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

Lyapunov exponents for products of truncated orthogonal matrices

Dong Qichao111dqccha@mail.ustc.edu.cn
Abstract

This article gives a non-asymptotic analysis of the largest Lyapunov exponent of truncated orthogonal matrix products. We prove that as long as NN, the number of terms in product, is sufficiently large, the largest Lyapunov exponent is asympototically Gaussian. Futhermore, the sum of finite largest Lyapunov exponents is asympototically Gaussian, where we use Weingarten Calculus.

1 Introduction

1.1 Main results

Let RiR_{i} be independent Haar distributed random real orthogonal matrices of size (li+n)×(li+n)(l_{i}+n)\times(l_{i}+n) and AiA_{i} be the top n×nn\times n subblock of RiR_{i}, where li>0l_{i}>0. We consider random matrix products

XN,n:=ANA1.X_{N,n}:=A_{N}\cdots A_{1}. (1)

Let s1sns_{1}\geq\cdots\geq s_{n} be the singular values of XN,nX_{N,n}, the Lyapunov exponents of XN,nX_{N,n} are defined as

λi=λi(XN,n):=1Nlogsi(XN,n).\lambda_{i}=\lambda_{i}\left(X_{N,n}\right):=\frac{1}{N}\log s_{i}\left(X_{N,n}\right). (2)

We prove that as long as NN is sufficiently large as a function of n,lin,l_{i}, the largest Lyapunov exponent

λ1=λ1(XN,n):=1Nlogs1(XN,n)\lambda_{1}=\lambda_{1}\left(X_{N,n}\right):=\frac{1}{N}\log s_{1}\left(X_{N,n}\right)

of XN,nX_{N,n} is asymptotically Gaussian (see Theorem 1). Our estimation provides quantitative concentration estimates when NN is large but finite even when nn grows with NN.

Let us give some notations. Define l=min1iNlil=\min\limits_{1\leq i\leq N}l_{i}, L=max1iNliL=\max\limits_{1\leq i\leq N}l_{i}, furthermore,

μn,li=𝔼log(Beta(n/2,l/2))=ψ(n/2)ψ((li+n)/2),\displaystyle\mu_{n,l_{i}}=\mathbb{E}\log\left(\operatorname{Beta}(n/2,l/2)\right)=\psi(n/2)-\psi((l_{i}+n)/2), (3)
σn,li=Varlog(Beta(n/2,li/2))=ψ1(n/2)ψ1((li+n)/2),\displaystyle\sigma_{n,l_{i}}=\mathrm{Var}\log\left(\operatorname{Beta}(n/2,l_{i}/2)\right)=\psi_{1}(n/2)-\psi_{1}((l_{i}+n)/2), (4)
μn,l\displaystyle\mu_{n,l} =1Ni=1Nμn,li,\displaystyle=\frac{1}{N}\sum_{i=1}^{N}\mu_{n,l_{i}}, (5)
Σn,l\displaystyle\Sigma_{n,l} =1N2i=1Nσn,li.\displaystyle=\frac{1}{N^{2}}\sum_{i=1}^{N}\sigma_{n,l_{i}}. (6)

Our main results are as follows

Theorem 1.1.

Suppose XN,nX_{N,n} is given as in (1), there exists constant C>0C>0 such that

dKS(λ1μn,lΣn,l,𝒩(0,1))(4Clog2nlog2(N/n)n(n+l)2lN)1/2,d_{KS}\left(\frac{\lambda_{1}-\mu_{n,l}}{\Sigma_{n,l}},\mathcal{N}\left(0,1\right)\right)\leq\left(\frac{4C\log^{2}n\log^{2}(N/n)n(n+l)}{2lN}\right)^{1/2}, (7)

then λ1\lambda_{1} is approximately Gaussian when N is sufficiently large as a function of n,ln,l. Here 𝒩(μ,Σ)\mathcal{N}(\mu,\Sigma) denotes a Gaussian with mean μ\mu and co-variance Σ\Sigma, dKSd_{KS} is the Kolmogorov-Smirnoff distance.

Futhermore, for kk dimension case we have

Theorem 1.2.

Suppose XN,nX_{N,n} is as in (1), k is finite, then the sum of the top k Lyapunov exponents λ1++λk\lambda_{1}+\cdots+\lambda_{k} is approximately Gaussian.

Remark: Our results can be extended to truncated unitary matrices.

1.2 Prior work

Furstenberg and Kesten first proved that λ1\lambda_{1} converges provided that 𝔼log+(Ai)<{\mathbb{E}}\log_{+}\left(\|A_{i}\|\right)<\infty in paper [10]. Oseledec proved convergence of other singular values later in paper [19], which is referred as multiplicative ergodic theorem. Cohen and Newman in [18] studied the behavior of the limit in the situation when N approaches infinity. Moreover, the work of LePage [15] and subsequent work [7] showed that the top Lyapunov exponent of matrix products such as XN,nX_{N,n} (not necessarily Gaussian) is asymptotically normal. To the best of our knowledge, all known mathematical proofs of asymptotic normality results hold only for finite fixed nn and do not include quantitative rates of convergence. For the case we study, we overcomes these deficiencies.

When we consider t=nNt=\frac{n}{N} as a time parameter in an interacting particle system, there are also interesting work. A number of remarkable articles [1, 2, 3] and especially [4] establish a correspondence between tt and the time parameter in the stochastic evolution of an interacting particle system. This correspondence between singular values for products of complex Ginibre matrices and DBM appears to be initially due to Maurice Duits.

A rigorous analysis of the determinental kernel for the joint distribution of singular values for products of complex Gaussian matrices was undertaken in a variety of articles.In particular, [16] shows that when NN is arbitrary but fixed and nn\rightarrow\infty, the determinental kernel for singular values in products of NN iid complex Gaussian matrices of size n×nn\times n converges to the familiar sine and Airy kernels that arise in the local spectral statistics of large GUE matrices in the bulk and edge, respectively. Moreover, [16] rigorously obtained an expression for the limiting determinental kernel when t=nNt=\frac{n}{N} is arbitrary in the context of products of complex Ginibre. XN,nX_{N,n} around the triangle law always converge to a Gaussian field.

We also refer the reader to [5], which obtains a CLT for linear statistics of top singular values when n/Nn/N is fixed and finite. For the real Gaussian case, [11] provides non-asymptotic analysis of the singular values, which inspires this article.

1.3 Strategy of proof

Basically, we follow the strategy in [11] . By reduction to small ball estimates for volumes of random projections ( Proposition 8.1 and Lemma 8.2 in [11]), we can estimate the difference

1NlogXN,n(Θ)supΘ1NlogXN,n(Θ)=1NlogXN,n(Θ)λ1.\frac{1}{N}\log\left\|X_{N,n}(\Theta)\right\|-\sup_{\Theta\in{\mathbb{R}}}\frac{1}{N}\log\left\|X_{N,n}(\Theta)\right\|=\frac{1}{N}\log\left\|X_{N,n}(\Theta)\right\|-\lambda_{1}. (8)
Proposition 1.3 ([11]Proposition 8.1).

There exists C>0C>0 with the following property. For any ε(0,1)\varepsilon\in(0,1) and any ΘFrn,k\Theta\in\operatorname{Fr}_{n,k} we have

(|1NlogXN,n(Θ)supΘFrn,k1NlogXN,n(Θ)||k2Nlog(nkε2))(Cε)k/2,\mathbb{P}\left(\left.\left|\frac{1}{N}\log\left\|X_{N,n}(\Theta)\right\|-\sup_{\Theta^{\prime}\in\operatorname{Fr}_{n,k}}\frac{1}{N}\log\|X_{N,n}\left(\Theta^{\prime}\right)\right|\right\rvert\,\geq\frac{k}{2N}\log\left(\frac{n}{k\varepsilon^{2}}\right)\right)\leq(C\varepsilon)^{k/2}, (9)

We use a high-dimensional version of Kolmogorov-Simirnov distance to measure normality,

d(X,Y):=supC𝒞k|(XC)(YC)|.d(X,Y):=\sup_{C\in\mathcal{C}_{k}}|\mathbb{P}(X\in C)-\mathbb{P}(Y\in C)|.

We have the following :

Proposition 1.4 ([11]Proposition 6.3).

There exists c>0c>0 with the following property. Suppose that X,YX,Y are k\mathbb{R}^{k}-valued random variables defined on the same probability space. For all μk\mu\in\mathbb{R}^{k}, invertible symmetric matrices ΣSymk+\Sigma\in\mathrm{Sym}_{k}^{+}, and δ>0\delta>0 we have

d(X+Y,𝒩(μ,Σ))3d(X,𝒩(μ,Σ))+cδΣ1HS+2(Y2>δ).d(X+Y,\mathcal{N}(\mu,\Sigma))\leq 3d(X,\mathcal{N}(\mu,\Sigma))+c\delta\sqrt{\left\|\Sigma^{-1}\right\|_{HS}}+2\mathbb{P}\left(\|Y\|_{2}>\delta\right).

We follow the notation in Bentkus [6] with one dimension case and define

S:=SN=X1++XN,S:=S_{N}=X_{1}+\cdots+X_{N},

where X1,,XNX_{1},\ldots,X_{N} are independent random varibles in k\mathbb{R}^{k} with common mean 𝔼Xj=\mathbb{E}X_{j}= 0 . We set

C:=cov(S)C:=\operatorname{cov}(S)

to be the covariance matrix of SS, which assumed to be invertible. With the definition

βj:=𝔼C12Xj23,β:=j=1Nβj,\beta_{j}:=\mathbb{E}\left\|C^{-\frac{1}{2}}X_{j}\right\|_{2}^{3},\beta:=\sum_{j=1}^{N}\beta_{j},

we have the following :

Theorem 1.5 ([6]Theorem 1.1).

There exists an absolute constant c>0c>0 such that

d(S,C12Z)ck14β,d\left(S,C^{\frac{1}{2}}Z\right)\leq ck^{\frac{1}{4}}\beta, (10)

where Z𝒩(0,Idk)Z\sim\mathcal{N}\left(0,\mathrm{Id}_{k}\right) denotes a standard Gaussian on k\mathbb{R}^{k}.

We use Proposition 1.4 and Theorem 1.5 to derive Theorem 1.1.

For Theorem 1.2, we use integration with respect to the Haar measure on orthogonal group, so-called Weingarten calculus, to estimate moments of the sum of Lyapunov exponents, which is new technique in this area.

