observe that

𝖱𝖾[π[logτ]1[logzlogw2mπ𝗂]]=πlog|z|log|w|logτπ,\mathsf{Re}\left[-\pi[\log\tau]^{-1}[\log z-\log w-2m\pi\mathsf{i}]\right]=-\pi\cdot\frac{\log|z|-\log|w|}{\log\tau}\not\in\pi\cdot\mathbb{Z},

which implies that each of the summands in (3.10) is well-defined and finite.

It follows from (2.2) that the series in (3.10) is absolutely convergent as we explain here. Put A=[logzlogw][logτ]1A=[\log z-\log w][\log\tau]^{-1}, B=2π[logτ]1B=-2\pi[\log\tau]^{-1} and u=Re𝗂ϕ-u=Re^{\mathsf{i}\phi} with ϕ(π,π)\phi\in(-\pi,\pi). From our assumption that |z|τn|w||z|\neq\tau^{n}|w| for any nn\in\mathbb{Z} and (2.2) we conclude that for any mm\in\mathbb{Z}

|π[u][logτ]1[logzlogw2mπ𝗂]sin(π[logτ]1[logzlogw2mπ𝗂])|πce|ϕ||A|+|logR||A|e(|ϕ|π)B|m|.\left|\frac{\pi\cdot[-u]^{[\log\tau]^{-1}[\log z-\log w-2m\pi\mathsf{i}]}}{\sin(-\pi[\log\tau]^{-1}[\log z-\log w-2m\pi\mathsf{i}])}\right|\leq\pi c^{\prime}\cdot e^{|\phi||A|+|\log R||A|}e^{(|\phi|-\pi)B|m|}.

The above shows that the sum in (3.10) is absolutely convergent by comparison with the geometric series e(|ϕ|π)B|m|.e^{(|\phi|-\pi)B|m|}.

Remark 3.6.

In equation (3.10) we chose the principal branch of the logarithm for expressing logw\log w and logz\log z. However, we could have chosen different branches for logw\log w and logz\log z, and note that then [logzlogw][logτ]1[\log z-\log w][\log\tau]^{-1} would shift by 2kπ𝗂[logτ]12k\pi\mathsf{i}[\log\tau]^{-1} for some kk\in\mathbb{Z}. Since the sum in the definition of S(w,z;u,τ)S(w,z;u,\tau) is over \mathbb{Z} we see that such a shift does not change the value of S(w,z;u,τ)S(w,z;u,\tau). So even though the logarithm is a multi-valued function for fixed u,τu,\tau the function S(w,z;u,τ)S(w,z;u,\tau) as a function of z,wz,w is single valued and well-defined as long as |z|τn|w||z|\neq\tau^{n}|w| for some nn\in\mathbb{Z}.

We next summarize the main properties we require for the function S(w,z;u,t)S(w,z;u,t) from Definition 3.5 in the following lemma.

Lemma 3.7.

[Dim20, Lemma 3.9] Fix τ(0,1)\tau\in(0,1) and R,r(0,)R,r\in(0,\infty) such that R>r>τRR>r>\tau R. Denote by A(r,R)A(r,R)\subset\mathbb{C} the annulus of inner radius rr and outer radius RR that has been centered at the origin. Then, the function S(w,z;u,τ)S(w,z;u,\tau) from Definition 3.5 is well-defined for (w,z,u)Y={(x1,x2,x3)A(r,R)×A(r,R)×([0,)):|x2|<|x1|}(w,z,u)\in Y=\{(x_{1},x_{2},x_{3})\in A(r,R)\times A(r,R)\times(\mathbb{C}\setminus[0,\infty)):|x_{2}|<|x_{1}|\} and is jointly continuous in those variables (for fixed τ\tau) over YY. If we fix u[0,)u\in\mathbb{C}\setminus[0,\infty) and wA(r,R)w\in A(r,R) then as a function of zz, S(w,z;u,τ)S(w,z;u,\tau) is analytic on {ζA(r,R):|ζ|<|w|}\{\zeta\in A(r,R):|\zeta|<|w|\}; analogously, if we fix u[0,)u\in\mathbb{C}\setminus[0,\infty) and zA(r,R)z\in A(r,R) then as a function of ww, S(w,z;u,τ)S(w,z;u,\tau) is analytic on {ζA(r,R):|ζ|>|z|}\{\zeta\in A(r,R):|\zeta|>|z|\}. Finally, if we fix w,zA(r,R)w,z\in A(r,R) with |w|>|z||w|>|z| then S(w,z;u,τ)S(w,z;u,\tau) is analytic in [0,)\mathbb{C}\setminus[0,\infty) as a function of uu.

With the above notation in place we can state the main result of this section.

Lemma 3.8.

Fix kk\in\mathbb{N}, ζ[0,)\zeta\in\mathbb{C}\setminus[0,\infty), set u=(1τ)ζu=(1-\tau)\zeta and let Hk(ζ)H_{k}(\zeta) be as in (1.13). Then

(3.11) Hk(ζ)=1k!(2π𝗂)2kCwk𝑑wCzk𝑑zT(w,z)D(w,z)B(w,z;u)G(w,z), where \begin{split}H_{k}(\zeta)=\frac{1}{k!(2\pi\mathsf{i})^{2k}}\oint_{C^{k}_{w}}d\vec{w}\oint_{C^{k}_{z}}d\vec{z}\hskip 1.42262ptT(\vec{w},\vec{z})D(\vec{w},\vec{z})B(\vec{w},\vec{z};u)G(\vec{w},\vec{z}),\mbox{ where }\end{split}
(3.12) T(w,z)=a=1kexp((qp)t1+wa(qp)t1+za)(1+za1+wa)x1(wa;τ)(za2;τ)(za;τ)(zawa;τ),D(w,z)=det[1wazb]a,b=1k,B(w,z;u)=a=1kS(wa,za;u,τ)[logτ]za, and G(w,z)=1a<bk(wawb;τ)(zazb;τ)(zawb;τ)(wazb;τ).\begin{split}&T(\vec{w},\vec{z})=\prod_{a=1}^{k}\exp\left(\frac{(q-p)t}{1+w_{a}}-\frac{(q-p)t}{1+z_{a}}\right)\left(\frac{1+z_{a}}{1+w_{a}}\right)^{x-1}\frac{(-w_{a};\tau)_{\infty}(z_{a}^{2};\tau)_{\infty}}{(-z_{a};\tau)_{\infty}(z_{a}w_{a};\tau)_{\infty}},\\ &D(\vec{w},\vec{z})=\det\left[\frac{1}{w_{a}-z_{b}}\right]_{a,b=1}^{k},\hskip 5.69054ptB(\vec{w},\vec{z};u)=\prod_{a=1}^{k}\frac{S(w_{a},z_{a};u,\tau)}{-[\log\tau]z_{a}},\mbox{ and }\\ &G(\vec{w},\vec{z})=\prod_{1\leq a<b\leq k}\frac{(w_{a}w_{b};\tau)_{\infty}(z_{a}z_{b};\tau)_{\infty}}{(z_{a}w_{b};\tau)_{\infty}(w_{a}z_{b};\tau)_{\infty}}.\end{split}

In equation (3.11) we have that Cw,CzC_{w},C_{z} are positively oriented circles, centered at the origin, of radii RwR_{w}, RzR_{z}, respectively, such that Rw>1>Rw1>Rz>τRwR_{w}>1>R_{w}^{-1}>R_{z}>\tau R_{w}.

Proof.

We first note that by Cauchy’s theorem we may deform CwC_{w} to γ1,0\gamma_{-1,0} as in Proposition 1.2, and CzC_{z} to τ1/2γ1,0\tau^{1/2}\cdot\gamma_{-1,0} without affecting the value of the integral. In the latter statement we used that S(w,z;u,τ)S(w,z;u,\tau) is analytic separately in ww and zz, in view of Lemma 3.7, and that in the process of deformation we do not cross any poles of the integrand. We proceed to denote γ1,0\gamma_{-1,0} by CwC_{w} and τ1/2γ1,0\tau^{1/2}\cdot\gamma_{-1,0} by CzC_{z} in the remainder of the proof.

Expanding the determinant D(w,z)D(\vec{w},\vec{z}) on the right side of (3.11) and the Cauchy determinant det[1waτsawb]a,b=1k\det\left[\frac{-1}{w_{a}\tau^{s_{a}}-w_{b}}\right]_{a,b=1}^{k} in the definition of Hk(ζ)H_{k}(\zeta), see (1.8) and (1.13), we see that to prove (3.11) it suffices to show that for each σSk\sigma\in S_{k} (the permutation group of kk elements) and wCwk\vec{w}\in C_{w}^{k} we have

(1/2+𝗂)k𝑑sa=1kπ(ζ)sa𝔣(wa;sa)𝔤(wa;sa)sin(πsa)(wawσ(a)τsσ(a))1a<bk𝔥(wa,wb;sa,sb)=Czk𝑑zT(w,z)B(w,z;u)G(w,z)a=1k1wazσ(a).\begin{split}&\int_{(1/2+\mathsf{i}\mathbb{R})^{k}}d\vec{s}\prod_{a=1}^{k}\frac{\pi(-\zeta)^{s_{a}}\mathfrak{f}(w_{a};s_{a})\mathfrak{g}(w_{a};s_{a})}{\sin(-\pi s_{a})(w_{a}-w_{\sigma(a)}\tau^{s_{\sigma(a)}})}\cdot\prod_{1\leq a<b\leq k}\mathfrak{h}(w_{a},w_{b};s_{a},s_{b})\\ &=\oint_{C^{k}_{z}}d\vec{z}\hskip 1.42262ptT(\vec{w},\vec{z})B(\vec{w},\vec{z};u)G(\vec{w},\vec{z})\cdot\prod_{a=1}^{k}\frac{1}{w_{a}-z_{\sigma(a)}}.\end{split}

Writing

𝔣~(w;n)=exp((qp)t1+w(qp)t1+τnw)(1+τnw1+w)x1,\tilde{\mathfrak{f}}(w;n)=\exp\left(\frac{(q-p)t}{1+w}-\frac{(q-p)t}{1+\tau^{n}w}\right)\cdot\left(\frac{1+\tau^{n}w}{1+w}\right)^{x-1},

and recalling the definition of 𝔣(w;n)\mathfrak{f}(w;n) from (1.7), and that u=(1τ)ζu=(1-\tau)\zeta, we see that to conclude (3.11) it suffices to show that for each σSk\sigma\in S_{k} and wCwk\vec{w}\in C_{w}^{k} we have

(3.13) (1/2+𝗂)k𝑑sa=1kπ(u)sa𝔣~(wa;sa)𝔤(wa;sa)sin(πsa)(wawσ(a)τsσ(a))1a<bk𝔥(wa,wb;sa,sb)=Czk𝑑zT(w,z)B(w,z;u)G(w,z)a=1k1wazσ(a).\begin{split}&\int_{(1/2+\mathsf{i}\mathbb{R})^{k}}d\vec{s}\prod_{a=1}^{k}\frac{\pi(-u)^{s_{a}}\tilde{\mathfrak{f}}(w_{a};s_{a})\mathfrak{g}(w_{a};s_{a})}{\sin(-\pi s_{a})(w_{a}-w_{\sigma(a)}\tau^{s_{\sigma(a)}})}\cdot\prod_{1\leq a<b\leq k}\mathfrak{h}(w_{a},w_{b};s_{a},s_{b})\\ &=\oint_{C^{k}_{z}}d\vec{z}\hskip 1.42262ptT(\vec{w},\vec{z})B(\vec{w},\vec{z};u)G(\vec{w},\vec{z})\cdot\prod_{a=1}^{k}\frac{1}{w_{a}-z_{\sigma(a)}}.\end{split}

We next note that

(3.14) (1/2+𝗂)k𝑑sa=1kπ(u)sa𝔣~(wa;sa)𝔤(wa;sa)sin(πsa)(wawσ(a)τsσ(a))1a<bk𝔥(wa,wb;sa,sb)=n1nk1/2+𝗂π[logτ]11/2𝗂π[logτ]1𝑑s11/2+𝗂π[logτ]11/2𝗂π[logτ]1𝑑ska=1kπ(u)sa2π𝗂na[logτ]1𝔣~(wa;sa2π𝗂na[logτ]1)𝔤(wa;sa2π𝗂na[logτ]1)sin(π[sa2π𝗂na[logτ]1])(wawσ(a)τsσ(a)2π𝗂nσ(a)[logτ]1)×1a<bk𝔥(wa,wb;sa2π𝗂na[logτ]1,sb2π𝗂nb[logτ]1)=n1nk1/2+𝗂π[logτ]11/2𝗂π[logτ]1𝑑s11/2+𝗂π[logτ]11/2𝗂π[logτ]1𝑑ska=1kπ(u)sa2π𝗂na[logτ]1𝔣~(wa;sa)𝔤(wa;sa)sin(π[sa2π𝗂na[logτ]1])(wawσ(a)τsσ(a))1a<bk𝔥(wa,wb;sa,sb).\begin{split}&\int_{(1/2+\mathsf{i}\mathbb{R})^{k}}d\vec{s}\prod_{a=1}^{k}\frac{\pi(-u)^{s_{a}}\tilde{\mathfrak{f}}(w_{a};s_{a})\mathfrak{g}(w_{a};s_{a})}{\sin(-\pi s_{a})(w_{a}-w_{\sigma(a)}\tau^{s_{\sigma(a)}})}\cdot\prod_{1\leq a<b\leq k}\mathfrak{h}(w_{a},w_{b};s_{a},s_{b})\\ &=\sum_{n_{1}\in\mathbb{Z}}\cdots\sum_{n_{k}\in\mathbb{Z}}\int_{1/2+\mathsf{i}\pi[\log\tau]^{-1}}^{1/2-\mathsf{i}\pi[\log\tau]^{-1}}ds_{1}\cdots\int_{1/2+\mathsf{i}\pi[\log\tau]^{-1}}^{1/2-\mathsf{i}\pi[\log\tau]^{-1}}ds_{k}\\ &\prod_{a=1}^{k}\frac{\pi(-u)^{s_{a}-2\pi\mathsf{i}n_{a}[\log\tau]^{-1}}\tilde{\mathfrak{f}}(w_{a};s_{a}-2\pi\mathsf{i}n_{a}[\log\tau]^{-1})\mathfrak{g}(w_{a};s_{a}-2\pi\mathsf{i}n_{a}[\log\tau]^{-1})}{\sin(-\pi[s_{a}-2\pi\mathsf{i}n_{a}[\log\tau]^{-1}])(w_{a}-w_{\sigma(a)}\tau^{s_{\sigma(a)}-2\pi\mathsf{i}n_{\sigma(a)}[\log\tau]^{-1}})}\\ &\times\prod_{1\leq a<b\leq k}\mathfrak{h}(w_{a},w_{b};s_{a}-2\pi\mathsf{i}n_{a}[\log\tau]^{-1},s_{b}-2\pi\mathsf{i}n_{b}[\log\tau]^{-1})\\ &=\sum_{n_{1}\in\mathbb{Z}}\cdots\sum_{n_{k}\in\mathbb{Z}}\int_{1/2+\mathsf{i}\pi[\log\tau]^{-1}}^{1/2-\mathsf{i}\pi[\log\tau]^{-1}}ds_{1}\cdots\int_{1/2+\mathsf{i}\pi[\log\tau]^{-1}}^{1/2-\mathsf{i}\pi[\log\tau]^{-1}}ds_{k}\\ &\prod_{a=1}^{k}\frac{\pi(-u)^{s_{a}-2\pi\mathsf{i}n_{a}[\log\tau]^{-1}}\tilde{\mathfrak{f}}(w_{a};s_{a})\mathfrak{g}(w_{a};s_{a})}{\sin(-\pi[s_{a}-2\pi\mathsf{i}n_{a}[\log\tau]^{-1}])(w_{a}-w_{\sigma(a)}\tau^{s_{\sigma(a)}})}\prod_{1\leq a<b\leq k}\mathfrak{h}(w_{a},w_{b};s_{a},s_{b}).\end{split}

We mention that in deriving the last equality we used that τ2π𝗂n[logτ]1=1\tau^{2\pi\mathsf{i}n[\log\tau]^{-1}}=1 for all nn\in\mathbb{Z} and the fact that 𝔣~(w;s)\tilde{\mathfrak{f}}(w;s), 𝔤(w;s)\mathfrak{g}(w;s) depend on ss only through τs\tau^{s}, while 𝔥(w1,w2;s1,s2)\mathfrak{h}(w_{1},w_{2};s_{1},s_{2}) depends on s1,s2s_{1},s_{2} only through τs1\tau^{s_{1}} and τs2\tau^{s_{2}}, see (1.7).

Applying the change of variables za=waτsaz_{a}=w_{a}\tau^{s_{a}} for a=1,,ka=1,\dots,k in (3.14) we conclude that

(3.15) (1/2+𝗂)k𝑑sa=1kπ(u)sa𝔣~(wa;sa)𝔤(wa;sa)sin(πsa)(wawσ(a)τsσ(a))1a<bk𝔥(wa,wb;sa,sb)=n1nkCzkdzT(w,z)G(w,z)a=1k1(wazσ(a))za[logτ]×a=1kπ(u)[logzalogwa][logτ]12π𝗂na[logτ]1sin(π[[logzalogwa][logτ]12π𝗂na[logτ]1]),\begin{split}&\int_{(1/2+\mathsf{i}\mathbb{R})^{k}}d\vec{s}\prod_{a=1}^{k}\frac{\pi(-u)^{s_{a}}\tilde{\mathfrak{f}}(w_{a};s_{a})\mathfrak{g}(w_{a};s_{a})}{\sin(-\pi s_{a})(w_{a}-w_{\sigma(a)}\tau^{s_{\sigma(a)}})}\prod_{1\leq a<b\leq k}\mathfrak{h}(w_{a},w_{b};s_{a},s_{b})\\ &=\sum_{n_{1}\in\mathbb{Z}}\cdots\sum_{n_{k}\in\mathbb{Z}}\int_{C_{z}^{k}}d\vec{z}\hskip 1.42262ptT(\vec{w},\vec{z})G(\vec{w},\vec{z})\prod_{a=1}^{k}\frac{1}{(w_{a}-z_{\sigma(a)})z_{a}[-\log\tau]}\times\\ &\prod_{a=1}^{k}\frac{\pi(-u)^{[\log z_{a}-\log w_{a}][\log\tau]^{-1}-2\pi\mathsf{i}n_{a}[\log\tau]^{-1}}}{\sin(-\pi[[\log z_{a}-\log w_{a}][\log\tau]^{-1}-2\pi\mathsf{i}n_{a}[\log\tau]^{-1}])},\end{split}

where we mention that the extra 1-1 sign in (3.15) comes from the fact that CzC_{z} is positively oriented while waτsaw_{a}\tau^{s_{a}} covers CzC_{z} with negative orientation as sas_{a} varies from 1/2+𝗂π[logτ]11/2+\mathsf{i}\pi[\log\tau]^{-1} upward to 1/2𝗂π[logτ]11/2-\mathsf{i}\pi[\log\tau]^{-1}. From Fubini’s theorem and (2.2) we can put the sums in (3.15) inside the integral at which point we obtain (3.13), as desired. ∎

4. Asymptotic analysis: Part I

The goal of this section is to prove Proposition 3.2. In Section 4.1 we use Lemma 3.8 to express I(k,t)I(k,t) from Proposition 3.2 as a 2k2k-fold contour integral, see (4.5), which is suitable for asymptotic analysis. The formula for I(k,t)I(k,t) in (4.5) involves several functions and in Section 4.2 we establish various estimates for these functions. The proof of Proposition 3.2 is presented in Section 4.3. It is based on the method of steepest descent and relies on the formula for I(k,t)I(k,t) from (4.5) and the results from Section 4.2. The lemmas in Section 4.1 and 4.2 are proved in Section 4.4.

4.1. Formula for I(k,t)I(k,t)

The goal of this section is to get a formula for I(k,t)I(k,t) from Proposition 3.2 that is suitable for asymptotic analysis. We begin with a lemma, which is proved in Section 4.4.

Lemma 4.1.

Let kk\in\mathbb{N}, z1,,zk,w1,,wkz_{1},\dots,z_{k},w_{1},\dots,w_{k}\in\mathbb{C} be such that ziwjz_{i}\neq w_{j} and ziwj1z_{i}\neq w_{j}^{-1} for 1i,jk1\leq i,j\leq k. Then we have

(4.1) 1i<jk(1wiwj)(1zizj)(wiwj)(zjzi)i,j=1k(1wizj)(wizj)=det[1(wizj)(1wizj)]i,j=1k.\frac{\prod_{1\leq i<j\leq k}(1-w_{i}w_{j})(1-z_{i}z_{j})(w_{i}-w_{j})(z_{j}-z_{i})}{\prod_{i,j=1}^{k}(1-w_{i}z_{j})(w_{i}-z_{j})}=\det\left[\frac{1}{(w_{i}-z_{j})(1-w_{i}z_{j})}\right]_{i,j=1}^{k}.

The next definition introduces various contours, which will be required in our arguments.

Definition 4.2.

