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

Finite size corrections relating to distributions of the length of longest increasing subsequences

Peter J. Forrester and Anthony Mays School of Mathematics and Statistics, ARC Centre of Excellence for Mathematical & Statistical Frontiers, University of Melbourne, Victoria 3010, Australia [email protected]
Abstract.

Considered are the large NN, or large intensity, forms of the distribution of the length of the longest increasing subsequences for various models. Earlier work has established that after centring and scaling, the limit laws for these distributions relate to certain distribution functions at the hard edge known from random matrix theory. By analysing the hard to soft edge transition, we supplement and extend results of Baik and Jenkins for the Hammersley model and symmetrisations, which give that the leading correction is proportional to z2/3z^{-2/3}, where z2z^{2} is the intensity of the Poisson rate, and provides a functional form as derivates of the limit law. Our methods give the functional form both in terms of Fredholm operator theoretic quantities, and in terms of Painlevé transcendents. For random permutations and their symmetrisations, numerical analysis of exact enumerations and simulations gives compelling evidence that the leading corrections are proportional to N1/3N^{-1/3}, and moreover provides an approximation to their graphical forms.

1. Introduction

Taking a viewpoint of random matrix theory in probability theory, it is very natural to ask about the rate of convergence to universal laws. Consider for example the spacing distribution, p2(s)p_{2}(s) say, between consecutive eigenvalues in ensembles with unitary symmetry. Here the subscript is the Dyson index β=2\beta=2 for unitary symmetry. The corresponding universal law, obtained by taking the large NN limit of an ensemble with unitary symmetry and scaling the mean spacing to unity, tells us that [16]

(1.1) p2(s)=d2ds2log(1𝕂(0,s)sine),p_{2}(s)={d^{2}\over ds^{2}}\log\Big{(}1-\mathbb{K}_{(0,s)}^{\rm sine}\Big{)},

where K(0,s)sineK_{(0,s)}^{\rm sine} is the Fredholm determinant of the integral operator on (0,s)(0,s) with the so-called sine kernel

(1.2) Ksine(x,y)=sinπ(xy)π(xy).K^{\rm sine}(x,y)={\sin\pi(x-y)\over\pi(x-y)}.

An example of an ensemble with unitary symmetry is the set of N×NN\times N unitary matrices chosen with Haar (uniform) measure. With p2,NU(N)(s)p_{2,N}^{U(N)}(s) denoting the spacing distribution between consecutive eigenvalues in this ensemble, the limit theorem relating to (1.1) is that

(1.3) limN(2πN)2p2,NU(N)(2πs/N)=p2(s).\lim_{N\to\infty}\Big{(}{2\pi\over N}\Big{)}^{2}p_{2,N}^{U(N)}(2\pi s/N)=p_{2}(s).

The rate of convergence question may be posed by asking for a tight bound on

(1.4) sup0sN|p2,NU(N)(2πs/N)p2(s)|.\sup_{0\leq s\leq N}\Big{|}p_{2,N}^{U(N)}(2\pi s/N)-p_{2}(s)\Big{|}.

Less ambitious, but more in keeping with an applied mathematics viewpoint on this aspect of random matrix theory, is to ask for the leading term in the large NN asymptotic expansion of the difference

(1.5) p2,NU(N)(2πs/N)p2(s)p_{2,N}^{U(N)}(2\pi s/N)-p_{2}(s)

occurring in (1.4) for ss fixed. Indeed this question is central to probing the Keating–Snaith hypothesis [33] relating the statistical distribution of the eigenvalues of Haar distributed random unitary matrices to the statistical distribution of the zeros of the Riemann zeta function on the critical line [35, 10, 24, 12]. Here one uses Odlyzko’s data set [35] of over 10910^{9} high precision consecutive zeros about a zero number near 102310^{23} to obtain the empirical spacing distribution. From a graphical viewpoint this appears identical to p2(s)p_{2}(s). This is in keeping with the Montgomery–Odlyzko law (see e.g. [41]) equating the scaled local statistics of the Riemann zeros infinitely high up the critical line to the limiting bulk scaled eigenvalue statistics from any random matrix ensemble with unitary symmetry. However there are finite size effects — even though 102310^{23} is huge on an absolute scale, it is the logarithm of the zero number which is the relevant measure of size. The extraordinary statistics provided by Odlyzko’s data set allows for the functional form of the analogue of the difference (1.5), where now p2,NU(N)(2πs/N)p_{2,N}^{U(N)}(2\pi s/N) is replaced by the empirically determined spacing distribution, to be accurately determined. The Keating–Snaith hypothesis predicts that this difference will be identical to the difference (1.5) for an appropriate value of NN and rescaling of ss. Hence the applied interest in (1.5) for fixed ss and large NN, a study of which was undertaken in [10, 24, 12].

Refer to captionlPr(l700l)\Pr\big{(}l^{\Box}_{700}\leq l\big{)}
Refer to captionlDifference (1.9)
Figure 1. On the left we have the empirical CDF of the longest increasing subsequences of 1,000,000 random permutations of length N=700N=700, along with the calculation of the exact CDF using c700(l)c_{700}^{\Box}(l) in (4.8) [black dots] and the limiting CDF given by the second term in (1.9) [red curve]. On the right is plotted the difference (1.9).

Our interest in the present work relates to the finite size corrections of another applied problem from random matrix theory, this time in the field of combinatorics. Take the set of the first NN positive integers, and choose a permutation uniformly at random. The question is to specify the statistics of the longest increasing subsequence length, lNl_{N}^{\square} say, from the permutation in the large NN limit. To explain the notion of the longest increasing subsequence, suppose N=8N=8 and the permutation is the ordered list 2,5,6,8,7,3,4,12,5,6,8,7,3,4,1. The longest subsequences of increasing numbers in this list have length 4: 2,5,6,82,5,6,8 and 2,5,6,72,5,6,7. A result of Logan and Shepp [34] (see the Introduction in [36] for more context) gives that asymptotically the expected length is 2N2\sqrt{N}. The tie in with random matrix theory is the limit theorem [5]

(1.6) limNPr(lN2NN1/6t)=E2soft(0;(t,)),\lim_{N\to\infty}{\rm Pr}\Big{(}{l_{N}^{\square}-2\sqrt{N}\over N^{1/6}}\leq t\Big{)}=E_{2}^{\rm soft}(0;(t,\infty)),

where E2soft(0,(t,))E_{2}^{\rm soft}(0,(t,\infty)) is the probability that, after centring and scaling about the largest eigenvalue, the interval (t,)(t,\infty) is free of eigenvalues in an Hermitian random matrix ensemble with unitary symmetry, or equivalently E2soft(0,(t,))E_{2}^{\rm soft}(0,(t,\infty)) is the distribution of the scaled largest eigenvalue. This quantity has the exact evaluations [18, 43]

E2soft(0;(t,))\displaystyle E_{2}^{\rm soft}(0;(t,\infty)) =det(𝕀𝕂(t,)soft)\displaystyle=\det\Big{(}\mathbb{I}-\mathbb{K}_{(t,\infty)}^{\rm soft}\Big{)}
(1.7) =exp(t(ts)q02(s)𝑑s).\displaystyle=\exp\bigg{(}-\int_{t}^{\infty}(t-s)q^{2}_{0}(s)\,ds\bigg{)}.

In the first of these expressions, 𝕂(t,)soft\mathbb{K}_{(t,\infty)}^{\rm soft} is the integral operator on (t,)(t,\infty) with kernel

(1.8) Ksoft(x,y)=Ai(x)Ai(y)Ai(y)Ai(x)xy,K^{\mathrm{soft}}(x,y)={{\rm Ai}(x){\rm Ai}^{\prime}(y)-{\rm Ai}(y){\rm Ai}^{\prime}(x)\over x-y},

where Ai(u){\rm Ai}(u) denotes the Airy function. In the second expression, q0(t)q_{0}(t) is the solution of the particular Painlevé II equation q′′=sq+2q3q^{\prime\prime}=sq+2q^{3} satisfying the boundary condition q0(s)Ai(s)q_{0}(s)\sim{\rm Ai}(s) as ss\to\infty. Much more about the mathematics relating to the longest increasing subsequence problem for a random permutation can be found in [2] and [39].

In analogy with our discussion of studies in random matrix theory motivated by Odlyzko’s data for the Riemann zeros, an immediate question is to inquire about the large NN form of the difference

(1.9) Pr(lNl)E2soft(0;(l2NN1/6,)).{\rm Pr}(l_{N}^{\square}\leq l)-E_{2}^{\rm soft}\left(0;\left(\frac{l-2\sqrt{N}}{N^{1/6}},\infty\right)\right).

There are various ways to generate exact and simulated data for this quantity; these are discussed in §4. In Figure 1, in the first panel we plot the histogram corresponding to the empirical cumulative distribution function (CDF) for l700l^{\Box}_{700}, computed from the longest increasing subsequence length of 10610^{6} random permutations. Also plotted, as black dots, is the exact CDF as a function of the positive integer ll, calculated as described in Section 4.2 below, and plotted as a red line is the quantity E2soft(0;((lN)/N1/6,))|N=700E_{2}^{\mathrm{soft}}(0;((l-\sqrt{N})/N^{1/6},\infty))|_{N=700} with ll varying continuously. In the second panel the difference (1.9) is displayed. In relation to the functional form in the second panel, we would like to know its dependence on NN, and its functional form at next-to-leading order. Our result of Conjecture 4.2 asserts the large NN asymptotic expansion

(1.10) Pr(lN2NN1/6t)=F2,0(t)+1N1/3F2,1(t)+,t:=([2N+tN1/6]2N)/N1/6,{\rm Pr}\Big{(}{l_{N}^{\square}-2\sqrt{N}\over N^{1/6}}\leq t\Big{)}=F_{2,0}(t^{*})+{1\over N^{1/3}}F_{2,1}(t)+\cdots,\quad t^{*}:=([2\sqrt{N}+tN^{1/6}]-2\sqrt{N})/N^{1/6},

where F2,0(t)=E2soft(0;(t,))F_{2,0}(t)=E_{2}^{\rm soft}(0;(t,\infty)) as known from (1.6), while the functional form F2,1(t)F_{2,1}(t) remains unknown as an analytic function, but can be approximated graphically; see Figure 7.

(0,0)(0,0)(1,0)(1,0)(0,1)(0,1)(1,1)(1,1)
11223344556677881122334455667788
Figure 2. An example of the Hammersley process, where the (random) number of points is N=8N=8. On the left we have eight points marked in the unit square, and the longest paths we can make from this arrangement of points (as measured by the number of points on the path) using only segments with positive slope have length four, i.e. l=4l^{\Box}=4. The mapping to the permutation (2,5,6,8,7,3,4,1)(2,5,6,8,7,3,4,1) is in the diagram on the right, where we number the points sequentially on the horizontal axis, and then move up the vertical axis listing the label of each point as we reach it. The two longest paths correspond to the subsequences (2,5,6,8)(2,5,6,8) and (2,5,6,7)(2,5,6,7).

There is a well known Poissonized form of the longest increasing subsequence problem known as the Hammersley process (see e.g. [22, §10.6]). Each permutation of length NN appears with probability z2N/N!z^{2N}/N! and is represented by NN points in the unit square, where zz is the Poisson rate of the number of these points. The length of the longest increasing subsequence is now the maximum over the number of points that can be joined using straight line segments starting from the origin and finishing at (1,1)(1,1), or equivalently the maximum number of segments in a path through the dots which always goes up and to the right. The example of the permutation for N=8N=8 given in the second paragraph is illustrated in Figure 2 from this viewpoint. Define this length to be the random variable l=l(z)l^{\square}=l^{\square}(z). From the definition,

(1.11) Pr(ll)=ez2N=0z2NN!Pr(lNl).{\rm Pr}(l^{\square}\leq l)=e^{-z^{2}}\sum_{N=0}^{\infty}{z^{2N}\over N!}{\rm Pr}(l_{N}^{\square}\leq l).

Analogous to (1.6), this quantity satisfies the limit law [5, 13]

(1.12) limzPr(l2zz1/3t)=E2soft(0;(t,)).\lim_{z\to\infty}\Pr\left(\frac{l^{\Box}-2z}{z^{1/3}}\leq t\right)=E_{2}^{\mathrm{soft}}\Big{(}0;(t,\infty)\Big{)}.

The study [17] showed the existence of an expansion analogous to (1.10) with NN replaced by z2z^{2}, while most significantly from the present viewpoint, the subsequent work of Baik and Jenkins [6, Theorem 1.3] provided an explicit functional form of the analogue of F2,1(s)F_{2,1}(s).

The reasons why Pr(ll){\rm Pr}(l^{\square}\leq l) is more tractable than Pr(lNl){\rm Pr}(l_{N}^{\square}\leq l) for asymptotic analysis are either of the formulas [30, 37]

(1.13) Pr(ll)=ez2e2zj=1lcosθjU(l),\displaystyle\Pr(l^{\Box}\leq l)=e^{-z^{2}}\left\langle e^{2z\sum_{j=1}^{l}\cos\theta_{j}}\right\rangle_{U(l)},

where U(l)\langle\cdot\rangle_{U(l)} is the average over the unitary group of degree ll with Haar measure, or [13]

(1.14) Pr(ll)=E2hard(0;(0,4z2);l),\displaystyle\Pr(l^{\Box}\leq l)=E_{2}^{\mathrm{hard}}\Big{(}0;(0,4z^{2});l\Big{)},

where E2hard(0;(0,s);a)E_{2}^{\mathrm{hard}}\Big{(}0;(0,s);a\Big{)} is the hard edge scaled probability that in the Laguerre unitary ensemble with parameter aa (recall the Laguerre weight is xaexx^{a}e^{-x}) the interval (0,s)(0,s) is free of eigenvalues. The first of these was used in [5] to prove (1.12), while an alternative proof of (1.12) using (1.14) was given in [13].

As with the work [6], the question of interest is now to quantify the large zz expansion of

(1.15) Pr(l2zz1/3t)E2soft(0;(t,)).\Pr\left(\frac{l^{\Box}-2z}{z^{1/3}}\leq t\right)-E_{2}^{\mathrm{soft}}\Big{(}0;(t,\infty)\Big{)}.

In this setting a precise statement can be made, supplementing the result known from [6]; in relation to the latter see §2.2. To state our result requires introducing the integral operator 𝕃(t,)\mathbb{L}_{(t,\infty)} on (t,)(t,\infty) with kernel

L(x,y):=121/3(xy)\displaystyle L(x,y):=-\frac{1}{2^{1/3}(x-y)} [xy5(Ai(x)Ai(y)+Ai(x)Ai(y))\displaystyle\Bigg{[}\frac{x-y}{5}\Big{(}{\rm Ai}(x){\rm Ai}^{\prime}(y)+{\rm Ai}^{\prime}(x){\rm Ai}(y)\Big{)}
(1.16) +x3y330Ai(x)Ai(y)x2y230Ai(x)Ai(y)].\displaystyle+\frac{x^{3}-y^{3}}{30}{\rm Ai}(x){\rm Ai}(y)-\frac{x^{2}-y^{2}}{30}{\rm Ai}^{\prime}(x){\rm Ai}^{\prime}(y)\Bigg{]}.
Proposition 1.1.