2 Proof for Theorem 1.1

Let Ai,i=1,,N{A_{i},i=1,\cdots,N}, is drawn from rotationally invariant ensemble, so [11]lemma 8 and proposition 8.1 still can be used. According to [11] lemma 9.5, which is standard technique in this area. We have

logXN,n(Θ)\displaystyle\log\left\|X_{N,n}(\Theta)\right\| =logANA1(Θ)\displaystyle=\log\left\|A_{N}\cdots A_{1}(\Theta)\right\| (11)
=logANA2A1(Θ)A1(Θ)+logA1(Θ).\displaystyle=\log\left\|A_{N}\cdots A_{2}\frac{A_{1}(\Theta)}{\left\|A_{1}(\Theta)\right\|}\right\|+\log\left\|A_{1}(\Theta)\right\|. (12)

Moreover, A2(A1(Θ)A1(Θ))A_{2}\left(\frac{A_{1}(\Theta)}{\left\|A_{1}(\Theta)\right\|}\right) is independent of A3,,ANA_{3},\cdots,A_{N} and

A2(A1(Θ)A1(Θ))=dA2(Θ),A_{2}\left(\frac{A_{1}(\Theta)}{\left\|A_{1}(\Theta)\right\|}\right)\stackrel{{\scriptstyle d}}{{=}}A_{2}(\Theta), (13)

the above equations use the rotationally invariance of Ai,i=1,,N{A_{i},i=1,\cdots,N}.
In conclusion, we have

logXN,n(Θ)=di=1NlogAi(Θ)\log\left\|X_{N,n}(\Theta)\right\|\stackrel{{\scriptstyle d}}{{=}}\sum_{i=1}^{N}\log\left\|A_{i}(\Theta)\right\| (14)

In order to do precise calculation, we assume Θ\Theta is standard orthogonal basis and consider Θ\Theta is 1 dimensional.
Here we recall a well-known result, Haar distributed orthogonal matrix can be derived from Gram-Schmidt transformation of Ginibre matrix, where entries are all real standard Gaussian variables, see [9] for more details. Then

logAi(Θ)=logζi,\log\left\|A_{i}(\Theta)\right\|=\log\|\zeta_{i}\|, (15)

where ζi\zeta_{i} is the first row of AiA_{i}. Furthermore,

ζi2=x12++xn2x2\displaystyle{\|\zeta_{i}\|}^{2}=\frac{{x_{1}}^{2}+\cdots+{x_{n}}^{2}}{{\|\vec{x}\|}^{2}}

where x\vec{x} is a vector of independent standard Gaussian variables. Then we have

ζ12Beta(n/2,l/2).\displaystyle{\|\zeta_{1}\|}^{2}\sim\operatorname{Beta}(n/2,l/2).

since if XGamma(a,θ)X\sim\operatorname{Gamma}(\mathrm{a},\theta) and YGamma(β,θ)Y\sim\operatorname{Gamma}(\beta,\theta) are independent, then XX+YBeta(α,β)\frac{X}{X+Y}\sim\operatorname{Beta}(\alpha,\beta).
According to [11] Lemma 9.5, we have in distribution that

2NlogXN,n(Θ)=1Ni=1NTi,\frac{2}{N}\log\left\|X_{N,n}(\Theta)\right\|=\frac{1}{N}\sum_{i=1}^{N}T_{i}, (16)

where TiT_{i} are independent and

Tilog(Beta(n/2,li/2)).T_{i}\sim\log\left(\operatorname{Beta}(n/2,l_{i}/2)\right). (17)

We already know that

μn,li=𝔼log(Beta(n/2,li/2))=ψ(n/2)ψ((li+n)/2),\displaystyle\mu_{n,l_{i}}=\mathbb{E}\log\left(\operatorname{Beta}(n/2,l_{i}/2)\right)=\psi(n/2)-\psi((l_{i}+n)/2), (18)
σn,li=Varlog(Beta(n/2,li/2))=ψ1(n/2)ψ1((li+n)/2),\displaystyle\sigma_{n,l_{i}}=\mathrm{Var}\log\left(\operatorname{Beta}(n/2,l_{i}/2)\right)=\psi_{1}(n/2)-\psi_{1}((l_{i}+n)/2), (19)

where ψ(z)=ddzlogΓ(z)=Γ(z)Γ(z),ψ1(z)=ddzψ(z)\psi(z)=\frac{\mathrm{d}}{\mathrm{d}z}\log\Gamma(z)=\frac{\Gamma^{\prime}(z)}{\Gamma(z)},\psi_{1}(z)=\frac{\mathrm{d}}{\mathrm{d}z}\psi(z).

We find in particular that

𝔼[1NlogXN,n(Θ)]=μn,l\displaystyle\mathbb{E}\left[\frac{1}{N}\log\left\|X_{N,n}(\Theta)\right\|\right]=\mu_{n,l} (20)

for any random variable YY and we will use the shorthand

Y¯:=Y𝔼[Y].\bar{Y}:=Y-\mathbb{E}[Y].

We will apply Markov’s inequality to bound moments of the sum of TiT_{i}’s, so we do the following estimate.

Proposition 2.1.

There exists a universal constant CC so that for any n,N,ln,N,l and p1p\geq 1

(𝔼[|i=1NT¯i|p])1/pC(pj=1N1MN+pMN)C(pNM+pM),\displaystyle\left(\mathbb{E}\left[\left|\sum_{i=1}^{N}\bar{T}_{i}\right|^{p}\right]\right)^{1/p}\leq C\left(\sqrt{p\sum_{j=1}^{N}\frac{1}{M_{N}}}+\frac{p}{M_{N}}\right)\leq C\left(\sqrt{\frac{pN}{M}}+\frac{p}{M}\right), (21)

where MN=n2(n+lN+2)(n+lN)2,M=n2(n+L+2)(n+L)2.M_{N}=\frac{n^{2}(n+l_{N}+2)}{(n+l_{N})^{2}},M=\frac{n^{2}(n+L+2)}{(n+L)^{2}}.

Lemma 2.2.

There exists a universal constant CC so that

(𝔼[|T¯i|p])1/pCpMiCpM.\left(\mathbb{E}\left[\left|\bar{T}_{i}\right|^{p}\right]\right)^{1/p}\leq C\sqrt{\frac{p}{M_{i}}}\leq C\sqrt{\frac{p}{M}}. (22)

Proof.

We first make a reduction. We now verify that the estimates established in Lemma 1 for T¯i=Ti𝔼[log(Beta(n/2,li/2))]\bar{T}_{i}=T_{i}-\mathbb{E}\left[\log\left(\operatorname{Beta}(n/2,l_{i}/2)\right)\right] can be derived directly from the corresponding estimates for

T^i:=Tilog(nn+li).\widehat{T}_{i}:=T_{i}-\log\left(\frac{n}{n+l_{i}}\right).

We have

𝔼[log(Beta(n/2,li/2))]=ψ(n/2)ψ((li+n)/2)=log(nn+li)+ε,ε=O(n1),\mathbb{E}\left[\log\left(\operatorname{Beta}(n/2,l_{i}/2)\right)\right]=\psi(n/2)-\psi((l_{i}+n)/2)=\log\left(\frac{n}{n+l_{i}}\right)+\varepsilon,\quad\varepsilon=O\left(n^{-1}\right),

where ψ\psi is the digamma function, and we have used its asymptotic expansion ψ(z)\psi(z)\sim log(z)+O(z1)\log(z)+O\left(z^{-1}\right) for large arguments. Thus, we have for each ii that

𝔼[|T¯i|p]=𝔼[|T^i+ε|p]k=0p(pk)𝔼[|T^i|k]|ε|pk.\mathbb{E}\left[\left|\bar{T}_{i}\right|^{p}\right]=\mathbb{E}\left[\left|\widehat{T}_{i}+\varepsilon\right|^{p}\right]\leq\sum_{k=0}^{p}\left(\begin{array}[]{l}p\\ k\end{array}\right)\mathbb{E}\left[\left|\widehat{T}_{i}\right|^{k}\right]\left|\varepsilon\right|^{p-k}.

So assuming that T^i\widehat{T}_{i} satisfy the conclusion of Lemma 1 , we find

𝔼[|T¯i|p]k=0p(pk)ζkk|ε|pk,ζk:=CkM.\mathbb{E}\left[\left|\bar{T}_{i}\right|^{p}\right]\leq\sum_{k=0}^{p}\left(\begin{array}[]{l}p\\ k\end{array}\right)\zeta_{k}^{k}\left|\varepsilon\right|^{p-k},\quad\zeta_{k}:=C\sqrt{\frac{k}{M}}.

Since for 0kp0\leq k\leq p we have ζkζp\zeta_{k}\leq\zeta_{p}, we see that

𝔼[|T¯i|p]k=0p(pk)ζpk|ε|pk(ζp+|ε|)p.\mathbb{E}\left[\left|\bar{T}_{i}\right|^{p}\right]\leq\sum_{k=0}^{p}\left(\begin{array}[]{l}p\\ k\end{array}\right)\zeta_{p}^{k}\left|\varepsilon\right|^{p-k}\leq\left(\zeta_{p}+\left|\varepsilon\right|\right)^{p}.

Finally, since ε=O(n1)=o(ζp)\varepsilon=O\left(n^{-1}\right)=o\left(\zeta_{p}\right), we find that there exists C>0C>0 so that

(𝔼[|T¯i|p])1/pCζpCpMi\left(\mathbb{E}\left[\left|\bar{T}_{i}\right|^{p}\right]\right)^{1/p}\leq C\zeta_{p}\leq C\sqrt{\frac{p}{M_{i}}}

as desired. It therefore remains to show that T^i=Tilog(nn+li)\widehat{T}_{i}=T_{i}-\log\left(\frac{n}{n+l_{i}}\right) satisfies the conclusion of Lemma 1. To do this, we begin by checking that with MiM_{i}, for all s0s\geq 0

(|Tilog(nn+li)|s)2eMis2.\mathbb{P}\left(\left|T_{i}-\log\left(\frac{n}{n+l_{i}}\right)\right|\geq s\right)\leq 2e^{-M_{i}s^{2}}. (23)

We have

(|Tilog(nn+li)|s)\displaystyle\mathbb{P}\left(\left|T_{i}-\log\left(\frac{n}{n+l_{i}}\right)\right|\geq s\right) =(|log(Beta(n/2,li/2))log(nn+li)|s)\displaystyle=\mathbb{P}\left(\left|\log\left(\operatorname{Beta}(n/2,l_{i}/2)\right)-\log\left(\frac{n}{n+l_{i}}\right)\right|\geq s\right)
=(|log(n+linBeta(n/2,li/2))|s)\displaystyle=\mathbb{P}\left(\left|\log\left(\frac{n+l_{i}}{n}\operatorname{Beta}(n/2,l_{i}/2)\right)\right|\geq s\right)
=(Beta(n/2,li/2)nn+lies)+(Beta(n/2,li/2)nn+lies).\displaystyle=\mathbb{P}\left(\operatorname{Beta}(n/2,l_{i}/2)\geq\frac{n}{n+l_{i}}e^{s}\right)+\mathbb{P}\left(\operatorname{Beta}(n/2,l_{i}/2)\leq\frac{n}{n+l_{i}}e^{-s}\right).