For each ϵ[0,1]\epsilon\in[0,1] and ρ0\rho\geq 0, we let γρ,ϵ+=γρ,ϵ+,0γρ,ϵ+,1\gamma^{+}_{\rho,\epsilon}=\gamma^{+,0}_{\rho,\epsilon}\cup\gamma^{+,1}_{\rho,\epsilon} and γρ,ϵ=γρ,ϵ,0γρ,ϵ,1\gamma^{-}_{\rho,\epsilon}=\gamma^{-,0}_{\rho,\epsilon}\cup\gamma^{-,1}_{\rho,\epsilon} be two contours, where

γρ,ϵ±,0:={z:z=ρ±(a+𝗂a) or z=ρ±(a𝗂a) for a[0,ϵ]},γρ,ϵ±,1:={z:ϵ|𝖨𝗆(z)|π,𝖱𝖾(z)=ρ±ϵ}.\begin{split}&\gamma^{\pm,0}_{\rho,\epsilon}:=\{z\in\mathbb{C}:z=\rho\pm(a+\mathsf{i}a)\mbox{ or }z=\rho\pm(a-\mathsf{i}a)\mbox{ for $a\in[0,\epsilon]$}\},\\ &\gamma^{\pm,1}_{\rho,\epsilon}:=\{z\in\mathbb{C}:\epsilon\leq|\mathsf{Im}(z)|\leq\pi,\mathsf{Re}(z)=\rho\pm\epsilon\}.\end{split}

The contours γρ,ϵ+,γρ,ϵ\gamma^{+}_{\rho,\epsilon},\gamma^{-}_{\rho,\epsilon} are oriented to have increasing imaginary part.

For t>0t>0 we define the contours

Γu,ϵ,t=t1/3γ0,ϵ+,Γv,ϵ,t=t1/3γt1/3,ϵ,Γu,ϵ,ti=t1/3γ0,ϵ+,i, and Γv,ϵ,ti=t1/3γt1/3,ϵ,i for i=0,1.\Gamma_{u,\epsilon,t}=t^{1/3}\cdot\gamma^{+}_{0,\epsilon},\hskip 5.69054pt\Gamma_{v,\epsilon,t}=t^{1/3}\cdot\gamma^{-}_{-t^{-1/3},\epsilon},\hskip 5.69054pt\Gamma^{i}_{u,\epsilon,t}=t^{1/3}\cdot\gamma^{+,i}_{0,\epsilon},\mbox{ and }\Gamma^{i}_{v,\epsilon,t}=t^{1/3}\cdot\gamma^{-,i}_{-t^{-1/3},\epsilon}\mbox{ for }i=0,1.

All of the above contours are oriented in the direction of increasing imaginary part.

We also define Cw,ϵ,tC_{w,\epsilon,t} and Cz,ϵ,tC_{z,\epsilon,t} to be the positively oriented contours obtained from γ0,ϵ+\gamma^{+}_{0,\epsilon} and γt1/3,ϵ\gamma^{-}_{-t^{-1/3},\epsilon}, respectively, under the map xexx\rightarrow e^{x}. Observe that Cw,ϵ,tC_{w,\epsilon,t} and Cz,ϵ,tC_{z,\epsilon,t} are piecewise smooth positively oriented contours, Cz,ϵ,tC_{z,\epsilon,t} is contained in the closed unit disc in \mathbb{C}, which in turn is contained in the region enclosed by Cw,ϵ,tC_{w,\epsilon,t}. Some of the contours in the definition are depicted in Figure 2.

Refer to caption
Figure 2. The left figure depicts the contours Cw,ϵ,tC_{w,\epsilon,t}, Cz,ϵ,tC_{z,\epsilon,t} as well as the circles of radii τ1/4\tau^{-1/4} and τ1/2\tau^{1/2}. The right figure depicts the contours γ0,ϵ+\gamma^{+}_{0,\epsilon} and γt1/3,ϵ\gamma^{-}_{-t^{-1/3},\epsilon}.

The following lemma introduces a certain analytic function F(z)F(z), which will appear in our analysis, and establishes a few of its properties. Its proof is given in Section 4.4.

Lemma 4.3.

For zz\in\mathbb{C} such that ez1e^{z}\neq-1 define the function

(4.2) F(z)=11+ez+z412.F(z)=\frac{1}{1+e^{z}}+\frac{z}{4}-\frac{1}{2}.

There exist universal constants 𝗋(0,1/2)\mathsf{r}\in(0,1/2), C1,C2>0C_{1},C_{2}>0, such that

(4.3) |F(z)z3/48|C1|z|4 for |z|3𝗋;𝖱𝖾[F(z)]|z|3200+C2|ρ|3 when zγρ,ϵ+,|𝖨𝗆(z)|ϵ with ρ0|ρ|ϵ𝗋;𝖱𝖾[F(z)]|z|3200C2|ρ|3 when zγρ,ϵ,|𝖨𝗆(z)|ϵ with ρ0|ρ|ϵ𝗋,\begin{split}&|F(z)-z^{3}/48|\leq C_{1}|z|^{4}\mbox{ for $|z|\leq 3\mathsf{r}$};\\ &\mathsf{Re}[F(z)]\leq-\frac{|z|^{3}}{200}+C_{2}|\rho|^{3}\mbox{ when }z\in\gamma_{\rho,\epsilon}^{+},|\mathsf{Im}(z)|\leq\epsilon\mbox{ with $\rho\in\mathbb{R}$, }0\leq|\rho|\leq\epsilon\leq\mathsf{r};\\ &\mathsf{Re}[F(z)]\geq\frac{|z|^{3}}{200}-C_{2}|\rho|^{3}\mbox{ when }z\in\gamma_{\rho,\epsilon}^{-},|\mathsf{Im}(z)|\leq\epsilon\mbox{ with $\rho\in\mathbb{R}$, }0\leq|\rho|\leq\epsilon\leq\mathsf{r},\end{split}

where γρ,ϵ±\gamma^{\pm}_{\rho,\epsilon} are as in Definition 4.2. Furthermore, for any x>0x>0 we have

(4.4) ddy𝖱𝖾[F(x±𝗂y)]0 and ddy𝖱𝖾[F(x±𝗂y)]0 for y[0,π].\begin{split}\frac{d}{dy}\mathsf{Re}[F(x\pm\mathsf{i}y)]\leq 0\mbox{ and }\frac{d}{dy}\mathsf{Re}[F(-x\pm\mathsf{i}y)]\geq 0\mbox{ for }y\in[0,\pi].\end{split}
Definition 4.4.

Assume the same notation as in Definition 3.1. We let ϵ0>0\epsilon_{0}>0 be sufficiently small so that τ1e20ϵ0\tau^{-1}\geq e^{20\epsilon_{0}} and ϵ0𝗋\epsilon_{0}\leq\mathsf{r}, where 𝗋\mathsf{r} is as in Lemma 4.3. We also let T0T_{0} be sufficiently large so that for tT0t\geq T_{0} we have t1/3ϵ0t^{-1/3}\leq\epsilon_{0}.

With the above notation, we are ready to state our formula for I(k,t)I(k,t). Let ϵ0,T0\epsilon_{0},T_{0} be as in Definition 4.4, and for tT0t\geq T_{0} let Γu,ϵ0,t\Gamma_{u,\epsilon_{0},t}, Γv,ϵ0,t\Gamma_{v,\epsilon_{0},t} be as in Definition 4.2. We then have the following formula for all tT0t\geq T_{0} and kk\in\mathbb{N}

(4.5) I(k,t)=1k!(2π𝗂)2kΓu,ϵ0,tk𝑑uΓv,ϵ0,tk𝑑vi=15Ai(u,v;t)i=12Bi(u,v;t),\begin{split}I(k,t)=\frac{1}{k!(2\pi\mathsf{i})^{2k}}\int_{\Gamma^{k}_{u,\epsilon_{0},t}}d\vec{u}\int_{\Gamma^{k}_{v,\epsilon_{0},t}}d\vec{v}&\prod_{i=1}^{5}A_{i}(\vec{u},\vec{v};t)\prod_{i=1}^{2}B_{i}(\vec{u},\vec{v};t),\end{split}

where

(4.6) A1(u,v;t)=a=1ket[F(t1/3ua)F(t1/3va)],A2(u,v;t)=a=1k(mt1/3[logτ]1πe2mπ𝗂[(1/4)t+(1/2)(t2/3α1)t1/3r~]sin(π[logτ]1[t1/3(vaua)2mπ𝗂])),A3(u,v;t)=a=1k((1+et1/3va)et1/3va/2(1+et1/3ua)et1/3ua/2)t2/3α1,A4(u,v;t)=a=1kexp(r~[vaua])A5(u,v;t)=tk/3a=1k(et1/3ua;τ)(e2t1/3va;τ)et1/3ua(et1/3va;τ)(τet1/3(ua+va);τ),\begin{split}&A_{1}(\vec{u},\vec{v};t)=\prod_{a=1}^{k}e^{t[F(t^{-1/3}u_{a})-F(t^{-1/3}v_{a})]},\\ &A_{2}(\vec{u},\vec{v};t)=\prod_{a=1}^{k}\left(\sum_{m\in\mathbb{Z}}\frac{t^{-1/3}[-\log\tau]^{-1}\pi\cdot e^{2m\pi\mathsf{i}[(1/4)t+(1/2)(\lfloor t^{2/3}\alpha\rfloor-1)-t^{1/3}\tilde{r}]}}{\sin(-\pi[\log\tau]^{-1}[t^{-1/3}(v_{a}-u_{a})-2m\pi\mathsf{i}])}\right),\\ &A_{3}(\vec{u},\vec{v};t)=\prod_{a=1}^{k}\left(\frac{(1+e^{t^{-1/3}v_{a}})e^{-t^{-1/3}v_{a}/2}}{(1+e^{t^{-1/3}u_{a}})e^{-t^{-1/3}u_{a}/2}}\right)^{\lfloor t^{2/3}\alpha\rfloor-1},\\ &A_{4}(\vec{u},\vec{v};t)=\prod_{a=1}^{k}\exp\left(\tilde{r}[v_{a}-u_{a}]\right)\\ &A_{5}(\vec{u},\vec{v};t)=t^{k/3}\cdot\prod_{a=1}^{k}\frac{(-e^{t^{-1/3}u_{a}};\tau)_{\infty}(e^{2t^{-1/3}v_{a}};\tau)_{\infty}e^{t^{-1/3}u_{a}}}{(-e^{t^{-1/3}v_{a}};\tau)_{\infty}(\tau e^{t^{-1/3}(u_{a}+v_{a})};\tau)_{\infty}},\end{split}
(4.7) B1(u,v;t)=det[t2/3(et1/3uaet1/3vb)(1et1/3(ua+vb))]a,b=1k,B2(u,v;t)=1a<bk(τet1/3(ua+ub);τ)(τet1/3(va+vb);τ)(τet1/3(ua+vb);τ)(τet1/3(va+ub);τ).\begin{split}&B_{1}(\vec{u},\vec{v};t)=\det\left[\frac{t^{-2/3}}{(e^{t^{-1/3}u_{a}}-e^{t^{-1/3}v_{b}})(1-e^{t^{-1/3}(u_{a}+v_{b})})}\right]_{a,b=1}^{k},\\ &B_{2}(\vec{u},\vec{v};t)=\prod_{1\leq a<b\leq k}\frac{(\tau e^{t^{-1/3}(u_{a}+u_{b})};\tau)_{\infty}(\tau e^{t^{-1/3}(v_{a}+v_{b})};\tau)_{\infty}}{(\tau e^{t^{-1/3}(u_{a}+v_{b})};\tau)_{\infty}(\tau e^{t^{-1/3}(v_{a}+u_{b})};\tau)_{\infty}}.\end{split}

Let us briefly explain the origin of (4.5). Our starting point is the formula for Hk(ζ)H_{k}(\zeta) from (3.11) in Lemma 3.8, where we take Rw=τ1/4R_{w}=\tau^{-1/4} and Rz=τ1/2R_{z}=\tau^{1/2}. At this point we can deform the contours CwC_{w} in (3.11) to Cw,ϵ,tC_{w,\epsilon,t} as in Definition 4.2 without crossing any poles and thus without affecting the value of the integral by Cauchy’s theorem. Here we used that S(w,z;u,τ)S(w,z;u,\tau) is analytic in ww for fixed zz, as follows from Lemma 3.7. We mention that the integrand in (3.11) has singularities at wa=τmzaw_{a}=\tau^{m}z_{a} for all mm\in\mathbb{Z}, coming from B(w,z;u)B(\vec{w},\vec{z};u), at wa=zbw_{a}=z_{b}, coming from D(w,z)D(\vec{w},\vec{z}), at wa=zb1τmw_{a}=z_{b}^{-1}\tau^{-m} for all m0m\in\mathbb{Z}_{\geq 0}, coming from T(w,z)T(\vec{w},\vec{z}) and G(w,z)G(\vec{w},\vec{z}), and all of these singularities lie outside of the annulus of inner radius 11 and outer radius τ1/4\tau^{-1/4}, where the deformation takes place. There is also a singularity at wa=1w_{a}=-1, coming from T(w,z)T(\vec{w},\vec{z}), which is also not crossed in the deformation, see the left part of Figure 2. After we deform all the CwC_{w} contours to Cw,ϵ,tC_{w,\epsilon,t} we proceed to deform the CzC_{z} contours to Cz,ϵ,tC_{z,\epsilon,t} as in Definition 4.2. Again, in the process of deformation we do not cross any poles, and by Cauchy’s theorem we do not change the value of the integral. We mention that the only poles in the annulus of inner radius τ1/2\tau^{1/2} and outer radius 11, where the deformation takes place, come from za=1z_{a}=-1 (from T(w,z)T(\vec{w},\vec{z})), za=wbz_{a}=w_{b} (from D(w,z)D(\vec{w},\vec{z})), and at za=wb1z_{a}=w_{b}^{-1} (from T(w,z)T(\vec{w},\vec{z}) and G(w,z)G(\vec{w},\vec{z})). As Cz,ϵ,tC_{z,\epsilon,t} is strictly contained in the unit circle, and Cw,ϵ,tC_{w,\epsilon,t} is outside of it we do not cross the za=1z_{a}=-1 and the za=wbz_{a}=w_{b} poles. On the other hand, the fact that γ0,ϵ+\gamma^{+}_{0,\epsilon} is to the left of γt1/3,ϵ-\gamma^{-}_{-t^{-1/3},\epsilon}, as in Definition 4.2, ensures that we also do not cross za=wb1z_{a}=w_{b}^{-1}.

Once we have deformed the contours in (3.11) to Cw,ϵ,tC_{w,\epsilon,t} and Cz,ϵ,tC_{z,\epsilon,t}, we may apply the Cauchy determinant formula (2.3) and Lemma 4.1 to get

I(k,t)=1k!(2π𝗂)2kCw,ϵ,tk𝑑wCz,ϵ,tk𝑑zdet[1(wazb)(1zawb)]a,b=1k1a<bk(τwawb;τ)(τzazb;τ)(τzawb;τ)(τwazb;τ)×a=1kS(wa,za;u,τ)[logτ]zaexp(t1+wat1+za)(1+za1+wa)αt2/31(wa;τ)(za2;τ)(za;τ)(τzawa;τ),\begin{split}&I(k,t)=\frac{1}{k!(2\pi\mathsf{i})^{2k}}\oint_{C^{k}_{w,\epsilon,t}}d\vec{w}\oint_{C^{k}_{z,\epsilon,t}}d\vec{z}\det\left[\frac{1}{(w_{a}-z_{b})(1-z_{a}w_{b})}\right]_{a,b=1}^{k}\cdot\prod_{1\leq a<b\leq k}\frac{(\tau w_{a}w_{b};\tau)_{\infty}(\tau z_{a}z_{b};\tau)_{\infty}}{(\tau z_{a}w_{b};\tau)_{\infty}(\tau w_{a}z_{b};\tau)_{\infty}}\\ &\times\prod_{a=1}^{k}\frac{S(w_{a},z_{a};u,\tau)}{-[\log\tau]z_{a}}\cdot\exp\left(\frac{t}{1+w_{a}}-\frac{t}{1+z_{a}}\right)\cdot\left(\frac{1+z_{a}}{1+w_{a}}\right)^{\lfloor\alpha t^{2/3}\rfloor-1}\cdot\frac{(-w_{a};\tau)_{\infty}(z_{a}^{2};\tau)_{\infty}}{(-z_{a};\tau)_{\infty}(\tau z_{a}w_{a};\tau)_{\infty}},\end{split}

where u=τ(1/4)t(1/2)(t2/3α1)+t1/3r~u=-\tau^{-(1/4)t-(1/2)(\lfloor t^{2/3}\alpha\rfloor-1)+t^{1/3}\tilde{r}}, and α,r~\alpha,\tilde{r} are as in Definition 3.1. Applying the change of variables wa=et1/3uaw_{a}=e^{t^{-1/3}u_{a}}, za=et1/3vaz_{a}=e^{t^{-1/3}v_{a}} for a=1,,ka=1,\dots,k and using that

a=1kS(wa,za;u,τ)[logτ]=a=1k(m[logτ]1πe2mπ𝗂[(1/4)t+(1/2)(t2/3α1)t1/3r~]sin(π[logτ]1[t1/3(vaua)2mπ𝗂]))×a=1kexp(t4(t1/3va+t1/3ua)+(t2/3α1)(t1/3va/2+t1/3ua/2)r~[vaua]),\begin{split}&\prod_{a=1}^{k}\frac{S(w_{a},z_{a};u,\tau)}{-[\log\tau]}=\prod_{a=1}^{k}\left(\sum_{m\in\mathbb{Z}}\frac{[-\log\tau]^{-1}\pi\cdot e^{2m\pi\mathsf{i}[(1/4)t+(1/2)(\lfloor t^{2/3}\alpha\rfloor-1)-t^{1/3}\tilde{r}]}}{\sin(-\pi[\log\tau]^{-1}[t^{-1/3}(v_{a}-u_{a})-2m\pi\mathsf{i}])}\right)\\ &\times\prod_{a=1}^{k}\exp\left(\frac{t}{4}(-t^{-1/3}v_{a}+t^{-1/3}u_{a})+(\lfloor t^{2/3}\alpha\rfloor-1)\left(-t^{-1/3}v_{a}/2+t^{-1/3}u_{a}/2\right)\tilde{r}[v_{a}-u_{a}]\right),\end{split}

which holds from (3.10), we arrive at (4.5).

4.2. Preliminary estimates

In this section we state various estimates for the functions Ai(u,v;t)A_{i}(\vec{u},\vec{v};t) for i=1,,5i=1,\dots,5 and Bj(u,v;t)B_{j}(\vec{u},\vec{v};t) for j=1,2j=1,2 from (4.6) and (4.7) in a sequence of lemmas, whose proofs are given in Section 4.4. These estimates will be required in the proof of Proposition 3.2 in Section 4.3. In the lemmas below Γu,ϵ,t,Γv,ϵ,t\Gamma_{u,\epsilon,t},\Gamma_{v,\epsilon,t} are as in Definition 4.2.

Lemma 4.5.

Let F(z)F(z) and 𝗋>0\mathsf{r}>0 be as in Lemma 4.3, and ϵ(0,𝗋]\epsilon\in(0,\mathsf{r}]. We can find Tϵ,1,Aϵ,1,aϵ,1>0T_{\epsilon,1},A_{\epsilon,1},a_{\epsilon,1}>0, depending on ϵ\epsilon alone, such that the following holds. If tTϵ,1t\geq T_{\epsilon,1}, uΓu,ϵ,t,vΓv,ϵ,tu\in\Gamma_{u,\epsilon,t},v\in\Gamma_{v,\epsilon,t}

(4.8) |et[F(t1/3u)F(t1/3v)]|Aϵ,1exp(aϵ,1|u|3aϵ,1|v|3).\left|e^{t[F(t^{-1/3}u)-F(t^{-1/3}v)]}\right|\leq A_{\epsilon,1}\cdot\exp\left(-a_{\epsilon,1}|u|^{3}-a_{\epsilon,1}|v|^{3}\right).
Lemma 4.6.

Let 𝗋>0\mathsf{r}>0 be as in Lemma 4.3, τ1(0,1)\tau_{1}\in(0,1) and ϵ(0,𝗋]\epsilon\in(0,\mathsf{r}] be sufficiently small so that e20ϵτ11e^{20\epsilon}\leq\tau_{1}^{-1}. We can find Tϵ,2,Aϵ,2>0T_{\epsilon,2},A_{\epsilon,2}>0, depending on ϵ\epsilon alone, such that the following holds. If tTϵ,2t\geq T_{\epsilon,2}, τ(0,τ1]\tau\in(0,\tau_{1}], α,r~\alpha,\tilde{r}\in\mathbb{R}, uΓu,ϵ,t,vΓv,ϵ,tu\in\Gamma_{u,\epsilon,t},v\in\Gamma_{v,\epsilon,t}

(4.9) |mt1/3[logτ]1πe2mπ𝗂[(1/4)t+(1/2)(t2/3α1)t1/3r~]sin(π[logτ]1[t1/3(vu)2mπ𝗂])|Aϵ,2.\left|\sum_{m\in\mathbb{Z}}\frac{t^{-1/3}[-\log\tau]^{-1}\pi\cdot e^{2m\pi\mathsf{i}[(1/4)t+(1/2)(\lfloor t^{2/3}\alpha\rfloor-1)-t^{1/3}\tilde{r}]}}{\sin(-\pi[\log\tau]^{-1}[t^{-1/3}(v-u)-2m\pi\mathsf{i}])}\right|\leq A_{\epsilon,2}.
Lemma 4.7.

Let 𝗋>0\mathsf{r}>0 be as in Lemma 4.3, α\alpha\in\mathbb{R} and ϵ(0,𝗋]\epsilon\in(0,\mathsf{r}]. We can find Tϵ,3,aϵ,3>0T_{\epsilon,3},a_{\epsilon,3}>0, depending on ϵ\epsilon alone, such that the following holds. If tTϵ,3t\geq T_{\epsilon,3}, uΓu,ϵ,t,vΓv,ϵ,tu\in\Gamma_{u,\epsilon,t},v\in\Gamma_{v,\epsilon,t}

(4.10) |((1+et1/3v)et1/3v/2(1+et1/3u)et1/3u/2)t2/3α1|exp(aϵ,3(1+|α|)(|u|2+|v|2)).\left|\left(\frac{(1+e^{t^{-1/3}v})e^{-t^{-1/3}v/2}}{(1+e^{t^{-1/3}u})e^{-t^{-1/3}u/2}}\right)^{\lfloor t^{2/3}\alpha\rfloor-1}\right|\leq\exp\left(a_{\epsilon,3}\cdot(1+|\alpha|)\cdot(|u|^{2}+|v|^{2})\right).
Lemma 4.8.