With [u][u] denoting the integer part of a positive real number uu, let

(1.17) t~=([2z+tz1/3]2z)/z1/3.\tilde{t}=([2z+tz^{1/3}]-2z)/z^{1/3}.

For large zz we have

(1.18) Pr(l2zz1/3t)=F2,0H(t~)+1(2z)2/3F2,1H(t)+O(z4/3),\Pr\left(\frac{l^{\Box}-2z}{z^{1/3}}\leq t\right)=F_{2,0}^{\rm H}(\tilde{t})+{1\over(2z)^{2/3}}F_{2,1}^{\rm H}(t)+{\rm O}(z^{-4/3}),

where F2,0H(t~)=E2soft(0,(t~,))F_{2,0}^{\rm H}(\tilde{t})=E_{2}^{\rm soft}(0,(\tilde{t},\infty)) and

(1.19) F2,1H(t)\displaystyle F_{2,1}^{\rm H}(t) =det(𝕀𝕂(t,)soft)Tr((𝕀𝕂(t,)soft)1𝕃(t,)).\displaystyle=-\det\Big{(}\mathbb{I}-\mathbb{K}^{\mathrm{soft}}_{(t,\infty)}\Big{)}\mathrm{Tr}\,\Big{(}(\mathbb{I}-\mathbb{K}^{\mathrm{soft}}_{(t,\infty)})^{-1}\mathbb{L}_{(t,\infty)}\Big{)}.

In keeping with the two characterisations of E2soft(0;(t,))E_{2}^{\rm soft}(0;(t,\infty)) given in (1), one as a Fredholm determinant and the other in terms of the Painlevé transcendent q0(s)q_{0}(s), the correction F2,1H(t)F_{2,1}^{\rm H}(t) can alternatively be written in terms of a second order linear differential equation with a Painlevé transcendent relating to q0(s)q_{0}(s) occurring in the coefficients; see Proposition 2.1 below. Moreover we show that this can be further simplified to give agreement with the result of Baik and Jenkins [6].

There are symmetrised versions of the longest increasing subsequence problem and the corresponding Hammersley process that permit analogues of the limit laws (1.6) and (1.12), and which also admit analogues of Proposition 1.1 [7, 8, 9]. To specify these, we first recall that a permutation of {1,,N}\{1,\dots,N\} can be represented as an N×NN\times N matrix, PP say, of zeros and ones with exactly NN ones, distributed so each row and column has a single one. For convenience, number the rows of PP starting from the bottom, and suppose that whenever there is an entry one in position (i,j)(i,j), there is also an entry one in position (j,i)(j,i). If furthermore there are no entries on the diagonal, NN must be even and the permutation consists entirely of two cycles. For such a permutation chosen uniformly at random, denote the random variable corresponding to the length of the longest increasing subsequence by lNl^{\boxslash}_{N}. Now, in the corresponding two line presentation, suppose the order of the second line is reversed, which is equivalent to rotating PP by ninety degrees clockwise. Assuming again that the original permutation of two cycles was chosen uniformly at random, the random variable corresponding to the longest increasing subsequence length of this rotated permutation (equivalently, the longest decreasing subsequence of the original permutation) is to be denoted by lNl^{\boxbslash}_{N}.

The Hammersley model relating to lNl^{\boxslash}_{N} has only the points in the unit square below the diagonal from (0,0)(0,0) to (1,1)(1,1) independent; the points above the diagonal are reflections of these points. The longest up/right path length in this setting will be denoted l=l(z)l^{\boxslash}=l^{\boxslash}(z). Similarly, for the rotated Hammersley model relating to lNl^{\boxbslash}_{N}, only the points in the unit square below the diagonal from (0,1)(0,1) to (1,0)(1,0) are chosen independently, with the remaining points the reflection in this diagonal of those points, and we denote by l=l(z)l^{\boxbslash}=l^{\boxbslash}(z) the longest up/right path length.

The analogue of the limit laws (1.6) and (1.12) are now [9]

(1.20) limNPr(lN2NN1/6t)=limzPr(l2zz1/3t)=E~4soft(0;(t,))\lim_{N\to\infty}{\rm Pr}\Big{(}{l^{\boxslash}_{N}-2\sqrt{N}\over N^{1/6}}\leq t\Big{)}=\lim_{z\to\infty}{\rm Pr}\Big{(}{l^{\boxslash}-2z\over z^{1/3}}\leq t\Big{)}=\tilde{E}_{4}^{\rm soft}(0;(t,\infty))

and

(1.21) limNPr(lN2NN1/6t)=limzPr(l2zz1/3t)=E1soft(0;(t,)).\lim_{N\to\infty}{\rm Pr}\Big{(}{l^{\boxbslash}_{N}-2\sqrt{N}\over N^{1/6}}\leq t\Big{)}=\lim_{z\to\infty}{\rm Pr}\Big{(}{l^{\boxbslash}-2z\over z^{1/3}}\leq t\Big{)}={E}_{1}^{\rm soft}(0;(t,\infty)).

The quantities E~4soft(0;(t,))\tilde{E}_{4}^{\rm soft}(0;(t,\infty)) and E1soft(0;(t,)){E}_{1}^{\rm soft}(0;(t,\infty)) denote the probability that upon a soft edge scaling in the neighbourhood of the largest eigenvalue in the Gaussian β\beta ensemble with β=4\beta=4 and β=1\beta=1, the interval (t,)(t,\infty) is free of eigenvalues. The tilde symbol on E~4soft(0;(t,))\tilde{E}_{4}^{\rm soft}(0;(t,\infty)) indicates a rescaling of the natural soft edge Gaussian β\beta ensemble variables; see [22, displayed equation below (9.139)]. As in the case of the limit laws associated with lNl_{N}^{\Box} and ll^{\Box}, we seek corrections to these limit laws. Our results are contained in Conjecture 4.4 in relation to lN,lNl^{\boxslash}_{N},l^{\boxbslash}_{N}, and in Propositions 3.1, 3.4 and Corollaries 3.2, 3.6 for l,ll^{\boxslash},l^{\boxbslash}. Baik and Jenkins [6, Theorem 1.2] contains a result that can be interpreted as corresponding to Proposition 3.1, but with a different functional form for F1,1H(t)F_{1,1}^{{\rm H}}(t). In Section 3.2 we discuss the latter in the context of our Painlevé characterisation of F1,1H(t)F_{1,1}^{{\rm H}}(t).

2. Large zz expansion of Pr(l2zz1/3t)\Pr\left(\frac{l^{\Box}-2z}{z^{1/3}}\leq t\right)

2.1. Proof of Proposition 1.1

Following [13] our strategy is to analyse the large zz form of the LHS of (1.18) by making use of (1.14). In the latter, ll is a simpler variable to work with than zz, so to begin we need to express zz in terms of ll. We see by matching the LHS of the respective equations that

(2.1) l=[2z+tz1/3],l=[2z+tz^{1/3}],

where [][\cdot] denotes the integer part. Introduce

(2.2) l~=2z+tz1/3.\tilde{l}=2z+tz^{1/3}.

It is convenient to consider (2.2) with tt replaced by XX so that z=z(l~;X)z=z(\tilde{l};X). This function of l~\tilde{l} and XX is uniquely determined by (2.2) and the requirement that zl~/2z\sim\tilde{l}/2, independent of XX, for l~\tilde{l} large. Furthermore we introduce notation for the square of z(l~;X)z(\tilde{l};X) and note the large l~\tilde{l} expansion of the latter

(2.3) Q(l~;X)=(2z(l~;X))2;2z(l~;X))=l~X(l~/2)1/3+X26(l~/2)1/3+O(l~5/3).Q(\tilde{l};X)=(2z(\tilde{l};X))^{2};\qquad 2z(\tilde{l};X))=\tilde{l}-X(\tilde{l}/2)^{1/3}+{X^{2}\over 6}(\tilde{l}/2)^{-1/3}+{\rm O}(\tilde{l}^{-5/3}).

We note furthermore that

(2.4) Q(l;t~)=4z2,Q(l;0)=l2,Q({l};\tilde{t})=4z^{2},\quad Q({l};0)={l}^{2},

where in the first of these zz refers to (2.1) and t~\tilde{t} is from (1.17), and we note too that Q(l~;X)Q(\tilde{l};X) is a decreasing function of XX.

The quantity E2hardE_{2}^{\rm hard} in (1.14) permits a Fredholm determinant form analogous to the first line in (1) [18]

(2.5) E2hard(0;(0,4z2);l)=det(𝕀𝕂(0,4z2)hard,l),E_{2}^{\rm hard}(0;(0,4z^{2});l)=\det\Big{(}\mathbb{I}-\mathbb{K}_{(0,4z^{2})}^{{\rm hard},l}\Big{)},

where 𝕂(0,4z2)hard,l\mathbb{K}_{(0,4z^{2})}^{{\rm hard},l} is the integral operator on (0,4z2)(0,4z^{2}) with kernel

(2.6) Khard,a(x,y)=Ja(x1/2)y1/2Ja(y1/2)Ja(x1/2)x1/2Ja(y1/2)2(xy)K^{\mathrm{hard},a}(x,y)=\frac{J_{a}(x^{1/2})y^{1/2}J_{a}^{\prime}(y^{1/2})-J_{a}^{\prime}(x^{1/2})x^{1/2}J_{a}(y^{1/2})}{2(x-y)}

and Ja(x)J_{a}(x) is the Bessel function of the first kind. It is a standard result in the theory of Fredholm integral equations [46] that the determinant in (2.5) can be expanded as a sum over kk-dimensional integrals, with the integrand a k×kk\times k determinant with entries (2.6)

(2.7) E2hard(0;(0,4z2);l)=1+n=1(1)nn!04z2𝑑x104z2𝑑xndet[Khard,l(xj,xk)]j,k=1n.E_{2}^{\mathrm{hard}}\Big{(}0;(0,4z^{2});l\Big{)}=1+\sum_{n=1}^{\infty}\frac{(-1)^{n}}{n!}\int_{0}^{4z^{2}}dx_{1}\cdots\int_{0}^{4z^{2}}dx_{n}\det\left[K^{\mathrm{hard},l}(x_{j},x_{k})\right]_{j,k=1}^{n}.

Now require that ll and zz are related by (2.1). We next change variables in each integrand of the series in (2.7), xl=Q(l;X)x_{l}=Q(l;X). Taking into consideration (2.4) the integral in the nn-th term reads

(2.8) (1)nt~l2𝑑X1Q(l;X1)t~l2𝑑XnQ(l;Xn)det[Khard,l(Q(l;Xj),Q(l;Xk))]j,k=1n.(-1)^{n}\int_{\tilde{t}}^{l^{2}}dX_{1}\,Q^{\prime}(l;X_{1})\cdots\int_{\tilde{t}}^{l^{2}}dX_{n}\,Q^{\prime}(l;X_{n})\det\Big{[}K^{\mathrm{hard},l}(Q(l;X_{j}),Q(l;X_{k}))\Big{]}_{j,k=1}^{n}.

The point here is that it follows from asymptotic expansions associated with the functional form (2.6) that for large ll the integrand is of order unity in the neighbourhood of the lower terminal of integration only. These asymptotic expansions [1, (9.3.23) & (9.3.27)] give that for large ν\nu

Jν(ν+uν1/3)\displaystyle J_{\nu}(\nu+u\nu^{1/3}) 21/3ν1/3Ai(21/3u)k=0Pk(u)ν2k/3+21/3νAi(21/3u)k=0Qk(u)ν2k/3\displaystyle\sim{2^{1/3}\over\nu^{1/3}}{\rm Ai}(-2^{1/3}u)\sum_{k=0}^{\infty}{P_{k}(u)\over\nu^{2k/3}}+{2^{1/3}\over\nu}{\rm Ai}^{\prime}(-2^{1/3}u)\sum_{k=0}^{\infty}{Q_{k}(u)\over\nu^{2k/3}}
(2.9) Jν(ν+uν1/3)\displaystyle J_{\nu}^{\prime}(\nu+u\nu^{1/3}) 22/3ν2/3Ai(21/3u)k=0Rk(u)ν2k/3+21/3ν4/3Ai(21/3u)k=0Sk(u)ν2k/3,\displaystyle\sim-{2^{2/3}\over\nu^{2/3}}{\rm Ai}^{\prime}(-2^{1/3}u)\sum_{k=0}^{\infty}{R_{k}(u)\over\nu^{2k/3}}+{2^{1/3}\over\nu^{4/3}}{\rm Ai}(-2^{1/3}u)\sum_{k=0}^{\infty}{S_{k}(u)\over\nu^{2k/3}},

for certain polynomials Pk(u),Qk(u),Rk(u),Sk(u)P_{k}(u),Q_{k}(u),R_{k}(u),S_{k}(u) of increasing degree. Moreover these expansions are uniform for u(,u0]u\in(-\infty,u_{0}] for any fixed u0u_{0}. Specifically, upon inserting the explicit values of these polynomials for low order, and slightly changing the notation,

Jl(lx(l/2)1/3)\displaystyle J_{l}\Big{(}l-x(l/2)^{1/3}\Big{)} l21/3l1/3Ai(x)+110l(2xAi(x)+3x2Ai(x))+O(1l5/3)O(ex)\displaystyle\mathop{\sim}\limits_{l\to\infty}\frac{2^{1/3}}{l^{1/3}}\mathrm{Ai}(x)+\frac{1}{10l}\left(2x\mathrm{Ai}(x)+3x^{2}\mathrm{Ai}^{\prime}(x)\right)+{\rm O}\Big{(}{1\over l^{5/3}}\Big{)}{\rm O}(e^{-x})
(2.10) Jl(lx(l/2)1/3)\displaystyle J_{l}^{\prime}\Big{(}l-x(l/2)^{1/3}\Big{)} l22/3l2/3Ai(x)21/310l4/3(8xAi(x)+(3x3+2)Ai(x))+O(1l2)O(ex),\displaystyle\mathop{\sim}\limits_{l\to\infty}-\frac{2^{2/3}}{l^{2/3}}\mathrm{Ai}^{\prime}(x)-\frac{2^{1/3}}{10l^{4/3}}\Big{(}8x\mathrm{Ai}^{\prime}(x)+\left(3x^{3}+2\right)\mathrm{Ai}(x)\Big{)}+{\rm O}\Big{(}{1\over l^{2}}\Big{)}{\rm O}(e^{-x}),

uniformly valid for x[x0,)x\in[x_{0},\infty). Recalling the form of the numerator in (2.6), we see in particular that

Jl(x1/2)y1/2Jl(y1/2)|x1/2lx(l/2)1/3y1/2ly(l/2)1/3l2Ai(x)Ai(y)+25l2/3[(yx)21/3Ai(x)Ai(y)\displaystyle J_{l}(x^{1/2})y^{1/2}J_{l}^{\prime}(y^{1/2})\Big{|}_{x^{1/2}\mapsto l-x(l/2)^{1/3}\atop y^{1/2}\mapsto l-y(l/2)^{1/3}}\mathop{\sim}\limits_{l\to\infty}-2\mathrm{Ai}(x)\mathrm{Ai}^{\prime}(y)+\frac{2}{5l^{2/3}}\Bigg{[}\frac{(y-x)}{2^{1/3}}\mathrm{Ai}(x)\mathrm{Ai}^{\prime}(y)
(2.11) (21/3+3y324/3)Ai(x)Ai(y)3x224/3Ai(x)Ai(y)]+O(1l4/3)O(ex)O(ey).\displaystyle-\left(2^{-1/3}+\frac{3y^{3}}{2^{4/3}}\right)\mathrm{Ai}(x)\mathrm{Ai}(y)-\frac{3x^{2}}{2^{4/3}}\mathrm{Ai}^{\prime}(x)\mathrm{Ai}^{\prime}(y)\Bigg{]}+{\rm O}\Big{(}{1\over l^{4/3}}\Big{)}{\rm O}(e^{-x}){\rm O}(e^{-y}).