Let us first bound (Beta(n/2,li/2)nn+lies)\mathbb{P}\left(\operatorname{Beta}(n/2,l_{i}/2)\geq\frac{n}{n+l_{i}}e^{s}\right). Notice that the mean of Beta(n/2,li/2)\operatorname{Beta}(n/2,l_{i}/2) is nn+li\frac{n}{n+l_{i}} and that Beta(n/2,li/2)\operatorname{Beta}(n/2,l_{i}/2) are subgaussian random variables. Thus, we can use Bernstein’s tail estimates for Beta variables. According to Theorem 1 [20],

Theorem 2.3 ([20]Theorem 1).

Let XBeta(α,β)X\sim\operatorname{Beta}(\alpha,\beta). Define the parameters

vαβ(α+β)2(α+β+1),\displaystyle v\triangleq\frac{\alpha\beta}{(\alpha+\beta)^{2}(\alpha+\beta+1)},
c2(βα)(α+β)(α+β+2).\displaystyle c\triangleq\frac{2(\beta-\alpha)}{(\alpha+\beta)(\alpha+\beta+2)}.

Then the upper tail of XX is bounded as

𝐏{X>𝐄[X]+ϵ}{exp(ϵ22(v+ϵϵ)3)),βα,exp(ϵ22v),β<α,\mathbf{P}\{X>\mathbf{E}[X]+\epsilon\}\leqslant\begin{cases}\exp\left(-\frac{\epsilon^{2}}{2\left(v+\frac{\epsilon\epsilon)}{3}\right)}\right),&\beta\geqslant\alpha,\\ \exp\left(-\frac{\epsilon^{2}}{2v}\right),&\beta<\alpha,\end{cases}

and the lower tail of XX is bounded as

𝐏{X<𝐄[X]ϵ}{exp(ϵ22(v+ϵϵ3)),αβ,exp(ϵ22v),α<β.\mathbf{P}\{X<\mathbf{E}[X]-\epsilon\}\leqslant\begin{cases}\exp\left(-\frac{\epsilon^{2}}{2\left(v+\frac{\epsilon\epsilon}{3}\right)}\right),&\alpha\geqslant\beta,\\ \exp\left(-\frac{\epsilon^{2}}{2v}\right),&\alpha<\beta.\end{cases}

We have vv for XBeta(n/2,li/2)X\sim\operatorname{Beta}(n/2,l_{i}/2) is 12(n+li+2)\frac{1}{2(n+l_{i}+2)}, then

𝐏{X>𝐄[X]+ϵ}exp(ϵ22v).\displaystyle\mathbf{P}\{X>\mathbf{E}[X]+\epsilon\}\leqslant\exp\left(-\frac{\epsilon^{2}}{2v}\right).

Thus

(Beta(n/2,li/2)nn+lies)\displaystyle\mathbb{P}\left(\operatorname{Beta}(n/2,l_{i}/2)\geq\frac{n}{n+l_{i}}e^{s}\right) =(Beta(n/2,l/2)nn+linn+li(es1))\displaystyle=\mathbb{P}\left(\operatorname{Beta}(n/2,l/2)-\frac{n}{n+l_{i}}\geq\frac{n}{n+l_{i}}\left(e^{s}-1\right)\right)
exp(Mis2),\displaystyle\leq\exp\left(-M_{i}s^{2}\right),

where we use es1se^{s}-1\geq s and define Mi=n2(n+li+2)(n+li)2M_{i}=\frac{n^{2}(n+l_{i}+2)}{(n+l_{i})^{2}}. (Beta(n/2,li/2)nn+lies)\mathbb{P}\left(\operatorname{Beta}(n/2,l_{i}/2)\leq\frac{n}{n+l_{i}}e^{-s}\right) is similar to bound, thus we have

(|Tilog(nn+li)|s)2eMis2.\mathbb{P}\left(\left|T_{i}-\log\left(\frac{n}{n+l_{i}}\right)\right|\geq s\right)\leq 2e^{-M_{i}s^{2}}.

To complete the proof of lemma 1, we can write

𝔼[|T^i|p]\displaystyle\mathbb{E}\left[\left|\widehat{T}_{i}\right|^{p}\right] =0(|T^i|>x)pxp1𝑑x\displaystyle=\int_{0}^{\infty}\mathbb{P}\left(\left|\widehat{T}_{i}\right|>x\right)px^{p-1}dx
p0ecMix2xp1𝑑x,\displaystyle\leq p\int_{0}^{\infty}e^{-cM_{i}x^{2}}x^{p-1}dx,

the term can be estimated by comparing to the moments of a Gaussian as follows:

p0ecMix2xp1𝑑x\displaystyle p\int_{0}^{\infty}e^{-cM_{i}x^{2}}x^{p-1}dx =p(2cMi)p/20ex2xp1𝑑x\displaystyle=p\left(2cM_{i}\right)^{-p/2}\int_{0}^{\infty}e^{-x^{2}}x^{p-1}dx
p(2cMi)p/20ex2xp1𝑑x\displaystyle\leq p\left(2cM_{i}\right)^{-p/2}\int_{0}^{\infty}e^{-x^{2}}x^{p-1}dx
p(2cMi)p/22p2Γ(p2)\displaystyle\leq p\left(2cM_{i}\right)^{-p/2}2^{\frac{p}{2}}\Gamma\left(\frac{p}{2}\right)
p(p2cMi)p/2,\displaystyle\leq p\left(\frac{p}{2cM_{i}}\right)^{p/2}, (24)

where we used that for z>0z>0 we have Γ(z)zz.\Gamma(z)\leq z^{z}. Taking 1/p1/p powers, we find that there exists C>0C>0 so that

(𝔼[|T^i|p])1/pCpMiCpM,\left(\mathbb{E}\left[\left|\widehat{T}_{i}\right|^{p}\right]\right)^{1/p}\leq C\sqrt{\frac{p}{M_{i}}}\leq C\sqrt{\frac{p}{M}}, (25)

for all p1p\geq 1, since MiM_{i} is decreasing function with respect to lil_{i}. This completes the proof.

We now in a position to prove proposition 3 with lemma 1 in hand. We introduce the following result of R.Latała.

Theorem 2.4.

Let X1,,XNX_{1},\cdots,X_{N} be mean zero, independent r.v. and p1p\geq 1. Then

(𝔼[|j=1NXj|p])1pinf{t>0:j=1Nlog[𝔼|1+Xjt|p]p},\left(\mathbb{E}\left[\left|\sum_{j=1}^{N}X_{j}\right|^{p}\right]\right)^{\frac{1}{p}}\simeq\inf\left\{t>0:\sum_{j=1}^{N}\log\left[\mathbb{E}\left|1+\frac{X_{j}}{t}\right|^{p}\right]\leq p\right\},

where aba\simeq b means there exist universal constants c1,c2c_{1},c_{2} so that c1abac2c_{1}a\leq b\leq ac_{2}. Moreover, if XiX_{i} are also identically distributed, then

(𝔼[|j=1NXj|p])1psupmax{2,pN}spps(Np)1nXis.\left(\mathbb{E}\left[\left|\sum_{j=1}^{N}X_{j}\right|^{p}\right]\right)^{\frac{1}{p}}\simeq\sup_{\max\left\{2,\frac{p}{N}\right\}\leq s\leq p}\frac{p}{s}\left(\frac{N}{p}\right)^{\frac{1}{n}}\left\|X_{i}\right\|_{s}.

We know that

(𝔼[|i=1NT¯i|p])1/pinf{t>0:j=1Nlog[𝔼|1+T¯it|p]p}.\displaystyle\left(\mathbb{E}\left[\left|\sum_{i=1}^{N}\bar{T}_{i}\right|^{p}\right]\right)^{1/p}\simeq\inf\left\{t>0:\sum_{j=1}^{N}\log\left[\mathbb{E}\left|1+\frac{\bar{T}_{i}}{t}\right|^{p}\right]\leq p\right\}. (26)

We will use the notation :

p0=MN2j=1NMj1.p_{0}=M_{N}^{2}\sum_{j=1}^{N}M_{j}^{-1}.

Since

pj=1NMN1pMNpp0,\sqrt{p\sum_{j=1}^{N}M_{N}^{-1}}\leq\frac{p}{M_{N}}\quad\Longleftrightarrow\quad p\geq p_{0}, (27)

we will show that there exists C>0C>0 so that

pp0(𝔼|i=1NT¯i|p)1pCpj=1NMj1p\leq p_{0}\quad\Longrightarrow\quad\left(\mathbb{E}\left|\sum_{i=1}^{N}\bar{T}_{i}\right|^{p}\right)^{\frac{1}{p}}\leq C\sqrt{p\sum_{j=1}^{N}M_{j}^{-1}} (28)

as well as

pp0(𝔼|i=1NT¯i|p)1pCpMN.p\geq p_{0}\quad\Longrightarrow\quad\left(\mathbb{E}\left|\sum_{i=1}^{N}\bar{T}_{i}\right|^{p}\right)^{\frac{1}{p}}\leq C\frac{p}{M_{N}}. (29)

We may assume that without loss of generality that p is even, since we can use higher even moments to control odd moments. Here we recall a well-known estimate

(nk)k(nk)(nk)kek,k1.\left(\frac{n}{k}\right)^{k}\leq\binom{n}{k}\leq\left(\frac{n}{k}\right)^{k}e^{k},\quad k\geq 1.