Let 𝗋>0\mathsf{r}>0 be as in Lemma 4.3, τ1(0,1)\tau_{1}\in(0,1) and ϵ(0,𝗋]\epsilon\in(0,\mathsf{r}] be sufficiently small so that e20ϵτ11e^{20\epsilon}\leq\tau_{1}^{-1}. We can find Tϵ,4,Aϵ,4>0T_{\epsilon,4},A_{\epsilon,4}>0, depending on ϵ\epsilon alone, such that the following holds. If tTϵ,4t\geq T_{\epsilon,4}, τ(0,τ1]\tau\in(0,\tau_{1}], uΓu,ϵ,t,vΓv,ϵ,tu\in\Gamma_{u,\epsilon,t},v\in\Gamma_{v,\epsilon,t}

(4.11) |t1/3(et1/3u;τ)(e2t1/3v;τ)et1/3u(et1/3v;τ)(τet1/3(u+v);τ)|Aϵ,4|v|.\left|\frac{t^{1/3}(-e^{t^{-1/3}u};\tau)_{\infty}(e^{2t^{-1/3}v};\tau)_{\infty}e^{t^{-1/3}u}}{(-e^{t^{-1/3}v};\tau)_{\infty}(\tau e^{t^{-1/3}(u+v)};\tau)_{\infty}}\right|\leq A_{\epsilon,4}\cdot|v|.
Lemma 4.9.

Let 𝗋>0\mathsf{r}>0 be as in Lemma 4.3, ϵ(0,𝗋]\epsilon\in(0,\mathsf{r}] and kk\in\mathbb{N}. Suppose that uaΓu,ϵ,t,vaΓv,ϵ,tu_{a}\in\Gamma_{u,\epsilon,t},v_{a}\in\Gamma_{v,\epsilon,t} for a{1,,k}a\in\{1,\dots,k\}. We can find Tϵ,5,Aϵ,5>0T_{\epsilon,5},A_{\epsilon,5}>0, depending on ϵ\epsilon alone, such that if B1(u,v;t)B_{1}(\vec{u},\vec{v};t) is as in (4.7) and tTϵ,5t\geq T_{\epsilon,5}, then

(4.12) |B1(u,v;t)|Aϵ,5kkk/2.\left|B_{1}(\vec{u},\vec{v};t)\right|\leq A_{\epsilon,5}^{k}\cdot k^{k/2}.
Lemma 4.10.

Let 𝗋>0\mathsf{r}>0 be as in Lemma 4.3, τ1(0,1)\tau_{1}\in(0,1) and ϵ(0,𝗋]\epsilon\in(0,\mathsf{r}] be sufficiently small so that e20ϵτ11e^{20\epsilon}\leq\tau_{1}^{-1}. We can find Tϵ,6,Aϵ,6>0T_{\epsilon,6},A_{\epsilon,6}>0, depending on ϵ\epsilon alone, such that the following holds. If tTϵ,6t\geq T_{\epsilon,6}, τ(0,τ1]\tau\in(0,\tau_{1}], uaΓu,ϵ,t,vaΓv,ϵ,tu_{a}\in\Gamma_{u,\epsilon,t},v_{a}\in\Gamma_{v,\epsilon,t} for a{1,,k}a\in\{1,\dots,k\} and B2(u,v;t)B_{2}(\vec{u},\vec{v};t) is as in (4.7), then

(4.13) |B2(u,v;t)|Aϵ,6k2.\left|B_{2}(\vec{u},\vec{v};t)\right|\leq A_{\epsilon,6}^{k^{2}}.

4.3. Proof of Proposition 3.2

In this section we present the proof of Proposition 3.2, and for clarity we split the proof into three steps.

Step 1. Let ϵ0\epsilon_{0} be as in Definition 4.4, and let T1>0T_{1}>0 be sufficiently large, so that T1T0T_{1}\geq T_{0} and T1maxi=1,,6Tϵ0,iT_{1}\geq\max_{i=1,\dots,6}T_{\epsilon_{0},i}, where T0T_{0} is as in Definition 4.4, and Tϵ0,iT_{\epsilon_{0},i} are as in Lemmas 4.5-4.10. For kk\in\mathbb{N} and tT1t\geq T_{1}, we define

(4.14) J(k,t):=(Γu,ϵ0,t0)k𝑑u(Γv,ϵ0,t0)k𝑑vi=15Ai(u,v;t)i=12Bi(u,v;t),J(k,t):=\int_{(\Gamma^{0}_{u,\epsilon_{0},t})^{k}}d\vec{u}\int_{(\Gamma^{0}_{v,\epsilon_{0},t})^{k}}d\vec{v}\prod_{i=1}^{5}A_{i}(\vec{u},\vec{v};t)\prod_{i=1}^{2}B_{i}(\vec{u},\vec{v};t),

where Γu,ϵ0,t0,Γv,ϵ0,t0\Gamma^{0}_{u,\epsilon_{0},t},\Gamma^{0}_{v,\epsilon_{0},t} are as in Definition 4.2, Ai(u,v;t)A_{i}(\vec{u},\vec{v};t) are as in (4.6), and Bi(u,v;t)B_{i}(\vec{u},\vec{v};t) are as in (4.7). We claim that

(4.15) limtJ(k,t)=C0,π/4k𝑑uC1,3π/4k𝑑vdet[2vaeua3/48ua2α/8uar~(uava)(vb2ua2)eva3/48va2α/8var~]a,b=1k.\lim_{t\rightarrow\infty}J(k,t)=\int_{C_{0,\pi/4}^{k}}d\vec{u}\int_{C_{-1,3\pi/4}^{k}}d\vec{v}\det\left[\frac{2v_{a}e^{u_{a}^{3}/48-u_{a}^{2}\alpha/8-u_{a}\tilde{r}}}{(u_{a}-v_{a})(v_{b}^{2}-u_{a}^{2})e^{v_{a}^{3}/48-v_{a}^{2}\alpha/8-v_{a}\tilde{r}}}\right]_{a,b=1}^{k}.

We prove (4.15) in the steps below. Here, we assume its validity and conclude the proof of (3.2).

It follows from (4.5) that

(4.16) k!(2π𝗂)2kI(k,t)J(k,t)=σ1,,σk,τ1,,τk{0,1}:iσi+τi>0Γu,ϵ0,tσ1Γu,ϵ0,tσk𝑑uΓv,ϵ0,tτ1Γv,ϵ0,tτk𝑑vi=15Ai(u,v;t)i=12Bi(u,v;t).\begin{split}&k!(2\pi\mathsf{i})^{2k}I(k,t)-J(k,t)=\sum_{\sigma_{1},\dots,\sigma_{k},\tau_{1},\dots,\tau_{k}\in\{0,1\}:\sum_{i}\sigma_{i}+\tau_{i}>0}\int_{\Gamma^{\sigma_{1}}_{u,\epsilon_{0},t}}\cdots\int_{\Gamma^{\sigma_{k}}_{u,\epsilon_{0},t}}d\vec{u}\\ &\int_{\Gamma^{\tau_{1}}_{v,\epsilon_{0},t}}\cdots\int_{\Gamma^{\tau_{k}}_{v,\epsilon_{0},t}}d\vec{v}\prod_{i=1}^{5}A_{i}(\vec{u},\vec{v};t)\prod_{i=1}^{2}B_{i}(\vec{u},\vec{v};t).\end{split}

It follows from Lemmas 4.5, 4.6, 4.7, 4.8, 4.9, and 4.10 (all lemmas are applied for ϵ=ϵ0\epsilon=\epsilon_{0} as in Definition 4.4 and τ=τ1\tau=\tau_{1}) that for uΓu,ϵ0,tk\vec{u}\in\Gamma^{k}_{u,\epsilon_{0},t}, vΓv,ϵ0,tk\vec{v}\in\Gamma^{k}_{v,\epsilon_{0},t}, and tT1t\geq T_{1}, we have

(4.17) |i=15Ai(u,v;t)i=12Bi(u,v;t)|Aϵ0,1kAϵ0,2kAϵ0,4kAϵ0,5kAϵ0,6k2kk/2×i=1k|vi|exp(aϵ0,1(|ui|3+|vi|3)+aϵ0,3(1+|α|)(|ui|2+|vi|2)+|r~|(|ui|+|vi|)).\begin{split}&\left|\prod_{i=1}^{5}A_{i}(\vec{u},\vec{v};t)\prod_{i=1}^{2}B_{i}(\vec{u},\vec{v};t)\right|\leq A^{k}_{\epsilon_{0},1}\cdot A^{k}_{\epsilon_{0},2}\cdot A^{k}_{\epsilon_{0},4}\cdot A^{k}_{\epsilon_{0},5}\cdot A^{k^{2}}_{\epsilon_{0},6}\cdot k^{k/2}\\ &\times\prod_{i=1}^{k}|v_{i}|\exp\left(-a_{\epsilon_{0},1}(|u_{i}|^{3}+|v_{i}|^{3})+a_{\epsilon_{0},3}(1+|\alpha|)(|u_{i}|^{2}+|v_{i}|^{2})+|\tilde{r}|(|u_{i}|+|v_{i}|)\right).\end{split}

Observe that if tT1t\geq T_{1}, we have

|z|ϵ0t1/3 for zΓu,ϵ0,t1Γv,ϵ0,t1, and |z|2πt1/3 for zΓu,ϵ0,tΓv,ϵ0,t.|z|\geq\epsilon_{0}t^{1/3}\mbox{ for }z\in\Gamma^{1}_{u,\epsilon_{0},t}\cup\Gamma^{1}_{v,\epsilon_{0},t},\mbox{ and }|z|\leq 2\pi t^{1/3}\mbox{ for }z\in\Gamma_{u,\epsilon_{0},t}\cup\Gamma_{v,\epsilon_{0},t}.

Using the latter and (4.17), we conclude that there exist constants A,a>0A,a>0, such that for tT1t\geq T_{1}

(4.18) |i=15Ai(u,v;t)i=12Bi(u,v;t)|Aexp(at2/3aϵ0,1ϵ03ti=1k(σi+τi)) if (u1,,uk)Γu,ϵ0,tσ1××Γu,ϵ0,tσk and (v1,,vk)Γv,ϵ0,tτ1××Γv,ϵ0,tτk.\begin{split}&\left|\prod_{i=1}^{5}A_{i}(\vec{u},\vec{v};t)\prod_{i=1}^{2}B_{i}(\vec{u},\vec{v};t)\right|\leq A\exp\left(at^{2/3}-a_{\epsilon_{0},1}\epsilon_{0}^{3}t\sum_{i=1}^{k}(\sigma_{i}+\tau_{i})\right)\mbox{ if }\\ &(u_{1},\dots,u_{k})\in\Gamma^{\sigma_{1}}_{u,\epsilon_{0},t}\times\cdots\times\Gamma^{\sigma_{k}}_{u,\epsilon_{0},t}\mbox{ and }(v_{1},\dots,v_{k})\in\Gamma^{\tau_{1}}_{v,\epsilon_{0},t}\times\cdots\times\Gamma^{\tau_{k}}_{v,\epsilon_{0},t}.\end{split}

Combining (4.16), (4.18) with the fact that the lengths of Γu,ϵ0,ti,Γv,ϵ0,ti\Gamma^{i}_{u,\epsilon_{0},t},\Gamma^{i}_{v,\epsilon_{0},t} for i=0,1i=0,1 are at most 4πt1/34\pi t^{1/3}, we conclude that for all large tt we have

|J(k,t)k!(2π𝗂)2kI(k,t)|exp(aϵ0,1ϵ03t/2).\left|J(k,t)-k!(2\pi\mathsf{i})^{2k}I(k,t)\right|\leq\exp\left(-a_{\epsilon_{0},1}\epsilon_{0}^{3}t/2\right).

The latter inequality and (4.15) together imply (3.2).

Step 2. From (4.14) we have that

(4.19) J(k,t)=C0,π/4kC1,3π/4kgt(u,v)𝑑v𝑑u, where gt(u,v)=i=1k𝟏{|𝖨𝗆(ui)|ϵ0t1/3,|𝖨𝗆(vi)|ϵ0t1/3}i=15Ai(u,v;t)i=12Bi(u,v;t).\begin{split}&J(k,t)=\int_{C^{k}_{0,\pi/4}}\int_{C_{-1,3\pi/4}^{k}}g_{t}(\vec{u},\vec{v})d\vec{v}d\vec{u},\mbox{ where }\\ &g_{t}(\vec{u},\vec{v})=\prod_{i=1}^{k}{\bf 1}\{|\mathsf{Im}(u_{i})|\leq\epsilon_{0}t^{1/3},|\mathsf{Im}(v_{i})|\leq\epsilon_{0}t^{1/3}\}\prod_{i=1}^{5}A_{i}(\vec{u},\vec{v};t)\prod_{i=1}^{2}B_{i}(\vec{u},\vec{v};t).\end{split}

We claim that for all uC0,π/4k\vec{u}\in C^{k}_{0,\pi/4} and vC1,3π/4k\vec{v}\in C_{-1,3\pi/4}^{k} we have

(4.20) limtgt(u,v)=det[2vaeua3/48ua2α/8uar~(uava)(vb2ua2)eva3/48va2α/8var~]a,b=1k.\begin{split}\lim_{t\rightarrow\infty}g_{t}(\vec{u},\vec{v})=\det\left[\frac{2v_{a}e^{u_{a}^{3}/48-u_{a}^{2}\alpha/8-u_{a}\tilde{r}}}{(u_{a}-v_{a})(v_{b}^{2}-u_{a}^{2})e^{v_{a}^{3}/48-v_{a}^{2}\alpha/8-v_{a}\tilde{r}}}\right]_{a,b=1}^{k}.\end{split}

We prove (4.20) in the next step. Here, we assume its validity and conclude the proof of (4.15).

In view of (4.20) and the dominated convergence theorem, with dominating function given by the right side of (4.17), we may take the tt\rightarrow\infty limit in (4.19) to get (4.15).

Step 3. In this step we prove (4.20), and in the sequel we assume that uC0,π/4k\vec{u}\in C^{k}_{0,\pi/4}, and vC1,3π/4k\vec{v}\in C_{-1,3\pi/4}^{k} are fixed. We begin to investigate the limits of Ai(u,v;t)A_{i}(\vec{u},\vec{v};t) and Bi(u,v;t)B_{i}(\vec{u},\vec{v};t) in the definition of gt(u,v)g_{t}(\vec{u},\vec{v}).

From the first line of (4.3) we have for all large tt and fixed zz\in\mathbb{C}

|tF(t1/3z)z3/48|C1t1/3|z|4, and so limttF(t1/3z)=z3/48.\left|tF(t^{-1/3}z)-z^{3}/48\right|\leq C_{1}t^{-1/3}|z|^{4},\mbox{ and so }\lim_{t\rightarrow\infty}tF(t^{-1/3}z)=z^{3}/48.

In particular, we conclude that

(4.21) limtA1(u,v;t)=a=1kexp(ua3/48va3/48).\lim_{t\rightarrow\infty}A_{1}(\vec{u},\vec{v};t)=\prod_{a=1}^{k}\exp\left(u_{a}^{3}/48-v_{a}^{3}/48\right).

We also observe that for each zz\in\mathbb{C}

limtmt1/3[logτ]1πe2mπ𝗂[(1/4)t+(1/2)(t2/3α1)t1/3r~]sin(π[logτ]1[t1/3z2mπ𝗂])=1z,\lim_{t\rightarrow\infty}\sum_{m\in\mathbb{Z}}\frac{t^{-1/3}[-\log\tau]^{-1}\pi\cdot e^{2m\pi\mathsf{i}[(1/4)t+(1/2)(\lfloor t^{2/3}\alpha\rfloor-1)-t^{1/3}\tilde{r}]}}{\sin(-\pi[\log\tau]^{-1}[t^{-1/3}z-2m\pi\mathsf{i}])}=\frac{1}{z},

where we can exchange the order of the sum and the limit by (2.2), and only the m=0m=0 summand contributes to the limit. We conclude that

(4.22) limtA2(u,v;t)=a=1k1vaua.\lim_{t\rightarrow\infty}A_{2}(\vec{u},\vec{v};t)=\prod_{a=1}^{k}\frac{1}{v_{a}-u_{a}}.

By a direct Taylor expansion we have for each α\alpha\in\mathbb{R} and zz\in\mathbb{C}

limt((1+et1/3z)et1/3z/22)t2/3α1=ez2α/8,\lim_{t\rightarrow\infty}\left(\frac{(1+e^{t^{-1/3}z})e^{-t^{-1/3}z/2}}{2}\right)^{\lfloor t^{2/3}\alpha\rfloor-1}=e^{z^{2}\alpha/8},

from which we conclude that

(4.23) limtA3(u,v;t)=a=1kexp(va2α/8ua2α/8).\lim_{t\rightarrow\infty}A_{3}(\vec{u},\vec{v};t)=\prod_{a=1}^{k}\exp\left(v_{a}^{2}\alpha/8-u_{a}^{2}\alpha/8\right).

For every z,wz,w\in\mathbb{C} we have

limtt1/3(et1/3z;τ)(e2t1/3w;τ)et1/3z(et1/3w;τ)(τet1/3(z+w);τ)=limtt1/3(1e2t1/3w)×limt(et1/3z;τ)(τe2t1/3w;τ)et1/3z(et1/3w;τ)(τet1/3(z+w);τ)=2w,\begin{split}&\lim_{t\rightarrow\infty}\frac{t^{1/3}(-e^{t^{-1/3}z};\tau)_{\infty}(e^{2t^{-1/3}w};\tau)_{\infty}e^{t^{-1/3}z}}{(-e^{t^{-1/3}w};\tau)_{\infty}(\tau e^{t^{-1/3}(z+w)};\tau)_{\infty}}=\lim_{t\rightarrow\infty}t^{1/3}(1-e^{2t^{-1/3}w})\\ &\times\lim_{t\rightarrow\infty}\frac{(-e^{t^{-1/3}z};\tau)_{\infty}(\tau e^{2t^{-1/3}w};\tau)_{\infty}e^{t^{-1/3}z}}{(-e^{t^{-1/3}w};\tau)_{\infty}(\tau e^{t^{-1/3}(z+w)};\tau)_{\infty}}=-2w,\end{split}

from which we conclude that

(4.24) limtA5(u,v;t)=a=1k(2va).\lim_{t\rightarrow\infty}A_{5}(\vec{u},\vec{v};t)=\prod_{a=1}^{k}(-2v_{a}).

Finally, we have the straightforward limits

(4.25) limtB1(u,v;t)=det[1vb2ua2]a,b=1k, and limtB2(u,v;t)=1.\lim_{t\rightarrow\infty}B_{1}(\vec{u},\vec{v};t)=\det\left[\frac{1}{v_{b}^{2}-u_{a}^{2}}\right]_{a,b=1}^{k}\mbox{, and }\lim_{t\rightarrow\infty}B_{2}(\vec{u},\vec{v};t)=1.

Combining (4.21), (4.22, (4.23), (4.24) and (4.25) we conclude that

limti=15Ai(u,v;t)i=12Bi(u,v;t)=a=1keua3/48ua2α/8uar~eva3/48va2α/8var~(2va)vaua×det[1vb2ua2]a,b=1k.\lim_{t\rightarrow\infty}\prod_{i=1}^{5}A_{i}(\vec{u},\vec{v};t)\prod_{i=1}^{2}B_{i}(\vec{u},\vec{v};t)=\prod_{a=1}^{k}\frac{e^{u_{a}^{3}/48-u_{a}^{2}\alpha/8-u_{a}\tilde{r}}}{e^{v_{a}^{3}/48-v_{a}^{2}\alpha/8-v_{a}\tilde{r}}}\cdot\frac{(-2v_{a})}{v_{a}-u_{a}}\times\det\left[\frac{1}{v_{b}^{2}-u_{a}^{2}}\right]_{a,b=1}^{k}.

The last equation implies (4.20) once we use the multilinearity of the determinant.

4.4. Proof of the lemmas from Sections 4.1 and 4.2

In this section we present the proofs of the eight lemmas from Sections 4.1 and 4.2.

Proof of Lemma 4.1.

The proof we present here is an adaptation of the one in [GT20, Section A]. We proceed to prove (4.1) by induction on kk with base case k=1k=1 being obvious. Assuming the result for kk we proceed to prove it for k+1k+1. Let us denote for convenience

m(w,z)=1(wz)(1wz) and g(w,z)=(wz)(1wz).m(w,z)=\frac{1}{(w-z)(1-wz)}\mbox{ and }g(w,z)=(w-z)(1-wz).

By dividing each row of (m(wi,zj))1i,jk+1\left(m(w_{i},z_{j})\right)_{1\leq i,j\leq k+1} by the entry in the first column, and subtracting the first row from all other rows, we obtain

det[m(wi,zj)]1i,jk+1=i=1k+1m(wi,z1)det[m(wi,zj)m(wi,z1)m(w1,zj)m(w1,z1)]2i,jk+1.\det\left[m(w_{i},z_{j})\right]_{1\leq i,j\leq k+1}=\prod_{i=1}^{k+1}m(w_{i},z_{1})\cdot\det\left[\frac{m(w_{i},z_{j})}{m(w_{i},z_{1})}-\frac{m(w_{1},z_{j})}{m(w_{1},z_{1})}\right]_{2\leq i,j\leq k+1}.