For applicability to (2.8), taking into consideration the second equation in (2.3) and (2.6), we see that we require in (2.10) that

(2.12) x=x(l)=X(1(X/6)(l/2)2/3+O(l2)).x=x(l)=X\Big{(}1-(X/6)(l/2)^{-2/3}+{\rm O}(l^{-2})\Big{)}.

To account for this in (2.1) we must use the Taylor expansions with bounds on error terms valid for x[x0,)x\in[x_{0},\infty)

Ai(x+a21/3l2/3)\displaystyle\mathrm{Ai}\left(x+\frac{a}{2^{1/3}l^{2/3}}\right) lAi(x)+a21/3l2/3Ai(x)+O(l4/3)O(ex),\displaystyle\mathop{\sim}\limits_{l\to\infty}\mathrm{Ai}(x)+\frac{a}{2^{1/3}l^{2/3}}\mathrm{Ai}^{\prime}(x)+{\rm O}(l^{-4/3}){\rm O}(e^{-x}),
(2.13) Ai(x+a21/3l2/3)\displaystyle\mathrm{Ai}^{\prime}\left(x+\frac{a}{2^{1/3}l^{2/3}}\right) lAi(x)+ax21/3l2/3Ai(x)+O(l4/3)O(ex),\displaystyle\mathop{\sim}\limits_{l\to\infty}\mathrm{Ai}^{\prime}(x)+\frac{ax}{2^{1/3}l^{2/3}}\mathrm{Ai}(x)+{\rm O}(l^{-4/3}){\rm O}(e^{-x}),

where we made use of the differential equation satisfied by the Airy function Ai′′(x)=xAi(x)\mathrm{Ai}^{\prime\prime}(x)=x\mathrm{Ai}(x) in deriving the second expression. We can now use (2.1) in (2.6) to conclude that for large ll

(2.14) (Q(l;Xj)Q(l;Xk))1/2Khard,l(Q(l;Xj),Q(l;Xk))lKsoft(Xj,Xk)+L(Xj,Xk)l2/3+O(l4/3)O(eXj)O(eXk).-(Q^{\prime}(l;X_{j})Q^{\prime}(l;X_{k}))^{1/2}K^{\mathrm{hard},l}(Q(l;X_{j}),Q(l;X_{k}))\\ \mathop{\sim}\limits_{l\to\infty}K^{\mathrm{soft}}(X_{j},X_{k})+L(X_{j},X_{k})l^{-2/3}+{\rm O}(l^{-4/3}){\rm O}(e^{-X_{j}}){\rm O}(e^{-X_{k}}).

Substituting in (2.8) with the upper terminals replaced by \infty (this is permissible by the error bounds) gives that for large ll, and zz related to ll by (2.1),

(2.15) E2hard(0;(0,4z2);l)=1+n=1(1)nn!t~𝑑X1t~𝑑Xndet[Ksoft(Xj,Xk)+l2/3L(Xj,Xk)]j,k=1n+O(l4/3).E_{2}^{\mathrm{hard}}\Big{(}0;(0,4z^{2});l\Big{)}=1+\\ \sum_{n=1}^{\infty}\frac{(-1)^{n}}{n!}\int_{\tilde{t}}^{\infty}dX_{1}\cdots\int_{\tilde{t}}^{\infty}dX_{n}\det\left[K^{\mathrm{soft}}(X_{j},X_{k})+l^{-2/3}L(X_{j},X_{k})\right]_{j,k=1}^{n}+{\rm O}(l^{-4/3}).

Recalling now (1.14), then rewriting the RHS of (2.15) as in the reverse of going from (2.5) to (2.7), this tells us that for large ll

(2.16) Pr(l2zz1/3t)=det(𝕀(𝕂(t~,)soft+l2/3𝕃(t~,)))+O(l4/3).\Pr\left(\frac{l^{\Box}-2z}{z^{1/3}}\leq t\right)=\det\Bigg{(}\mathbb{I}-\Big{(}\mathbb{K}^{\mathrm{soft}}_{(\tilde{t},\infty)}+l^{-2/3}\mathbb{L}_{(\tilde{t},\infty)}\Big{)}\Bigg{)}+{\rm O}(l^{-4/3}).

The stated result (1.18) now follows from [12, Lemma 1], and in the term proportional to z2/3z^{-2/3} replacing t~\tilde{t} by tt, which is valid since they are equal to leading order in zz.

2.2. A differential equation characterisation of F2,1H(t)F_{2,1}^{\rm H}(t)

As mentioned below (1.19), the quantity F2,1H(t)F_{2,1}^{\rm H}(t) in (1.18), defined in terms of Fredholm integral operators in (1.19), also permits a characterisation as the solution of a particular second order linear differential equation, with coefficients given in terms of a particular (σ\sigma form) Painlevé II transcendent. In preparation, we first recall that an alternative to the second expression in (1) is the evaluation [43] (see also [22, §8.3.2])

(2.17) E2soft(0;(s,))=exp(su0(r)𝑑r),E_{2}^{\mathrm{soft}}\Big{(}0;(s,\infty)\Big{)}=\exp\left(-\int_{s}^{\infty}u_{0}(r)dr\right),

where u0(r)u_{0}(r) satisfies the particular σ\sigma-PII equation and boundary condition

(2.18) (u′′)2+4u((u)2ru+u)=0,u0(r)rAi(r)2rAi(r)2.\displaystyle(u^{\prime\prime})^{2}+4u^{\prime}\Big{(}(u^{\prime})^{2}-ru^{\prime}+u\Big{)}=0,\quad u_{0}(r)\mathop{\sim}\limits_{r\to\infty}{\rm Ai}^{\prime}(r)^{2}-r{\rm Ai}(r)^{2}.
Proposition 2.1.

Consider the quantity F2,1H(t)F_{2,1}^{\rm H}(t) in the expansion (1.18). As an alternative to (1.19) we have

(2.19) F2,1H(t)=exp(tu0(r)𝑑r)(tu1(r)𝑑r).F_{2,1}^{\rm H}(t)=-\exp\left(-\int_{t}^{\infty}u_{0}(r)dr\right)\left(\int_{t}^{\infty}u_{1}(r)dr\right).

Here u0u_{0} is specified by (2.17). The function u1u_{1} is specified as the solution of the inhomogeneous second order linear differential equation

(2.20) A2(r)u1′′+B2(r)u1+C2(r)u1=D2(r)\displaystyle A_{2}(r)u_{1}^{\prime\prime}+B_{2}(r)u_{1}^{\prime}+C_{2}(r)u_{1}=D_{2}(r)

with the coefficients

A2(r)\displaystyle A_{2}(r) =u0′′(r),B2(r)=2u0(r)4ru0(r)+6(u0(r))2,C2(r)=2u0(r),\displaystyle=u_{0}^{\prime\prime}(r),\quad B_{2}(r)=2u_{0}(r)-4ru_{0}^{\prime}(r)+6(u_{0}^{\prime}(r))^{2},\quad C_{2}(r)=2u_{0}^{\prime}(r),
(2.21) D2(r)\displaystyle D_{2}(r) =13(21/3)[u0(r)(r2u0(r)+6u0(r)u0(r)2ru0(r)+3u0′′(r))2(u0(r))2],\displaystyle=-\frac{1}{3(2^{1/3})}\Bigg{[}u_{0}^{\prime}(r)\Big{(}r^{2}u_{0}^{\prime}(r)+6u_{0}(r)u_{0}^{\prime}(r)-2ru_{0}(r)+3u_{0}^{\prime\prime}(r)\Big{)}-2(u_{0}(r))^{2}\Bigg{]},

and with boundary condition

(2.22) u1(r)r1(21/3)30[12Ai(r)Ai(r)+3r2(Ai(r))22r(Ai(r))2].\displaystyle u_{1}(r)\mathop{\sim}\limits_{r\to\infty}\frac{1}{(2^{1/3})30}\Bigg{[}12\mathrm{Ai}(r)\mathrm{Ai}^{\prime}(r)+3r^{2}(\mathrm{Ai}(r))^{2}-2r(\mathrm{Ai}^{\prime}(r))^{2}\Bigg{]}.
Proof.

We require knowledge [45] (see also [22, §8.3.3]) of an alternative to (2.5), telling us that

(2.23) E2hard(0;(0,s);a)=exp(0sv(r;a)r𝑑r),\displaystyle E_{2}^{\mathrm{hard}}(0;(0,s);a)=\exp\left(\int_{0}^{s}\frac{v(r;a)}{r}dr\right),

where vv satisfies the particular σ\sigma-PIII equation and boundary condition

(2.24) (rv′′)2(av)2v(4v+1)(vrv)=0,v(r;a)r0+r1+a22(1+a)Γ(1+a)Γ(2+a).\displaystyle(rv^{\prime\prime})^{2}-(av^{\prime})^{2}-v^{\prime}(4v^{\prime}+1)(v-rv^{\prime})=0,\quad v(r;a)\mathop{\sim}\limits_{r\to 0^{+}}-\frac{r^{1+a}}{2^{2(1+a)}\Gamma(1+a)\Gamma(2+a)}.

With zz related to tt and ll by (2.1), and with Q(l;X)Q(l;X) given by (2.3), we can change variables in (2.23) to obtain

(2.25) E2hard(0;(0,4z2);l)=exp(t~l2v(Q(l;s))Q(l;s)Q(l;s)𝑑s),E_{2}^{\mathrm{hard}}(0;(0,4z^{2});l)=\exp\bigg{(}-\int_{\tilde{t}}^{l^{2}}{v(Q(l;s))\over Q(l;s)}Q^{\prime}(l;s)\,ds\bigg{)},

where t~\tilde{t} is defined in (1.17). To be consistent with (1.18) and (2.17) we must have that for large ll

(2.26) v(Q(l;s))Q(l;s)Q(l;s)=u0(s)+u1(s)l2/3+{v(Q(l;s))\over Q(l;s)}Q^{\prime}(l;s)=u_{0}(s)+{u_{1}(s)\over l^{2/3}}+\cdots

Rearranging this gives a functional form for v(Q(l;s))v(Q(l;s)), which is to be substituted in (2.24) after first changing variables r=Q(l;s)r=Q(l;s). These steps are readily carried out using computer algebra. Equating terms at leading powers of ll gives the equation (2.18) for u0u_{0} at order l0l^{0}, and the differential equation (2.20) at order l2/3l^{-2/3}.

In relation to the boundary condition, we reconsider the above working, and the working which gave (2.16), in the case of ll continuous. The only difference is that the discrete variable t~\tilde{t} should be replaced by the continuous variable tt. Taking the logarithmic derivative of the RHS of (2.16) in this setting gives

(2.27) ddtlogdet(𝕀(𝕂(t,)soft+l2/3𝕃(t,)))=ddtTrlog(𝕀(𝕂(t,)soft+l2/3𝕃(t,)))t,lddtt(Ksoft(x,x)+L(x,x)l2/3)𝑑x=Ksoft(t,t)+L(t,t)l2/3.\frac{d}{dt}\log\det\Bigg{(}\mathbb{I}-\Big{(}\mathbb{K}^{\rm soft}_{(t,\infty)}+l^{-2/3}\mathbb{L}_{(t,\infty)}\Big{)}\Bigg{)}=\frac{d}{dt}\mathrm{Tr}\,\log\Bigg{(}\mathbb{I}-\Big{(}\mathbb{K}^{\rm soft}_{(t,\infty)}+l^{-2/3}\mathbb{L}_{(t,\infty)}\Big{)}\Bigg{)}\\ \mathop{\sim}\limits_{t,l\to\infty}-\frac{d}{dt}\int_{t}^{\infty}\Big{(}K^{\mathrm{soft}}(x,x)+L(x,x)l^{-2/3}\Big{)}dx=K^{\mathrm{soft}}(t,t)+L(t,t)l^{-2/3}.

On the other hand, it follows by substituting (2.26) in (2.25) that this same quantity is also equal to

(2.28) u0(t)+u1(t)l2/3.\displaystyle u_{0}(t)+\frac{u_{1}(t)}{l^{2/3}}.

Comparing (2.27) and (2.28) at leading order gives u0(t)tKsoft(t,t)u_{0}(t)\mathop{\sim}\limits_{t\to\infty}K^{\mathrm{soft}}(t,t), and thus the boundary condition for u0u_{0} in (2.18). At O(l2/3){\rm O}(l^{-2/3}) this comparison gives u1(t)tL(t,t)u_{1}(t)\mathop{\sim}\limits_{t\to\infty}L(t,t). Taking the limit xy=rx\to y=r in (1.16) we then obtain (2.22) for u1u_{1}.

\Box

As noted in the Introduction, earlier Baik and Jenkins [6] obtained an evaluation of F2,1H(t)F_{2,1}^{\rm H}(t) relating to Painlevé transcendents. This is simpler than our (2.19) as it involves only u0(t)u_{0}(t),

(2.29) F2,1H(t)=22/310(d2dt2+t26ddt)exp(tu0(r)𝑑r).F_{2,1}^{\rm H}(t)=-{2^{2/3}\over 10}\bigg{(}{d^{2}\over dt^{2}}+{t^{2}\over 6}{d\over dt}\bigg{)}\exp\Big{(}-\int_{t}^{\infty}u_{0}(r)\,dr\Big{)}.

Comparing with (2.19), it follows that we must have

(2.30) u1(r)=22/310(d2u0(r)dr2+(2u0(r)+r26)du0(r)dr+r3u0(r)).u_{1}(r)=-{2^{2/3}\over 10}\bigg{(}{d^{2}u_{0}(r)\over dr^{2}}+\Big{(}2u_{0}(r)+{r^{2}\over 6}\Big{)}{du_{0}(r)\over dr}+{r\over 3}u_{0}(r)\bigg{)}.

A similar circumstance arose in the study [25, discussion below (3.42)], which suggests how (2.30) can be verified from the characterisation of u1(r)u_{1}(r) in Proposition 2.1.

First, we verify from the boundary condition of u0(r)u_{0}(r) in (2.18), and that of u1(r)u_{1}(r) in (2.22) that they are compatible with (2.30). It remains then to verify that (2.30) satisfies the differential equation (2.20). This can be done be direct substitution, then substituting for the third and fourth derivatives of u0(r)u_{0}(r). In relation to the latter, we note that differentiating the differential equation (2.18) for u0u_{0}, and simplifying, gives us

(2.31) u0′′′(r)=2u0(r)+4ru0(r)6(u0(r))2.u_{0}^{\prime\prime\prime}(r)=-2u_{0}(r)+4ru_{0}^{\prime}(r)-6(u_{0}^{\prime}(r))^{2}.