Since p is even, we can drop the absolute value,

𝔼[(1+T¯it)p]\displaystyle\mathbb{E}\left[\left(1+\frac{\bar{T}_{i}}{t}\right)^{p}\right] =1+l=2p(pl)𝔼[T¯il]tl\displaystyle=1+\sum_{l=2}^{p}\binom{p}{l}\frac{\mathbb{E}\left[\bar{T}_{i}^{l}\right]}{t^{l}}
1+l=2p(pl)1tl(ClM)l\displaystyle\leq 1+\sum_{l=2}^{p}\binom{p}{l}\frac{1}{t^{l}}\left(C\frac{l}{M}\right)^{l}
1+l=2p(pl)l2(Cpt2M)l2.\displaystyle\leq 1+\sum_{l=2}^{p}\left(\frac{p}{l}\right)^{\frac{l}{2}}\left(\frac{Cp}{t^{2}M}\right)^{\frac{l}{2}}. (30)

We now bound the term in the previous line by breaking into two terms where l is even and odd. When l is even, the term in (30) can be bounded by

1+l=2l evenp(pl)l2(Cpt2M)l21+l=2l evenp(p2l2)(Cpt2M)l2(1+Cpt2M)l2.\displaystyle 1+\sum_{\begin{subarray}{c}l=2\\ l\text{ even}\end{subarray}}^{p}\left(\frac{p}{l}\right)^{\frac{l}{2}}\left(\frac{Cp}{t^{2}M}\right)^{\frac{l}{2}}\leq 1+\sum_{\begin{subarray}{c}l=2\\ l\text{ even}\end{subarray}}^{p}\binom{\frac{p}{2}}{\frac{l}{2}}\left(\frac{Cp}{t^{2}M}\right)^{\frac{l}{2}}\leq\left(1+\frac{Cp}{t^{2}M}\right)^{\frac{l}{2}}. (31)

When l is odd, we have fact that for any 0mlp0\leq m\leq l\leq p

(p)pm(pm).\left(\frac{p}{\ell}\right)^{\ell}\leq p^{m}\binom{p}{\ell-m}. (32)

Thus the odd term in (30) is bounded by

=3 odd p(p)/2(Cpt2M)/2minm=1,3{(Cp2t2M)m2=31p1(pm)12(Cpt2M)m2}.\sum_{\begin{subarray}{c}\ell=3\\ \ell\text{ odd }\end{subarray}}^{p}\left(\frac{p}{\ell}\right)^{\ell/2}\left(\frac{Cp}{t^{2}M}\right)^{\ell/2}\leq\min_{m=1,3}\left\{\left(\frac{Cp^{2}}{t^{2}M}\right)^{\frac{m}{2}}\sum_{\begin{subarray}{c}\ell=3\\ \ell-1\end{subarray}}^{p-1}\binom{p}{\ell-m}^{\frac{1}{2}}\left(\frac{Cp}{t^{2}M}\right)^{\frac{\ell-m}{2}}\right\}. (33)

To proceed, note that for any 0ba0\leq b\leq a

(2a2b)2b(ab)2.\binom{2a}{2b}\leq 2^{b}\binom{a}{b}^{2}.

This inequality follows by observing that for any j=0,,b1j=0,\ldots,b-1, we have

(2a2j)(2a2j1)(2b2j1)=(aj)(aj1/2)(bj)(bj1/2)2(ajbj)2,\frac{(2a-2j)(2a-2j-1)}{(2b-2j-1)}=\frac{(a-j)(a-j-1/2)}{(b-j)(b-j-1/2)}\leq 2\left(\frac{a-j}{b-j}\right)^{2},

and repeatedly applying this estimate to the terms in (2a2b)\binom{2a}{2b}. Thus, we obtain

=3 odd p(p)/2(Cpt2M)/2\displaystyle\sum_{\begin{subarray}{c}\ell=3\\ \ell\text{ odd }\end{subarray}}^{p}\left(\frac{p}{\ell}\right)^{\ell/2}\left(\frac{Cp}{t^{2}M}\right)^{\ell/2} minm=1,3{(Cp2t2M)m2}=0p/2(p/2l)(Cpt2M)l\displaystyle\leq\min_{m=1,3}\left\{\left(\frac{Cp^{2}}{t^{2}M}\right)^{\frac{m}{2}}\right\}\sum_{\ell=0}^{p/2}\binom{p/2}{l}\left(\frac{Cp}{t^{2}M}\right)^{l}
=minm=1,3{(Cp2t2M)m2}(1+Cpt2M)p/2\displaystyle=\min_{m=1,3}\left\{\left(\frac{Cp^{2}}{t^{2}M}\right)^{\frac{m}{2}}\right\}\left(1+\frac{Cp}{t^{2}M}\right)^{p/2}
(Cp2t2M)(1+Cpt2M)p/2,\displaystyle\leq\left(\frac{Cp^{2}}{t^{2}M}\right)\left(1+\frac{Cp}{t^{2}M}\right)^{p/2}, (34)

where in the last inequality we’ve used that min{x1/2,x3/2}x\min\left\{x^{1/2},x^{3/2}\right\}\leq x for all x0x\geq 0. In conclusion, we see that there exists C>0C>0 so that

𝔼[(1+T¯it)p](1+Cp2t2M)(1+Cpt2M)p/2.\mathbb{E}\left[\left(1+\frac{\bar{T}_{i}}{t}\right)^{p}\right]\leq\left(1+\frac{Cp^{2}}{t^{2}M}\right)\left(1+\frac{Cp}{t^{2}M}\right)^{p/2}. (35)

Hence, since log(a+b)(loga)+b\log(a+b)\leq(\log a)+b for a1a\geq 1 and b>0b>0,

j=1Nlog𝔼[(1+(T¯it)p)]\displaystyle\sum_{j=1}^{N}\log\mathbb{E}\left[\left(1+\left(\frac{\bar{T}_{i}}{t}\right)^{p}\right)\right] p2j=1Nlog(1+Cpt2M)+j=1Nlog(1+Cp2t2M)\displaystyle\leq\frac{p}{2}\sum_{j=1}^{N}\log\left(1+\frac{Cp}{t^{2}M}\right)+\sum_{j=1}^{N}\log\left(1+\frac{Cp^{2}}{t^{2}M}\right)
p2j=1NCpt2M+j=1NCp2t2M.\displaystyle\leq\frac{p}{2}\sum_{j=1}^{N}\frac{Cp}{t^{2}M}+\sum_{j=1}^{N}\frac{Cp^{2}}{t^{2}M}. (36)

When pp0p\leq p_{0}, set t=Cpi=1NMi1t=\sqrt{C^{\prime}p\sum_{i=1}^{N}M_{i}^{-1}} where

C=max{(16C)2,2C1/2},C^{\prime}=\max\left\{(16C)^{2},2C^{1/2}\right\},

when pp0p\geq p_{0}, set

t=CpMNt=\frac{C^{\prime}p}{M_{N}}

with

C=max{4C,2C1/2},C^{\prime}=\max\left\{4C,2C^{1/2}\right\},

we have the terms in (36) is less than p, which completes the proof of Proposition 2.1.

Proposition 2.5.

There exists a universal constant c>0c>0 with the following property. Fix any vector θ\theta in n\mathbb{R}^{n}, we have

(|1NlogXN,n(θ)μn,l|s)2exp{cNmin{M^s2,MNs}},s>0.\mathbb{P}\left(\left|\frac{1}{N}\log\left\|X_{N,n}(\theta)\right\|-\mu_{n,l}\right|\geq s\right)\leq 2\exp\left\{-cN\min\{\hat{M}s^{2},M_{N}s\}\right\},\quad s>0. (37)

Proof.

To prove Proposition 4, we can see that Proposition 4 is equivalent to showing that for any s>0s>0

(|1Ni=1NT¯i|s)2exp{cNMs2}.\mathbb{P}\left(\left|\frac{1}{N}\sum_{i=1}^{N}\bar{T}_{i}\right|\geq s\right)\leq 2\exp\left\{-cNMs^{2}\right\}. (38)

We remind the readers here that we will use the notation :

p0=MN2j=1NMj1,p_{0}=M_{N}^{2}\sum_{j=1}^{N}M_{j}^{-1},

and note that

pj=1NMj1pMNpp0.\sqrt{p\sum_{j=1}^{N}M_{j}^{-1}}\leq\frac{p}{M_{N}}\quad\Longleftrightarrow\quad p\geq p_{0}.

Thus, applying Markov’s inequality to Proposition 2.1 shows that there exists C>0C>0 so that for 1pp01\leq p\leq p_{0}

(|1Ni=1NT¯i|CNpj=1N1Mj)ep.\mathbb{P}\left(\left|\frac{1}{N}\sum_{i=1}^{N}\bar{T}_{i}\right|\geq\frac{C}{N}\sqrt{p\sum_{j=1}^{N}\frac{1}{M_{j}}}\right)\leq e^{-p}. (39)

Equivalently, set

M^=j=1N1Mj,\hat{M}=\sum_{j=1}^{N}\frac{1}{M_{j}},

we see that there exists c>0c>0 so that

(|1Ni=1NT¯i|s)2ecNM^s2,0sCNMNM^.\mathbb{P}\left(\left|\frac{1}{N}\sum_{i=1}^{N}\bar{T}_{i}\right|\geq s\right)\leq 2e^{-cN\hat{M}s^{2}},\quad 0\leq s\leq\frac{C}{N}M_{N}\hat{M}. (40)

This establishes (37) in this range of ss. To treat sCNMNM^s\geq\frac{C}{N}M_{N}\hat{M}, we again apply Markov’s inequality to Proposition 2.1 to see that there exists C>0C>0 so that

pp0(|i=1NT¯i|>CpMN)ep.p\geq p_{0}\Longrightarrow\mathbb{P}\left(\left|\sum_{i=1}^{N}\bar{T}_{i}\right|>C\frac{p}{M_{N}}\right)\leq e^{-p}.

Hence, there exists c>0c>0 so that

(|1Ni=1NT¯i|s)ecNMNs,sCNMNM^,\mathbb{P}\left(\left|\frac{1}{N}\sum_{i=1}^{N}\bar{T}_{i}\right|\geq s\right)\leq e^{-cNM_{N}s},\quad s\geq\frac{C}{N}M_{N}\hat{M}, (41)

completing the proof.

Turning to the probability in (9), recall that in [11]Proposition 8.1, we have shown that for every ε(0,1)\varepsilon\in(0,1),

(|1nNlogXN,n(Θ)1ni=1kλi|k2Nnlog(nkε2))(Cε)k2.\mathbb{P}\left(\left|\frac{1}{nN}\log\left\|X_{N,n}(\Theta)\right\|-\frac{1}{n}\sum_{i=1}^{k}\lambda_{i}\right|\geq\frac{k}{2Nn}\log\left(\frac{n}{k\varepsilon^{2}}\right)\right)\leq(C\varepsilon)^{\frac{k}{2}}. (42)

If we set s:=knNlogenkε2s:=\frac{k}{nN}\log\frac{en}{k\varepsilon^{2}}, then

(Cε)k/2=exp[14snN+k4log(enk)+k2log(C)].(C\varepsilon)^{k/2}=\exp\left[-\frac{1}{4}snN+\frac{k}{4}\log\left(\frac{en}{k}\right)+\frac{k}{2}\log(C)\right].