We next note by a direct computation that

m(wi,zj)m(wi,z1)m(w1,zj)m(w1,z1)=g(z1,zj)g(w1,wi)m(w1,zj)m(wi,zj).\frac{m(w_{i},z_{j})}{m(w_{i},z_{1})}-\frac{m(w_{1},z_{j})}{m(w_{1},z_{1})}=-g(z_{1},z_{j})g(w_{1},w_{i})m(w_{1},z_{j})m(w_{i},z_{j}).

Combining the last two equalities and the multi-linearity of the determinant we get

det[m(wi,zj)]1i,jk+1=(1)ki=1k+1m(wi,z1)j=2k+1g(z1,zj)m(w1,zj)×i=2k+1g(w1,wi)det[m(wi,zj)]2i,jk+1.\begin{split}\det\left[m(w_{i},z_{j})\right]_{1\leq i,j\leq k+1}=\hskip 5.69054pt&(-1)^{k}\prod_{i=1}^{k+1}m(w_{i},z_{1})\prod_{j=2}^{k+1}g(z_{1},z_{j})m(w_{1},z_{j})\\ &\times\prod_{i=2}^{k+1}g(w_{1},w_{i})\cdot\det\left[m(w_{i},z_{j})\right]_{2\leq i,j\leq k+1}.\end{split}

Applying the induction hypothesis to the last determinant we arrive at (4.1) for k+1k+1, which completes the induction step. ∎

Proof of Lemma 4.3.

We note that for any a0a\neq 0

𝖱𝖾[F(a±𝗂y)]=𝖱𝖾11+ea[cos(y)±𝗂sin(y)]+a412=eacos(y)+1e2a+1+2eacos(y)+a412.\mathsf{Re}[F(a\pm\mathsf{i}y)]=\mathsf{Re}\frac{1}{1+e^{a}[\cos(y)\pm\mathsf{i}\sin(y)]}+\frac{a}{4}-\frac{1}{2}=\frac{e^{a}\cos(y)+1}{e^{2a}+1+2e^{a}\cos(y)}+\frac{a}{4}-\frac{1}{2}.

By direct computation, we conclude that

ddy𝖱𝖾[F(a+𝗂y)]=easin(y)(e2a1)(e2a+1+2eacos(y))2.\frac{d}{dy}\mathsf{Re}[F(a+\mathsf{i}y)]=-\frac{e^{a}\sin(y)(e^{2a}-1)}{(e^{2a}+1+2e^{a}\cos(y))^{2}}.

The last equality for a=xa=x and a=xa=-x implies (4.4). In the remainder we prove (4.3).

We note that F(z)F(z) is a meromorphic function on \mathbb{C} with simple poles at π𝗂+2π𝗂\pi\mathsf{i}+2\pi\mathsf{i}\cdot\mathbb{Z}, and so it is analytic in the zero-centered disc of radius π\pi. If F(z)=n=0anznF(z)=\sum_{n=0}^{\infty}a_{n}z^{n} is the Taylor expansion of F(z)F(z) at the origin, we directly compute a0=a1=a2=0a_{0}=a_{1}=a_{2}=0 and a3=1/48a_{3}=1/48. The latter suggests that z4(F(z)z3/48)z^{-4}\cdot(F(z)-z^{3}/48) is an analytic function in the zero-centered disc of radius π\pi, so that there is a constant C1>0C_{1}>0, such that for all |z|3/2|z|\leq 3/2 we have

|F(z)z3/48z4|C1.\left|\frac{F(z)-z^{3}/48}{z^{4}}\right|\leq C_{1}.

This specifies our choice of C1C_{1} and proves the first inequality in (4.3) for all 𝗋1/2\mathsf{r}\leq 1/2.

Let 𝗋(0,1/2)\mathsf{r}\in(0,1/2) be sufficiently small so that 3C1𝗋10133C_{1}\cdot\mathsf{r}\leq 101^{-3}. We also let C2=106C_{2}=10^{6}. This specifies our choice of 𝗋\mathsf{r} and C2C_{2} and we proceed to prove the second and third line in (4.3). As the proofs of the second and third inequality in (4.3) are quite similar, we only establish the second.

We fix ρ,ϵ0\rho\in\mathbb{R},\epsilon\geq 0, such that 0|ρ|ϵ𝗋0\leq|\rho|\leq\epsilon\leq\mathsf{r}. We also fix zγρ,ϵ+z\in\gamma_{\rho,\epsilon}^{+}, such that |𝖨𝗆(z)|ϵ|\mathsf{Im}(z)|\leq\epsilon, and note that z=ρ+re𝗂ϕz=\rho+re^{\mathsf{i}\phi}, where ϕ{π/4,π/4}\phi\in\{-\pi/4,\pi/4\}, while 0r2ϵ0\leq r\leq\sqrt{2}\epsilon. In particular, we see that |z||ρ|+r3𝗋|z|\leq|\rho|+r\leq 3\mathsf{r}, so that from the first inequality in (4.3) we have

(4.26) 𝖱𝖾[F(z)]𝖱𝖾[z3/48]+C1|z|4𝖱𝖾[z3/48]+C13𝗋|z|3𝖱𝖾[z3/48]+1013|z|3.\mathsf{Re}[F(z)]\leq\mathsf{Re}[z^{3}/48]+C_{1}|z|^{4}\leq\mathsf{Re}[z^{3}/48]+C_{1}\cdot 3\mathsf{r}|z|^{3}\leq\mathsf{Re}[z^{3}/48]+101^{-3}|z|^{3}.

We also have from ϕ{π/4,π/4}\phi\in\{-\pi/4,\pi/4\} that

(4.27) 𝖱𝖾[z3]=𝖱𝖾[(ρ+re𝗂ϕ)3]=ρ323/2r3+321/2ρ2r.\mathsf{Re}[z^{3}]=\mathsf{Re}[(\rho+re^{\mathsf{i}\phi})^{3}]=\rho^{3}-2^{-3/2}\cdot r^{3}+3\cdot 2^{-1/2}\cdot\rho^{2}r.

If r100|ρ|r\leq 100|\rho|, then (4.27) implies that

(4.28) 𝖱𝖾[z3/48]5|ρ|3, and |z|3(101)3|ρ|3.\mathsf{Re}[z^{3}/48]\leq 5\cdot|\rho|^{3},\mbox{ and }|z|^{3}\leq(101)^{3}|\rho|^{3}.

In particular, from (4.26) and (4.28) we get

𝖱𝖾[F(z)]𝖱𝖾[z3/48]+1013|z|36|ρ|3C2|ρ|32001(101)3|ρ|3|z|3200+C2|ρ|3,\mathsf{Re}[F(z)]\leq\mathsf{Re}[z^{3}/48]+101^{-3}|z|^{3}\leq 6\cdot|\rho|^{3}\leq C_{2}|\rho|^{3}-200^{-1}\cdot(101)^{3}|\rho|^{3}\leq-\frac{|z|^{3}}{200}+C_{2}|\rho|^{3},

which proves the second line in (4.3) if r100|ρ|r\leq 100|\rho|.

Finally, we suppose that r100|ρ|r\geq 100|\rho|. In this case, we have

(101/100)rr+|ρ||z|r|ρ|(99/100)r.(101/100)\cdot r\geq r+|\rho|\geq|z|\geq r-|\rho|\geq(99/100)\cdot r.

The latter and (4.27) imply

𝖱𝖾[z3/48]r348(23/2+106+321/2104)r3144|z|3(2001+1013).\mathsf{Re}[z^{3}/48]\leq\frac{r^{3}}{48}\cdot(-2^{-3/2}+10^{-6}+3\cdot 2^{-1/2}\cdot 10^{-4})\leq-\frac{r^{3}}{144}\leq-|z|^{3}\cdot\left(200^{-1}+101^{-3}\right).

The last inequality and (4.26) prove (4.3) if r100|ρ|r\geq 100|\rho|. ∎

Proof of Lemma 4.5.

Let C2C_{2} be as in Lemma 4.3. We will prove that if tϵ3t\geq\epsilon^{-3}

(4.29) |etF(t1/3u)|exp(C2|u|3ϵ3800π3) for uΓu,ϵ,t, and |etF(t1/3v)|exp(C2|v|3ϵ3800π3) for vΓv,ϵ,t.\begin{split}&\left|e^{tF(t^{-1/3}u)}\right|\leq\exp\left(C_{2}-\frac{|u|^{3}\epsilon^{3}}{800\pi^{3}}\right)\mbox{ for $u\in\Gamma_{u,\epsilon,t}$, and }\\ &\left|e^{-tF(t^{-1/3}v)}\right|\leq\exp\left(C_{2}-\frac{|v|^{3}\epsilon^{3}}{800\pi^{3}}\right)\mbox{ for $v\in\Gamma_{v,\epsilon,t}$}.\end{split}

Notice that (4.29) implies the statement of the lemma with Tϵ,1=ϵ3T_{\epsilon,1}=\epsilon^{-3}, Aϵ,1=exp(2C2)A_{\epsilon,1}=\exp(2C_{2}), and aϵ,1=ϵ3800π3a_{\epsilon,1}=\frac{\epsilon^{3}}{800\pi^{3}}. As the proofs of the two lines of (4.29) are quite similar, we only establish the second.

Notice that if |𝖨𝗆(v)|ϵt1/3|\mathsf{Im}(v)|\leq\epsilon t^{1/3} and vΓv,ϵ,tv\in\Gamma_{v,\epsilon,t}, we have from the third line of (4.3) with ρ=t1/3\rho=-t^{-1/3}

(4.30) 𝖱𝖾[tF(t1/3v)]|v|3200+C2.\mathsf{Re}\left[-tF(t^{-1/3}v)\right]\leq-\frac{|v|^{3}}{200}+C_{2}.

Suppose that vΓv,ϵ,tv\in\Gamma_{v,\epsilon,t} is such that |𝖨𝗆(v)|ϵt1/3|\mathsf{Im}(v)|\geq\epsilon t^{1/3}, and let v0=1ϵt1/3±𝗂ϵt1/3v_{0}=-1-\epsilon t^{1/3}\pm\mathsf{i}\epsilon t^{1/3}, where the sign is chosen to agree with the imaginary part of vv. Then, we observe that

(4.31) 𝖱𝖾[tF(t1/3v)]𝖱𝖾[tF(t1/3v0)]|v0|3200+C2|v|3ϵ3800π3+C2,\mathsf{Re}\left[-tF(t^{-1/3}v)\right]\leq\mathsf{Re}\left[-tF(t^{-1/3}v_{0})\right]\leq-\frac{|v_{0}|^{3}}{200}+C_{2}\leq-\frac{|v|^{3}\epsilon^{3}}{800\pi^{3}}+C_{2},

where in the first inequality we used the second inequality in (4.4) with x=t1/3+ϵx=t^{-1/3}+\epsilon, in the second inequality we used (4.30) with v=v0v=v_{0}, and in the last inequality we used that |v0|ϵ|v_{0}|\geq\epsilon, |v|2π|v|\leq\sqrt{2}\pi. Equations (4.30) and (4.31) imply the second line of (4.29), as |ez|=exp(𝖱𝖾(z))|e^{z}|=\exp(\mathsf{Re}(z)) for all zz\in\mathbb{C}. ∎

Proof of Lemma 4.6.

We set Tϵ,2=ϵ3T_{\epsilon,2}=\epsilon^{-3} and note that for tTϵ,2t\geq T_{\epsilon,2}, uΓu,ϵ,tu\in\Gamma_{u,\epsilon,t}, and vΓv,ϵ,tv\in\Gamma_{v,\epsilon,t} we have

2ϵt1/3𝖱𝖾(v)t1/3 and ϵt1/3𝖱𝖾(u)0.-2\epsilon\leq t^{-1/3}\mathsf{Re}(v)\leq-t^{-1/3}\mbox{ and }\epsilon\geq t^{-1/3}\mathsf{Re}(u)\geq 0.

The latter implies that the conditions of [Dim18, Lemma 4.5] are satisfied with t=τt=\tau, u=t1/3u=t^{-1/3}, U=2ϵU=2\epsilon (here we used that ϵ𝗋<1/2\epsilon\leq\mathsf{r}<1/2 as in Lemma 4.3, and e20ϵτ11e^{20\epsilon}\leq\tau_{1}^{-1}). From the proof of [Dim18, Lemma 4.5], see the displayed equation after [Dim18, (6.15)], we conclude that

m|1sin(π[logτ]1[t1/3(vu)2mπ𝗂])|2[logτ]t1/3k0τ2kπ2.\sum_{m\in\mathbb{Z}}\left|\frac{1}{\sin(-\pi[\log\tau]^{-1}[t^{-1/3}(v-u)-2m\pi\mathsf{i}])}\right|\leq 2[-\log\tau]t^{1/3}\sum_{k\geq 0}\tau^{2k\pi^{2}}.

The latter, the triangle inequality, and the fact that |e𝗂x|=1|e^{\mathsf{i}x}|=1 for xx\in\mathbb{R}, imply that the left side of (4.9) is bounded by

2πk0τ2kπ22πk0e40kϵπ2,2\pi\cdot\sum_{k\geq 0}\tau^{2k\pi^{2}}\leq 2\pi\cdot\sum_{k\geq 0}e^{-40k\epsilon\pi^{2}},

which proves (4.9) with Aϵ,2=2πk0e40kϵπ2A_{\epsilon,2}=2\pi\cdot\sum_{k\geq 0}e^{-40k\epsilon\pi^{2}} and Tϵ,2=ϵ3T_{\epsilon,2}=\epsilon^{-3}. ∎

Proof of Lemma 4.7.

Let Kϵ={z=x+𝗂y:|x|2ϵ,|y|π and |z±𝗂π|ϵ}K_{\epsilon}=\{z=x+\mathsf{i}y\in\mathbb{C}:|x|\leq 2\epsilon,|y|\leq\pi\mbox{ and }|z\pm\mathsf{i}\pi|\geq\epsilon\}. We claim that we can find aϵ,3>0a_{\epsilon,3}>0, such that for zKϵz\in K_{\epsilon} we have

(4.32) e(aϵ,3/2)|z|2|ez/2+ez/22|e(aϵ,3/2)|z|2.\begin{split}e^{-(a_{\epsilon,3}/2)|z|^{2}}\leq\left|\frac{e^{z/2}+e^{-z/2}}{2}\right|\leq e^{(a_{\epsilon,3}/2)|z|^{2}}.\end{split}

If (4.32) holds, then (4.10) would follow with the same choice of aϵ,3a_{\epsilon,3} and Tϵ,3=ϵ3T_{\epsilon,3}=\epsilon^{-3}, since |t2/3α1|t2/3|α|+2|\lfloor t^{2/3}\alpha\rfloor-1|\leq t^{2/3}|\alpha|+2 and for tTϵ,3t\geq T_{\epsilon,3} we have t1/3Γu,ϵ,tKϵt^{-1/3}\Gamma_{u,\epsilon,t}\subseteq K_{\epsilon}, t1/3Γv,ϵ,tKϵt^{-1/3}\Gamma_{v,\epsilon,t}\subseteq K_{\epsilon}.

In the remainder we prove (4.32). By Taylor expansion, we can find δ,A>0\delta,A>0, such that for |z|δ|z|\leq\delta the function log(ez+ez2)\log\left(\frac{e^{z}+e^{-z}}{2}\right) is well-defined, and

|log(ez+ez2)|A|z|2.\left|\log\left(\frac{e^{z}+e^{-z}}{2}\right)\right|\leq A|z|^{2}.

As always, the logarithm is with respect to the principal branch. This shows that for |z|δ|z|\leq\delta

(4.33) eA|z|2|ez/2+ez/22|eA|z|2.\begin{split}e^{-A|z|^{2}}\leq\left|\frac{e^{z/2}+e^{-z/2}}{2}\right|\leq e^{A|z|^{2}}.\end{split}

Since ez/2+ez/22\frac{e^{z/2}+e^{-z/2}}{2} is continuous and does not vanish on KϵK_{\epsilon} (note that the zeros are at π𝗂+2π𝗂\pi\mathsf{i}+2\pi\mathsf{i}\mathbb{Z}), we conclude that there is a constant M>0M>0 such that for zKϵz\in K_{\epsilon}

eM|ez/2+ez/22|eM.e^{-M}\leq\left|\frac{e^{z/2}+e^{-z/2}}{2}\right|\leq e^{M}.

In particular, we see that if zKϵz\in K_{\epsilon} and |z|δ|z|\geq\delta we have

(4.34) eMδ2|z|2eM|ez/2+ez/22|eMeMδ2|z|2.\begin{split}e^{-M\delta^{-2}|z|^{2}}\leq e^{-M}\leq\left|\frac{e^{z/2}+e^{-z/2}}{2}\right|\leq e^{M}\leq e^{M\delta^{-2}|z|^{2}}.\end{split}

Combining, (4.33) and (4.34) we conclude (4.32) with aϵ,3=2max(A,Mδ2)a_{\epsilon,3}=2\cdot\max(A,M\delta^{-2}). ∎

Proof of Lemma 4.8.

We set Tϵ,4=ϵ3T_{\epsilon,4}=\epsilon^{-3} and proceed to estimate the various terms that appear in (4.11) for tTϵ,4t\geq T_{\epsilon,4}. Throughout we will use frequently that ττ1e20ϵ\tau\leq\tau_{1}\leq e^{-20\epsilon}, which follows from our assumptions.

We note that if |z|=r[0,1)|z|=r\in[0,1) and τ(0,τ1]\tau\in(0,\tau_{1}], then

(4.35) m=0(1rτ1m)m=0(1rτm)|(z;τ)|m=0(1+rτm)m=0(1+rτ1m).\prod_{m=0}^{\infty}(1-r\cdot\tau_{1}^{m})\leq\prod_{m=0}^{\infty}(1-r\cdot\tau^{m})\leq|(z;\tau)_{\infty}|\leq\prod_{m=0}^{\infty}(1+r\cdot\tau^{m})\leq\prod_{m=0}^{\infty}(1+r\cdot\tau_{1}^{m}).

Since 1e2zz\frac{1-e^{2z}}{z} is entire, we can find a constant M1M_{1}, such that if |z|2π|z|\leq 2\pi we have

|1e2z|M1|z|.\left|1-e^{2z}\right|\leq M_{1}|z|.

Since for tTϵ,4t\geq T_{\epsilon,4} we have |vt1/3|2π|vt^{-1/3}|\leq 2\pi for all vΓv,ϵ,tv\in\Gamma_{v,\epsilon,t}, we conclude that

(4.36) |t1/3(e2t1/3v;τ)|=|t1/3(1e2t1/3v)||(τe2t1/3v;τ)|M1|v|m=1(1+τ1m|e2mt1/3v|)M1|v|m=1(1+τ1m)M1|v|m=0(1+e20(m+1)ϵ),\begin{split}&\left|t^{1/3}(e^{2t^{-1/3}v};\tau)_{\infty}\right|=\left|t^{1/3}(1-e^{2t^{-1/3}v})\right|\cdot\left|(\tau e^{2t^{-1/3}v};\tau)_{\infty}\right|\\ &\leq M_{1}|v|\cdot\prod_{m=1}^{\infty}(1+\tau_{1}^{m}|e^{2mt^{-1/3}v}|)\leq M_{1}|v|\cdot\prod_{m=1}^{\infty}(1+\tau_{1}^{m})\leq M_{1}|v|\cdot\prod_{m=0}^{\infty}(1+e^{-20(m+1)\epsilon}),\end{split}

where in going to the second line we also used (4.35), in the second inequality we used that 𝖱𝖾(v)0\mathsf{Re}(v)\leq 0, and in the last one we used that τ1e20ϵ\tau_{1}\leq e^{-20\epsilon}.

Since 𝖱𝖾(ut1/3)ϵ\mathsf{Re}(ut^{-1/3})\leq\epsilon for uΓu,ϵ,tu\in\Gamma_{u,\epsilon,t}, we see that

(4.37) |(et1/3u;τ)|m=0(1+eϵτm)m=0(1+eϵ20mϵ),\begin{split}&\left|(-e^{t^{-1/3}u};\tau)_{\infty}\right|\leq\prod_{m=0}^{\infty}(1+e^{\epsilon}\tau^{m})\leq\prod_{m=0}^{\infty}(1+e^{\epsilon-20m\epsilon}),\end{split}

where again we used ττ1e20ϵ\tau\leq\tau_{1}\leq e^{-20\epsilon}.

In addition, for tTϵ,4t\geq T_{\epsilon,4}, uΓu,ϵ,t,vΓv,ϵ,tu\in\Gamma_{u,\epsilon,t},v\in\Gamma_{v,\epsilon,t} we note that t1/3𝖱𝖾(u+v)ϵt^{-1/3}\mathsf{Re}(u+v)\leq\epsilon and hence from (4.35) we get

(4.38) |(τet1/3(u+v);τ)|m=0(1τ1mτ|et1/3(u+v)|)m=0(1e19ϵ20mϵ),\begin{split}\left|(\tau e^{t^{-1/3}(u+v)};\tau)_{\infty}\right|\geq\prod_{m=0}^{\infty}(1-\tau_{1}^{m}\tau|e^{t^{-1/3}(u+v)}|)\geq\prod_{m=0}^{\infty}(1-e^{-19\epsilon-20m\epsilon}),\end{split}

where again we used ττ1e20ϵ\tau\leq\tau_{1}\leq e^{-20\epsilon}.