Again differentiating this equation, and making further use of (2.18), we can express the fourth derivative of u0(r)u_{0}(r) in terms of u0(r)u_{0}(r) and u0(r)u_{0}^{\prime}(r). Once the substitutions have been performed, the resulting equation only involves the second derivative of u0(r)u_{0}(r) in the form of (u0′′(r))2(u_{0}^{\prime\prime}(r))^{2}, which we eliminate in favour of u0(r)u_{0}(r) and u0(r)u_{0}^{\prime}(r) using (2.18). These steps, performed using computer algebra, verify that (2.30) solves (2.20), as required.

2.3. Comparison with numerical calculations

We numerically calculate the correction term F2,1H(t)F_{2,1}^{\rm H}(t) in (1.18) using both the integral operator characterisation of Proposition 1.1 and the expression (2.19) in terms of the solution to a differential equation. We compare these to calculations of the difference

(2.32) δ2H(t):=l2/3(E2hard(0;(0,Q(l;t));l)E2soft(0;(t,)))\displaystyle\delta_{2}^{\rm H}(t):=l^{2/3}\Bigg{(}E_{2}^{\mathrm{hard}}\Big{(}0;(0,Q(l;t));l\Big{)}-E_{2}^{\mathrm{soft}}\Big{(}0;(t,\infty)\Big{)}\Bigg{)}

for l=20l=20, where, for the purposes of comparison, we use the continuous variable tt.

To numerically evaluate the integral operators we use the Fredholm determinant Matlab toolbox by Folkmar Bornemann [11], and a Mathematica implementation by Allan Trinh, coauthor on some related works along the theme of finite size corrections to limit formulas in random matrix theory [26, 27, 28, 23]. For the DE solutions u0(r),u1(r)u_{0}(r),u_{1}(r) needed for (2.19) we use a sequence of Taylor series expanded about various rr points, beginning on the right (near ++\infty, to match the DE boundary conditions) and proceeding to the left. For u0(r)u_{0}(r) we use a sequence of 600 series of degree 11, while for u1(r)u_{1}(r) we use a sequence of 500 series of degree 6. To calculate the finite l=20l=20 correction (2.32) we also need a sequence of Taylor series solutions for v(r;20)v(r;20) from (2.24) — we use a sequence of 15,446 series of degree 1010. The results are plotted in the left panel of Figure 3. In the right panel we plot a numerical estimate of the next order correction in (1.18) by calculating

(2.33) E2hard(0;(0,Q(l;t));l)E2soft(0;(t,))1l2/3F2,1H(t).\displaystyle E_{2}^{\mathrm{hard}}\Big{(}0;(0,Q(l;t));l\Big{)}-E_{2}^{\mathrm{soft}}\Big{(}0;(t,\infty)\Big{)}-\frac{1}{l^{2/3}}F_{2,1}^{\rm H}(t).
Refer to captiontF2,1H(t)F_{2,1}^{\rm H}(t)
Refer to captiontExpression (2.33)
Figure 3. In the left panel we have the correction term F2,1H(t)F_{2,1}^{\rm H}(t) in (1.18) calculated using (2.19) [blue crosses], and using (1.19) [red dashed line]. On the same axes we also plot δ2H(t)\delta_{2}^{\rm H}(t), the difference (2.32) for l=20l=20, using the Fredholm determinant expressions (2.5) and the first equality in (1) [black line] and also using the expressions in terms of solutions to differential equations (2.23) and the second equality in (1) [black dots]. In the right panel we plot the difference (2.33), which is a numerical approximation to the unknown O(z4/3)O(z^{-4/3}) term in (1.18).

3. Large zz expansion of Pr(l2zz1/3t)\Pr\left(\frac{{l^{\boxslash}}-2z}{z^{1/3}}\leq t\right) and Pr(l2zz1/3t)\Pr\left(\frac{{l^{\boxbslash}}-2z}{z^{1/3}}\leq t\right)

3.1. Fredholm determinant form

The analogues of (1.13) and (1.14) are the formulas

(3.1) Pr(ll)\displaystyle\Pr(l^{\boxslash}\leq l) =ez2/2ezTr𝐔𝐔O(l)=E~4hard(0;(0,4z2);l)\displaystyle=e^{-z^{2}/2}\left\langle e^{z\mathrm{Tr}\,\mathbf{U}}\right\rangle_{\mathbf{U}\in O(l)}=\tilde{E}_{4}^{\mathrm{hard}}\Big{(}0;(0,4z^{2});l\Big{)}
(3.2) Pr(l2l)\displaystyle\Pr(l^{\boxbslash}\leq 2l) =ez2/2ezj=1l2cosθjSp(2l)=E1hard(0;(0,4z2);l).\displaystyle=e^{-z^{2}/2}\left\langle e^{z\sum_{j=1}^{l}2\cos\theta_{j}}\right\rangle_{\mathrm{Sp}(2l)}=E_{1}^{\mathrm{hard}}\Big{(}0;(0,4z^{2});l\Big{)}.

The first equality in both is due to Rains [37], while the second were found in [29]. We note too that the validity of the second formula in (3.1) as derived in [29] is restricted to ll even. However, by the different strategy of expressing both the average over 𝐔O(l)\mathbf{U}\in O(l) and E~4hard\tilde{E}_{4}^{\mathrm{hard}} in terms of a generalised hypergeometric function of ll variables based on zonal polynomials — see [31] in relation to the former and [19] in relation to the latter — the validity can be established independent of the parity of ll. The use of a tilde in the notation E~4hard\tilde{E}_{4}^{\mathrm{hard}} indicates the use of a rescaling of the natural hard edge Laguerre β\beta ensemble variables; see [22, second displayed equation §9.8] or (3.25) below. The analogues of (2.5) are the formulas [14, 21]

(3.3) E1hard(0;(0,s);a12)\displaystyle E_{1}^{\mathrm{hard}}\left(0;(0,s);\frac{a-1}{2}\right) =det(𝕀𝕍s,(0,1)hard,a)\displaystyle=\det\left(\mathbb{I}-\mathbb{V}^{{\rm hard},a}_{s,(0,1)}\right)
(3.4) E~4hard(0;(0,s);a+1)\displaystyle\tilde{E}_{4}^{\mathrm{hard}}\Big{(}0;(0,s);a+1\Big{)} =12[det(𝕀𝕍s,(0,1)hard,a)+det(𝕀+𝕍s,(0,1)hard,a)],\displaystyle=\frac{1}{2}\left[\det\left(\mathbb{I}-\mathbb{V}^{{\rm hard},a}_{s,(0,1)}\right)+\det\left(\mathbb{I}+\mathbb{V}^{{\rm hard},a}_{s,(0,1)}\right)\right],

where 𝕍s,(0,1)hard,a\mathbb{V}^{{\rm hard},a}_{s,(0,1)} is the integral operator on (0,1)(0,1) with kernel

(3.5) Vshard,a(x,y):=s2Ja(xys).\displaystyle V_{s}^{{\rm hard},a}(x,y):=\frac{\sqrt{s}}{2}J_{a}(\sqrt{xys}).

In relation to the probabilities in (3.1) and (3.2) we have the known limit theorems (1.20) and (1.21) in terms of certain soft edge gap probabilities. As for E2softE_{2}^{\rm soft} the latter admit evaluations in terms of Fredholm determinants, and Painlevé transcendents. The Fredholm determinant forms read [42, 21]

(3.6) E1soft(0;(s,))\displaystyle E_{1}^{\mathrm{soft}}\Big{(}0;(s,\infty)\Big{)} =det(𝕀𝕍s,(0,)soft)\displaystyle=\det\left(\mathbb{I}-\mathbb{V}_{s,(0,\infty)}^{\mathrm{soft}}\right)
(3.7) E~4soft(0;(s,))\displaystyle\tilde{E}_{4}^{\mathrm{soft}}\Big{(}0;(s,\infty)\Big{)} =12[det(𝕀𝕍s,(0,)soft)+det(𝕀+𝕍s,(0,)soft)],\displaystyle=\frac{1}{2}\left[\det\left(\mathbb{I}-\mathbb{V}_{s,(0,\infty)}^{\mathrm{soft}}\right)+\det\left(\mathbb{I}+\mathbb{V}_{s,(0,\infty)}^{\mathrm{soft}}\right)\right],

where 𝕍s,(0,)soft\mathbb{V}_{s,(0,\infty)}^{\mathrm{soft}} is the integral operator on (0,)(0,\infty) with kernel

(3.8) Vssoft(x,y):=Ai(x+y+s).\displaystyle V^{\mathrm{soft}}_{s}(x,y):={\rm Ai}(x+y+s).

The forms in terms of Painlevé transcendents, assuming (2.17) or the second expression in (1), are [44]

(3.9) E1soft(0;(s,))\displaystyle E_{1}^{\mathrm{soft}}\Big{(}0;(s,\infty)\Big{)} =E2soft(0;(s,))1/2exp(12sq0(r)𝑑r)\displaystyle=E_{2}^{\mathrm{soft}}\Big{(}0;(s,\infty)\Big{)}^{1/2}\exp\left(-\frac{1}{2}\int_{s}^{\infty}q_{0}(r)dr\right)
(3.10) E~4soft(0;(s,))\displaystyle\tilde{E}_{4}^{\mathrm{soft}}\Big{(}0;(s,\infty)\Big{)} =E2soft(0;(s,))1/2cosh(12sq0(r)𝑑r),\displaystyle=E_{2}^{\mathrm{soft}}\Big{(}0;(s,\infty)\Big{)}^{1/2}\cosh\left(\frac{1}{2}\int_{s}^{\infty}q_{0}(r)dr\right),

where q0(r)q_{0}(r) satisfies the particular PII equation and boundary condition as noted below (1.8).

We will first consider the Fredholm determinant forms and obtain the analogues of Proposition 1.1.

Proposition 3.1.

For large zz we have

(3.11) Pr(l+12zz1/3t)=F1,0H(t~)+1(2z)2/3F1,1H(t)+O(z4/3)\Pr\left(\frac{{l^{\boxbslash}}+1-2z}{z^{1/3}}\leq t\right)=F_{1,0}^{{\rm H}}(\tilde{t})+{1\over(2z)^{2/3}}F_{1,1}^{{\rm H}}(t)+{\rm O}(z^{-4/3})

with t~\tilde{t} specified by (1.17), F1,0H(t~)=E1soft(0,(t~,))F_{1,0}^{{\rm H}}(\tilde{t})=E_{1}^{\rm soft}(0,(\tilde{t},\infty)) and

(3.12) F1,1H(t)\displaystyle F_{1,1}^{{\rm H}}(t) =det(𝕀𝕍t,(0,)soft)Tr((𝕀𝕍t,(0,)soft)1𝕄t,(0,)),\displaystyle=-\det\left(\mathbb{I}-\mathbb{V}_{t,(0,\infty)}^{\mathrm{soft}}\right)\mathrm{Tr}\,\left((\mathbb{I}-\mathbb{V}_{t,(0,\infty)}^{\mathrm{soft}})^{-1}\mathbb{M}_{t,(0,\infty)}\right),

where 𝕍t,(0,)soft\mathbb{V}_{t,(0,\infty)}^{\mathrm{soft}} is specified as in (3.6) and 𝕄t,(0,)\mathbb{M}_{t,(0,\infty)} is the integral operator on (0,)(0,\infty) with kernel

Mt(x,y)\displaystyle M_{t}(x,y) =1(21/3)10[(2x+2y8t)Ai(t+x+y)\displaystyle=\frac{1}{(2^{1/3})10}\Bigg{[}\Big{(}2x+2y-8t\Big{)}\mathrm{Ai}(t+x+y)
(3.13) +13(24x2+24y212xt12xy12ytt2)Ai(t+x+y)].\displaystyle+\frac{1}{3}\Big{(}24x^{2}+24y^{2}-12xt-12xy-12yt-t^{2}\Big{)}\mathrm{Ai}^{\prime}(t+x+y)\Bigg{]}.
Proof.

We start with the Fredholm determinant (3.3)

(3.14) E1hard(0;(0,4z2);l12)=det(𝕀𝕍4z2,(0,1)hard,l)=1+n=0(1)nn!01𝑑x101𝑑xndet[zJl(2zxjxk)]j,k=1n.E_{1}^{\mathrm{hard}}\left(0;(0,4z^{2});\frac{l-1}{2}\right)=\det\left(\mathbb{I}-\mathbb{V}^{\mathrm{hard},l}_{4z^{2},(0,1)}\right)\\ =1+\sum_{n=0}^{\infty}\frac{(-1)^{n}}{n!}\int_{0}^{1}dx_{1}\dots\int_{0}^{1}dx_{n}\det\left[zJ_{l}(2z\sqrt{x_{j}x_{k}})\right]_{j,k=1}^{n}.

In each variable in the integrand of the series we change variables xj=1Xj(2/l)2/3x_{j}=1-X_{j}(2/l)^{2/3}, which transforms the corresponding multi-dimensional integral to read

(3.15) (1)n0l𝑑X10l𝑑Xndet[(2/l)2/3zJl(2z(1Xj(2/l)2/3)(1Xk(2/l)2/3))]j,k=1n.(-1)^{n}\int_{0}^{l}dX_{1}\,\cdots\int_{0}^{l}dX_{n}\,\det\left[(2/l)^{2/3}zJ_{l}\Big{(}2z\sqrt{(1-X_{j}(2/l)^{2/3})(1-X_{k}(2/l)^{2/3})}\Big{)}\right]_{j,k=1}^{n}.

Introducing now z=z(l;t~)z=z(l;\tilde{t}) as expanded for large ll according to (2.3) with X=t~X=\tilde{t} shows that the argument of the Bessel function in (3.15) has the large ll expansion

(3.16) l(l2)1/3(t~+Xj+Xk+γ21/3l2/3)+O(l1),\displaystyle l-\left(\frac{l}{2}\right)^{1/3}\left(\tilde{t}+X_{j}+X_{k}+\frac{\gamma}{2^{1/3}l^{2/3}}\right)+O(l^{-1}),

with

(3.17) γ:=Xj22+Xk22XjXkXjt~Xkt~t~23,\displaystyle\gamma:=\frac{X_{j}^{2}}{2}+\frac{X_{k}^{2}}{2}-X_{j}X_{k}-X_{j}\tilde{t}-X_{k}\tilde{t}-\frac{\tilde{t}^{2}}{3},

which from the first formula in (2.10) gives the large ll behaviour of the Bessel function in (3.15) itself

(3.18) (2l)1/3Ai(t~+Xj+Xk+γ21/3l2/3)+110l[2(t~+Xj+Xk)Ai(t~+Xj+Xk)+3(t~+Xj+Xk)2Ai(t~+Xj+Xk)]+O(l5/3)O(eXjXk).\left(\frac{2}{l}\right)^{1/3}\mathrm{Ai}\left(\tilde{t}+X_{j}+X_{k}+\frac{\gamma}{2^{1/3}l^{2/3}}\right)+\frac{1}{10l}\bigg{[}2\left(\tilde{t}+X_{j}+X_{k}\right)\rm Ai(\tilde{t}+X_{j}+X_{k})\\ +3\left(\tilde{t}+X_{j}+X_{k}\right)^{2}\mathrm{Ai}^{\prime}(\tilde{t}+X_{j}+X_{k})\bigg{]}+{\rm O}(l^{-5/3}){\rm O}(e^{-X_{j}-X_{k}}).