Hence, assuming that

sCknNlog(enk)s\geq C^{\prime}\frac{k}{nN}\log\left(\frac{en}{k}\right)

for CC^{\prime} sufficiently large, we arrive to the following expression:

(|1ni=1kλilogXN,n(Θ)nN|s)esnN4,sCknNlog(enk).\mathbb{P}\left(\left|\frac{1}{n}\sum_{i=1}^{k}\lambda_{i}-\frac{\log\left\|X_{N,n}(\Theta)\right\|}{nN}\right|\geq s\right)\leq e^{-\frac{snN}{4}},\quad s\geq C^{\prime}\frac{k}{nN}\log\left(\frac{en}{k}\right). (43)

Now we prove Theorem 1, we follow the proof of Theorem 1.3 in [11], where μn,Σn,l=1Nσn,l2\mu_{n},\Sigma_{n,l}=\frac{1}{N}\sigma_{n,l}^{2} are the mean and variance of 12log(Beta(n/2,l/2)),\frac{1}{2}\log\left(\operatorname{Beta}(n/2,l/2)\right),

d(λ1,𝒩(μn,Σn,l))\displaystyle d\left(\lambda_{1},\mathcal{N}\left(\mu_{n},\Sigma_{n,l}\right)\right)\leq 3d(S^1,𝒩(μn,Σn,l))+c0δ(Σn,l)1HS1/2\displaystyle 3d\left(\widehat{S}_{1},\mathcal{N}\left(\mu_{n},\Sigma_{n,l}\right)\right)+c_{0}\delta\left\|\left(\Sigma_{n,l}\right)^{-1}\right\|_{HS}^{1/2}
+2(S1S^1>δ)\displaystyle+2\mathbb{P}\left(\left\|S_{1}-\widehat{S}_{1}\right\|>\delta\right)
=\displaystyle= 3d(S^1,𝒩(μn,Σn,l))+c0δ(Σn,l)1HS1/2\displaystyle 3d\left(\widehat{S}_{1},\mathcal{N}\left(\mu_{n},\Sigma_{n,l}\right)\right)+c_{0}\delta\left\|\left(\Sigma_{n,l}\right)^{-1}\right\|_{HS}^{1/2}
+2(S1S^1>δ),\displaystyle+2\mathbb{P}\left(\left\|S_{1}-\widehat{S}_{1}\right\|>\delta\right), (44)

where S^1=1NlogXN,n(Θ).\widehat{S}_{1}=\frac{1}{N}\log\left\|X_{N,n}(\Theta)\right\|. For part 1, we use Theorem 1.5 and k=1k=1,

1NlogXN,n(Θ)=1Ni=1NYi,\frac{1}{N}\log\left\|X_{N,n}(\Theta)\right\|=\frac{1}{N}\sum_{i=1}^{N}Y_{i},

where Yi12logBeta(n/2,l/2)Y_{i}\sim\frac{1}{2}\log\operatorname{Beta}(n/2,l/2), and YiY_{i} is log concave variable. Furthermore,

(𝔼(Σn,l)12Yj¯23)13C(𝔼(Σn,l)12Yj¯22)12=C.\left(\mathbb{E}\|(\Sigma_{n,l})^{-\frac{1}{2}}\bar{Y_{j}}\|_{2}^{3}\right)^{\frac{1}{3}}\leq C\left(\mathbb{E}\|(\Sigma_{n,l})^{-\frac{1}{2}}\bar{Y_{j}}\|_{2}^{2}\right)^{\frac{1}{2}}=C. (45)

Thus we have

βi=1N32𝔼(Σn,l)12Yj¯23C3N321iN.\beta_{i}=\frac{1}{N^{\frac{3}{2}}}\mathbb{E}\|(\Sigma_{n,l})^{-\frac{1}{2}}\bar{Y_{j}}\|_{2}^{3}\leq\frac{C^{3}}{N^{\frac{3}{2}}}1\leq i\leq N. (46)

Therefore,

β:=j=1NβjC3N12,\beta:=\sum_{j=1}^{N}\beta_{j}\leq\frac{C^{3}}{N^{\frac{1}{2}}},

and we conclude that there exists an absolute constant c>0c>0 so that

d(S^1,𝒩(μn,Σn,l))cN1/2.d\left(\widehat{S}_{1},\mathcal{N}\left(\mu_{n},\Sigma_{n,l}\right)\right)\leq cN^{-1/2}. (47)

For part 2,

(Σn,l)1HSNψ1(n/2)ψ1((n+l)/2)n(n+l)N2l,\displaystyle\left\|\left(\Sigma_{n,l}\right)^{-1}\right\|_{HS}\leq\frac{N}{\psi_{1}(n/2)-\psi_{1}((n+l)/2)}\sim\frac{n(n+l)N}{2l}, (48)

where we have used its asymptotic expansion ψ1(z)1z+O(z2)\psi_{1}(z)\sim\frac{1}{z}+O\left(z^{-2}\right) for large arguments and ψ1(z)\psi_{1}(z) is deceasing function for z>0z>0, then

(Σn,l)1HS12n(n+l)N2l.\displaystyle\left\|\left(\Sigma_{n,l}\right)^{-1}\right\|_{HS}^{\frac{1}{2}}\sim\sqrt{\frac{n(n+l)N}{2l}}. (49)

For part 3, it is same as [2]. For any collection positive real numbers δ>C1Nlog(en)\delta>C\frac{1}{N}\log\left(en\right) , we therefore have

(S1S^12δ)2eδN/4.\displaystyle\mathbb{P}\left(\left\|S_{1}-\widehat{S}_{1}\right\|_{2}\geq\delta\right)\leq 2e^{-\delta N/4}.

Setting

δ:=CNlog(en)log(Nn)\delta:=\frac{C}{N}\log\left(en\right)\log\left(\frac{N}{n}\right)

for a sufficiently large constant CC we find

(|S1S^1|δj)2eClog(en)log(N/n)2(n/N)1/2.\mathbb{P}\left(\left|S_{1}-\widehat{S}_{1}\right|\geq\delta_{j}\right)\leq 2e^{-C\log(en)\log(N/n)}\leq 2(n/N)^{1/2}. (50)

Hence, as soon as N>nN>n, we have

(SkS^k2δ)C(nN)1/2,\mathbb{P}\left(\left\|\mid S_{k}-\widehat{S}_{k}\right\|_{2}\geq\delta\right)\leq C\left(\frac{n}{N}\right)^{1/2}, (51)

where

δClog(n)log(N/n)N.\delta\leq\frac{C\log(n)\log(N/n)}{N}.

In conclusion,

d(λ1,𝒩(μn,Σn,l))\displaystyle d\left(\lambda_{1},\mathcal{N}\left(\mu_{n},\Sigma_{n,l}\right)\right) C1N+(Clog2nlog2(N/n)n(n+l)2lN)1/2+C(nN)1/2\displaystyle\leq C\sqrt{\frac{1}{N}}+\left(\frac{C\log^{2}n\log^{2}(N/n)n(n+l)}{2lN}\right)^{1/2}+C\left(\frac{n}{N}\right)^{1/2} (52)
(4Clog2nlog2(N/n)n(n+l)2lN)1/2.\displaystyle\leq\left(\frac{4C\log^{2}n\log^{2}(N/n)n(n+l)}{2lN}\right)^{1/2}. (53)

3 Proof for Theorem 1.2

For k dimensions, we remind the reader that

Ai(Θ)=Aiθ1Aiθk,A_{i}(\Theta)=A_{i}\theta_{1}\wedge\cdots\wedge A_{i}\theta_{k}, (54)

where θi\theta_{i} is a fixed k-frame in n{\mathbb{R}}^{n}. Here we set Θ\Theta is the standard k-frames, then the Gram identity reads

A(Θ)=det(A(k)A(k))=σSk(1)|σ|j1,j2jk=1nΠi=1kAjiiAjiσ(i),\displaystyle\left\|A(\Theta)\right\|=\det(A_{(k)}^{*}A_{(k)})=\sum_{\sigma\in S_{k}}(-1)^{\left|\sigma\right|}\sum_{j_{1},j_{2}\cdots j_{k}=1}^{n}\Pi_{i=1}^{k}A_{j_{i}i}A_{j_{i}\sigma(i)}, (55)

where A(k)A_{(k)} denotes the matrix composed of the first k columns of matrix AA and the second equation comes from the definition of matrix determinant. With some observation, we have the following lemma

Lemma 3.1.

For any k

Z=det(A(k)A(k))=σSk(1)|σ|j1j2jkΠi=1kAjiiAjiσ(i)\displaystyle Z=\det(A_{(k)}^{*}A_{(k)})=\sum_{\sigma\in S_{k}}(-1)^{\left|\sigma\right|}\sum_{j_{1}\neq j_{2}\neq\cdots\neq j_{k}}\Pi_{i=1}^{k}A_{j_{i}i}A_{j_{i}\sigma(i)}

Proof.

Without loss of generation, assume j1=j2j_{1}=j_{2}. We can add cycle (12) on σ\sigma, define

σ:=σ(12),\sigma^{{}^{\prime}}:=\sigma(12),

then we have

AjiisjiAjiσ(i)=AjiiAjiσ(i).A_{j_{i}i}s_{j_{i}}A_{j_{i}\sigma^{{}^{\prime}}(i)}=A_{j_{i}i}A_{j_{i}\sigma(i)}.

Moreover, take

B\displaystyle B =σSk(1)|σ|j1,,jk,j1=j2Πi=1kAjiAjiσ(i)\displaystyle=\sum_{\sigma\in S_{k}}(-1)^{\left|\sigma\right|}\sum_{j_{1},\cdots,j_{k},j_{1}=j_{2}}\Pi_{i=1}^{k}A_{j_{i}}A_{j_{i}\sigma(i)}
=σSk(1)|σ|j1,,jk,j1=j2Πi=1kAjiAjiσ(i)\displaystyle=\sum_{\sigma\in S_{k}}(-1)^{\left|\sigma\right|}\sum_{j_{1},\cdots,j_{k},j_{1}=j_{2}}\Pi_{i=1}^{k}A_{j_{i}}A_{j_{i}\sigma^{{}^{\prime}}(i)}
=σSk(1)|σ|j1,,jk,j1=j2Πi=1kAjiAjiσ(i)\displaystyle=\sum_{\sigma^{{}^{\prime}}\in S_{k}}(-1)^{\left|\sigma\right|}\sum_{j_{1},\cdots,j_{k},j_{1}=j_{2}}\Pi_{i=1}^{k}A_{ji}A_{j_{i}\sigma^{{}^{\prime}}(i)}
=σSk(1)|σ|j1,,jk,j1=j2Πi=1kAjiAjiσ(i)\displaystyle=-\sum_{\sigma^{{}^{\prime}}\in S_{k}}(-1)^{\left|\sigma^{{}^{\prime}}\right|}\sum_{j_{1},\cdots,j_{k},j_{1}=j_{2}}\Pi_{i=1}^{k}A_{ji}A_{j_{i}\sigma^{{}^{\prime}}(i)}
=B,\displaystyle=-B,

then B=0B=0, which ends the proof.