Finally, let KϵK_{\epsilon} be the compact set Kϵ={z=x+𝗂y:|x|2ϵ,|y|π and |z±𝗂π|ϵ}K_{\epsilon}=\{z=x+\mathsf{i}y\in\mathbb{C}:|x|\leq 2\epsilon,|y|\leq\pi\mbox{ and }|z\pm\mathsf{i}\pi|\geq\epsilon\}. Notice that ez+1e^{z}+1 does not vanish on KϵK_{\epsilon} and so we can find Lϵ>0L_{\epsilon}>0 such that for zKϵz\in K_{\epsilon} we have

|ez+1|Lϵ|e^{z}+1|\geq L_{\epsilon}

For tTϵ,4t\geq T_{\epsilon,4}, and vΓv,ϵ,tv\in\Gamma_{v,\epsilon,t} we have that t1/3vKϵt^{-1/3}v\in K_{\epsilon} and so we get

(4.39) |(et1/3v;τ)||1+et1/3v|m=0(1τ1mτ|et1/3v|)Lϵm=0(1e20(m+1)ϵ),\begin{split}\left|(-e^{t^{-1/3}v};\tau)_{\infty}\right|\geq\left|1+e^{t^{-1/3}v}\right|\cdot\prod_{m=0}^{\infty}\left(1-\tau_{1}^{m}\tau\left|e^{t^{-1/3}v}\right|\right)\geq L_{\epsilon}\cdot\prod_{m=0}^{\infty}\left(1-e^{-20(m+1)\epsilon}\right),\end{split}

where we used (4.35), ττ1e20ϵ\tau\leq\tau_{1}\leq e^{-20\epsilon} and that 𝖱𝖾(v)0\mathsf{Re}(v)\leq 0. Combining (4.36), (4.37), (4.38) and (4.39), we obtain (4.11) with

Aϵ,4=1Lϵm=0(1+e20(m+1)ϵ)(1+eϵ20mϵ)(1e19ϵ20mϵ)(1e20(m+1)ϵ).A_{\epsilon,4}=\frac{1}{L_{\epsilon}}\prod_{m=0}^{\infty}\frac{(1+e^{-20(m+1)\epsilon})(1+e^{\epsilon-20m\epsilon})}{(1-e^{-19\epsilon-20m\epsilon})(1-e^{-20(m+1)\epsilon})}.

Proof of Lemma 4.9.

Let Tϵ,5=ϵ3T_{\epsilon,5}=\epsilon^{-3} and δ>0\delta>0 be small enough so that for x[0,1]x\in[0,1] we have

1exδx.1-e^{-x}\geq\delta x.

We note that if tTϵ,5t\geq T_{\epsilon,5} we have t1/3𝖱𝖾(u)[0,ϵ]t^{-1/3}\mathsf{Re}(u)\in[0,\epsilon] and t1/3𝖱𝖾(v)[2ϵ,t1/3]t^{-1/3}\mathsf{Re}(v)\in[-2\epsilon,-t^{-1/3}] for uΓu,ϵ,tu\in\Gamma_{u,\epsilon,t} and vΓv,ϵ,tv\in\Gamma_{v,\epsilon,t}. The latter implies that for tTϵ,5t\geq T_{\epsilon,5}, uΓu,ϵ,tu\in\Gamma_{u,\epsilon,t} and vΓv,ϵ,tv\in\Gamma_{v,\epsilon,t} we have

(4.40) |t1/3et1/3uet1/3v|t1/3|et1/3u||et1/3v|t1/31et1/3δ1.\left|\frac{t^{-1/3}}{e^{t^{-1/3}u}-e^{t^{-1/3}v}}\right|\leq\frac{t^{-1/3}}{|e^{t^{-1/3}u}|-|e^{t^{-1/3}v}|}\leq\frac{t^{-1/3}}{1-e^{-t^{-1/3}}}\leq\delta^{-1}.

Next, we let KϵK_{\epsilon} be the compact set Kϵ={z=x+𝗂y:|x|2ϵ,|y|π}K_{\epsilon}=\{z=x+\mathsf{i}y\in\mathbb{C}:|x|\leq 2\epsilon,|y|\leq\pi\}. We observe that ez1z\frac{e^{z}-1}{z} does not vanish on KϵK_{\epsilon}, and so we can find Lϵ>0L_{\epsilon}>0, such that for all zKϵz\in K_{\epsilon}

|ez1z|Lϵ.\left|\frac{e^{z}-1}{z}\right|\geq L_{\epsilon}.

Notice that for tTϵ,5t\geq T_{\epsilon,5}, uΓu,ϵ,tu\in\Gamma_{u,\epsilon,t} and vΓv,ϵ,tv\in\Gamma_{v,\epsilon,t} we have t1/3(u+v)t^{-1/3}(u+v), upto a shift by 2mπ𝗂2m\pi\mathsf{i} for some m{1,0,1}m\in\{-1,0,1\}, belongs to Kϵ{z:|z|t1/3}K_{\epsilon}\cap\{z\in\mathbb{C}:|z|\geq t^{-1/3}\}. The latter implies that for tTϵ,5t\geq T_{\epsilon,5}, uΓu,ϵ,tu\in\Gamma_{u,\epsilon,t} and vΓv,ϵ,tv\in\Gamma_{v,\epsilon,t} we have

(4.41) |t1/31et1/3(u+v)|t1/3Lϵt1/3=Lϵ1.\left|\frac{t^{-1/3}}{1-e^{t^{-1/3}(u+v)}}\right|\leq\frac{t^{-1/3}}{L_{\epsilon}t^{-1/3}}=L^{-1}_{\epsilon}.

Combining (4.40), (4.41) and Hadamard’s inequality from Lemma 2.3, we get for tTϵ,5t\geq T_{\epsilon,5}

|det[t2/3(et1/3uaet1/3vb)(1et1/3(ua+vb))]a,b=1k|kk/2(δLϵ)k,\left|\det\left[\frac{t^{-2/3}}{(e^{t^{-1/3}u_{a}}-e^{t^{-1/3}v_{b}})(1-e^{t^{-1/3}(u_{a}+v_{b})})}\right]_{a,b=1}^{k}\right|\leq k^{k/2}(\delta L_{\epsilon})^{-k},

which proves (4.12) with Aϵ,5=(δLϵ)1A_{\epsilon,5}=(\delta L_{\epsilon})^{-1}. ∎

Proof of Lemma 4.10.

Let Tϵ,6=ϵ3T_{\epsilon,6}=\epsilon^{-3}. Note that for tTϵ,6t\geq T_{\epsilon,6}, uΓu,ϵ,tu\in\Gamma_{u,\epsilon,t}, and vΓv,ϵ,tv\in\Gamma_{v,\epsilon,t} we have

2ϵt1/3𝖱𝖾(v)0 and ϵt1/3𝖱𝖾(u)0.-2\epsilon\leq t^{-1/3}\mathsf{Re}(v)\leq 0\mbox{ and }\epsilon\geq t^{-1/3}\mathsf{Re}(u)\geq 0.

The latter inequalities with (4.35) and ττ1e20ϵ\tau\leq\tau_{1}\leq e^{-20\epsilon} (which follows from the assumptions in the lemma) together imply for tTϵ,6t\geq T_{\epsilon,6}, uΓu,ϵ,tu\in\Gamma_{u,\epsilon,t}, and vΓv,ϵ,tv\in\Gamma_{v,\epsilon,t} that

|(τet1/3(ua+ub);τ)(τet1/3(va+vb);τ)(τet1/3(ua+vb);τ)(τet1/3(va+ub);τ)|m=0(1+τm+1e2ϵ)(1+τm+1)(1τm+1eϵ)2m=0(1+e2ϵ20(m+1)ϵ)2(1eϵ20(m+1)ϵ)2.\begin{split}\left|\frac{(\tau e^{t^{-1/3}(u_{a}+u_{b})};\tau)_{\infty}(\tau e^{t^{-1/3}(v_{a}+v_{b})};\tau)_{\infty}}{(\tau e^{t^{-1/3}(u_{a}+v_{b})};\tau)_{\infty}(\tau e^{t^{-1/3}(v_{a}+u_{b})};\tau)_{\infty}}\right|\leq\prod_{m=0}^{\infty}\frac{(1+\tau^{m+1}e^{2\epsilon})(1+\tau^{m+1})}{(1-\tau^{m+1}e^{\epsilon})^{2}}\leq\prod_{m=0}^{\infty}\frac{(1+e^{2\epsilon-20(m+1)\epsilon})^{2}}{(1-e^{\epsilon-20(m+1)\epsilon})^{2}}.\end{split}

The latter proves (4.13) with Aϵ,6=m=01+e2ϵ20(m+1)ϵ1eϵ20(m+1)ϵA_{\epsilon,6}=\prod_{m=0}^{\infty}\frac{1+e^{2\epsilon-20(m+1)\epsilon}}{1-e^{\epsilon-20(m+1)\epsilon}}. ∎

5. Asymptotic analysis: Part II

In this section we present the proof of Proposition 3.3. In Section 5.1 we present several lemmas, which will be required for our arguments, and whose proofs are given in Section 5.4. The proof of Proposition 3.3 is established by considering the cases 1kt1\leq k\leq t and tkt\leq k separately in Sections 5.2 and 5.3, respectively.

5.1. Preliminary results

In this section we summarize various estimates that will be required in the proof of Proposition 3.3. As we explained earlier, the estimate of I(k,t)I(k,t) in Proposition 3.3 is established by considering the cases 1kt1\leq k\leq t and tkt\geq k separately. In the case 1kt1\leq k\leq t, we will use the formula for I(k,t)I(k,t) from (4.5) and to upper bound it we will use the estimates in Lemmas 4.5-4.9. In addition, we will use Lemma 5.1 below in place of Lemma 4.10 to upper bound the term B2(u,v;t)B_{2}(\vec{u},\vec{v};t) from (4.7). For the case ktk\geq t, we use a different formula for I(k,t)I(k,t), established in (5.3). This formula involves a different set of contours and functions A~i(w,z;t)\tilde{A}_{i}(\vec{w},\vec{z};t) and B~i(w,z;t)\tilde{B}_{i}(\vec{w},\vec{z};t) and we estimate these new functions in Lemmas 5.3-5.7. The proofs of all lemmas from this section can be found in Section 5.4.

Lemma 5.1.

There is a universal constant 𝖼>0\mathsf{c}>0 such that the following holds for τ(0,e8π]\tau\in(0,e^{-8\pi}]. If kk\in\mathbb{N} and Wi,ZiW_{i},Z_{i}\in\mathbb{C} with |Wi|2π,|Zi|2π|W_{i}|\leq 2\pi,|Z_{i}|\leq 2\pi for i=1,,ki=1,\dots,k, then

(5.1) |1a<bk(τeWa+Wb;τ)(τeZa+Zb;τ)(τeWa+Zb;τ)(τeZa+Wb;τ)|exp(A(τ)𝖱𝖾[(a=1k(WaZa))2])exp(A(τ)𝖱𝖾[a=1k(WaZa)2])a=1kexp(𝖼τka=1k(|Za|3+|Wa|3)),\begin{split}&\left|\prod_{1\leq a<b\leq k}\frac{(\tau e^{W_{a}+W_{b}};\tau)_{\infty}(\tau e^{Z_{a}+Z_{b}};\tau)_{\infty}}{(\tau e^{W_{a}+Z_{b}};\tau)_{\infty}(\tau e^{Z_{a}+W_{b}};\tau)_{\infty}}\right|\leq\exp\left(-A(\tau)\cdot\mathsf{Re}\left[\left(\sum_{a=1}^{k}(W_{a}-Z_{a})\right)^{2}\right]\right)\cdot\\ &\exp\left(A(\tau)\cdot\mathsf{Re}\left[\sum_{a=1}^{k}(W_{a}-Z_{a})^{2}\right]\right)\cdot\prod_{a=1}^{k}\exp\left(\mathsf{c}\tau k\cdot\sum_{a=1}^{k}(|Z_{a}|^{3}+|W_{a}|^{3})\right),\end{split}

where

(5.2) A(τ):=12n=1k=1kτnk=12k=1kτk1τk(0,).A(\tau):=\frac{1}{2}\sum_{n=1}^{\infty}\sum_{k=1}^{\infty}k\tau^{nk}=\frac{1}{2}\sum_{k=1}^{\infty}\frac{k\tau^{k}}{1-\tau^{k}}\in(0,\infty).
Definition 5.2.

Let τ(0,e8π]\tau\in(0,e^{-8\pi}], and let Cw,CzC_{w},C_{z} be the positively oriented circles, centered at the origin, of radii Rw=eR_{w}=e and Rz=τ3/4R_{z}=\tau^{3/4}. We observe that the choice of Rw,RzR_{w},R_{z} satisfies Rw>1>Rw1>Rz>τRwR_{w}>1>R_{w}^{-1}>R_{z}>\tau R_{w}, since τ(0,e8π]\tau\in(0,e^{-8\pi}].

Using the definition of I(k,t)I(k,t) from Definition 3.1 and Lemma 3.8 we conclude that

(5.3) I(k,t)=1k!(2π𝗂)2kCwk𝑑wCzk𝑑zi=14A~i(w,z;t)i=12B~i(w,z;t),\begin{split}I(k,t)=\frac{1}{k!(2\pi\mathsf{i})^{2k}}\oint_{C_{w}^{k}}d\vec{w}\oint_{C_{z}^{k}}d\vec{z}&\prod_{i=1}^{4}\tilde{A}_{i}(\vec{w},\vec{z};t)\prod_{i=1}^{2}\tilde{B}_{i}(\vec{w},\vec{z};t),\end{split}

where Cw,CzC_{w},C_{z} are as in Definition 5.2, and

(5.4) B~1(w,z;t)=det[1wazb]a,b=1k,B~2(w,z;t)=1a<bk(wawb;τ)(zazb;τ)(zawb;τ)(wazb;τ).\begin{split}&\tilde{B}_{1}(\vec{w},\vec{z};t)=\det\left[\frac{1}{w_{a}-z_{b}}\right]_{a,b=1}^{k},\hskip 5.69054pt\tilde{B}_{2}(\vec{w},\vec{z};t)=\prod_{1\leq a<b\leq k}\frac{(w_{a}w_{b};\tau)_{\infty}(z_{a}z_{b};\tau)_{\infty}}{(z_{a}w_{b};\tau)_{\infty}(w_{a}z_{b};\tau)_{\infty}}.\end{split}
(5.5) A~1(w,z;t)=a=1kexp(t[11+wa+logwa411+zalogza4]),A~2(w,z;t)=a=1k(m[logτ]1πe2mπ𝗂[(1/4)t+(1/2)(t2/3α1)t1/3r~]sin(π[logτ]1[logzalogwa2mπ𝗂])),A~3(w,z;t)=a=1k((1+za)elogza/2(1+wa)elogwa/2)t2/3α1,A~4(w,z;t)=a=1k(wa;τ)(za2;τ)(za;τ)(zawa;τ)zaexp(t1/3r~[logzalogwa]),\begin{split}&\tilde{A}_{1}(\vec{w},\vec{z};t)=\prod_{a=1}^{k}\exp\left(t\left[\frac{1}{1+w_{a}}+\frac{\log w_{a}}{4}-\frac{1}{1+z_{a}}-\frac{\log z_{a}}{4}\right]\right),\\ &\tilde{A}_{2}(\vec{w},\vec{z};t)=\prod_{a=1}^{k}\left(\sum_{m\in\mathbb{Z}}\frac{[-\log\tau]^{-1}\pi\cdot e^{2m\pi\mathsf{i}[(1/4)t+(1/2)(\lfloor t^{2/3}\alpha\rfloor-1)-t^{1/3}\tilde{r}]}}{\sin(-\pi[\log\tau]^{-1}[\log z_{a}-\log w_{a}-2m\pi\mathsf{i}])}\right),\\ &\tilde{A}_{3}(\vec{w},\vec{z};t)=\prod_{a=1}^{k}\left(\frac{(1+z_{a})e^{-\log z_{a}/2}}{(1+w_{a})e^{-\log w_{a}/2}}\right)^{\lfloor t^{2/3}\alpha\rfloor-1},\\ &\tilde{A}_{4}(\vec{w},\vec{z};t)=\prod_{a=1}^{k}\frac{(-w_{a};\tau)_{\infty}(z_{a}^{2};\tau)_{\infty}}{(-z_{a};\tau)_{\infty}(z_{a}w_{a};\tau)_{\infty}\cdot z_{a}}\cdot\exp\left(t^{1/3}\tilde{r}[\log z_{a}-\log w_{a}]\right),\end{split}

In equations (5.5), and (5.4) we take the principal branch of the logarithm everywhere.

In the remainder of this section, we estimate the functions that appear in A~i(w,z;t)\tilde{A}_{i}(\vec{w},\vec{z};t), B~i(w,z;t)\tilde{B}_{i}(\vec{w},\vec{z};t) over the contours Cw,CzC_{w},C_{z} in a sequence of lemmas.

Lemma 5.3.

Let τ,Cw,Cz\tau,C_{w},C_{z} be as in Definition 5.2. For wCww\in C_{w}, and zCzz\in C_{z} we have

(5.6) |exp(11+w+logw411+zlogz4)|τ1/4.\left|\exp\left(\frac{1}{1+w}+\frac{\log w}{4}-\frac{1}{1+z}-\frac{\log z}{4}\right)\right|\leq\tau^{-1/4}.
Lemma 5.4.

Let τ,Cw,Cz\tau,C_{w},C_{z} be as in Definition 5.2. There exists a constant Bτ,2>0B_{\tau,2}>0, depending on τ\tau alone, such that for all t>0t>0, α,r~\alpha,\tilde{r}\in\mathbb{R}, wCww\in C_{w}, and zCzz\in C_{z} we have

(5.7) |m[logτ]1πe2mπ𝗂[(1/4)t+(1/2)(t2/3α1)t1/3r~]sin(π[logτ]1[logzlogw2mπ𝗂])|Bτ,2.\left|\sum_{m\in\mathbb{Z}}\frac{[-\log\tau]^{-1}\pi\cdot e^{2m\pi\mathsf{i}[(1/4)t+(1/2)(\lfloor t^{2/3}\alpha\rfloor-1)-t^{1/3}\tilde{r}]}}{\sin(-\pi[\log\tau]^{-1}[\log z-\log w-2m\pi\mathsf{i}])}\right|\leq B_{\tau,2}.
Lemma 5.5.

Let τ,Cw,Cz\tau,C_{w},C_{z} be as in Definition 5.2. There exists a constant bτ,3>0b_{\tau,3}>0, depending on τ\tau alone, such that for all t1t\geq 1, α\alpha\in\mathbb{R}, wCww\in C_{w}, and zCzz\in C_{z} we have

(5.8) |((1+z)elogz/2(1+w)elogw/2)t2/3α1|exp(bτ,3t2/3(1+|α|)).\left|\left(\frac{(1+z)e^{-\log z/2}}{(1+w)e^{-\log w/2}}\right)^{\lfloor t^{2/3}\alpha\rfloor-1}\right|\leq\exp\left(b_{\tau,3}\cdot t^{2/3}\cdot(1+|\alpha|)\right).
Lemma 5.6.

Let τ,Cw,Cz\tau,C_{w},C_{z} be as in Definition 5.2. There exists a constant bτ,4>0b_{\tau,4}>0, depending on τ\tau alone, such that for all t1t\geq 1, r~\tilde{r}\in\mathbb{R}, wCww\in C_{w}, and zCzz\in C_{z} we have

(5.9) |(w;τ)(z2;τ)(z;τ)(zw;τ)zexp(t1/3r~[logzlogw])|exp(bτ,4t1/3(1+|r~|)).\left|\frac{(-w;\tau)_{\infty}(z^{2};\tau)_{\infty}}{(-z;\tau)_{\infty}(zw;\tau)_{\infty}\cdot z}\cdot\exp\left(t^{1/3}\tilde{r}[\log z-\log w]\right)\right|\leq\exp\left(b_{\tau,4}\cdot t^{1/3}\cdot(1+|\tilde{r}|)\right).
Lemma 5.7.

Let τ,Cw,Cz\tau,C_{w},C_{z} be as in Definition 5.2, and fix kk\in\mathbb{N}. Suppose that t>0t>0, waCww_{a}\in C_{w}, zaCzz_{a}\in C_{z} for a=1,,ka=1,\dots,k. Then, we have

(5.10) |B~1(w,z;t)|kk/2τ3k/8[2τ3/8]k2.\left|\tilde{B}_{1}(\vec{w},\vec{z};t)\right|\leq k^{k/2}\cdot\tau^{-3k/8}\cdot[2\tau^{3/8}]^{k^{2}}.
Lemma 5.8.

Let τ,Cw,Cz\tau,C_{w},C_{z} be as in Definition 5.2. We can find a universal constant B6>0B_{6}>0, such that for t>0t>0, kk\in\mathbb{N}, waCww_{a}\in C_{w}, zaCzz_{a}\in C_{z} for a=1,,ka=1,\dots,k

(5.11) |B~2(w,z;t)|B6k2.\left|\tilde{B}_{2}(\vec{w},\vec{z};t)\right|\leq B_{6}^{k^{2}}.

5.2. Proof of Proposition 3.3: Part I

In this section we specify the choice of τ0\tau_{0}, AA and TT as in the statement of Proposition 3.3 and prove (3.3) when 1kt1\leq k\leq t.

Let 𝗋\mathsf{r} be as in Lemma 4.3 and 𝖼\mathsf{c} as in Lemma 5.1. We assume that τ0(0,1)\tau_{0}\in(0,1) is sufficiently small so that the following all hold for τ(0,τ0]\tau\in(0,\tau_{0}]

(5.12) τe8π,𝖼τa𝗋,1/2,64πA(τ)𝗋2a𝗋,1,4τ1/8B61,\begin{split}\tau\leq e^{-8\pi},\hskip 5.69054pt\mathsf{c}\tau\leq a_{\mathsf{r},1}/2,\hskip 5.69054pt64\pi A(\tau)\leq\mathsf{r}^{2}a_{\mathsf{r},1},\hskip 5.69054pt4\tau^{1/8}B_{6}\leq 1,\end{split}

where a𝗋,1a_{\mathsf{r},1} as in Lemma 4.5, A()A(\cdot) is as in (5.2), and B6B_{6} is as in Lemma 5.8. This specifies τ0\tau_{0} in the statement of the proposition. Below we fix τ(0,τ0]\tau\in(0,\tau_{0}].