Now making further use of the large ll expansion of z=z(l;t~)z=z(l;\tilde{t}) we see from this that the argument of the determinant in (3.15) has the large ll expansion

(3.19) Ai(t~+Xj+Xk+γ21/3l2/3)+1(21/3)10l2/3[(2Xj+2Xk8t~)Ai(t~+Xj+Xk)+3(t~+Xj+Xk)2Ai(t~+Xj+Xk)]+O(l4/3)O(eXjXk).\mathrm{Ai}\left(\tilde{t}+X_{j}+X_{k}+\frac{\gamma}{2^{1/3}l^{2/3}}\right)+\frac{1}{(2^{1/3})10l^{2/3}}\bigg{[}(2X_{j}+2X_{k}-8\tilde{t})\mathrm{Ai}(\tilde{t}+X_{j}+X_{k})\\ +3\left(\tilde{t}+X_{j}+X_{k}\right)^{2}\mathrm{Ai}^{\prime}(\tilde{t}+X_{j}+X_{k})\bigg{]}+{\rm O}(l^{-4/3}){\rm O}(e^{-X_{j}-X_{k}}).

Expanding the argument of the Airy function in the first term according to the first formula in (2.1) shows that this reduces to

(3.20) Vt~soft(Xj,Xk)+Mt~(Xj,Xk)l2/3+O(l4/3)O(eXjXk).V_{\tilde{t}}^{\mathrm{soft}}(X_{j},X_{k})+M_{\tilde{t}}(X_{j},X_{k})l^{-2/3}+{\rm O}(l^{-4/3}){\rm O}(e^{-X_{j}-X_{k}}).

The result (3.11) now follows from [12, Lemma 1], where, as in the proof of Proposition 1.1, we replace t~\tilde{t} by tt in the second-order term since they are of the same order in ll.

We see from (3.4) that knowledge of the scaled asymptotics of the Fredholm determinant in (3.3) is sufficient to compute the same for the quantity E~4hard\tilde{E}_{4}^{\mathrm{hard}} and thus from (3.1) the scaled asymptotics of Pr(ll)(l^{\boxslash}\leq l).

Corollary 3.2.

For large zz we have

(3.21) Pr(l12zz1/3t)=F4,0H(t~)+1(2z)2/3F4,1H(t)+O(z4/3),\Pr\left(\frac{l^{\boxslash}-1-2z}{z^{1/3}}\leq t\right)=F_{4,0}^{{\rm H}}(\tilde{t})+{1\over(2z)^{2/3}}F_{4,1}^{{\rm H}}(t)+{\rm O}(z^{-4/3}),

where F4,0H(t~)=E~4soft(0;(t~,))F_{4,0}^{{\rm H}}(\tilde{t})=\tilde{E}_{4}^{\rm soft}(0;(\tilde{t},\infty)) and

(3.22) F4,1H(t)=12[det(𝕀+𝕍t,(0,)soft)Tr((𝕀+𝕍t,(0,)soft)1𝕄t,(0,))det(𝕀𝕍t,(0,)soft)Tr((𝕀𝕍t,(0,)soft)1𝕄t,(0,))].F_{4,1}^{{\rm H}}(t)=\frac{1}{2}\left[\det\left(\mathbb{I}+\mathbb{V}_{t,(0,\infty)}^{\mathrm{soft}}\right)\mathrm{Tr}\,\left((\mathbb{I}+\mathbb{V}_{t,(0,\infty)}^{\mathrm{soft}})^{-1}\mathbb{M}_{t,(0,\infty)}\right)\right.\\ \left.-\det\left(\mathbb{I}-\mathbb{V}_{t,(0,\infty)}^{\mathrm{soft}}\right)\mathrm{Tr}\,\left((\mathbb{I}-\mathbb{V}_{t,(0,\infty)}^{\mathrm{soft}})^{-1}\mathbb{M}_{t,(0,\infty)}\right)\right].
Remark 3.3.

For general β>0\beta>0, define the Laguerre β\beta ensemble in terms of the eigenvalue PDF proportional to

(3.23) l=1Nxlaeβxl/21j<kN|xkxj|β,xl0.\prod_{l=1}^{N}x_{l}^{a}e^{-\beta x_{l}/2}\prod_{1\leq j<k\leq N}|x_{k}-x_{j}|^{\beta},\qquad x_{l}\geq 0.

Let Eβ,NL(0;(0,t);a)E_{\beta,N}^{\rm L}(0;(0,t);a) denote the probability that the interval (0,t)(0,t) has no eigenvalues in this ensemble. The corresponding hard edge scaled limit is specified by

(3.24) Eβhard(0;(0,t);a):=limNEβ,NL(0;(0,t/4N);a).\displaystyle E_{\beta}^{\rm hard}(0;(0,t);a):=\lim_{N\to\infty}E_{\beta,N}^{\rm L}(0;(0,t/4N);a).

With β=1,2\beta=1,2 this agrees with the meaning of E1hardE_{1}^{\rm hard} and E2hardE_{2}^{\rm hard} as appear above, while

(3.25) E~4hard(0;(0,t);a):=limNE4,N/2L(0;(0,t/4N);a),\tilde{E}_{4}^{\rm hard}(0;(0,t);a):=\lim_{N\to\infty}E_{4,N/2}^{\rm L^{*}}(0;(0,t/4N);a),

where L{\rm L}^{*} refers to the ensemble with eigenvalue PDF proportional to (3.23) but with eβxl/2e^{-\beta x_{l}/2} replaced by exle^{-x_{l}}. As used in [15] in the context of the spectral density, specify the soft edge scaled limit by

(3.26) Eβsoft(0;(t,)):=limNEβ,NL(0;(4N+2(2N)1/3t,);a).{E}_{\beta}^{\rm soft}(0;(t,\infty)):=\lim_{N\to\infty}E_{\beta,N}^{\rm L}(0;(4N+2(2N)^{1/3}t,\infty);a).

Note here that the quantity on the RHS is the probability that the interval at the far end of the spectrum (4N+2(2N)1/3t,)(4N+2(2N)^{1/3}t,\infty) contains no eigenvalues, and that the limit is independent of aa. The significance of the value 4N+2(2N)1/3t4N+2(2N)^{1/3}t is that this centres and scales the coordinates so that in the variable tt the largest eigenvalue is near the origin, and has spacing of order unity with its neighbours.

In this random matrix setting, our results suggest that in relation to the hard to soft edge transition, we have that for large α\alpha

(3.27) Eβhard(0;(0,4z2);β(α+12/β))|α=2z+tz1/3=Eβsoft(0;(t,))+O(1α2/3),{E}_{\beta}^{\rm hard}(0;(0,4z^{2});\beta(\alpha+1-2/\beta))\Big{|}_{\alpha=2z+tz^{1/3}}=E_{\beta}^{\rm soft}(0;(t,\infty))+{\rm O}\Big{(}{1\over\alpha^{2/3}}\Big{)},

with the main point being the order of the correction term. The limit law itself was established in [13] for β=1,2\beta=1,2 and 4, and, using different techniques, for general β>0\beta>0 in [38]. In relation to the correction, as already pointed out in [6] in the context of the Hammersley process corresponding to the β=1\beta=1 case, the shift αα+12/β\alpha\mapsto\alpha+1-2/\beta on the LHS is crucial for the optimal rate of convergence O(1/α2/3){\rm O}(1/\alpha^{2/3}).

3.2. Differential equation form

The scaled asymptotics of E2hardE_{2}^{\rm hard} were obtained in the proof of Proposition 2.1 in terms of the solution of a differential equation, starting from knowledge of the Painlevé transcendent evaluation (2.23). For E1hardE_{1}^{\rm hard} and E~4hard\tilde{E}_{4}^{\rm hard} the analogues of (2.23) are [20]

(3.28) E1hard(0;(0,s);a12)\displaystyle E_{1}^{\mathrm{hard}}\left(0;(0,s);\frac{a-1}{2}\right) =E2hard(0;(0,s);a)1/2exp(140sphard(r;a)r𝑑r)\displaystyle=E_{2}^{\mathrm{hard}}(0;(0,s);a)^{1/2}\exp\left(-\frac{1}{4}\int_{0}^{s}\frac{p_{\mathrm{hard}}(r;a)}{\sqrt{r}}dr\right)
(3.29) E~4hard(0;(0,s);a+1)\displaystyle\tilde{E}_{4}^{\mathrm{hard}}(0;(0,s);a+1) =E2hard(0;(0,s);a)1/2cosh(140sphard(r;a)r𝑑r),\displaystyle=E_{2}^{\mathrm{hard}}(0;(0,s);a)^{1/2}\cosh\left(\frac{1}{4}\int_{0}^{s}\frac{p_{\mathrm{hard}}(r;a)}{\sqrt{r}}dr\right),

where phard(r;a)p_{\mathrm{hard}}(r;a) satisfies the particular Painlevé III equation and boundary condition

(3.30) r(1p2)(rp)+p(rp)2+14(ra2)p+14rp3(p22)=0,phard(r;a)r0+ra/22aΓ(1+a).\displaystyle r(1-p^{2})(rp^{\prime})^{\prime}+p(rp^{\prime})^{2}+\frac{1}{4}(r-a^{2})p+\frac{1}{4}rp^{3}(p^{2}-2)=0,\quad p_{\mathrm{hard}}(r;a)\mathop{\sim}\limits_{r\to 0^{+}}\frac{r^{a/2}}{2^{a}\Gamma(1+a)}.

From Proposition 2.1 we already know how to express the leading two terms in the scaled limit of the factors E2hard(0;(0,s);a)E_{2}^{\mathrm{hard}}(0;(0,s);a) in (3.28) and (3.29) in terms of quantities satisfying differential equations. Our primary task then is to do the same for the second factor in (3.28).

Proposition 3.4.

Let z,lz,l and tt be related by (2.1), and t~\tilde{t} be defined by (1.17). We have the large zz and ll expansion

(3.31) exp(1404z2phard(r;l)r𝑑r)=exp(12t~q0(r)𝑑r)(112l2/3q1(t)+).\exp\left(-\frac{1}{4}\int_{0}^{4z^{2}}\frac{p_{\mathrm{hard}}(r;l)}{\sqrt{r}}dr\right)=\exp\Big{(}-{1\over 2}\int_{\tilde{t}}^{\infty}q_{0}(r)\,dr\Big{)}\Big{(}1-{1\over 2l^{2/3}}q_{1}(t)+\cdots\Big{)}.

Here q0q_{0} is specified as in (1) and q1(r)q_{1}(r) satisfies the DE

(3.32) A1(r)q1′′+B1(r)q1+C1(r)q1=D1(r),\displaystyle A_{1}(r)q_{1}^{\prime\prime}+B_{1}(r)q_{1}^{\prime}+C_{1}(r)q_{1}=D_{1}(r),

where

A1(r)\displaystyle A_{1}(r) :=12,B1(r):=0,C1(r):=r23q02(r),\displaystyle:=\frac{1}{2},\quad B_{1}(r):=0,\quad C_{1}(r):=-\frac{r}{2}-3q_{0}^{2}(r),
(3.33) D1(r)\displaystyle D_{1}(r) :=121/3(r2q0(r)12+rq0(r)3+q0(r)5q0(r)2q0(r)q0(r)2),\displaystyle:=\frac{1}{2^{1/3}}\left(-\frac{r^{2}q_{0}(r)}{12}+rq_{0}(r)^{3}+q_{0}(r)^{5}-\frac{q_{0}^{\prime}(r)}{2}-q_{0}(r)q_{0}^{\prime}(r)^{2}\right),

with boundary condition

(3.34) q1(r)r130(21/3)(14rAi(r)+r2Ai(r)).\displaystyle q_{1}(r)\mathop{\sim}\limits_{r\to\infty}-\frac{1}{30(2^{1/3})}\Big{(}14r\mathrm{Ai}(r)+r^{2}\mathrm{Ai}^{\prime}(r)\Big{)}.

Furthermore, for large l,zl,z

(3.35) E1hard(0;(0,4z2);l12)=Pr(l+12zz1/3t)=exp(12t~(u0(r)+q0(r))𝑑r)(112l2/3t(u1(r)+q1(r))𝑑r+),E_{1}^{\rm hard}\left(0;(0,4z^{2});\frac{l-1}{2}\right)=\Pr\left(\frac{l^{\boxslash}+1-2z}{z^{1/3}}\leq t\right)\\ =\exp\Big{(}-{1\over 2}\int_{\tilde{t}}^{\infty}(u_{0}(r)+q_{0}(r))\,dr\Big{)}\Big{(}1-{1\over 2l^{2/3}}\int_{{t}}^{\infty}(u_{1}(r)+q_{1}(r))\,dr+\cdots\Big{)},

with u0u_{0} and u1u_{1} from Proposition 2.1, and hence

(3.36) F1,1H(t)=12exp(12t(u0(r)+q0(r))𝑑r)t(u1(r)+q1(r))𝑑r.F_{1,1}^{\rm H}(t)=-{1\over 2}\exp\Big{(}-{1\over 2}\int_{{t}}^{\infty}(u_{0}(r)+q_{0}(r))\,dr\Big{)}\int_{t}^{\infty}(u_{1}(r)+q_{1}(r))\,dr.
Proof.

Analogous to (2.25), with Q(l;X)Q(l;X) given by (2.3), we can change variables to obtain

(3.37) exp(1404z2phard(r;l)r𝑑r)=exp(14t~l2phard(Q(l;s);l)Q(l;s)Q(l;s)𝑑s).\exp\left(-\frac{1}{4}\int_{0}^{4z^{2}}\frac{p_{\mathrm{hard}}(r;l)}{\sqrt{r}}dr\right)=\exp\bigg{(}{1\over 4}\int_{\tilde{t}}^{l^{2}}{p_{\rm hard}(Q(l;s);l)\over\sqrt{Q(l;s)}}Q^{\prime}(l;s)\,ds\bigg{)}.

To be consistent with (3.11) and (3.9) we must have that for large ll

(3.38) 12phard(Q(l;s);l)Q(l;s)Q(l;s)=q0(s)q1(s)l2/3+{1\over 2}{p_{\rm hard}(Q(l;s);l)\over\sqrt{Q(l;s)}}Q^{\prime}(l;s)=-q_{0}(s)-{q_{1}(s)\over l^{2/3}}+\cdots

Rearranging this gives a particular functional form for phard(Q(l;s);l)p_{\rm hard}(Q(l;s);l). This is to be substituted in the differential equation (3.30) with the change of variable r=Q(l;s)r=Q(l;s), which we do using computer algebra. Equating terms at leading powers of ll gives the differential equation stated below (1.8) for q0q_{0} at order ll, and equation (3.32) at order l1/3l^{1/3}.