We want to control the moments of Z.

𝔼Z\displaystyle{\mathbb{E}}Z =𝔼σSk(1)|σ|j1j2jkΠi=1kAjiiAjiσ(i)\displaystyle={\mathbb{E}}\sum_{\sigma\in S_{k}}(-1)^{\left|\sigma\right|}\sum_{j_{1}\neq j_{2}\neq\cdots\neq j_{k}}\Pi_{i=1}^{k}A_{j_{i}i}A_{j_{i}\sigma(i)}
=σSk(1)|σ|j1j2jk𝔼Πi=1kAjiiAjiσ(i).\displaystyle=\sum_{\sigma\in S_{k}}(-1)^{\left|\sigma\right|}\sum_{j_{1}\neq j_{2}\neq\cdots\neq j_{k}}{\mathbb{E}}\Pi_{i=1}^{k}A_{j_{i}i}A_{j_{i}\sigma(i)}. (56)

Here, we need to recall a main result about integration with respect to the Haar measure on orthogonal group,(see [8] Theorem3.13)

Proposition 3.2.

Suppose NnN\geq n. Let g=(gij)1i,jNg=\left(g_{ij}\right)_{1\leq i,j\leq N} be a Haar-distributed random matrix from O(N)O(N) and let dg the normalized Haar measure on O(N)O(N). Given two functions 𝒊,𝒋\bm{i},\bm{j} from {1,2,,2n}\{1,2,\ldots,2n\} to {1,2,,N}\{1,2,\ldots,N\}, we have

gO(N)gi(1)j(1)gi(2)j(2)gi(2n)j(2n)𝑑g\displaystyle\int_{g\in O(N)}g_{i(1)j(1)}g_{i(2)j(2)}\cdots g_{i(2n)j(2n)}dg
=\displaystyle= m,n(2n)WgnO(N)(m1n)k=1nδi(m(2k1)),i(m(2k))δj(m(2k1)),j(m(2k)).\displaystyle\sum_{\mathrm{m},\mathrm{n}\in\mathcal{M}(2n)}\mathrm{Wg}_{n}^{O(N)}\left(\mathrm{m}^{-1}\mathrm{n}\right)\prod_{k=1}^{n}\delta_{i(\mathrm{~{}m}(2k-1)),i(\mathrm{~{}m}(2k))}\delta_{j(\mathrm{~{}m}(2k-1)),j(\mathrm{~{}m}(2k))}.

Here we regard (2n)\mathcal{M}(2n) as a subset of S2nS_{2n}. As a special case , we obtain an integral expression for WgnO(N)(σ)\mathrm{Wg}_{n}^{O(N)}(\sigma) :

WgnO(N)(σ)=gO(N)g1j1g1j2g2j3g2j4gnj2n1gnj2n𝑑g,σS2n\mathrm{Wg}_{n}^{O(N)}(\sigma)=\int_{g\in O(N)}g_{1j_{1}}g_{1j_{2}}g_{2j_{3}}g_{2j_{4}}\cdots g_{nj_{2n-1}}g_{nj_{2n}}dg,\quad\sigma\in S_{2n}

with

(j1,j2,,j2n)=(σ(1)2,σ(2)2,,σ(2n)2).\left(j_{1},j_{2},\ldots,j_{2n}\right)=\left(\left\lceil\frac{\sigma(1)}{2}\right\rceil,\left\lceil\frac{\sigma(2)}{2}\right\rceil,\ldots,\left\lceil\frac{\sigma(2n)}{2}\right\rceil\right).

Collins and Śniady [8] obtained

WgnO(N)(σ)=(1)|μ|i1CatμiNn|μ|+O(Nn|μ|1),N,\mathrm{Wg}_{n}^{O(N)}(\sigma)=(-1)^{|\mu|}\prod_{i\geq 1}\mathrm{Cat}_{\mu_{i}}\cdot N^{-n-|\mu|}+\mathrm{O}\left(N^{-n-|\mu|-1}\right),\quad N\rightarrow\infty,

where σ\sigma is a permutation in S2nS_{2n} of reduced coset-type μ\mu, which implies the permutation σ\sigma for which Wg(σ)\mathrm{Wg}(\sigma) will have the largest order is the only one satisfying |σ|=0|\sigma|=0, i.e. σ=id\sigma=id.
Furthermore, S. Matsumoto obtain a more precise result for the expansion of Wg(σ)\mathrm{Wg}(\sigma)([17]). Given a partition μ\mu, we define WgO(N)(μ;n)=Wn(N)(σ)\mathrm{Wg}^{O(N)}(\mu;n)=\mathrm{W}_{n}^{(N)}(\sigma), where σ\sigma is a permutation in S2nS_{2n} of reduced coset-type μ\mu. For example,

WgO(N)((0);n)\displaystyle\mathrm{Wg}^{O(N)}((0);n) =WgO(N)(id2n)=Nn+n(n1)Nn2n(n1)Nn3+O(Nn4)\displaystyle=\mathrm{Wg}^{O(N)}\left(\mathrm{id}_{2n}\right)=N^{-n}+n(n-1)N^{-n-2}-n(n-1)N^{-n-3}+\mathrm{O}\left(N^{-n-4}\right) (57)
WgO(N)((1);n)\displaystyle\mathrm{Wg}^{O(N)}((1);n) =Nn1+Nn2(n2+3n7)Nn3+O(Nn4)\displaystyle=-N^{-n-1}+N^{-n-2}-\left(n^{2}+3n-7\right)N^{-n-3}+\mathrm{O}\left(N^{-n-4}\right) (58)

In our case, only when σ(i)=i\sigma(i)=i, Wg(σ)\mathrm{Wg}(\sigma) will have the largest order, then

1122j1j_{1}j1j_{1}j2j_{2}j2j_{2}
Figure 3.1:
𝔼Πi=1kAjiiAjii=WgO(n)((0);k)=nk+k(k1)nk2k(k1)nk3+O(nk4).\displaystyle{\mathbb{E}}\Pi_{i=1}^{k}A_{j_{i}i}A_{j_{i}i}=\mathrm{Wg}^{O(n)}((0);k)=n^{-k}+k(k-1)n^{-k-2}-k(k-1)n^{-k-3}+\mathrm{O}\left(n^{-k-4}\right). (59)

If |σ|=1|\sigma|=1, we have

WgO(n)((1);k)\displaystyle\mathrm{Wg}^{O(n)}((1);k) =nk1+nk2(k2+3k7)nk3+O(nk4).\displaystyle=-n^{-k-1}+n^{-k-2}-\left(k^{2}+3k-7\right)n^{-k-3}+\mathrm{O}\left(n^{-k-4}\right).

Directly, when k is finite we have

𝔼Z=n(n1)(nk+1)nk+O(n1)=1+O(n1),{\mathbb{E}}Z=\frac{n(n-1)\cdots(n-k+1)}{n^{k}}+{\mathrm{O}}(n^{-1})=1+{\mathrm{O}}(n^{-1}), (60)

which comes from the fact tha the size of set {j1j2jk}\{j_{1}\neq j_{2}\neq\cdots\neq j_{k}\} is n(n1)(nk+1)n(n-1)\cdots(n-k+1). Next, we need to estimate the variance and fourth moment of Z.

Proposition 3.3.

For finite k and large n, we have

𝐕𝐚𝐫Z=O(n1)\displaystyle{\mathbf{Var}}Z={\mathrm{O}}(n^{-1})

Proof.
𝔼Z2\displaystyle{\mathbb{E}}Z^{2} =𝔼(σSk(1)|σ|j1j2jkΠi=1kAjiiAjiσ(i))2\displaystyle={\mathbb{E}}\left(\sum_{\sigma\in S_{k}}(-1)^{\left|\sigma\right|}\sum_{j_{1}\neq j_{2}\neq\cdots\neq j_{k}}\Pi_{i=1}^{k}A_{j_{i}i}A_{j_{i}\sigma(i)}\right)^{2}
=𝔼(σSk(1)|σ|j1j2jkΠi=1kAjiiAjiσ(i))(σSk(1)|σ|l1l2lkΠi=1kAliiAliσ(i)),\displaystyle={\mathbb{E}}\left(\sum_{\sigma\in S_{k}}(-1)^{\left|\sigma\right|}\sum_{j_{1}\neq j_{2}\neq\cdots\neq j_{k}}\Pi_{i=1}^{k}A_{j_{i}i}A_{j_{i}\sigma(i)}\right)\left(\sum_{\sigma\in S_{k}}(-1)^{\left|\sigma\right|}\sum_{l_{1}\neq l_{2}\neq\cdots\neq l_{k}}\Pi_{i=1}^{k}A_{l_{i}i}A_{l_{i}\sigma(i)}\right),

we need to find the leading term in the above equation.
Case ii: If the k-tuple j:=(j1,,jk)\stackrel{{\scriptstyle\rightharpoonup}}{{j}}:=(j_{1},\cdots,j_{k}) and l:=(l1,,lk)\stackrel{{\scriptstyle\rightharpoonup}}{{l}}:=(l_{1},\cdots,l_{k}) are totally different,