We let A1A\geq 1 be sufficiently large so that for t1t\geq 1 we have

(5.13) A𝗋,1A𝗋,2A𝗋,4A𝗋,5Γu,𝗋,texp((a𝗋,3+1)|u|2+|r~||u|(a𝗋,1/4)|u|3)|du|×Γv,𝗋,t|v|exp((a𝗋,3+1)|v|2+|r~||v|(a𝗋,1/4)|v|3)|dv|A,\begin{split}&A_{\mathsf{r},1}A_{\mathsf{r},2}A_{\mathsf{r},4}A_{\mathsf{r},5}\int_{\Gamma_{u,\mathsf{r},t}}\exp\left((a_{\mathsf{r},3}+1)|u|^{2}+|\tilde{r}||u|-(a_{\mathsf{r},1}/4)|u|^{3}\right)|du|\\ &\times\int_{\Gamma_{v,\mathsf{r},t}}|v|\exp\left((a_{\mathsf{r},3}+1)|v|^{2}+|\tilde{r}||v|-(a_{\mathsf{r},1}/4)|v|^{3}\right)|dv|\leq A,\end{split}

where 𝗋\mathsf{r} is as in Lemma 4.3, Γu,𝗋,t\Gamma_{u,\mathsf{r},t} and Γv,𝗋,t\Gamma_{v,\mathsf{r},t} are as in Definition 4.2, A𝗋,1,a𝗋,1A_{\mathsf{r},1},a_{\mathsf{r},1} is as in Lemma 4.5, A𝗋,2A_{\mathsf{r},2} is as in Lemma 4.6, a𝗋,3a_{\mathsf{r},3} is as in Lemma 4.7, A𝗋,4A_{\mathsf{r},4} is as in Lemma 4.8, A𝗋,5A_{\mathsf{r},5} is as in Lemma 4.9, and r~\tilde{r} is as in the statement of the proposition. In (5.13) the integrals are with respect to arc-length. This specifies AA in the statement of the proposition.

We now fix T1T\geq 1 sufficiently large so that for tTt\geq T we have

(5.14) tmax1i5T𝗋,i,t1/3𝗋,t2/3A(e8π)1/2, and for kT we have kk/2k!ek2k2Bτ,2kexp(k5/3bτ,3(1+|α|)+k4/3bτ,4(1+|r~|))kk/2,\begin{split}&t\geq\max_{1\leq i\leq 5}T_{\mathsf{r},i},\hskip 5.69054ptt^{-1/3}\leq\mathsf{r},\hskip 5.69054ptt^{-2/3}A(e^{-8\pi})\leq 1/2,\mbox{ and for $k\geq T$ we have }\\ &\frac{k^{k/2}}{k!}\cdot e^{k}\cdot 2^{-k^{2}}\cdot B_{\tau,2}^{k}\cdot\exp\left(k^{5/3}b_{\tau,3}(1+|\alpha|)+k^{4/3}b_{\tau,4}(1+|\tilde{r}|)\right)\leq k^{-k/2},\end{split}

where T𝗋,iT_{\mathsf{r},i} for i=1,,5i=1,\dots,5 are as in Lemmas 4.5 - 4.9 (here τ1=e8π\tau_{1}=e^{-8\pi} and so e20𝗋e10τ11e^{20\mathsf{r}}\leq e^{10}\leq\tau_{1}^{-1}). In addition, A()A(\cdot) is as in (5.2), Bτ,2B_{\tau,2} is as in Lemma 5.4, bτ,3b_{\tau,3} is as in Lemma 5.5, bτ,4b_{\tau,4} is as in Lemma 5.6, α,r~\alpha,\tilde{r} are as in the statement of the proposition. This specifies TT in the statement of the proposition.

In the remainder of this section we assume tTt\geq T, τ(0,τ0]\tau\in(0,\tau_{0}], 1kt1\leq k\leq t and proceed to prove (3.3). It follows from (4.5) that

(5.15) |I(k,t)|1k!(2π)2kΓu,𝗋,tk|du|Γv,𝗋,tk|dv||i=15Ai(u,v;t)i=12Bi(u,v;t)|,\begin{split}\left|I(k,t)\right|\leq\frac{1}{k!(2\pi)^{2k}}\int_{\Gamma^{k}_{u,\mathsf{r},t}}|d\vec{u}|\int_{\Gamma^{k}_{v,\mathsf{r},t}}|d\vec{v}|&\left|\prod_{i=1}^{5}A_{i}(\vec{u},\vec{v};t)\prod_{i=1}^{2}B_{i}(\vec{u},\vec{v};t)\right|,\end{split}

where Ai(u,v;t)A_{i}(\vec{u},\vec{v};t) for i=1,,5i=1,\dots,5 are as in (4.6), Bj(u,v;t)B_{j}(\vec{u},\vec{v};t) for j=1,2j=1,2 are as in (4.7), |du|=|du1||duk||d\vec{u}|=|du_{1}|\cdots|du_{k}|, and |dv|=|dv1||dvk||d\vec{v}|=|dv_{1}|\cdots|dv_{k}|. We mention that in obtaining (5.15) we implicitly used that t1/3𝗋t^{-1/3}\leq\mathsf{r}, as follows from the second inequality in (5.14), and that e20𝗋e10τ1e^{20\mathsf{r}}\leq e^{10}\leq\tau^{-1} as follows from the first inequality in (5.12).

It follows from Lemmas 4.5-4.9 that if uΓu,𝗋,tk\vec{u}\in\Gamma^{k}_{u,\mathsf{r},t} and vΓv,𝗋,tk\vec{v}\in\Gamma^{k}_{v,\mathsf{r},t} we have

|i=15Ai(u,v;t)B1(u,v;t)|A𝗋,1kA𝗋,2kA𝗋,4kA𝗋,5kkk/2a=1k|va|×exp(a=1ka𝗋,3|ua|2+|r~||ua|a𝗋,1|ua|3+a𝗋,3|va|2+|r~||va|a𝗋,1|va|3).\begin{split}&\left|\prod_{i=1}^{5}A_{i}(\vec{u},\vec{v};t)\cdot B_{1}(\vec{u},\vec{v};t)\right|\leq A_{\mathsf{r},1}^{k}A_{\mathsf{r},2}^{k}A_{\mathsf{r},4}^{k}A_{\mathsf{r},5}^{k}\cdot k^{k/2}\cdot\prod_{a=1}^{k}|v_{a}|\\ &\times\exp\left(\sum_{a=1}^{k}a_{\mathsf{r},3}|u_{a}|^{2}+|\tilde{r}||u_{a}|-a_{\mathsf{r},1}|u_{a}|^{3}+a_{\mathsf{r},3}|v_{a}|^{2}+|\tilde{r}||v_{a}|-a_{\mathsf{r},1}|v_{a}|^{3}\right).\end{split}

In addition, from Lemma 5.1 we have

|B2(u,v;t)|exp(A(τ)t2/3𝖱𝖾[(a=1k(uava))2])exp(A(τ)t2/3𝖱𝖾[a=1k(uava)2])a=1kexp(𝖼τkt1a=1k(|ua|3+|va|3)).\begin{split}&\left|B_{2}(\vec{u},\vec{v};t)\right|\leq\exp\left(-A(\tau)t^{-2/3}\cdot\mathsf{Re}\left[\left(\sum_{a=1}^{k}(u_{a}-v_{a})\right)^{2}\right]\right)\cdot\\ &\exp\left(A(\tau)t^{-2/3}\cdot\mathsf{Re}\left[\sum_{a=1}^{k}(u_{a}-v_{a})^{2}\right]\right)\cdot\prod_{a=1}^{k}\exp\left(\mathsf{c}\tau kt^{-1}\cdot\sum_{a=1}^{k}(|u_{a}|^{3}+|v_{a}|^{3})\right).\end{split}

We next observe that our assumptions give the inequalities

A(τ)t2/31/2,𝖼τa𝗋,1/2,kt,|𝖱𝖾[z1z2]2|2(|z1|2+|z2|2),A(\tau)t^{-2/3}\leq 1/2,\hskip 5.69054pt\mathsf{c}\tau\leq a_{\mathsf{r},1}/2,\hskip 5.69054ptk\leq t,\hskip 5.69054pt|\mathsf{Re}[z_{1}-z_{2}]^{2}|\leq 2(|z_{1}|^{2}+|z_{2}|^{2}),

where the first inequality used the first inequality in (5.12) and the third inequality in (5.14).

Combining the last three inequalities, we get

(5.16) |i=15Ai(u,v;t)i=12B2(u,v;t)|A𝗋,1kA𝗋,2kA𝗋,4kA𝗋,5kkk/2a=1k|va|×exp(a=1k(a𝗋,3+1)(|ua|2+|va|2)+|r~|(|ua|+|va|)(a𝗋,1/2)(|ua|3+|va|3))×exp(A(τ)t2/3𝖱𝖾[(a=1k(uava))2]).\begin{split}&\left|\prod_{i=1}^{5}A_{i}(\vec{u},\vec{v};t)\prod_{i=1}^{2}B_{2}(\vec{u},\vec{v};t)\right|\leq A_{\mathsf{r},1}^{k}A_{\mathsf{r},2}^{k}A_{\mathsf{r},4}^{k}A_{\mathsf{r},5}^{k}\cdot k^{k/2}\cdot\prod_{a=1}^{k}|v_{a}|\\ &\times\exp\left(\sum_{a=1}^{k}(a_{\mathsf{r},3}+1)(|u_{a}|^{2}+|v_{a}|^{2})+|\tilde{r}|(|u_{a}|+|v_{a}|)-(a_{\mathsf{r},1}/2)(|u_{a}|^{3}+|v_{a}|^{3})\right)\\ &\times\exp\left(-A(\tau)t^{-2/3}\cdot\mathsf{Re}\left[\left(\sum_{a=1}^{k}(u_{a}-v_{a})\right)^{2}\right]\right).\end{split}

We next claim that

(5.17) A(τ)t2/3𝖱𝖾[(a=1k(uava))2]a𝗋,14a=1k(|ua|3+|va|3).\begin{split}-A(\tau)t^{-2/3}\cdot\mathsf{Re}\left[\left(\sum_{a=1}^{k}(u_{a}-v_{a})\right)^{2}\right]\leq\frac{a_{\mathsf{r},1}}{4}\sum_{a=1}^{k}(|u_{a}|^{3}+|v_{a}|^{3}).\end{split}

To see why (5.17) holds let us denote by SS the set indices ss in {1,,k}\{1,\dots,k\} such that usΓu,𝗋,t0u_{s}\in\Gamma_{u,\mathsf{r},t}^{0} and vsΓv,𝗋,t0v_{s}\in\Gamma_{v,\mathsf{r},t}^{0}, where we recall that Γu,𝗋,t0,Γv,𝗋,t0\Gamma_{u,\mathsf{r},t}^{0},\Gamma_{v,\mathsf{r},t}^{0} were defined in Definition 4.2. We also set Sc={1,,k}SS^{c}=\{1,\dots,k\}\setminus S. Using that |t1/3ua|2π|t^{-1/3}u_{a}|\leq 2\pi and |t1/3va|2π|t^{-1/3}v_{a}|\leq 2\pi, we see that

t2/3𝖱𝖾[(a=1k(uava))2]t2/3𝖱𝖾[(aS(uava))2]+8kπt1/3aSc(|ua|+|va|).\begin{split}-t^{-2/3}\cdot\mathsf{Re}\left[\left(\sum_{a=1}^{k}(u_{a}-v_{a})\right)^{2}\right]\leq-t^{-2/3}\cdot\mathsf{Re}\left[\left(\sum_{a\in S}(u_{a}-v_{a})\right)^{2}\right]+8k\pi t^{-1/3}\sum_{a\in S^{c}}\left(|u_{a}|+|v_{a}|\right).\end{split}

We also have for aSca\in S^{c} that |ua|𝗋t1/3|u_{a}|\geq\mathsf{r}\cdot t^{1/3} or |va|𝗋t1/3|v_{a}|\geq\mathsf{r}\cdot t^{1/3} (or both), which implies that for aSca\in S^{c}

8A(τ)kπt1/3(|ua|+|va|)8A(τ)πt2/3(|ua|+|va|)a1,𝗋4(|ua|3+|va|3).\begin{split}8A(\tau)k\pi t^{-1/3}(|u_{a}|+|v_{a}|)\leq 8A(\tau)\pi t^{2/3}(|u_{a}|+|v_{a}|)\leq\frac{a_{1,\mathsf{r}}}{4}(|u_{a}|^{3}+|v_{a}|^{3}).\end{split}

We remark that in the first inequality we used that ktk\leq t and in the second one we used the third inequality in (5.12) and the fact that 2(x3+y3)(x+y)(x2+y2)2(x^{3}+y^{3})\geq(x+y)(x^{2}+y^{2}) for x,y0x,y\geq 0.

Finally, we note that for aSa\in S we have ua=xa+𝗂ϵaxau_{a}=x_{a}+\mathsf{i}\epsilon_{a}x_{a} and va=1ya+𝗂δayav_{a}=-1-y_{a}+\mathsf{i}\delta_{a}y_{a} where xa,ya0x_{a},y_{a}\geq 0 and ϵa,δa{1,1}\epsilon_{a},\delta_{a}\in\{-1,1\}. The latter implies that

𝖱𝖾[(aS(uava))2]=(aS(xa+ya+1))2(aS(ϵaxaδaya))20.\mathsf{Re}\left[\left(\sum_{a\in S}(u_{a}-v_{a})\right)^{2}\right]=\left(\sum_{a\in S}\left(x_{a}+y_{a}+1\right)\right)^{2}-\left(\sum_{a\in S}\left(\epsilon_{a}x_{a}-\delta_{a}y_{a}\right)\right)^{2}\geq 0.

Combining the last three inequalities, we deduce (5.17).

We now combine (5.16) and (5.17) to conclude that

(5.18) |i=15Ai(u,v;t)i=12B2(u,v;t)|A𝗋,1kA𝗋,2kA𝗋,4kA𝗋,5kkk/2a=1k|va|×exp(a=1k(a𝗋,3+1)(|ua|2+|va|2)+|r~|(|ua|+|va|)(a𝗋,1/4)(|ua|3+|va|3)).\begin{split}&\left|\prod_{i=1}^{5}A_{i}(\vec{u},\vec{v};t)\prod_{i=1}^{2}B_{2}(\vec{u},\vec{v};t)\right|\leq A_{\mathsf{r},1}^{k}A_{\mathsf{r},2}^{k}A_{\mathsf{r},4}^{k}A_{\mathsf{r},5}^{k}\cdot k^{k/2}\cdot\prod_{a=1}^{k}|v_{a}|\\ &\times\exp\left(\sum_{a=1}^{k}(a_{\mathsf{r},3}+1)(|u_{a}|^{2}+|v_{a}|^{2})+|\tilde{r}|(|u_{a}|+|v_{a}|)-(a_{\mathsf{r},1}/4)(|u_{a}|^{3}+|v_{a}|^{3})\right).\end{split}

From (5.13), (5.15) and (5.18) we conclude that

|I(k,t)|kk/2k!(2π)2kAkkk/2kkek(2π)2kAkkk/2Ak,\left|I(k,t)\right|\leq\frac{k^{k/2}}{k!(2\pi)^{2k}}\cdot A^{k}\leq\frac{k^{k/2}}{k^{k}e^{-k}(2\pi)^{2k}}\cdot A^{k}\leq k^{-k/2}\cdot A^{k},

where in the middle inequality we used that k!kkekk!\geq k^{k}e^{-k}, see (2.9). The last inequality proves (3.3) when 1kt1\leq k\leq t.

5.3. Proof of Proposition 3.3: Part II

In this section we assume ktTk\geq t\geq T, τ(0,τ0]\tau\in(0,\tau_{0}], and proceed to prove (3.3). It follows from (5.3) that

(5.19) |I(k,t)|1k!(2π)2kCwk|dw|Czk|dz||i=14A~i(w,z;t)i=12B~i(w,z;t)|,\begin{split}\left|I(k,t)\right|\leq\frac{1}{k!(2\pi)^{2k}}\int_{C_{w}^{k}}|d\vec{w}|\int_{C_{z}^{k}}|d\vec{z}|&\left|\prod_{i=1}^{4}\tilde{A}_{i}(\vec{w},\vec{z};t)\prod_{i=1}^{2}\tilde{B}_{i}(\vec{w},\vec{z};t)\right|,\end{split}

where A~i(w,z;t)\tilde{A}_{i}(\vec{w},\vec{z};t) for i=1,,4i=1,\dots,4 are as in (5.5), B~j(w,z;t)\tilde{B}_{j}(\vec{w},\vec{z};t) for j=1,2j=1,2 are as in (5.4), |dw|=|dw1||dwk||d\vec{w}|=|dw_{1}|\cdots|dw_{k}|, and |dz|=|dz1||dzk||d\vec{z}|=|dz_{1}|\cdots|dz_{k}| and |dzi||dz_{i}|, |dwi||dw_{i}| denote integration with respect to arc-length. We mention that in obtaining (5.19) we implicitly used that τe8π\tau\leq e^{-8\pi} as follows from the first inequality in (5.12).

From Lemmas 5.3-5.8, and the fact that ktk\geq t, we have for wCwk\vec{w}\in C_{w}^{k} and zCzk\vec{z}\in C_{z}^{k} that

|i=14A~i(w,z;t)i=12B~i(w,z;t)|τk2/4Bτ,2kexp(k5/3bτ,3(1+|α|)+k4/3bτ,4(1+|r~|))×kk/2τ3k/8[2τ3/8]k2B6k2.\begin{split}\left|\prod_{i=1}^{4}\tilde{A}_{i}(\vec{w},\vec{z};t)\prod_{i=1}^{2}\tilde{B}_{i}(\vec{w},\vec{z};t)\right|\leq\hskip 5.69054pt&\tau^{-k^{2}/4}\cdot B_{\tau,2}^{k}\cdot\exp\left(k^{5/3}b_{\tau,3}(1+|\alpha|)+k^{4/3}b_{\tau,4}(1+|\tilde{r}|)\right)\\ &\times k^{k/2}\cdot\tau^{-3k/8}\cdot[2\tau^{3/8}]^{k^{2}}\cdot B_{6}^{k^{2}}.\end{split}

We mention that in applying Lemmas 5.3-5.8 we used that 0<τe8π0<\tau\leq e^{-8\pi}, and t1t\geq 1. Combining the latter with (5.19), the fourth inequality in (5.12), and the fact that the length of CwC_{w} is 2πe2\pi e and that of CzC_{z} is 2πτ3/82\pi\tau^{3/8}, we conclude that

(5.20) |I(k,t)|kk/2k!ek2k2Bτ,2kexp(k5/3bτ,3(1+|α|)+k4/3bτ,4(1+|r~|)).\begin{split}&\left|I(k,t)\right|\leq\frac{k^{k/2}}{k!}\cdot e^{k}\cdot 2^{-k^{2}}\cdot B_{\tau,2}^{k}\cdot\exp\left(k^{5/3}b_{\tau,3}(1+|\alpha|)+k^{4/3}b_{\tau,4}(1+|\tilde{r}|)\right).\end{split}

In view of (5.20) and the second line in (5.14), we conclude that |I(k,t)|kk/2Akkk/2\left|I(k,t)\right|\leq k^{-k/2}\leq A^{k}k^{-k/2} as A1A\geq 1 by construction. This proves (3.3) when ktk\geq t.

5.4. Proof of the lemmas from Section 5.1

In this section we present the proofs of the seven lemmas from Section 5.1.

Proof of Lemma 5.1.

We first note that if λ[0,e8π]\lambda\in[0,e^{-8\pi}] and |z|4π|z|\leq 4\pi then

log(1λez)=k=1λkekzk=k=1λkkm=0zmkmm!=m=0zmm!k=1km1λk,\log\left(1-\lambda e^{z}\right)=-\sum_{k=1}^{\infty}\frac{\lambda^{k}e^{kz}}{k}=-\sum_{k=1}^{\infty}\frac{\lambda^{k}}{k}\sum_{m=0}^{\infty}\frac{z^{m}k^{m}}{m!}=-\sum_{m=0}^{\infty}\frac{z^{m}}{m!}\sum_{k=1}^{\infty}k^{m-1}\lambda^{k},

where in the last line the order of the sums can be exchanged by Fubini’s theorem as

k=1m=0λkk|z|mkmm!k=1λkek|z|k<.\sum_{k=1}^{\infty}\sum_{m=0}^{\infty}\frac{\lambda^{k}}{k}\cdot\frac{|z|^{m}k^{m}}{m!}\leq\sum_{k=1}^{\infty}\frac{\lambda^{k}e^{k|z|}}{k}<\infty.

In particular, if we set

Am(λ)=1m!k=1km1λk1 and Bm=1m!k=1km1e8π(k1)A_{m}(\lambda)=\frac{1}{m!}\sum_{k=1}^{\infty}k^{m-1}\lambda^{k-1}\mbox{ and }B_{m}=\frac{1}{m!}\sum_{k=1}^{\infty}k^{m-1}e^{-8\pi(k-1)}

we see that for λ[0,e8π]\lambda\in[0,e^{-8\pi}] and |z|4π|z|\leq 4\pi we have

(5.21) 0Am(λ)Bm,B:=m=0Bm(4π)m<, and log(1λez)=λm=0Am(λ)zm,0\leq A_{m}(\lambda)\leq B_{m},\hskip 5.69054ptB:=\sum_{m=0}^{\infty}B_{m}\cdot(4\pi)^{m}<\infty\mbox{, and }\log\left(1-\lambda e^{z}\right)=-\lambda\cdot\sum_{m=0}^{\infty}A_{m}(\lambda)z^{m},

and the latter is an absolutely convergent in the closed disc of radius 4π4\pi.