From the working of the proof of Proposition 2.1 we have

(3.39) E2hard(0;(0,4z2);l)=exp(t~u0(r)𝑑r)(11l2/3t~u1(r)𝑑r+O(l1)).E_{2}^{\rm hard}(0;(0,4z^{2});l)=\exp\Big{(}-\int_{\tilde{t}}^{\infty}u_{0}(r)\,dr\Big{)}\Big{(}1-{1\over l^{2/3}}\int_{\tilde{t}}^{\infty}u_{1}(r)\,dr+{\rm O}(l^{-1})\Big{)}.

The expansion (3.35) now follows from this, (3.31) and (3.28).

In relation to the boundary condition, we allow ll in both the above working, and that of the proof of Proposition 3.1 to be continuous. The effect is to replace the discrete variable t~\tilde{t} by the continuous variable tt. The working of the proof of Proposition 3.1 then tells us

(3.40) E1hard(0;(0,4z2);(l1)/2)=det(𝕀(𝕍t,(0,)soft+(1/l2/3)𝕄t,(0,))+),E_{1}^{\rm hard}(0;(0,4z^{2});(l-1)/2)=\det\Big{(}\mathbb{I}-\big{(}\mathbb{V}_{{t},(0,\infty)}^{\rm soft}+(1/l^{2/3})\mathbb{M}_{t,(0,\infty)}\big{)}+\cdots\Big{)},

From this latter formula, it follows that for large tt

(3.41) logE1hard(0;(0,4z2);(l1)/2)0(Vtsoft(x,x)+(1/l2/3)Mt(x,x))𝑑x.\log E_{1}^{\rm hard}(0;(0,4z^{2});(l-1)/2)\sim-\int_{0}^{\infty}\Big{(}V_{t}^{\rm soft}(x,x)+(1/l^{2/3})M_{t}(x,x)\Big{)}\,dx.

Differentiating (3.41), simplifying using the explicit forms of the kernels (3.8) and (3.13), then comparing to the logarithmic derivative of (3.35) with t~\tilde{t} replaced by tt we obtain

q0(r)u0(r)\displaystyle-q_{0}(r)-u_{0}(r) rAi(r),\displaystyle\mathop{\sim}\limits_{r\to\infty}-\mathrm{Ai}(r),
(3.42) q1(r)u1(r)\displaystyle-q_{1}(r)-u_{1}(r) r1(21/3)30(14rAi(r)+r2Ai(r)).\displaystyle\mathop{\sim}\limits_{r\to\infty}\frac{1}{(2^{1/3})30}\Big{(}14r\mathrm{Ai}(r)+r^{2}\mathrm{Ai}^{\prime}(r)\Big{)}.

We already have the asymptotic behaviour for u0u_{0} in (2.18) and u1u_{1} in (2.22), from which we can check that u0u_{0} and u1u_{1} fall off faster than the RHS’s in (3.2). This implies the boundary conditions as stated below (1.8) for q0q_{0}, and (3.34) for q1q_{1}. ∎

Remark 3.5.

It is known that [43] that

(3.43) tu0(r)𝑑r=t(rt)q0(r)2𝑑r.\displaystyle\int_{t}^{\infty}u_{0}(r)\,dr=\int_{t}^{\infty}(r-t)q_{0}(r)^{2}\,dr.

Using this in (2.29) and comparing with (2.19) shows

(3.44) tu1(r)𝑑r=22/310(q0(t)2+(tq0(x)2𝑑x)2+t26tq0(x)2𝑑x).\displaystyle\int_{t}^{\infty}u_{1}(r)\,dr={2^{2/3}\over 10}\Big{(}-q_{0}(t)^{2}+\Big{(}\int_{t}^{\infty}q_{0}(x)^{2}\,dx\Big{)}^{2}+{t^{2}\over 6}\int_{t}^{\infty}q_{0}(x)^{2}\,dx\Big{)}.

Hence F1,1H(t)F_{1,1}^{\rm H}(t) can be expressed entirely in terms of q0(r)q_{0}(r) and q1(r)q_{1}(r).

As for F2,1H(t)F_{2,1}^{\rm H}(t), the earlier work of [6] gives an expression for F1,1H(t)F_{1,1}^{\rm H}(t) in terms of Painlevé transcendents simpler than (3.36). This reads [6, Th. 1.2]

(3.45) F1,1H(t)=22/310(2d2dt2+t26ddt)exp(12t((tr)q0(r)2+q0(r))𝑑r).F_{1,1}^{\rm H}(t)=-{2^{2/3}\over 10}\bigg{(}2{d^{2}\over dt^{2}}+{t^{2}\over 6}{d\over dt}\bigg{)}\exp\bigg{(}-{1\over 2}\int_{t}^{\infty}\Big{(}(t-r)q_{0}(r)^{2}+q_{0}(r)\Big{)}\,dr\bigg{)}.

A direct verification of the consistency of (3.45) and (3.36) can, upon use of Remark 3.5, be attempted along the same lines of that used in relation to verifying the identity (2.30). However the implied identity for q1(r)q_{1}(r) has extra terms and a more complex structure relative to (2.30), and this has not been carried through. We remark that graphical agreement is readily verified.

Knowledge of (3.35) and the structural relation between (3.28) and (3.29) allows for the differential equation companion to Corollary 3.2 to be presented.

Corollary 3.6.

Let {u0,u1,q0,q1}\{u_{0},u_{1},q_{0},q_{1}\} be as in Proposition 3.4. For large l,zl,z we have

(3.46) F4,1H(t)=12exp(12tu0(r)dr)[sinh(12tq0(r)dr)tq1(r)drcosh(12tq0(r)dr)tu1(r)dr].F_{4,1}^{{\rm H}}(t)=\frac{1}{2}\exp\left(-\frac{1}{2}\int_{t}^{\infty}u_{0}(r)dr\right)\Bigg{[}\sinh\left(\frac{1}{2}\int_{t}^{\infty}q_{0}(r)dr\right)\int_{t}^{\infty}q_{1}(r)dr\\ -\cosh\left(\frac{1}{2}\int_{t}^{\infty}q_{0}(r)dr\right)\int_{t}^{\infty}u_{1}(r)dr\Bigg{]}.
Remark 3.7.

This quantity was not considered in [6]. There is no evidence of an analogue of the simplified formulas (2.29) and (3.45).

3.3. Comparison with numerical calculations

Here we perform similar numerical calculations to those in Section 2.3 above, calculating the corrections F1,1HF_{1,1}^{{\rm H}} and F4,1HF_{4,1}^{{\rm H}}, from (3.11) and (3.21) respectively, using both the Fredholm determinant expressions and the expressions in terms of solutions to differential equations. We compare them to the differences

(3.47) δ1H(t):=l2/3(E1hard(0;(0,Q(l;t));l12)E1soft(0;(t,)))\displaystyle\delta_{1}^{\rm H}(t):=l^{2/3}\Bigg{(}E_{1}^{\mathrm{hard}}\Bigg{(}0;(0,Q(l;t));\frac{l-1}{2}\Bigg{)}-E_{1}^{\mathrm{soft}}\Big{(}0;(t,\infty)\Big{)}\Bigg{)}

and

(3.48) δ4H(t):=l2/3(E~4hard(0;(0,Q(l;t));l+1)E~4soft(0;(t,)))\displaystyle\delta_{4}^{\rm H}(t):=l^{2/3}\Bigg{(}\tilde{E}_{4}^{\mathrm{hard}}\Big{(}0;(0,Q(l;t));l+1\Big{)}-\tilde{E}_{4}^{\mathrm{soft}}\Big{(}0;(t,\infty)\Big{)}\Bigg{)}

for l=20l=20.

For the Fredholm determinants expressions we again use the toolbox of [11]. For the differential equation solutions q0q_{0} and q1q_{1} we obtain a sequence of Taylor series solutions; 500500 series of degree 66 for q0q_{0} and 1,0001,000 series of degree 88 for q1q_{1}. Lastly, we compute a sequence of 5,4005,400 series of degree 66 for phard(r,20)p_{\mathrm{hard}}(r,20). With these sequences, and the corresponding sequences of DE solutions from Section 2.3 we obtain the graphs in the left panels of Figure 4 for F1,1HF_{1,1}^{{\rm H}} and of Figure 5 for F4,1HF_{4,1}^{{\rm H}}. In the right panels of these figures we present plots of

(3.49) E1hard(0;(0,Q(l;t));l12)E1soft(0;(t,))1l2/3F1,1H(t)\displaystyle E_{1}^{\mathrm{hard}}\Bigg{(}0;(0,Q(l;t));\frac{l-1}{2}\Bigg{)}-E_{1}^{\mathrm{soft}}\Big{(}0;(t,\infty)\Big{)}-\frac{1}{l^{2/3}}F_{1,1}^{\rm H}(t)

and

(3.50) E~4hard(0;(0,Q(l;t));l+1)E~4soft(0;(t,))1l2/3F4,1H(t).\displaystyle\tilde{E}_{4}^{\mathrm{hard}}\Big{(}0;(0,Q(l;t));l+1\Big{)}-\tilde{E}_{4}^{\mathrm{soft}}\Big{(}0;(t,\infty)\Big{)}-\frac{1}{l^{2/3}}F_{4,1}^{\rm H}(t).

as approximations to the higher order corrections in (3.11) and (3.21) respectively.

Refer to captiontF1,1H(t)F_{1,1}^{\rm H}(t)
Refer to captiontExpression (3.49)
Figure 4. In the left panel we have the correction term F1,1H(t)F_{1,1}^{\rm H}(t) in (3.11) calculated using (3.36) [blue crosses], and using (3.12) [red dashed line]. On the same axes we also plot δ1H(t)\delta_{1}^{\rm H}(t) from (3.47) for l=20l=20 using the Fredholm determinant expressions (3.3) and (3.6) [black line] and also using the expressions in terms of solutions to differential equations (3.9) and (3.28) [black dots]. In the right panel we plot the difference (3.49), which is a numerical approximation to the remaining terms in (3.11).
Refer to captiontF4,1H(t)F_{4,1}^{\rm H}(t)
Refer to captiontExpression (3.50)
Figure 5. These are the plots for F4,1H(t)F_{4,1}^{\rm H}(t) from (3.22) and (3.46) analogous to those in Figure 4, using the same Fredholm determinant calculations and differential equation solutions. We also have δ4H(t)\delta_{4}^{\rm H}(t) from (3.48) for l=20l=20 again using the Fredholm determinant expressions and the expressions in terms of solutions to differential equations. Similarly, in the right panel is the difference (3.50).

4. Large NN expansion of Pr(lN2NN1/6t){\rm Pr}\Big{(}{l_{N}^{\square}-2\sqrt{N}\over N^{1/6}}\leq t\Big{)} and symmetrised analogues

4.1. Relationship to large zz form of Pr(l2zz1/3t)\Pr\left(\frac{l^{\Box}-2z}{z^{1/3}}\leq t\right) and conjecture

The longest increasing subsequence problem has been described in the paragraph including (1.6). Equating the latter with (1.12) shows the coincidence of limit laws with the maximal up/right path length in the Hammersley process,

(4.1) limNPr(lN2NN1/6t)=limzPr(l2zz1/3t)=E2soft(0;(t,)).\lim_{N\to\infty}{\rm Pr}\Big{(}{l_{N}^{\square}-2\sqrt{N}\over N^{1/6}}\leq t\Big{)}=\lim_{z\to\infty}\Pr\left(\frac{l^{\Box}-2z}{z^{1/3}}\leq t\right)=E_{2}^{\rm soft}\Big{(}0;(t,\infty)\Big{)}.

To understand why these two limits coincide, first recall from (1.11) that Pr(ll){\rm Pr}(l^{\square}\leq l) is an exponential generating function for Pr(lNl){\rm Pr}(l_{N}^{\square}\leq l). Furthermore the latter is a decreasing function of NN that takes values between 0 and 11. In this general setting Johansson [32] proved what has been referred to as a de-Poissonisation lemma.

Proposition 4.1.

Let the sequence {qn}n=0,1,\{q_{n}\}_{n=0,1,\dots} satisfy the bounds 0qn10\leq q_{n}\leq 1 and be monotonically decreasing so that qnqn+1q_{n}\geq q_{n+1}. Let

(4.2) ϕ(ξ):=eξn=0qnξnn!\displaystyle\phi(\xi):=e^{-\xi}\sum_{n=0}^{\infty}q_{n}{\xi^{n}\over n!}

and for given d>0d>0 write

(4.3) μn(d)=n+(2d+1+1)nlogn,νn(d)=n(2d+1+1)nlogn.\displaystyle\mu_{n}^{(d)}=n+(2\sqrt{d+1}+1)\sqrt{n\log n},\qquad\nu_{n}^{(d)}=n-(2\sqrt{d+1}+1)\sqrt{n\log n}.

One has

(4.4) ϕ(μn(d))Cndqnϕ(νn(d))+Cnd\phi(\mu_{n}^{(d)})-Cn^{-d}\leq q_{n}\leq\phi(\nu_{n}^{(d)})+Cn^{-d}

for all nn0n\geq n_{0}, where CC is some positive constant.

Rewriting (1.11) so that it reads

(4.5) Pr(l2zz1/3t)=ez2N=0z2NN!Pr(lN2zz1/3t),\Pr\left(\frac{l^{\Box}-2z}{z^{1/3}}\leq t\right)=e^{-z^{2}}\sum_{N=0}^{\infty}{z^{2N}\over N!}{\rm Pr}\left({l_{N}^{\square}-2z\over z^{1/3}}\leq t\right),

then applying Proposition 4.1 establishes the first equality in (4.1). From Proposition 1.1 we know details of further terms in the large zz expansion of Pr(l2zz1/3t){\rm Pr}\Big{(}{l^{\square}-2z\over z^{1/3}}\leq t\Big{)}, with the leading correction to the limit formula (1.12) being O(1/z2/3){\rm O}(1/z^{2/3}). However, this knowledge used in Proposition 4.1 does not give information on details of further terms in the large NN expansion of Pr(lN2NN1/6t){\rm Pr}\Big{(}{l_{N}^{\square}-2\sqrt{N}\over N^{1/6}}\leq t\Big{)}. In fact we don’t know of any analytic approach to this question. Nonetheless there are numerical methods that allow for data to be obtained leading to a conjecture.

Conjecture 4.2.

Set F2,0(t)=E2soft(0;(t,))F_{2,0}(t)=E_{2}^{\rm soft}(0;(t,\infty)). For some F2,1(t)F_{2,1}(t) we have

(4.6) Pr(lN2NN1/6t)=F2,0(t)+1N1/3F2,1(t)+,{\rm Pr}\left({l_{N}^{\square}-2\sqrt{N}\over N^{1/6}}\leq t\right)=F_{2,0}(t^{*})+{1\over N^{1/3}}F_{2,1}(t)+\cdots,

where tt^{*} is defined in (1.10).

Remark 4.3.

With N\sqrt{N} identified as zz the expansion (4.6) is consistent with (1.18).