𝔼Z2=𝔼σ(1)|σj|+|σl|j,lΠi=1kAjiiAjiσ(i)Πi=1kAliiAliσ(i),\displaystyle{\mathbb{E}}Z^{2}={\mathbb{E}}\sum_{\sigma}(-1)^{\left|\sigma_{j}\right|+\left|\sigma_{l}\right|}\sum_{\stackrel{{\scriptstyle\rightharpoonup}}{{j}},\stackrel{{\scriptstyle\rightharpoonup}}{{l}}}\Pi_{i=1}^{k}A_{j_{i}i}A_{j_{i}\sigma(i)}\Pi_{i=1}^{k}A_{l_{i}i}A_{l_{i}\sigma(i)}, (61)

only when σ=id\sigma=id, we have the leading term

1111j1j_{1}j1j_{1}j1j_{1}j1j_{1}
Figure 3.2:
𝔼Πi=1kAjiiAjiσ(i)Πi=1kAliiAliσ(i)=n2k2k(2k1)n2k2+O(n2k3).\displaystyle{\mathbb{E}}\Pi_{i=1}^{k}A_{j_{i}i}A_{j_{i}\sigma(i)}\Pi_{i=1}^{k}A_{l_{i}i}A_{l_{i}\sigma(i)}=n^{-2k}-2k(2k-1)n^{-2k-2}+\mathrm{O}(n^{-2k-3}). (62)

If σ\sigma has one cycle,

𝔼Πi=1kAjiiAjiσ(i)Πi=1kAliiAliσ(i)=n2k1+n2k2+O(n2k3).\displaystyle{\mathbb{E}}\Pi_{i=1}^{k}A_{j_{i}i}A_{j_{i}\sigma(i)}\Pi_{i=1}^{k}A_{l_{i}i}A_{l_{i}\sigma(i)}=-n^{-2k-1}+n^{-2k-2}+\mathrm{O}(n^{-2k-3}). (63)

If σ\sigma has two cycles,

𝔼Πi=1kAjiiAjiσ(i)Πi=1kAliiAliσ(i)=O(n2k2).\displaystyle{\mathbb{E}}\Pi_{i=1}^{k}A_{j_{i}i}A_{j_{i}\sigma(i)}\Pi_{i=1}^{k}A_{l_{i}i}A_{l_{i}\sigma(i)}={\mathrm{O}}(n^{-2k-2}). (64)

Case iiii: If sequences j\stackrel{{\scriptstyle\rightharpoonup}}{{j}} and l\stackrel{{\scriptstyle\rightharpoonup}}{{l}} has exactly one equal number where the index is same, w.l.o.g we assume j1=l1j_{1}=l_{1}.Since j2j3jkj_{2}\neq j_{3}\neq\cdots\neq j_{k}, we need σ(i)=i,i2\sigma(i)=i,i\geq 2 to get the largest order, which means σ=id\sigma=id. In conclusion the total number of pairing is 3k.

1111j1j_{1}j1j_{1}j1j_{1}j1j_{1}
Figure 3.3:

If the index is different, for example j1=l2j_{1}=l_{2}, σ\sigma need to be identity to get the largest order.

111122j1j_{1}j1j_{1}l1l_{1}l1l_{1}j1j_{1}j1j_{1}
Figure 3.4:

the total number of pairing is 2(k2)2\binom{k}{2}.

Case iiiiii: j\stackrel{{\scriptstyle\rightharpoonup}}{{j}} and l\stackrel{{\scriptstyle\rightharpoonup}}{{l}} has two equal numbers, the term is negligible O(n2){\mathrm{O}}(n^{-2}).

𝔼Z2\displaystyle{\mathbb{E}}Z^{2} =(n2k+O(1n2k+2))(n(n1)(n2k+1))2(k2)(n2k1+n2k2)(n(n1)(n2k+1))\displaystyle=\left(n^{-2k}+{\mathrm{O}}\left(\frac{1}{n^{2k+2}}\right)\right)*\left(n(n-1)\cdots(n-2k+1)\right)-2\binom{k}{2}(-n^{-2k-1}+n^{-2k-2})\left(n(n-1)\cdots(n-2k+1)\right)
+(3k+2(k2))1n2k(n(n1)(n2k+2))\displaystyle+(3k+2\binom{k}{2})*\frac{1}{n^{2k}}\left(n(n-1)\cdots(n-2k+2)\right)
=1+O(kn)+O(1n2).\displaystyle=1+{\mathrm{O}}(\frac{k}{n})+{\mathrm{O}}(\frac{1}{n^{2}}).

Moreover,

𝐕𝐚𝐫Z=𝔼Z2(𝔼Z)2=O(kn)+O(1n2).\displaystyle{\mathbf{Var}}Z={\mathbb{E}}Z^{2}-\left({\mathbb{E}}Z\right)^{2}={\mathrm{O}}\left(\frac{k}{n}\right)+{\mathrm{O}}\left(\frac{1}{n^{2}}\right). (65)

Next, we consider the fourth central moment of Z.

Proposition 3.4.
𝔼(Z𝔼Z)4=O(1n2)\displaystyle{\mathbb{E}}\left(Z-{\mathbb{E}}Z\right)^{4}={\mathrm{O}}(\frac{1}{n^{2}})

Proof.

Define Dj=σSk(1)|σ|j1j2jkΠi=1kAjiiAjiσ(i)D_{\stackrel{{\scriptstyle\rightharpoonup}}{{j}}}=\sum_{\sigma\in S_{k}}(-1)^{\left|\sigma\right|}\sum_{j_{1}\neq j_{2}\neq\cdots\neq j_{k}}\Pi_{i=1}^{k}A_{j_{i}i}A_{j_{i}\sigma(i)}.

𝔼(Z𝔼Z)4\displaystyle{\mathbb{E}}\left(Z-{\mathbb{E}}Z\right)^{4} =𝔼(Dj𝔼Dj)4\displaystyle={\mathbb{E}}\left(D_{\stackrel{{\scriptstyle\rightharpoonup}}{{j}}}-{\mathbb{E}}D_{\stackrel{{\scriptstyle\rightharpoonup}}{{j}}}\right)^{4}
=𝔼(Dj𝔼Dj)(Dk𝔼Dk)(Dl𝔼Dl)(Dm𝔼Dm)\displaystyle={\mathbb{E}}\left(D_{\stackrel{{\scriptstyle\rightharpoonup}}{{j}}}-{\mathbb{E}}D_{\stackrel{{\scriptstyle\rightharpoonup}}{{j}}}\right)\left(D_{\stackrel{{\scriptstyle\rightharpoonup}}{{k}}}-{\mathbb{E}}D_{\stackrel{{\scriptstyle\rightharpoonup}}{{k}}}\right)\left(D_{\stackrel{{\scriptstyle\rightharpoonup}}{{l}}}-{\mathbb{E}}D_{\stackrel{{\scriptstyle\rightharpoonup}}{{l}}}\right)\left(D_{\stackrel{{\scriptstyle\rightharpoonup}}{{m}}}-{\mathbb{E}}D_{\stackrel{{\scriptstyle\rightharpoonup}}{{m}}}\right)
=𝔼DjDkDlDm4𝔼DjDkDl𝔼Dm+6𝔼DjDk𝔼DlDm4𝔼(Dj)4+𝔼(Dj)4.\displaystyle={\mathbb{E}}D_{\stackrel{{\scriptstyle\rightharpoonup}}{{j}}}D_{\stackrel{{\scriptstyle\rightharpoonup}}{{k}}}D_{\stackrel{{\scriptstyle\rightharpoonup}}{{l}}}D_{\stackrel{{\scriptstyle\rightharpoonup}}{{m}}}-4{\mathbb{E}}D_{\stackrel{{\scriptstyle\rightharpoonup}}{{j}}}D_{\stackrel{{\scriptstyle\rightharpoonup}}{{k}}}D_{\stackrel{{\scriptstyle\rightharpoonup}}{{l}}}{\mathbb{E}}D_{\stackrel{{\scriptstyle\rightharpoonup}}{{m}}}+6{\mathbb{E}}D_{\stackrel{{\scriptstyle\rightharpoonup}}{{j}}}D_{\stackrel{{\scriptstyle\rightharpoonup}}{{k}}}{\mathbb{E}}D_{\stackrel{{\scriptstyle\rightharpoonup}}{{l}}}D_{\stackrel{{\scriptstyle\rightharpoonup}}{{m}}}-4{\mathbb{E}}(D_{\stackrel{{\scriptstyle\rightharpoonup}}{{j}}})^{4}+{\mathbb{E}}(D_{\stackrel{{\scriptstyle\rightharpoonup}}{{j}}})^{4}. (66)

Step1, if there are same index, the number of free index will decrease, the term above will be smaller. W.l.o.g we assume j1=l1j_{1}=l_{1} and show that O(n4k){\mathrm{O}}(n^{-4k}) term vanish. The leading term only happens when σ=id\sigma=id similarly.

1111j1j_{1}j1j_{1}j1j_{1}j1j_{1}
Figure 3.5:

The number of O(n4k){\mathrm{O}}(n^{-4k}) term is

3(3×2+1×2)+(3×1+5)4+1=0.\displaystyle 3-(3\times 2+1\times 2)+(3\times 1+5)-4+1=0.

Multiply the number of index n4k1n^{4k-1}, then we know O(1n){\mathrm{O}}(\frac{1}{n}) term vanishes in the fourth moment of Z. Step 2, we prove that for totally different 4 k-tuple , the term O(1n4k),O(1n4k+1){\mathrm{O}}(\frac{1}{n^{4k}}),{\mathrm{O}}(\frac{1}{n^{4k+1}}) both vanish. Firstly, for O(1n4k){\mathrm{O}}(\frac{1}{n^{4k}}) term, σ\sigma must be identity. The number of O(1n4k){\mathrm{O}}(\frac{1}{n^{4k}}) term is

14+64+1=0.1-4+6-4+1=0. (67)

In addition (30) gives the fact that O(1n4k+1){\mathrm{O}}(\frac{1}{n^{4k+1}}) vanish.