We next observe that if z{Wa+Wb,Za+Zb,Wa+Zb,Za+Wb}z\in\{W_{a}+W_{b},Z_{a}+Z_{b},W_{a}+Z_{b},Z_{a}+W_{b}\} we have |z|4π|z|\leq 4\pi and so from (5.21) we conclude for any λ[0,e8π]\lambda\in[0,e^{-8\pi}] that

|(1λeWa+Wb)(1λeZa+Zb)(1λeWa+Zb)(1λeZa+Wb)|=exp(λm=2Am(λ)𝖱𝖾[(Wa+Wb)m+(Za+Zb)m(Wa+Zb)m(Za+Wb)m]).\begin{split}&\left|\frac{(1-\lambda e^{W_{a}+W_{b}})(1-\lambda e^{Z_{a}+Z_{b}})}{(1-\lambda e^{W_{a}+Z_{b}})(1-\lambda e^{Z_{a}+W_{b}})}\right|\\ &=\exp\Big{(}-\lambda\sum_{m=2}^{\infty}A_{m}(\lambda)\mathsf{Re}\left[(W_{a}+W_{b})^{m}+(Z_{a}+Z_{b})^{m}-(W_{a}+Z_{b})^{m}-(Z_{a}+W_{b})^{m}\right]\Big{)}.\end{split}

In particular, the last equation and (5.21) imply

(5.22) |(1λeWa+Wb)(1λeZa+Zb)(1λeWa+Zb)(1λeZa+Wb)|exp(2λA2(λ)𝖱𝖾[(WaZa)(WbZb)])×exp(λB(|Wa|3+|Wb|3+|Za|3+|Zb|3)).\begin{split}\left|\frac{(1-\lambda e^{W_{a}+W_{b}})(1-\lambda e^{Z_{a}+Z_{b}})}{(1-\lambda e^{W_{a}+Z_{b}})(1-\lambda e^{Z_{a}+W_{b}})}\right|\leq\hskip 5.69054pt&\exp\left(-2\lambda A_{2}(\lambda)\cdot\mathsf{Re}[(W_{a}-Z_{a})(W_{b}-Z_{b})]\right)\\ &\times\exp\left(\lambda B\cdot(|W_{a}|^{3}+|W_{b}|^{3}+|Z_{a}|^{3}+|Z_{b}|^{3})\right).\end{split}

Taking a product of (5.22) over λ=τ,τ2,\lambda=\tau,\tau^{2},\cdots as well as 1a<bk1\leq a<b\leq k we see that

|1a<bk(τeWa+Wb;τ)(τeZa+Zb;τ)(τeWa+Zb;τ)(τeZa+Wb;τ)|exp(2A(τ)𝖱𝖾[1a<bk(WaZa)(WbZb)])×exp(kτB1τ(a=1k|Wa|3+|Za|3)),\begin{split}&\left|\prod_{1\leq a<b\leq k}\frac{(\tau e^{W_{a}+W_{b}};\tau)_{\infty}(\tau e^{Z_{a}+Z_{b}};\tau)_{\infty}}{(\tau e^{W_{a}+Z_{b}};\tau)_{\infty}(\tau e^{Z_{a}+W_{b}};\tau)_{\infty}}\right|\leq\exp\left(-2A(\tau)\cdot\mathsf{Re}\left[\sum_{1\leq a<b\leq k}(W_{a}-Z_{a})(W_{b}-Z_{b})\right]\right)\\ &\times\exp\left(\frac{k\cdot\tau B}{1-\tau}\cdot\left(\sum_{a=1}^{k}|W_{a}|^{3}+|Z_{a}|^{3}\right)\right),\end{split}

where A(τ)A(\tau) is as in (5.2). The last inequality implies (5.1) with 𝖼=B/(1e8π)\mathsf{c}=B/(1-e^{-8\pi}) once we use

21a<bk(WaZa)(WbZb)+a=1k(WaZa)2=(a=1k(WaZa))2.2\sum_{1\leq a<b\leq k}(W_{a}-Z_{a})(W_{b}-Z_{b})+\sum_{a=1}^{k}(W_{a}-Z_{a})^{2}=\left(\sum_{a=1}^{k}(W_{a}-Z_{a})\right)^{2}.

Proof of Lemma 5.3.

Notice that as uu varies over {x+𝗂y:x=1,y[π,π]}\{x+\mathsf{i}y\in\mathbb{C}:x=1,y\in[-\pi,\pi]\}, the variable w=euw=e^{u} covers CwC_{w}. Similarly, as vv varies over {x+𝗂y:x=(3/4)logτ,y[π,π]}\{x+\mathsf{i}y\in\mathbb{C}:x=(3/4)\log\tau,y\in[-\pi,\pi]\}, the variable z=evz=e^{v} covers CzC_{z}. In particular, we see that to prove (5.6) it suffices to show that

(5.23) 𝖱𝖾[F(u)]𝖱𝖾[F(v)](1/4)logτ,\mathsf{Re}[F(u)]-\mathsf{Re}[F(v)]\leq(-1/4)\log\tau,

where FF is as in (4.2). From (4.4) we conclude that

𝖱𝖾[F(u)]𝖱𝖾[F(v)]F(1)F((3/4)logτ)=11+e+1411+τ3/4(3/16)logτ,\mathsf{Re}[F(u)]-\mathsf{Re}[F(v)]\leq F(1)-F((3/4)\log\tau)=\frac{1}{1+e}+\frac{1}{4}-\frac{1}{1+\tau^{3/4}}-(3/16)\log\tau,

which implies (5.23) since τe8π\tau\leq e^{-8\pi}. ∎

Proof of Lemma 5.4.

Notice that for wCww\in C_{w} and zCzz\in C_{z} we have logw=1+𝗂x\log w=1+\mathsf{i}x, and logz=(3/4)logτ+𝗂y\log z=(3/4)\log\tau+\mathsf{i}y for some x,y[π,π]x,y\in[-\pi,\pi]. The latter, the fact that |e𝗂h|=1|e^{\mathsf{i}h}|=1 for hh\in\mathbb{R}, and [Dim18, Equation (6.14)] together imply

|[logτ]1πe2mπ𝗂[(1/4)t+(1/2)(t2/3α1)t1/3r~]sin(π[logτ]1[logzlogw2mπ𝗂])|=[logτ]1π|sin(π[logτ]1[logzlogw2mπ𝗂])|[logτ]1πexp(π[logτ]1|yx+2πm|)|sin(3π/22π[logτ]1)|2π[logτ]1τ2π2τ2π2|m|.\begin{split}&\left|\frac{[-\log\tau]^{-1}\pi\cdot e^{2m\pi\mathsf{i}[(1/4)t+(1/2)(\lfloor t^{2/3}\alpha\rfloor-1)-t^{1/3}\tilde{r}]}}{\sin(-\pi[\log\tau]^{-1}[\log z-\log w-2m\pi\mathsf{i}])}\right|=\frac{[-\log\tau]^{-1}\pi}{|\sin(-\pi[\log\tau]^{-1}[\log z-\log w-2m\pi\mathsf{i}])|}\\ &\leq\frac{[-\log\tau]^{-1}\pi\exp(-\pi[-\log\tau]^{-1}|y-x+2\pi m|)}{|\sin(3\pi/2-2\pi[-\log\tau]^{-1})|}\leq 2\pi[-\log\tau]^{-1}\tau^{-2\pi^{2}}\cdot\tau^{2\pi^{2}|m|}.\end{split}

We mention that in the last inequality we used that τe8π\tau\leq e^{-8\pi} and so |sin(3π/22π[logτ]1)|1/2|\sin(3\pi/2-2\pi[-\log\tau]^{-1})|\geq 1/2, and also that x,y[π,π]x,y\in[-\pi,\pi]. Summing over mm\in\mathbb{Z}, we see that (5.7) holds with Bτ,2=2π[logτ]1τ2π2mτ2π2|m|.B_{\tau,2}=2\pi[-\log\tau]^{-1}\tau^{-2\pi^{2}}\cdot\sum_{m\in\mathbb{Z}}\tau^{2\pi^{2}|m|}.

Proof of Lemma 5.5.

Using that τ(0,e8π],\tau\in(0,e^{-8\pi}], |z|=τ3/4|z|=\tau^{3/4} and |w|=e|w|=e, we see that

1(1τ3/4)τ3/8e1/2e+1|1+z|τ3/8e1/2|1+w|=|(1+z)elogz/2(1+w)elogw/2|(1+τ3/4)τ3/8e1/2e1τ1/2.1\leq\frac{(1-\tau^{3/4})\tau^{-3/8}e^{1/2}}{e+1}\leq\frac{|1+z|\tau^{-3/8}e^{1/2}}{|1+w|}=\left|\frac{(1+z)e^{-\log z/2}}{(1+w)e^{-\log w/2}}\right|\leq\frac{(1+\tau^{3/4})\tau^{-3/8}e^{1/2}}{e-1}\leq\tau^{-1/2}.

Combining the latter with the fact that for t1t\geq 1 we have |t2/3α1|2t2/3(1+|α|)|\lfloor t^{2/3}\alpha\rfloor-1|\leq 2t^{2/3}(1+|\alpha|) , we conclude (5.8) with bτ,3=logτb_{\tau,3}=-\log\tau. ∎

Proof of Lemma 5.6.

Using that τ(0,e8π],\tau\in(0,e^{-8\pi}], |z|=τ3/4|z|=\tau^{3/4} and |w|=e|w|=e, we see that

(5.24) |exp(t1/3r~[logzlogw])|=exp(t1/3r~[(3/4)logτ1])exp(t1/3|r~|logτ).\left|\exp\left(t^{1/3}\tilde{r}[\log z-\log w]\right)\right|=\exp\left(t^{1/3}\tilde{r}[(3/4)\log\tau-1]\right)\leq\exp\left(-t^{1/3}|\tilde{r}|\log\tau\right).

In addition, from (4.35) and τ(0,e8π],\tau\in(0,e^{-8\pi}], we have

(5.25) |(w;τ)(z2;τ)(z;τ)(zw;τ)z|(e;τ)(τ3/2;τ)(τ3/4;τ)(eτ3/4;τ)τ3/4=:Bτ,4.\left|\frac{(-w;\tau)_{\infty}(z^{2};\tau)_{\infty}}{(-z;\tau)_{\infty}(zw;\tau)_{\infty}\cdot z}\right|\leq\frac{(-e;\tau)_{\infty}(-\tau^{3/2};\tau)_{\infty}}{(\tau^{3/4};\tau)_{\infty}(e\tau^{3/4};\tau)_{\infty}\cdot\tau^{3/4}}=:B_{\tau,4}.

Equations (5.24), (5.25), and the fact that t1t\geq 1 imply (5.9) with bτ,4=logτ+logBτ,4b_{\tau,4}=-\log\tau+\log B_{\tau,4}. ∎

Proof of Lemma 5.7.

From [Dim20, Lemma 8.5] with R=eR=e, r=τ3/4r=\tau^{3/4}, α=e1τ3/8\alpha=e^{-1}\tau^{3/8} we have

|det[1wazb]a,b=1k|kk/2τ3k/8(e1)k(τ3/8(1+e)e+τ3/4)k2.\left|\det\left[\frac{1}{w_{a}-z_{b}}\right]_{a,b=1}^{k}\right|\leq\frac{k^{k/2}}{\tau^{3k/8}(e-1)^{k}}\cdot\left(\frac{\tau^{3/8}(1+e)}{e+\tau^{3/4}}\right)^{k^{2}}.

The latter inequality implies (5.10). ∎

Proof of Lemma 5.8.

Using that |za|=τ3/4|z_{a}|=\tau^{3/4}, |wa|=e|w_{a}|=e for a=1,,ka=1,\dots,k, τ(0,e8π]\tau\in(0,e^{-8\pi}] and (4.35), we see that for each 1a<bk1\leq a<b\leq k we have

|(wawb;τ)(zazb;τ)(zawb;τ)(wazb;τ)|(e2;τ)(τ3/2;τ)(eτ3/4;τ)(eτ3/4;τ)(e2;e8π)(e12π;e8π)(e16π;e8π)(e16π;e8π).\begin{split}&\left|\frac{(w_{a}w_{b};\tau)_{\infty}(z_{a}z_{b};\tau)_{\infty}}{(z_{a}w_{b};\tau)_{\infty}(w_{a}z_{b};\tau)_{\infty}}\right|\leq\frac{(-e^{2};\tau)_{\infty}(-\tau^{3/2};\tau)_{\infty}}{(e\tau^{3/4};\tau)_{\infty}(e\tau^{3/4};\tau)_{\infty}}\leq\frac{(-e^{2};e^{-8\pi})_{\infty}(-e^{-12\pi};e^{-8\pi})_{\infty}}{(e^{1-6\pi};e^{-8\pi})_{\infty}(e^{1-6\pi};e^{-8\pi})_{\infty}}.\end{split}

The latter inequality implies (5.11) with B6=(e2;e8π)(e12π;e8π)(e16π;e8π)(e16π;e8π)B_{6}=\frac{(-e^{2};e^{-8\pi})_{\infty}(-e^{-12\pi};e^{-8\pi})_{\infty}}{(e^{1-6\pi};e^{-8\pi})_{\infty}(e^{1-6\pi};e^{-8\pi})_{\infty}}. ∎

6. Appendix: Proof of Lemma 1.13

In the proof we use the same notation as in Definition 1.10. For clarity, we split the proof into four steps.

Step 1. For y1,t1y_{1},t_{1}\in\mathbb{R} we define

(6.1) Gt121(y1)=det(IχyKχy)L2({t1}×).G_{t_{1}}^{2\rightarrow 1}(y_{1})=\det\left(I-\chi_{y}K_{\infty}\chi_{y}\right)_{L^{2}(\{t_{1}\}\times\mathbb{R})}.

In this step we show that Gt121G_{t_{1}}^{2\rightarrow 1} is a distribution function (DF) on \mathbb{R}.

From the displayed equation preceding Remark 1.1 in [QR13], we see that Gt121G_{t_{1}}^{2\rightarrow 1} as in (6.1) is the same as Gt121G_{t_{1}}^{2\rightarrow 1} from [QR13, (1.7)]. In particular, from [QR13, Theorem 1] we have for y1,t1y_{1},t_{1}\in\mathbb{R}

(6.2) Gt121(y1)=(supxt1[𝒜2(x)x2]y1t12𝟏{t10}),G_{t_{1}}^{2\rightarrow 1}(y_{1})=\mathbb{P}\left(\sup_{x\leq t_{1}}[\mathcal{A}_{2}(x)-x^{2}]\leq y_{1}-t_{1}^{2}\cdot{\bf 1}\{t_{1}\leq 0\}\right),

where 𝒜2\mathcal{A}_{2} is the Airy2 process. From the almost sure continuity of the Airy2 process, see [Joh03, Theorem 1.2], we conclude that Gt121G_{t_{1}}^{2\rightarrow 1} is the DF of an extended real-valued random variable taking values in (,](-\infty,\infty], and so it suffices to show that

(6.3) limyGt121(y)=1.\lim_{y\rightarrow\infty}G_{t_{1}}^{2\rightarrow 1}(y)=1.

To prove (6.3), we note from (6.2) that for each yy\in\mathbb{R} and y~=yt12𝟏{t10}\tilde{y}=y-t_{1}^{2}\cdot{\bf 1}\{t_{1}\leq 0\}

(6.4) F1(41/3y~)=(supx[𝒜2(x)x2]y~)Gt121(y),\begin{split}&F_{1}(4^{1/3}\tilde{y})=\mathbb{P}\left(\sup_{x\in\mathbb{R}}[\mathcal{A}_{2}(x)-x^{2}]\leq\tilde{y}\right)\leq G_{t_{1}}^{2\rightarrow 1}(y),\end{split}

where F1F_{1} is the GOE Tracy-Widom distribution [TW96]. We mention that the equality in (6.4) was derived by Johansson [Joh03] (the factor 41/34^{1/3} is omitted in that paper, see [CQR13, Theorem 1] for the correct statement). Using that F1F_{1} is a DF on \mathbb{R}, we have limyF1(y)=1\lim_{y\rightarrow\infty}F_{1}(y)=1. The latter statement and (6.4) imply (6.3), and so Gt121G_{t_{1}}^{2\rightarrow 1} is a DF on \mathbb{R}. We mention that (6.3) also follows from [QR13, Proposition 1.2].

Step 2. In this step we prove the third part of the lemma. By using the Fredholm determinant expansion, we have for all y1,t1y_{1},t_{1}\in\mathbb{R}

(6.5) Gt121(y1)=1+k=1(1)kk!(y1,)k𝑑λdet[K(t1,λa;t1,λb)]a,b=1k=1+k=1Hk, where Hk=(1)k(2π𝗂)2kk!(y1,)k𝑑λC1,π/4k𝑑wC0,3π/4k𝑑zdet[2wazb2wa2]a,b=1ka=1kewa3/3+t1wa2λ~awaeza3/3+t1za2λ~aza,\begin{split}&G_{t_{1}}^{2\rightarrow 1}(y_{1})=1+\sum_{k=1}^{\infty}\frac{(-1)^{k}}{k!}\int_{(y_{1},\infty)^{k}}d\vec{\lambda}\det\left[K_{\infty}(t_{1},\lambda_{a};t_{1},\lambda_{b})\right]_{a,b=1}^{k}=1+\sum_{k=1}^{\infty}H_{k},\mbox{ where }\\ &H_{k}=\frac{(-1)^{k}}{(2\pi\mathsf{i})^{2k}k!}\int_{(y_{1},\infty)^{k}}d\vec{\lambda}\int_{C_{1,\pi/4}^{k}}d\vec{w}\int_{C_{0,3\pi/4}^{k}}d\vec{z}\det\left[\frac{-2w_{a}}{z_{b}^{2}-w_{a}^{2}}\right]_{a,b=1}^{k}\prod_{a=1}^{k}\frac{e^{w_{a}^{3}/3+t_{1}w_{a}^{2}-\tilde{\lambda}_{a}w_{a}}}{e^{z_{a}^{3}/3+t_{1}z_{a}^{2}-\tilde{\lambda}_{a}z_{a}}},\end{split}

and λ~a=λat12𝟏{t10}\tilde{\lambda}_{a}=\lambda_{a}-t_{1}^{2}\cdot{\bf 1}\{t_{1}\leq 0\}. We mention that in deriving the formula for HkH_{k} we used the definition of K(x,s;y,t)K_{\infty}(x,s;y,t) from (1.17) and the multilinearity of the determinant function.

We next seek to exchange the order of the integrals in HkH_{k}. We observe that we can find a constant A(0,)A\in(0,\infty), depending on y1,t1y_{1},t_{1}, such that

(6.6) C1,π/4|dw|C0,3π/4|dz||2w||ew3/3+t1w2ez3/3+t1z2|e(t12+|y1|)(|z|+|w|)A,\begin{split}&\int_{C_{1,\pi/4}}|dw|\int_{C_{0,3\pi/4}}|dz||2w|\cdot\left|\frac{e^{w^{3}/3+t_{1}w^{2}}}{e^{z^{3}/3+t_{1}z^{2}}}\right|\cdot e^{(t_{1}^{2}+|y_{1}|)(|z|+|w|)}\leq A,\end{split}

where |dz||dz|, |dw||dw| denote integration with respect to arc length. The latter follows from the cubic terms in the exponential functions.

Note that for zC0,3π/4z\in C_{0,3\pi/4}, wC1,π/4w\in C_{1,\pi/4} we have |z±w|1|z\pm w|\geq 1. The latter and Hadamard’s inequality from Lemma 2.3 imply

(6.7) |det[2wazb2wa2]a,b=1k|kk/2a=1k|2wa|.\left|\det\left[\frac{-2w_{a}}{z_{b}^{2}-w_{a}^{2}}\right]_{a,b=1}^{k}\right|\leq k^{k/2}\cdot\prod_{a=1}^{k}|2w_{a}|.

In addition, since 𝖱𝖾(waza)1\mathsf{Re}(w_{a}-z_{a})\geq 1, we have

(6.8) y1|eλ~a(zawa)|𝑑λa=exp(t12𝟏{t10}𝖱𝖾(waza)+y1𝖱𝖾(zawa))𝖱𝖾(waza).\int_{y_{1}}^{\infty}\left|e^{\tilde{\lambda}_{a}(z_{a}-w_{a})}\right|d\lambda_{a}=\frac{\exp\left(t_{1}^{2}\cdot{\bf 1}\{t_{1}\leq 0\}\cdot\mathsf{Re}(w_{a}-z_{a})+y_{1}\cdot\mathsf{Re}(z_{a}-w_{a})\right)}{\mathsf{Re}(w_{a}-z_{a})}.