4.2. Data from the Painlevé characterisation

Refer to captionlPr(l105l)\Pr\big{(}l^{\Box}_{10^{5}}\leq l\big{)}
Refer to captionlδ2(l)\delta_{2}(l)
Figure 6. Here we have the analogous plots as those in Figure 1, only now with 5×1065\times 10^{6} samples of random permutations with length N=105N=10^{5}, again with the limiting CDF given by the second term in (1.9) [red curve]. On the right is plotted an approximation to the quantity δ2(l)\delta_{2}(l) from (4.10), where we estimate Pr(lNl)\Pr\left(l^{\Box}_{N}\leq l\right) by using the empirical CDF from the left panel.

Use of (2.23) in (1.14) and recalling (1.11) tells us that

(4.7) N=0z2NN!Pr(lNl)=ez2exp(04z2v(r;l)rdr)=:G(z;l),\displaystyle\sum_{N=0}^{\infty}\frac{z^{2N}}{N!}\Pr\left(l_{N}^{\Box}\leq l\right)=e^{z^{2}}\exp\left(\int_{0}^{4z^{2}}\frac{v(r;l)}{r}dr\right)=:G^{\Box}(z;l),

where v(r;l)v(r;l) is the solution of the particular σ\sigma-PIII equation specified by (2.24). This shows that if we expand G(z;l)G^{\Box}(z;l) in powers of z2z^{2}

(4.8) G(z;l)=N=0z2NcN(l),cN(l)N!:=Pr(lNl),\displaystyle G^{\Box}(z;l)=\sum_{N=0}^{\infty}z^{2N}c_{N}^{\Box}(l),\qquad c_{N}^{\Box}(l)N!:=\Pr\left(l_{N}^{\Box}\leq l\right),

then we have a practical method to compute {Pr(lNl)}\{\Pr\left(l_{N}^{\Box}\leq l\right)\}. Thus our approach is to use the characterisation (2.24) to carry out the series expansion

(4.9) v(r;l)=rl+1k=0Mak(l)rk,\displaystyle v(r;l)=r^{l+1}\sum_{k=0}^{M}a_{k}^{\Box}(l)r^{k},

up to some cutoff MM. We were able to carry out the computation for M=700M=700, allowing for the computation of the CDF for all lNl_{N}^{\Box} up to N=700N=700. The data for this quantity can be stored as exact integers, by multiplying each of the probabilities by N!N!; see [36, Table 2–4] for some examples, with the largest value of NN there being N=60N=60. Recall Figure 1, where on the left we displayed the data for the case N=700N=700 in a graphical form — the histogram is the empirical CDF while the black dots are calculated using the cN(l)c_{N}^{\Box}(l) from (4.8). On the right of the figure we plotted the difference (1.9). Multiplying this difference by the conjectured order of the correction term in (4.6) we obtain the scaled difference

(4.10) δ2(l):=N1/3[Pr(lNl)E2soft(0;(l2NN1/6,))],\delta_{2}(l):=N^{1/3}\left[\Pr\left(l^{\Box}_{N}\leq l\right)-E_{2}^{\mathrm{soft}}\left(0;\left(\frac{l-2\sqrt{N}}{N^{1/6}},\infty\right)\right)\right],

which will allow us to compare the data to other values of NN.

4.3. Data from simulations

For values of NN beyond N=700N=700 data can be generated by Monte Carlo simulations. The C code used to generate samples of lNl_{N}^{\Box} was given to the authors by Eric Rains, based on the code used for the simulations in [36], which uses the algorithm of [4]. The most expensive part of the code is the pseudo random number generation, for which Rains’ code uses a Marsaglia-style multiply-with-carry bit-shifting algorithm. To generate the value of lNl_{N}^{\Box} from 5×1065\times 10^{6} trials with N=105N=10^{5} took approximately 51,00051,000 seconds. Without this code, an alternative method to generate the simulation data is to simply use the inbuilt Mathematica command LongestOrderedSequence, which has comparable runtime for this value of NN.

In Figure 6 we display the data in a graphical form, along with the estimate of the scaled difference δ2(l)\delta_{2}(l) from (4.10), where Pr(lNl)\Pr\left(l^{\Box}_{N}\leq l\right) is taken to be the empirical CDF. In Figure 7 we compare δ2(t)\delta_{2}(t) with N=700, 20000N=700,\,20000 and 10510^{5}, where we have rescaled t=(l2N)/N1/6t=(l-2\sqrt{N})/N^{1/6} — the agreement in the plots suggests that N1/3N^{-1/3} is indeed the correct order of the next-to-leading term in (4.6).

Refer to captiontδ2(t)\delta_{2}(t)
Figure 7. Comparison of δ2(t)\delta_{2}(t) from (4.10) with N=700N=700 [black dots], 2000020000 [blue crosses] and 10510^{5} [red circles], where the horizontal axis has been rescaled by t=(l2N)/N1/6t=(l-2\sqrt{N})/N^{1/6}.

4.4. Large NN expansion of the mean and variance of lNl^{\Box}_{N}

We turn our attention now to the large NN form of the mean and variance. From the leading term in (4.6) it follows that for large NN [5]

(4.11) 𝔼[lN]N2N+m2(1)N1/6+,m2(1)1.771086807,\displaystyle\mathbb{E}[l^{\Box}_{N}]\mathop{\sim}\limits_{N\to\infty}2\sqrt{N}+m_{2}^{(1)}N^{1/6}+\cdots,\quad m_{2}^{(1)}\approx-1.771086807,

where, with dF2(r)dr=ddrE2soft(0;(r,)){dF_{2}(r)\over dr}={d\over dr}E_{2}^{\rm soft}(0;(r,\infty)),

(4.12) m2(k):=rk𝑑F2(r),\displaystyle m_{2}^{(k)}:=\int_{-\infty}^{\infty}r^{k}dF_{2}(r),

and the numerical value follows from a computation based on (1) in [43]. On the other hand the correction term in (4.6) does not immediately reveal information about higher order terms in (4.11), the reason being that lNl^{\Box}_{N} is a discrete quantity, while the right hand side of (4.6) corresponds to rescaling and smoothing of the discrete distribution. This is similarly true of the variance, for which the limit theorem (1.6) gives that

(4.13) Var[lN]N(m2(2)(m2(1))2)N1/3,m2(2)(m2(1))20.81319,\displaystyle{\rm Var}[l^{\Box}_{N}]\mathop{\sim}\limits_{N\to\infty}\Big{(}m_{2}^{(2)}-(m_{2}^{(1)})^{2}\Big{)}N^{1/3},\qquad m_{2}^{(2)}-(m_{2}^{(1)})^{2}\approx 0.81319,

with the nature of higher order terms not immediately determined by the correction term in (4.6). Our data can be used to investigate the corrections to 𝔼[lN]\mathbb{E}[l^{\Box}_{N}] and Var[lN]{\rm Var}[l^{\Box}_{N}] at a numerical level.

For this purpose, we note from elementary probability theory that

(4.14) 𝔼[lN]=k=0N1(k+1)(Pr(lNk+1)Pr(lNk))=k=0N(1Pr(lNk))\displaystyle\mathbb{E}[l^{\Box}_{N}]=\sum_{k=0}^{N-1}(k+1)\Big{(}{\rm Pr}(l^{\Box}_{N}\leq k+1)-{\rm Pr}(l^{\Box}_{N}\leq k)\Big{)}=\sum_{k=0}^{N}\Big{(}1-\Pr(l^{\Box}_{N}\leq k)\Big{)}

and

Var[lN]\displaystyle{\rm Var}[l^{\Box}_{N}] =k=0N1(k+1)2(Pr(lNk+1)Pr(lNk))(𝔼[lN])2\displaystyle=\sum_{k=0}^{N-1}(k+1)^{2}\Big{(}{\rm Pr}(l^{\Box}_{N}\leq k+1)-{\rm Pr}(l^{\Box}_{N}\leq k)\Big{)}-\Big{(}\mathbb{E}[l^{\Box}_{N}]\Big{)}^{2}
(4.15) =1+k=1N(2k+1)(1Pr(lNk))(𝔼[lN])2.\displaystyle=1+\sum_{k=1}^{N}(2k+1)\Big{(}1-\Pr(l^{\Box}_{N}\leq k)\Big{)}-\Big{(}\mathbb{E}[l^{\Box}_{N}]\Big{)}^{2}.

From the theory of Section 4.2, for NN up to 700700 we have exact knowledge of {Pr(lNk)}\{{\rm Pr}(l^{\Box}_{N}\leq k)\}. Already a consequence of these distributions being discrete shows itself. Thus if we approximate the CDF Pr(lNl)\Pr(l_{N}^{\Box}\leq l) by the limiting expression E2soft(0;((l2N)/N1/6,))E_{2}^{\mathrm{soft}}\Big{(}0;\big{(}(l-2\sqrt{N})/N^{1/6},\infty\big{)}\Big{)}, and substitute this into (4.14) and (4.15) for the expected mean and variance, which we denote 𝔼[lN]\mathbb{E}_{\infty}[l_{N}^{\Box}] and Var[lN]\mathrm{Var}_{\infty}[l_{N}^{\Box}] respectively, then we obtain the numerical estimates

(4.16) 𝔼[lN](2N+m2(1)N1/6)\displaystyle\mathbb{E}_{\infty}[l^{\Box}_{N}]-\Big{(}2\sqrt{N}+m_{2}^{(1)}N^{1/6}\Big{)} N12\displaystyle\mathop{\to}\limits_{N\to\infty}\frac{1}{2}
(4.17) Var[lN](m2(2)(m2(1))2)N1/3\displaystyle\mathrm{Var}_{\infty}[l^{\Box}_{N}]-\Big{(}m_{2}^{(2)}-(m_{2}^{(1)})^{2}\Big{)}N^{1/3} N112.\displaystyle\mathop{\to}\limits_{N\to\infty}\frac{1}{12}.

We recognise the values 1/21/2 and 1/121/12 as the mean and variance of the continuous uniform distribution on [0,1][0,1].

Tabulating the quantities

(4.18) μ^2(N):=𝔼[lN](2N+m2(1)N1/6),σ^22(N):=Var[lN](m2(2)(m2(1))2)N1/3,\displaystyle\hat{\mu}_{2}(N):=\mathbb{E}[l^{\Box}_{N}]-\Big{(}2\sqrt{N}+m_{2}^{(1)}N^{1/6}\Big{)},\quad\hat{\sigma}^{2}_{2}(N):={\rm Var}[l^{\Box}_{N}]-\Big{(}m_{2}^{(2)}-(m_{2}^{(1)})^{2}\Big{)}N^{1/3},

leads us to believe that for large NN both these quantities are of order unity. Making an ansatz μ^2(N)=c+dNα\hat{\mu}_{2}(N)=c+dN^{-\alpha} and choosing between α=1/6\alpha=1/6 or 1/31/3 as suggested by their appearance already in this problem, we found that the choice α=1/3\alpha=1/3 gives the better fit. Notice that the latter exponent is precisely the one appearing in Conjecture 4.2 for the CDF. Performing a least squares analysis from our tabulation with NN from 10 up to 700 then gives

(4.19) μ^2(N)0.5065+0.222N1/3,σ^22(N)1.206+0.545N1/3.\displaystyle\hat{\mu}_{2}(N)\approx 0.5065+{0.222\over N^{1/3}},\qquad\hat{\sigma}^{2}_{2}(N)\approx-1.206+{0.545\over N^{1/3}}.

In keeping with (4.16), we expect the value 0.50650.5065 in relation to μ^2(N)\hat{\mu}_{2}(N) is exactly 1/21/2. Note that the value 1.206-1.206, being distinct from 1/121/12 in (4.17), can be understood as being due to the square of the mean occurring in (4.15). This gives a mechanism for the coupling of terms decaying in NN in the expansion of the mean, with terms that increase, which are not taken into consideration in deriving (4.17).

In Figures 8 and 9 we plot μ^2(N)\hat{\mu}_{2}(N) and σ^22(N)\hat{\sigma}_{2}^{2}(N) respectively, along with the conjectures in (4.19). In the right panel of each we also plot the differences

(4.20) μ^2(N)(0.5065+0.222N1/3),σ^22(N)(1.206+0.545N1/3).\displaystyle\hat{\mu}_{2}(N)-\left(0.5065+{0.222\over N^{1/3}}\right),\qquad\hat{\sigma}^{2}_{2}(N)-\left(-1.206+{0.545\over N^{1/3}}\right).
Refer to captionNμ^2(N)\hat{\mu}_{2}(N)
Refer to captionNDifference (4.20)
Figure 8. On the left we have the exact values of μ^2(N)\hat{\mu}_{2}(N) from (4.18) for N=10,11,,700N=10,11,\dots,700 calculated using (4.14) [black dots] compared to the conjectured form in (4.19) [red line]. The difference between these, the quantity (4.20), is plotted on the right.
Refer to captionNσ^22(N)\hat{\sigma}^{2}_{2}(N)
Refer to captionNDifference (4.20)
Figure 9. As in Figure 8, on the left we have the exact values of σ^22(N)\hat{\sigma}^{2}_{2}(N) from (4.18) for N=10,11,,700N=10,11,\dots,700 calculated using (4.15) [black dots] compared to the conjectured form in (4.19) [red line]. On the right is plotted the difference (4.20).

4.5. The quantities Pr(lN2NN1/6t){\rm Pr}\Big{(}{l^{\boxslash}_{N}-2\sqrt{N}\over N^{1/6}}\leq t\Big{)} and Pr(lN2NN1/6t){\rm Pr}\Big{(}{l^{\boxbslash}_{N}-2\sqrt{N}\over N^{1/6}}\leq t\Big{)}

Analogous to (4.7) we have

(4.21) N=0zNN!!Pr(lNl)=ez2/2exp(1204z2v(r;l1)r𝑑r)cosh(1404z2phard(r;l1)r𝑑r):=G(z;l)\sum_{N=0}^{\infty}\frac{z^{N}}{N!!}{\rm Pr}\Big{(}l^{\boxslash}_{N}\leq l\Big{)}\\ =e^{z^{2}/2}\exp\left(\frac{1}{2}\int_{0}^{4z^{2}}\frac{v(r;l-1)}{r}dr\right)\cosh\left(-\frac{1}{4}\int_{0}^{4z^{2}}\frac{p_{\mathrm{hard}}(r;l-1)}{\sqrt{r}}dr\right):=G^{\boxslash}(z;l)

and

(4.22) N=0zNN!!Pr(lN2l)=ez2/2exp(1204z2v(r;2l+1)r𝑑r)exp(1404z2phard(r;2l+1)r𝑑r):=G(z;l)\sum_{N=0}^{\infty}\frac{z^{N}}{N!!}{\rm Pr}\Big{(}l^{\boxbslash}_{N}\leq 2l\Big{)}\\ =e^{z^{2}/2}\exp\left(\frac{1}{2}\int_{0}^{4z^{2}}\frac{v(r;2l+1)}{r}dr\right)\exp\left(-\frac{1}{4}\int_{0}^{4z^{2}}\frac{p_{\mathrm{hard}}(r;2l+1)}{\sqrt{r}}dr\right):=G^{\boxbslash}(z;l)

These follow from (3.1), (3.2), (3.28) and (3.29); for example see [22, §10.7]. Hence

(4.23) G(z;l)=N=0z2NcN(l),cN(l)N!!:=Pr(lNl),\displaystyle G^{\boxslash}(z;l)=\sum_{N=0}^{\infty}z^{2N}c_{N}^{\boxslash}(l),\qquad c_{N}^{\boxslash}(l)N!!:=\Pr\left(l_{N}^{\boxslash}\leq l\right),

and

(4.24) G(z;l)=N=0z2NcN(l),cN(2l)N!!:=Pr(lN2l),\displaystyle G^{\boxbslash}(z;l)=\sum_{N=0}^{\infty}z^{2N}c_{N}^{\boxbslash}(l),\qquad c_{N}^{\boxbslash}(2l)N!!:=\Pr\left(l_{N}^{\boxbslash}\leq 2l\right),

We now proceed as detailed in Section 4.2, which provides us with the exact values of {cN(l)}\{c_{N}^{\boxbslash}(l)\} and {cN(l)}\{c_{N}^{\boxslash}(l)\} for NN up to 400400. That is, we find a series solution of degree 400400 to the differential equation in (3.30), and use it (along with the v(r;l)v(r;l) from Section 4.2) to expand (4.21) and (4.22) in powers of zz.