Next consider the O(1n4k+1){\mathrm{O}}(\frac{1}{n^{4k+1}}) term , Weingarten calculus says that we have the sub-leading term when σ\sigma has exactly one transposition. To simplify the proof , we omit the error term

𝔼Dj\displaystyle{\mathbb{E}}D_{\stackrel{{\scriptstyle\rightharpoonup}}{{j}}} =nk(k2)nk1,\displaystyle=n^{-k}-\binom{k}{2}n^{-k-1}, (68)
𝔼DjDk\displaystyle{\mathbb{E}}D_{\stackrel{{\scriptstyle\rightharpoonup}}{{j}}}D_{\stackrel{{\scriptstyle\rightharpoonup}}{{k}}} =n2k+2kn2k12(k2)n2k1=n2k+(2k2(k2))n2k1.\displaystyle=n^{-2k}+2kn^{-2k-1}-2\binom{k}{2}n^{-2k-1}=n^{-2k}+(2k-2\binom{k}{2})n^{-2k-1}. (69)

The first part of second equation comes from the case σj=σk=id\sigma_{j}=\sigma_{k}=id, the second part comes from the case σj\sigma_{j} and σk\sigma_{k} only has one transposition. For example the reason that 2k2k in first part occurs is as follows,

11221122j1j_{1}j1j_{1}j2j_{2}j2j_{2}l1l_{1}l1l_{1}l2l_{2}l2l_{2}
Figure 3.6:
11221122j1j_{1}j1j_{1}j2j_{2}j2j_{2}l1l_{1}l1l_{1}l2l_{2}l2l_{2}
Figure 3.7:

Similarly, we have

𝔼DjDkDl\displaystyle{\mathbb{E}}D_{\stackrel{{\scriptstyle\rightharpoonup}}{{j}}}D_{\stackrel{{\scriptstyle\rightharpoonup}}{{k}}}D_{\stackrel{{\scriptstyle\rightharpoonup}}{{l}}} =n3k+(6k3(k2))n3k1,\displaystyle=n^{-3k}+(6k-3\binom{k}{2})n^{-3k-1}, (70)
𝔼DjDkDlDm\displaystyle{\mathbb{E}}D_{\stackrel{{\scriptstyle\rightharpoonup}}{{j}}}D_{\stackrel{{\scriptstyle\rightharpoonup}}{{k}}}D_{\stackrel{{\scriptstyle\rightharpoonup}}{{l}}}D_{\stackrel{{\scriptstyle\rightharpoonup}}{{m}}} =n4k+(2(42)k4(k2))n4k1.\displaystyle=n^{-4k}+(2\binom{4}{2}k-4\binom{k}{2})n^{-4k-1}. (71)

Put the above four equations back into equation (3), we prove that O(n4k1){\mathrm{O}}(n^{-4k-1}) term vanishes. Since the number of free index is 4k, which means that O(nk1){\mathrm{O}}(n^{-k-1}) term vanish in the fourth moment of Z.

Let μk\mu_{k} denote the k-th absolute central moment of Z. Consider Taylor expansion of logZ\log Z at 𝔼Z\mathbb{E}Z, which is a special case of the delta method. See [12] p166 for more details.

𝔼[f(Z)]\displaystyle{\mathbb{E}}[f(Z)] =𝔼[f(𝔼Z+(Z𝔼Z))]\displaystyle={\mathbb{E}}\left[f\left({\mathbb{E}}Z+\left(Z-{\mathbb{E}}Z\right)\right)\right]
𝔼[f(𝔼Z)+f(𝔼Z)(Z𝔼Z)+12f′′(𝔼Z)(Z𝔼Z)2]\displaystyle\approx{\mathbb{E}}\left[f\left({\mathbb{E}}Z\right)+f^{\prime}\left({\mathbb{E}}Z\right)\left(Z-{\mathbb{E}}Z\right)+\frac{1}{2}f^{\prime\prime}\left({\mathbb{E}}Z\right)\left(Z-{\mathbb{E}}Z\right)^{2}\right]
=f(𝔼Z)+12f′′(𝔼Z)𝔼[(Z𝔼Z)2].\displaystyle=f\left({\mathbb{E}}Z\right)+\frac{1}{2}f^{\prime\prime}\left({\mathbb{E}}Z\right){\mathbb{E}}\left[\left(Z-{\mathbb{E}}Z\right)^{2}\right]. (72)

Similarly,

𝐕𝐚𝐫[f(Z)](f(𝔼[Z]))2𝐕𝐚𝐫[X]=(f(𝔼Z)))2μ2(Z)214(f′′(𝔼Z))2μ2(Z)4.{\mathbf{Var}}[f(Z)]\approx\left(f^{\prime}({\mathbb{E}}[Z])\right)^{2}{\mathbf{Var}}[X]=\left(f^{\prime}\left({\mathbb{E}}Z)\right)\right)^{2}\mu_{2}(Z)^{2}-\frac{1}{4}\left(f^{\prime\prime}\left({\mathbb{E}}Z\right)\right)^{2}\mu_{2}(Z)^{4}.

To do more

logZ\displaystyle\log Z log𝔼Z+Z𝔼Z𝔼Z,\displaystyle\approx\log\mathbb{E}Z+\frac{Z-\mathbb{E}Z}{\mathbb{E}Z}, (73)
𝐕𝐚𝐫logZ\displaystyle\mathbf{Var}\log Z 𝐕𝐚𝐫Z(𝔼Z)2=O(1n),\displaystyle\approx\frac{\mathbf{Var}Z}{(\mathbb{E}Z)^{2}}={\mathrm{O}}(\frac{1}{n}), (74)
μ4(logZ)\displaystyle\mu_{4}(\log Z) μ4(Z)(𝔼Z)4=O(1n2).\displaystyle\approx\frac{\mu_{4}(Z)}{(\mathbb{E}Z)^{4}}={\mathrm{O}}(\frac{1}{n^{2}}). (75)

We follow the notation in Bentkus [6] and define

S:=SN=X1++XN,S:=S_{N}=X_{1}+\cdots+X_{N},

where X1,,XNX_{1},\ldots,X_{N} are independent random vectors in k\mathbb{R}^{k} with common mean 𝔼Xj=\mathbb{E}X_{j}= 0 . We set

C:=cov(S)C:=\operatorname{cov}(S)

to be the covariance matrix of SS, which we assume is invertible. With the definition

βj:=𝔼C12Xj23,β:=j=1Nβj,\beta_{j}:=\mathbb{E}\left\|C^{-\frac{1}{2}}X_{j}\right\|_{2}^{3},\beta:=\sum_{j=1}^{N}\beta_{j},

we have the following [6]: Theorem 6.4 (Multivariate CLT with Rate).
There exists an absolute constant c>0c>0 such that

d(S,C12Z)ck14βd\left(S,C^{\frac{1}{2}}Z\right)\leq ck^{\frac{1}{4}}\beta

where Z𝒩(0,Idk)Z\sim\mathcal{N}\left(0,\operatorname{Id}_{k}\right) denotes a standard Gaussian on k\mathbb{R}^{k}.
We have

βj\displaystyle\beta_{j} =𝔼|[N𝐕𝐚𝐫(logZ)]12logZ|3\displaystyle=\mathbb{E}\left|[N\mathbf{Var}(\log Z)]^{-\frac{1}{2}}\log Z\right|^{3} (76)
CN32[𝐕𝐚𝐫(logZ)]32μ4(logZ)\displaystyle\leq CN^{-\frac{3}{2}}[\mathbf{Var}(\log Z)]^{-\frac{3}{2}}\mu_{4}(\log Z) (77)
=C1N32n12\displaystyle=C\frac{1}{N^{\frac{3}{2}}n^{\frac{1}{2}}} (78)

Therefore,

β:=j=1NβjC1nN,\displaystyle\beta:=\sum_{j=1}^{N}\beta_{j}\leq C\frac{1}{\sqrt{nN}}, (79)

then similar to proof of Theorem 1.1 , divide into 3 parts we can prove λ1++λk\lambda_{1}+\cdots+\lambda_{k} is convergent to Gaussian.

Acknowledgements

I am grateful to Professor Dang-Zheng Liu for his guidance. I would also like to thank Yandong Gu, Guangyi Zou, Ruohan Geng for their useful advice.

References

  • [1] G. Akemann G, Burda Z. Universal microscopic correlation functions for products of independent Ginibre matrices. J. Phys. A: Math. Theor. 45(46) (2012), 465201.
  • [2] G. Akemann, Z. Burda, M. Kieburg, and T. Nagao. Universal microscopic correlation functions for products of truncated unitary matrices. Journal of Physics A: Mathematical and Theoretical, (25)47 (2014), 255202.
  • [3] G. Akemann, Z. Burda, and M. Kieburg. Universal distribution of Lyapunov exponents for products of Ginibre matrices. Journal of Physics A: Mathematical and Theoretical, (39)47 (2014), 395202.
  • [4] G. Akemann, Z. Burda, and M. Kieburg. From integrable to chaotic systems: universal local statistics of Lyapunov exponents. EPL (Europhysics Letters), (4)126 (2019), 40001.
  • [5] Ahn A. Fluctuations of beta-Jacobi product processes. Probability Theory and Related Fields, 183(1)(2022), 257-123.
  • [6] V. Bentkus. A Lyapunov-type bound in RD. Theory of Probability Its Applications, (2)49(2005), 311–323.
  • [7] R. Carmona. Exponential localization in one dimensional disordered systems. Duke Mathematical Journal, (1)49 (1982), 191-213.
  • [8] Collins B, Śniady P. Integration with respect to the Haar measure on unitary, orthogonal and symplectic group. Communications in Mathematical Physics, 264(3)(2006), 773-795.
  • [9] P. Diaconis, Patterns in eigenvalues: the 70th Josian Willard Gibbs Lecture, Bull. of Amer. Math. Soc. 40 (2003), No. 2, 155-178.
  • [10] H.Furstenberg and H.Kesten. Products of random matrices. Ann. Math. Statist, 31(1960), 457–469.
  • [11] Hanin B, Paouris G. Non-asymptotic results for singular values of gaussian matrix products. Geometric and Functional Analysis, 31(2)(2021), 268-324.
  • [12] Haym Benaroya, Seon Mi Han, and Mark Nagurka. Probability Models in Engineering and Science. CRC Press, (2005).
  • [13] Kargin V. On the largest Lyapunov exponent for products of Gaussian matrices. Journal of Statistical Physics, 157(1)(2014), 70-83.
  • [14] Latała R. Estimation of moments of sums of independent real random variables. The Annals of Probability, 25(3)(1997), 1502-1513.
  • [15] É. Le Page. Théoremes limites pour les produits de matrices aléatoires. In: Probability Measures on Groups. Springer, Berlin (1982), pp. 258-303.
  • [16] Liu D Z, Wang D, Wang Y. Lyapunov exponent, universality and phase transition for products of random matrices. Communications in Mathematical Physics, 399(3)(2023), 1811-1855.
  • [17] Matsumoto, Sho. ”Jucys–Murphy elements, orthogonal matrix integrals, and Jack measures.” The Ramanujan Journal 26(2011), 69-107.
  • [18] Cohen J E, Newman C M. The stability of large random matrices and their products. The Annals of Probability, (1984), 283-310.
  • [19] Oseledec, V. I. A multiplicative ergodic theorem. Ljapunov characteristic numbers for dynamical systems. Trans. Moscow Math. Soc. 19(1968) 197–231.
  • [20] Skorski M. Bernstein-type bounds for beta distribution. Modern Stochastics: Theory and Applications, 10(2)(2023), 211-228.