Combining (6.7) and (6.8) with the fact that 𝖱𝖾(waza)1\mathsf{Re}(w_{a}-z_{a})\geq 1, we conclude that

(6.9) (y1,)k𝑑λC1,π/4k|dw|C0,3π/4k|dz||det[2wazb2wa2]a,b=1ka=1kewa3/3+t1wa2λ~awaeza3/3+t1za2λ~aza|kk/2C1,π/4k|dw|C0,3π/4k|dz|a=1k|2wa||ewa3/3+t1wa2eza3/3+t1za2|e(t12+|y1|)(|za|+|wa|)Ak<,\begin{split}&\int_{(y_{1},\infty)^{k}}d\vec{\lambda}\int_{C_{1,\pi/4}^{k}}\left|d\vec{w}\right|\int_{C_{0,3\pi/4}^{k}}\left|d\vec{z}\right|\left|\det\left[\frac{-2w_{a}}{z_{b}^{2}-w_{a}^{2}}\right]_{a,b=1}^{k}\prod_{a=1}^{k}\frac{e^{w_{a}^{3}/3+t_{1}w_{a}^{2}-\tilde{\lambda}_{a}w_{a}}}{e^{z_{a}^{3}/3+t_{1}z_{a}^{2}-\tilde{\lambda}_{a}z_{a}}}\right|\\ &\leq k^{k/2}\cdot\int_{C_{1,\pi/4}^{k}}\left|d\vec{w}\right|\int_{C_{0,3\pi/4}^{k}}\left|d\vec{z}\right|\prod_{a=1}^{k}|2w_{a}|\cdot\left|\frac{e^{w_{a}^{3}/3+t_{1}w_{a}^{2}}}{e^{z_{a}^{3}/3+t_{1}z_{a}^{2}}}\right|\cdot e^{(t_{1}^{2}+|y_{1}|)(|z_{a}|+|w_{a}|)}\leq A^{k}<\infty,\end{split}

where |dz|=|dz1||dzk||d\vec{z}|=|dz_{1}|\cdots|dz_{k}|, |dw|=|dw1||dwk||d\vec{w}|=|dw_{1}|\cdots|dw_{k}|.

One consequence of (6.9) is that

(6.10) |Hk|Akkk/2(2π)2kk!,\begin{split}|H_{k}|\leq\frac{A^{k}k^{k/2}}{(2\pi)^{2k}k!},\end{split}

which is summable over kk. Another consequence of (6.9) is that we can exchange the order of the integrals in HkH_{k} without affecting the value of the integral by Fubini’s theorem. The result is

(6.11) Hk=(1)k(2π𝗂)2kk!C1,π/4k𝑑wC0,3π/4k𝑑z(y1,)k𝑑λdet[2wazb2wa2]a,b=1ka=1kewa3/3+t1wa2λ~awaeza3/3+t1za2λ~aza=1(2π𝗂)2kk!C1,π/4k𝑑wC0,3π/4k𝑑zdet[2waewa3/3+t1wa2+𝟏{t10}t12way1wa(za2wb2)(waza)eza3/3+t1za2+𝟏{t10}t12zay1za]a,b=1k,\begin{split}&H_{k}=\frac{(-1)^{k}}{(2\pi\mathsf{i})^{2k}k!}\int_{C_{1,\pi/4}^{k}}d\vec{w}\int_{C_{0,3\pi/4}^{k}}d\vec{z}\int_{(y_{1},\infty)^{k}}d\vec{\lambda}\det\left[\frac{-2w_{a}}{z_{b}^{2}-w_{a}^{2}}\right]_{a,b=1}^{k}\prod_{a=1}^{k}\frac{e^{w_{a}^{3}/3+t_{1}w_{a}^{2}-\tilde{\lambda}_{a}w_{a}}}{e^{z_{a}^{3}/3+t_{1}z_{a}^{2}-\tilde{\lambda}_{a}z_{a}}}\\ &=\frac{1}{(2\pi\mathsf{i})^{2k}k!}\int_{C_{1,\pi/4}^{k}}d\vec{w}\int_{C_{0,3\pi/4}^{k}}d\vec{z}\det\left[\frac{2w_{a}e^{w_{a}^{3}/3+t_{1}w_{a}^{2}+{\bf 1}\{t_{1}\leq 0\}t_{1}^{2}w_{a}-y_{1}w_{a}}}{(z_{a}^{2}-w_{b}^{2})(w_{a}-z_{a})e^{z_{a}^{3}/3+t_{1}z_{a}^{2}+{\bf 1}\{t_{1}\leq 0\}t_{1}^{2}z_{a}-y_{1}z_{a}}}\right]_{a,b=1}^{k},\end{split}

where we used the multilinearity of the determinant, the fact that the determinant of a matrix is equal to that of its transpose, and that for 𝖱𝖾(waza)1\mathsf{Re}(w_{a}-z_{a})\geq 1 we have

y1eλ~a(zawa)𝑑λa=exp(t12𝟏{t10}(waza)+y1(zawa))1waza.\int_{y_{1}}^{\infty}e^{\tilde{\lambda}_{a}(z_{a}-w_{a})}d\lambda_{a}=\exp\left(t_{1}^{2}{\bf 1}\{t_{1}\leq 0\}(w_{a}-z_{a})+y_{1}(z_{a}-w_{a})\right)\cdot\frac{1}{w_{a}-z_{a}}.

Equations (6.5) and (6.11) establish (1.20), and so we conclude the third part of the lemma.

Step 3. In this step we prove the second part of the lemma. We first show that for each t1t_{1}\in\mathbb{R} the function Gt121G_{t_{1}}^{2\rightarrow 1} from (6.1) is continuous. Let yNy_{N}\in\mathbb{R} for N{}N\in\mathbb{N}\cup\{\infty\} be such that limNyN=y\lim_{N\rightarrow\infty}y_{N}=y_{\infty}. We seek to show that

(6.12) limNGt121(yN)=Gt121(y).\lim_{N\rightarrow\infty}G_{t_{1}}^{2\rightarrow 1}(y_{N})=G_{t_{1}}^{2\rightarrow 1}(y_{\infty}).

We define for w1,w2C1,π/4w_{1},w_{2}\in C_{1,\pi/4}, zC0,3π/4z\in C_{0,3\pi/4} and N{}N\in\mathbb{N}\cup\{\infty\}

gw1,w2N(z)=2w1ew13/3+t1w12+𝟏{t10}t12w1yNw1(z2w22)(w1z)ez3/3+t1z2+𝟏{t10}t12zyNz.g_{w_{1},w_{2}}^{N}(z)=\frac{2w_{1}e^{w_{1}^{3}/3+t_{1}w_{1}^{2}+{\bf 1}\{t_{1}\leq 0\}t_{1}^{2}w_{1}-y_{N}w_{1}}}{(z^{2}-w_{2}^{2})(w_{1}-z)e^{z^{3}/3+t_{1}z^{2}+{\bf 1}\{t_{1}\leq 0\}t_{1}^{2}z-y_{N}z}}.

Let R>0R>0 be sufficiently large so that |yN|R|y_{N}|\leq R for all N{}N\in\mathbb{N}\cup\{\infty\}. We observe that the functions gw1,w2Ng_{w_{1},w_{2}}^{N} satisfy the conditions of [Dim18, Lemma 2.3] with Γ1=C1,π/4\Gamma_{1}=C_{1,\pi/4}, Γ2=C0,3π/4\Gamma_{2}=C_{0,3\pi/4} and functions

F1(w)=|2wew3/3+t1w2+𝟏{t10}t12w|e|R||w|,F2(z)=|ez3/3t1z2𝟏{t10}t12z|e|R||z|.F_{1}(w)=\left|2we^{w^{3}/3+t_{1}w^{2}+{\bf 1}\{t_{1}\leq 0\}t_{1}^{2}w}\right|\cdot e^{|R||w|},\hskip 5.69054ptF_{2}(z)=\left|e^{-z^{3}/3-t_{1}z^{2}-{\bf 1}\{t_{1}\leq 0\}t_{1}^{2}z}\right|\cdot e^{|R||z|}.

From [Dim18, Lemmas 2.2 and 2.3] we conclude that

(6.13) limNdet(I+KN)L2(C1,π/4)=det(I+K)L2(C1,π/4), where KN(w1,w2)=C0,3π/4gw1,w2N(z)𝑑z.\lim_{N\rightarrow\infty}\det(I+K^{N})_{L^{2}(C_{1,\pi/4})}=\det(I+K^{\infty})_{L^{2}(C_{1,\pi/4})},\mbox{ where }K^{N}(w_{1},w_{2})=\int_{C_{0,3\pi/4}}g_{w_{1},w_{2}}^{N}(z)dz.

From (1.20), which we proved in Step 2 above, and (6.1) we see that

det(I+KN)L2(C1,π/4)=Gt121(yN) for all N{},\det(I+K^{N})_{L^{2}(C_{1,\pi/4})}=G_{t_{1}}^{2\rightarrow 1}(y_{N})\mbox{ for all $N\in\mathbb{N}\cup\{\infty\}$,}

which in view of (6.13) proves (6.12).

Let us fix mm\in\mathbb{N}, t1<<tmt_{1}<\cdots<t_{m} and define for y=(y1,,ym)m\vec{y}=(y_{1},\dots,y_{m})\in\mathbb{R}^{m} the function

(6.14) fm(y;t)=fm(y1,,ym;t1,,tm):=det(IχyKχy)L2({t1,,tm}×).f_{m}(\vec{y};\vec{t})=f_{m}(y_{1},\dots,y_{m};t_{1},\dots,t_{m}):=\det\left(I-\chi_{y}K_{\infty}\chi_{y}\right)_{L^{2}(\{t_{1},\dots,t_{m}\}\times\mathbb{R})}.

In the remainder of this step we prove that, for fixed t\vec{t}, fm(;t)f_{m}(\cdot;\vec{t}) is a continuous function on m\mathbb{R}^{m}. Let yNm\vec{y}^{N}\in\mathbb{R}^{m} for N{}N\in\mathbb{N}\cup\{\infty\} be such that limNyN=y\lim_{N\rightarrow\infty}\vec{y}^{N}=\vec{y}^{\infty}. We seek to show that

(6.15) limNfm(yN;t)=fm(y;t).\lim_{N\rightarrow\infty}f_{m}(\vec{y}^{N};\vec{t})=f_{m}(\vec{y}^{\infty};\vec{t}).

In principle, it is possible to use the Fredholm determinant expansion formula in the right side of (6.14) and repeat our work from Step 2 and the first part of Step 3 to prove (6.15). As the computations are quite involved, we instead deduce the continuity of fmf_{m} from the continuity of f1f_{1} (i.e. Gt121G_{t_{1}}^{2\rightarrow 1}), which we showed in (6.12) and the fact that the fmf_{m} arise as limits of distribution functions of the process XtX_{t} in [BFS08]. We provide the details below.

Let XtX_{t} be the process, defined in [BFS08, Equation (2.5)]. From the work in [BFS08, Section 4] we have that for each mm\in\mathbb{N}, t1<t2<<tmt_{1}<t_{2}<\cdots<t_{m} and y1,,ymy_{1},\dots,y_{m}\in\mathbb{R}

(6.16) limn(k=1m{Xn(tk)yk})=fm(y;t).\lim_{n\rightarrow\infty}\mathbb{P}\left(\cap_{k=1}^{m}\{X_{n}(t_{k})\leq y_{k}\}\right)=f_{m}(\vec{y};\vec{t}).

Equation (6.16) shows that fm(y;t)[0,1]f_{m}(\vec{y};\vec{t})\in[0,1] for each ym\vec{y}\in\mathbb{R}^{m}, and for fixed t\vec{t} is non-decreasing in the yy variables. Let us define the sequences xN,XNm\vec{x}^{N},\vec{X}^{N}\in\mathbb{R}^{m} through

xN=(min(y1N,y1),,min(ymN,ym)), and XN=(max(y1N,y1),,max(ymN,ym)).\vec{x}^{N}=\left(\min(y_{1}^{N},y_{1}^{\infty}),\dots,\min(y_{m}^{N},y_{m}^{\infty})\right),\mbox{ and }\vec{X}^{N}=\left(\max(y_{1}^{N},y_{1}^{\infty}),\dots,\max(y_{m}^{N},y_{m}^{\infty})\right).

As fmf_{m} is non-decreasing, we see that for each NN\in\mathbb{N}

(6.17) |fm(yN;t)fm(y;t)|f(XN;t)f(xN;t).\left|f_{m}(\vec{y}^{N};\vec{t})-f_{m}(\vec{y}^{\infty};\vec{t})\right|\leq f(\vec{X}^{N};\vec{t})-f(\vec{x}^{N};\vec{t}).

In addition, by subadditivity we have for each N{}N\in\mathbb{N}\cup\{\infty\}

(k=1m{Xn(tk)XkN})(k=1m{Xn(tk)xkN})k=1m(xkN<Xn(tk)XkN).\mathbb{P}\left(\cap_{k=1}^{m}\{X_{n}(t_{k})\leq X^{N}_{k}\}\right)-\mathbb{P}\left(\cap_{k=1}^{m}\{X_{n}(t_{k})\leq x^{N}_{k}\}\right)\leq\sum_{k=1}^{m}\mathbb{P}\left(x^{N}_{k}<X_{n}(t_{k})\leq X^{N}_{k}\right).

Taking the limit nn\rightarrow\infty in the last inequality, using (6.16), we get

(6.18) f(XN;t)f(xN;t)k=1m[f1(XkN;tk)f1(xkN;tk)].f(\vec{X}^{N};\vec{t})-f(\vec{x}^{N};\vec{t})\leq\sum_{k=1}^{m}[f_{1}(X_{k}^{N};t_{k})-f_{1}(x_{k}^{N};t_{k})].

Since f1f_{1} is continuous from (6.12), and limNXN=limNxN=y\lim_{N\rightarrow\infty}\vec{X}^{N}=\lim_{N\rightarrow\infty}\vec{x}^{N}=\vec{y}^{\infty}, we may conclude (6.15) by combining (6.17) and (6.18).

Step 4. In this final step we prove the first part of the lemma. From our work in Steps 1 and 3, we know that for each t1t_{1}\in\mathbb{R}, the function f1(;t1)f_{1}(\cdot;t_{1}) as in (6.14) is a continuous DF on \mathbb{R}. This and (6.16) show that Xn(t1)X_{n}(t_{1}) weakly converge to a random variable with DF f1(;t1)f_{1}(\cdot;t_{1}) as nn\rightarrow\infty. In particular, we conclude that Xn(t1)X_{n}(t_{1}) is a tight sequence of random variables. The latter implies that for all mm\in\mathbb{N} and t1<t2<<tmt_{1}<t_{2}<\cdots<t_{m}, we have (Xn(t1),,Xn(tm))(X_{n}(t_{1}),\dots,X_{n}(t_{m})) is a tight sequence of random vectors in m\mathbb{R}^{m}. The tightness of (Xn(t1),,Xn(tm))(X_{n}(t_{1}),\dots,X_{n}(t_{m})), equation (6.16) and the continuity of fm(;t)f_{m}(\cdot;\vec{t}) together imply that fm(;t)f_{m}(\cdot;\vec{t}) is a continuous DF on m\mathbb{R}^{m} and (Xn(t1),,Xn(tm))(X_{n}(t_{1}),\dots,X_{n}(t_{m})) converge weakly to a random vector in m\mathbb{R}^{m} with DF fm(;t)f_{m}(\cdot;\vec{t}) as nn\rightarrow\infty.

For xx\in\mathbb{R}, we let x\mathbb{R}_{x} denote a copy of \mathbb{R} and endow it with the Borel σ\sigma-algebra (x)\mathcal{B}(\mathbb{R}_{x}). If II\subset\mathbb{R} is a finite set, and I={t1,,tm}I=\{t_{1},\dots,t_{m}\}, with t1<<tmt_{1}<\cdots<t_{m}, we let SI=×xIiS_{I}=\times_{x\in I}\mathbb{R}_{i} and endow SIS_{I} with the product σ\sigma-algebra xI(x)\otimes_{x\in I}\mathcal{B}(\mathbb{R}_{x}). We also let μI\mu_{I} denote the unique measure on (×xIx,xI(x))\left(\times_{x\in I}\mathbb{R}_{x},\otimes_{x\in I}\mathcal{B}(\mathbb{R}_{x})\right), whose DF is given by

μI(i=1m(,yi]ti)=fm(y1,,ym;t).\mu_{I}\left(\prod_{i=1}^{m}(-\infty,y_{i}]_{t_{i}}\right)=f_{m}(y_{1},\dots,y_{m};\vec{t}).

Here, we have placed the subscript tit_{i} to indicate that (,yi]titi(-\infty,y_{i}]_{t_{i}}\subset\mathbb{R}_{t_{i}}. Since (6.16) holds for each mm\in\mathbb{N}, t1<t2<<tmt_{1}<t_{2}<\cdots<t_{m} and y1,,ymy_{1},\dots,y_{m}\in\mathbb{R}, we conclude that for finite sets IJI\subseteq J\subset\mathbb{R}, we have

μJ(×SJI)=μI.\mu_{J}(\cdot\times S_{J\setminus I})=\mu_{I}.

From the Kolmogorov existence theorem, see [Kal97, Theorem 5.16], we conclude that there is a probability space (Ω,,)(\Omega,\mathcal{F},\mathbb{P}) and a real-valued process {𝒜21(t):t}\{\mathcal{A}_{2\rightarrow 1}(t):t\in\mathbb{R}\} on that space, whose finite-dimensional distribution functions are given by fmf_{m}. This proves the first part of the lemma.

References

  • [AAR99] G. Andrews, R. Askey, and R. Roy. Special functions. Cambridge University Press, Cambridge, 1999.
  • [BC14] A. Borodin and I. Corwin. Macdonald processes. Probab. Theory Relat. Fields, 158:225–400, 2014.
  • [BCS14] A. Borodin, I. Corwin, and T. Sasamoto. From duality to determinants for qq-TASEP and ASEP. Ann. Probab., 42(6):2314–2382, 2014.
  • [BFS08] A. Borodin, P.L. Ferrari, and T. Sasamoto. Transition between Airy1 and Airy2 processes and TASEP fluctuations. Commun. Pure Appl. Math., 61(11):1603–1629, 2008.
  • [CLDR10] P. Calabrese, P. Le Doussal, and A. Rosso. Free-energy distribution of the directed polymer at high temperature. EPL (Europhysics Letters), 90:p.20002, 2010.
  • [Cor12] I. Corwin. The Kardar-Parisi-Zhang equation and universality class. Random Matrices: Theory Appl., 1, 2012.
  • [CQR13] I. Corwin, J. Quastel, and D. Remenik. Continuum statistics of the Airy2 process. Commun. Math. Phys., 317:347–362, 2013.
  • [Dim18] E. Dimitrov. KPZ and Airy limits of Hall-Littlewood random plane partitions. Ann. Inst. H. Poincaré Probab. Statist., 54:640–693, 2018.
  • [Dim20] E. Dimitrov. Two-point convergence of the stochastic six-vertex model to the Airy process. 2020. Preprint, arXiv:2006.15934.
  • [FV15] P. L. Ferrari and B. Vető. Tracy-Widom asymptotics for qq-TASEP. Ann. Inst. H. Poincaré Probab. Statist., 51:1465–1485, 2015.
  • [GT20] J.M.N. Garcia and A. Torrielli. Norms and scalar products for AdS3{A}d{S}_{3}. J. Phys. A Math., 53(14):145401, 2020.
  • [Har72] T.E. Harris. Nearest-neighbor Markov interaction processes on multidimensional lattices. Adv. Math., 9:66–89, 1972.
  • [HHT15] T. Halpin-Healy and K. Takeuchi. A KPZ cocktail-shaken, not stirred: Toasting 30 years of kinetically roughened surfaces. J. Stat. Phys., 160:794–814, 2015.
  • [HO97] G.J. Heckman and E.M. Opdam. Yang’s system of particles and Hecke algebras. Ann. Math., 145(1):139–173, 1997.
  • [Hol70] R. Holley. A class of interactions in an infinite particle system. Adv. Math., 5:291–309, 1970.
  • [Joh03] K. Johansson. Discrete polynuclear growth and determinantal processes. Comm. Math. Phys., 242:277–329, 2003.
  • [Kal97] O. Kallenberg. Foundations of Modern Probability. Springer-Verlag, New York, 1997.
  • [Lig72] T. M. Liggett. Existence theorems for infinite particle systems. Trans. Amer. Math. Soc., 165:471–481, 1972.
  • [MGP68] C.T. MacDonald, J.H. Gibbs, and A.C. Pipkin. Kinetics of biopolymerization on nucleic acid templates. Biopolymers, 6:1–25, 1968.
  • [MQR21] K. Matetski, J. Quastel, and D. Remenik. The KPZ fixed point. Acta Math., 227:115–203, 2021.
  • [OQR16] J. Ortmann, J. Quastel, and D. Remenik. Exact formulas for random growth with half-flat initial data. Ann. Appl. Probab., 26(1):507–548, 2016.
  • [Pra94] V. Prasolov. Problems and Theorems in Linear Algebra. Amer. Math. Soc., Providence, RI, 1994.
  • [PS02] M. Prähofer and H. Spohn. Scale invariance of the PNG droplet and the Airy process. J. Stat. Phys., 108:1071–1106, 2002.
  • [QR13] J. Quastel and D. Remenik. Supremum of the Airy2 process minus a parabola on a half line. J. Stat. Phys., 150:442–456, 2013.
  • [QR14] J. Quastel and D. Remenik. Airy processes and variational problems. In Topics in percolative and disordered systems, pages 121–171. Springer, New York, 2014.
  • [QS15] J. Quastel and H. Spohn. The one-dimensional KPZ equation and its universality class. J. Stat. Phys., 160:965–984, 2015.
  • [QS22] J. Quastel and S. Sarkar. Convergence of exclusion processes and the KPZ equation to the KPZ fixed point. J. Amer. Math. Soc., 2022. https://doi.org/10.1090/jams/999.
  • [Rob55] H. Robbins. A remark on Stirling’s formula. Amer. Math. Monthly, 62:26–29, 1955.
  • [Sas05] T. Sasamoto. Spatial correlations of the 11D KPZ surface on a flat substrate. J. Phys. A Math. Gen., 38(33):L549, 2005.
  • [Spi70] F. Spitzer. Interaction of Markov processes. Adv. Math., 5:246–290, 1970.
  • [SS03] E. Stein and R. Shakarchi. Complex analysis. Princeton University Press, Princeton, NJ, 2003.
  • [TW96] C.A. Tracy and H. Widom. On orthogonal and symplectic matrix ensembles. Comm. Math. Phys., 177(3):727–754, 1996.