In Figures 10 and 11 we display the cases N=400N=400 in graphical form, along with the scaled differences

(4.25) δ1(l):=N1/3[Pr(lNl)E1soft(0;(l+12NN1/6,))],\displaystyle\delta_{1}(l):=N^{1/3}\left[\Pr\left(l^{\boxbslash}_{N}\leq l\right)-E_{1}^{\mathrm{soft}}\left(0;\left(\frac{l+1-2\sqrt{N}}{N^{1/6}},\infty\right)\right)\right],
(4.26) δ4(l):=N1/3[Pr(lNl)E~4soft(0;(l12NN1/6,))].\displaystyle\delta_{4}(l):=N^{1/3}\left[\Pr\left(l^{\boxslash}_{N}\leq l\right)-\tilde{E}_{4}^{\mathrm{soft}}\left(0;\left(\frac{l-1-2\sqrt{N}}{N^{1/6}},\infty\right)\right)\right].
Refer to captionlPr(l400l)\Pr\big{(}l^{\boxbslash}_{400}\leq l\big{)}
Refer to captionlδ1(l)\delta_{1}(l)
Figure 10. On the left we have the empirical CDF of the longest decreasing subsequences of 1,000,000 random permutations (which consist entirely of two cycles) of length N=400N=400 along with the calculation of the exact CDF using c400(l)c_{400}^{\boxbslash}(l) in (4.24) [black dots] and the limiting CDF given by the right hand side of (1.21) with t=(l12N)/N1/6t=(l-1-2\sqrt{N})/N^{1/6} [red curve]. On the right is plotted the quantity δ1(l)\delta_{1}(l) from (4.25).
Refer to captionlPr(l400l)\Pr\big{(}l^{\boxslash}_{400}\leq l\big{)}
Refer to captionlδ4(l)\delta_{4}(l)
Figure 11. Here we have the plots analogous to those in Figure 10, now counting the longest increasing subsequences, where the limiting CDF is given by the right hand side of (1.20) with t=(l+12N)/N1/6t=(l+1-2\sqrt{N})/N^{1/6} [red curve]. The right panel is again the scaled difference between the red curve and the black dots, i.e. δ4(l)\delta_{4}(l) from (4.26)

The exact values for Pr(lNl)\Pr\Big{(}l^{\boxbslash}_{N}\leq l\Big{)} and Pr(lNl)\Pr\Big{(}l^{\boxslash}_{N}\leq l\Big{)} can be supplemented by simulations as in Section 4.3. For this we use the C++ code for generating self-inverse permutations from [3], which samples the permutations uniformly for very large NN. To find the longest increasing subsequence we use the C++ implementation from [40] of an optimal algorithm. Plotting the scaled differences (4.25) and (4.26) for N=400,20000N=400,20000 and 10510^{5}, with the horizontal axis rescaled by t=(l±12N)/N1/6t=(l\pm 1-2\sqrt{N})/N^{1/6} — see Figure 12 — gives evidence for the analogue of Conjecture 4.2.

Conjecture 4.4.

Specify tt^{*} as in (1.10). Set F1,0(t)=E1soft(0;(t,))F_{1,0}(t)=E_{1}^{\rm soft}(0;(t,\infty)). For some F1,1(t)F_{1,1}(t) we have

(4.27) Pr(lN+12NN1/6t)=F1,0(t)+1N1/3F1,1(t)+{\rm Pr}\left({l_{N}^{\boxbslash}+1-2\sqrt{N}\over N^{1/6}}\leq t\right)=F_{1,0}(t^{*})+{1\over N^{1/3}}F_{1,1}(t)+\cdots

Similarly, with F4,0(t)=E~4soft(0;(t,))F_{4,0}(t)=\tilde{E}_{4}^{\rm soft}(0;(t,\infty)), for some F4,1(t)F_{4,1}(t) we have

(4.28) Pr(lN12NN1/6t)=F4,0(t)+1N1/3F4,1(t)+{\rm Pr}\left({l_{N}^{\boxslash}-1-2\sqrt{N}\over N^{1/6}}\leq t\right)=F_{4,0}(t^{*})+{1\over N^{1/3}}F_{4,1}(t)+\cdots
Refer to captiontδ1(t)\delta_{1}(t)
Refer to captiontδ4(t)\delta_{4}(t)
Figure 12. Comparison of δ1(t)\delta_{1}(t) from (4.25) and δ4(t)\delta_{4}(t) from (4.26) rescaled by t=(l±12N)/N1/6t=(l\pm 1-2\sqrt{N})/N^{1/6}, with N=400N=400 [black dots], 2000020000 [blue crosses] and 10510^{5} [red circles].

Finally, we comment on the analogues of (4.18) in relation to the large NN expansion of the mean and variance. We proceeded as for lNl_{N}^{\Box} and postulated that the quantities μ^1(N),σ^12(N),μ^4(N),σ^42(N)\hat{\mu}_{1}(N),\,\hat{\sigma}^{2}_{1}(N),\,\hat{\mu}_{4}(N),\,\hat{\sigma}^{2}_{4}(N) each have a large NN expansion of the form c+dN1/3+c+dN^{-1/3}+\cdots. The exact data for lN,lNl^{\boxbslash}_{N},\,l^{\boxslash}_{N} available for NN up to 400 was then used to find best fits for the corresponding values of cc and dd. However, as distinct from our findings in the case of lNl_{N}^{\Box} seen in Figures 8 and 9, when calculating the differences analogous to (4.20) a decrease to zero as NN increased was not observed. Hence, as yet we do not have convincing evidence for NN dependence of higher order terms in the large NN expansion of the mean and standard deviation of lN,lNl^{\boxbslash}_{N},\,l^{\boxslash}_{N}.

Acknowledgements

This research is part of the program of study supported by the Australian Research Council Centre of Excellence ACEMS and the Discovery Project grant DP210102887. We thank Eric Rains for providing the computer program as referenced in the text. We are most grateful to Jinho Baik for bringing to our attention the crucial reference [6], which unfortunately was missed when we prepared the first draft of this work. Also, we acknowledge the contribution of Allan Trinh in collaborating in the early stages of this project.

References

  • [1] M. Abramowitz and I. A. Stegun, editors. Handbook of Mathematical Functions, Dover, New York (1972).
  • [2] D. Aldous and P. Diaconis, Longest increasing subsequences: from patience sorting to the Baik-Deift-Johansson theorem, Bull. Amer. Math. Soc. 36 (1999), 413–432.
  • [3] J. Arndt, Matters Computational: Ideas, algorithms and source code, Springer, Heidelberg (2011).
  • [4] R.M. Baer and P. Brock, Natural sorting over permutation spaces, Math. Comp. 22 (1968), 385–410.
  • [5] J. Baik, P. Deift, and K. Johansson, On the distribution of the length of the longest increasing subsequence of random permutations, J. Amer. Math. Soc. 12 (1999), 1119–1178.
  • [6] J. Baik and R. Jenkins, Limiting distribution of maximal crossing and nesting of Poissonized random matchings, Ann. Probab., 41 (2013), 4359–4406.
  • [7] J. Baik and E.M. Rains, Symmetrized random permutations, Random matrix models and their applications (P.M. Bleher and A.R. Its, eds.), Mathematical Sciences Research Institute Publications, vol. 40, Cambridge University Press, Cambridge (2001), pp. 171–208.
  • [8] J. Baik and E.M. Rains, Algebraic aspects of increasing subsequences, Duke Math. J. 109 (2001), 1–65.
  • [9] J. Baik and E.M. Rains, The asymptotics of monotone subsequences of involutions, Duke Math. J. 109 (2001), 205–281.
  • [10] E. Bogomolny, O. Bohigas, P. Leboeuf, and A.C. Monastra, On the spacing distribution of the Riemann zeros: corrections to the asymptotic result, J. Phys. A 39 (2006), 10743–10754.
  • [11] F. Bornemann, On the numerical evaluation of distributions in random matrix theory: a review, Markov Processes Relat. Fields 16 (2010), 803–866.
  • [12] F. Bornemann, P. Forrester and A. Mays. Finite size effects for spacing distributions in random matrix theory: circular ensembles and Riemann zeros. Stud. Appl. Math., 138 (2017), 401–437.
  • [13] A. Borodin and P.J. Forrester, Increasing subsequences and the hard-to-soft transition in matrix ensembles, J.Phys. A 36 (2003), 2963–2981.
  • [14] P. Desrosiers and P.J. Forrester, Relationships between τ\tau-functions and Fredholm determinant expressions for gap probabilities in random matrix theory, Nonlinearity 19 (2006), 1643–1656.
  • [15] P. Desrosiers and P.J. Forrester, Hermite and Laguerre β\beta-ensembles: asymptotic corrections to the eigenvalue density, Nucl. Phys. B 743 (2006), 307–332.
  • [16] F.J. Dyson, Statistical theory of energy levels of complex systems III, J. Math. Phys. 3 (1962), 166–175.
  • [17] P.L. Ferrari and R. Frings, Finite time corrections in KPZ growth models, J. Stat. Phys. 144 (2011), 1123–1150.
  • [18] P.J. Forrester, The spectrum edge of random matrix ensembles, Nucl. Phys. B 402 (1993), 709–728.
  • [19] P.J. Forrester, Exact results and universal asymptotics in the Laguerre random matrix ensemble, J. Math. Phys. 35 (1993), 2539–2551.
  • [20] P.J. Forrester, Painlevé transcendent evaluation of the scaled distribution of the smallest eigenvalue in the Laguerre orthogonal and symplectic ensembles, arXiv:nlin.SI/0005064 (2000).
  • [21] P.J. Forrester, Hard and soft edge spacing distributions for random matrix ensembles with orthogonal and symplectic symmetry, Nonlinearity 19 (2006), 2989–3002.
  • [22] P.J. Forrester, Log-gases and random matrices, Princeton University Press, Princeton, NJ (2010).
  • [23] P.J. Forrester, S.-H. Li and A.K. Trinh, Asymptotic correlations with corrections for the circular Jacobi β\beta-ensemble, J. Approximation Th. 271 (2021), 105633.
  • [24] P.J. Forrester and A. Mays. Finite-size corrections in random matrix theory and Odlyzko’s dataset for the Riemann zeros. Proc. R. Soc. A, 471 (2015), 20150436.
  • [25] P.J. Forrester, J.H.H. Perk, A.K. Trinh and N.S. Witte, Leading corrections to the scaling function on the diagonal for the two-dimensional Ising model, J. Stat. Mech. 2019 (2019), 023106.
  • [26] P.J. Forrester and A.K. Trinh. Functional form for the leading correction to the distribution of the largest eigenvalue in the GUE and LUE. J. Math. Phys., 59(5) (2018), 053302.
  • [27] P.J. Forrester and A.K. Trinh, Finite-size corrections at the hard edge for the Laguerre β\beta ensemble, Stud. Applied Math. 143 (2019), 315–336.
  • [28] P.J. Forrester and A. Trinh. Optimal soft edge scaling variables for the Gaussian and Laguerre even β\beta ensembles. Nuclear Phys. B, 938 (2019), 621–639.
  • [29] P.J. Forrester and N.S. Witte, Application of the τ\tau-function theory of Painlevé equations to random matrices: PVI, the JUE,CyUE, cJUE and scaled limits, Nagoya Math. J. 174 (2004), 29–114.
  • [30] I.M. Gessel, Symmetric functions and pp-recursiveness, J. Comb. Th. A 53 (1990), 257–285.
  • [31] A.T. James, Distributions of matrix variate and latent roots derived from normal samples, Ann. Math. Statist. 35 (1964), 475–501.
  • [32] K. Johansson, The longest increasing subsequence in a random permutation and a unitary random matrix model, Math. Research Lett. 5 (1998), 63–82.
  • [33] J.P. Keating and N.C. Snaith, Random matrix theory and ζ(1/2+it)\zeta(1/2+it), Commun. Math. Phys. 214 (2001), 57–89.
  • [34] B.F. Logan and L.A. Shepp, A variational problem for random Young tableaux, Advances in Math., 26 (1977), 206–222.
  • [35] A.M. Odlyzko, The 102210^{22}-nd zero of the Riemann zeta function, Dynamical, Spectral, and Arithmeitc Zeta Functions (M. van Frankenhuysen and M.L. Lapidus, eds.), Contemporary Math. 2001, Amer. Math. Soc, Providence, RI, (2001), 139–144.
  • [36] A.M. Odlyzko and E.M. Rains, On Longest Increasing Subsequences in Random Permutations, Amer. Math. Soc., Contemp. Math. 251 (2000), 439–451.
  • [37] E.M. Rains, Increasing subsequences and the classical groups, Elect. J. of Combinatorics 5 (1998), #R12.
  • [38] J. Ramirez and B. Rider, Diffusion at the random matrix hard edge, Commun. Math. Phys. 288 (2009), 887–906.
  • [39] D. Romik, The surprising mathematics of longest increasing subsequences, Institute of Mathematical Statistics Textbooks. Cambridge University Press (2015).
  • [40] V. Sanaka, Longest Increasing Subsequence Size (N log N), Geeks for Geeks (2012), url:
    www.geeksforgeeks.org/
    longest-monotonically-increasing-subsequence-size-n-log-n/.
  • [41] P. Sarnak, Problems of the Millennium: The Riemann Hypothesis, Clay Mathematics Institute Annual Report (2004), 5–21.
  • [42] T. Sasamoto, Spatial correlations of the 1D KPZ surface on a flat substrate, J. Phys. A 38 (2005), L549–L556.
  • [43] C.A. Tracy and H. Widom, Level-spacing distributions and the Airy kernel, Commun. Math. Phys. 159 (1994), 151–174.
  • [44] C.A. Tracy and H. Widom, On orthogonal and symplectic matrix ensembles, Commun. Math. Phys. 177 (1996), 727–754.
  • [45] C.A. Tracy and H. Widom, Level-spacing distributions and the Bessel kernel, Commun. Math. Phys. 161 (1994), 289–309.
  • [46] E.T. Whittaker and G.N. Watson, A course of modern analysis, 2nd ed., Cambridge University Press, Cambridge (1965).