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

Necessary and Sufficient Conditions for Convergence to the Semicircle Distribution

Calvin Wooyoung Chin
Abstract.

We consider random Hermitian matrices with independent upper triangular entries. Wigner’s semicircle law says that under certain additional assumptions, the empirical spectral distribution converges to the semicircle distribution. We characterize convergence to semicircle in terms of the variances of the entries, under natural assumptions such as the Lindeberg condition. The result extends to certain matrices with entries having infinite second moments. As a corollary, another characterization of semicircle convergence is given in terms of convergence in distribution of the row sums to the standard normal distribution.

1. Introduction

Let WnW_{n}, for each n𝐍n\in\mathbf{N}, be a random n×nn\times n Hermitian matrix whose upper triangular entries are independent. We call (Wn)n𝐍(W_{n})_{n\in\mathbf{N}} a Hermitian Wigner ensemble. In case WnW_{n} is real symmetric for all n𝐍n\in\mathbf{N}, we call (Wn)n𝐍(W_{n})_{n\in\mathbf{N}} a symmetric Wigner ensemble. We write Wn=(wij)i,j=1nW_{n}=(w_{ij})_{i,j=1}^{n} throughout this paper. If λ1(Wn)λn(Wn)\lambda_{1}(W_{n})\geq\cdots\geq\lambda_{n}(W_{n}) are the eigenvalues of WnW_{n} counted with multiplicity, then the empirical spectral distribution μWn\mu_{W_{n}} of WnW_{n} is defined by

μWn:=1ni=1nδλi(Wn).\mu_{W_{n}}:=\frac{1}{n}\sum_{i=1}^{n}\delta_{\lambda_{i}(W_{n})}.

Since μWn\mu_{W_{n}} is a random measure, we can think of the mean measure 𝐄μWn\operatorname{\mathbf{E}}\mu_{W_{n}}, which is defined and treated in Appendix A.

Let us use the term semicircle law to refer to a class of theorems that state, under certain conditions, that μWn\mu_{W_{n}} converges in some sense to the semicircle distribution μsc\mu_{\mathrm{sc}} on 𝐑\mathbf{R} given by

μsc(dx):=(4x2)+dx.\mu_{\mathrm{sc}}(dx):=\sqrt{(4-x^{2})_{+}}\,dx.

(We let x+:=max{x,0}x_{+}:=\max\{x,0\}.) Wigner initiated the spectral study of random matrices by proving the following very first version of the semicircle law in [Wig55, Wig58].

Theorem 1.1 (semicircle law, Wigner).

Let (Wn)n𝐍(W_{n})_{n\in\mathbf{N}} be a symmetric Wigner ensemble such that the upper triangular entries of WnW_{n} have identical symmetric distribution with mean zero and variance 1/n1/n. If for each k𝐍k\in\mathbf{N} we have

(1.1) 𝐄|wij|kBknk/2for some Bk< independent of n,i,j,\operatorname{\mathbf{E}}|w_{ij}|^{k}\leq B_{k}n^{-k/2}\qquad\text{for some $B_{k}<\infty$ independent of $n,i,j$,}

then 𝐄μWnμsc\operatorname{\mathbf{E}}\mu_{W_{n}}\Rightarrow\mu_{\mathrm{sc}}. Here \Rightarrow denotes convergence in distribution.

Subsequent works by [Arn71], [Pas73], and others led to the following much more general semicircle law.

Theorem 1.2 (semicircle law, [BS10, Theorem 2.9]).

Let (Wn)n𝐍(W_{n})_{n\in\mathbf{N}} be a Hermitian Wigner ensemble such that the upper triangular entries of WnW_{n} are of mean zero and variance 1/n1/n. If

(1.2) 1ni,j=1n𝐄[|wij|2;|wij|>ϵ]0for all ϵ>0,\frac{1}{n}\sum_{i,j=1}^{n}\operatorname{\mathbf{E}}[|w_{ij}|^{2};|w_{ij}|>\epsilon]\to 0\qquad\text{for all $\epsilon>0$,}

then μWnμsc\mu_{W_{n}}\Rightarrow\mu_{\mathrm{sc}} a.s.

Note that (1.1) for k=3k=3 implies (1.2) since

1ni,j=1n𝐄[|wij|2;|wij|>ϵ]1ϵni,j=1n𝐄|wij|3B3ϵn0.\frac{1}{n}\sum_{i,j=1}^{n}\operatorname{\mathbf{E}}[|w_{ij}|^{2};|w_{ij}|>\epsilon]\leq\frac{1}{\epsilon n}\sum_{i,j=1}^{n}\operatorname{\mathbf{E}}|w_{ij}|^{3}\leq\frac{B_{3}}{\epsilon\sqrt{n}}\to 0.

Let us call (1.2) the Lindeberg condition, following the Lindeberg–Feller central limit theorem. Girko [Gir90, Theorem 9.4.1] states that the converse of Theorem 1.2 holds.

Rather surprisingly, we have the following:

Lemma 1.3 (a.s. convergence).

Let (Wn)n𝐍(W_{n})_{n\in\mathbf{N}} be a Hermitian Wigner ensemble. Then

𝐄μWnμscif and only ifμWnμsc a.s.\operatorname{\mathbf{E}}\mu_{W_{n}}\Rightarrow\mu_{\mathrm{sc}}\quad\text{if and only if}\quad\mu_{W_{n}}\Rightarrow\mu_{\mathrm{sc}}\text{ a.s.}

A proof of this fact using a concentration-of-measure inequality is given in Appendix A. Thanks to this equivalence, we will be able to go back and forth freely between the two types of convergences throughout the paper.

Theorem 1.2 suggests an extension of the semicircle law to the case where the entries of WnW_{n} have variances other than 1/n1/n. Here is one possible approach to such an extension. Assume that the underlying probability space is the product

(Ω1×Ω2,1×2,𝐏1×𝐏2)(\Omega_{1}\times\Omega_{2},\mathcal{F}_{1}\times\mathcal{F}_{2},\operatorname{\mathbf{P}}_{1}\times\operatorname{\mathbf{P}}_{2})

of two probability spaces (Ω1,1,𝐏1)(\Omega_{1},\mathcal{F}_{1},\operatorname{\mathbf{P}}_{1}) and (Ω2,2,𝐏2)(\Omega_{2},\mathcal{F}_{2},\operatorname{\mathbf{P}}_{2}). Then let Xn=(xij)i,j=1nX_{n}=(x_{ij})_{i,j=1}^{n} and Yn=(yij)i,j=1nY_{n}=(y_{ij})_{i,j=1}^{n} be random real symmetric matrices defined on Ω1\Omega_{1} and Ω2\Omega_{2} having i.i.d. upper triangular entries. If x11x_{11} is standard normal and

𝐏2(y11=1)=𝐏2(y11=1)=1/2,\operatorname{\mathbf{P}}_{2}(y_{11}=1)=\operatorname{\mathbf{P}}_{2}(y_{11}=-1)=1/2,

then it is not difficult to show that (Wn)n𝐍(W_{n})_{n\in\mathbf{N}} given by

wij(ω1,ω2):=xij(ω1)yij(ω2)/nw_{ij}(\omega_{1},\omega_{2}):=x_{ij}(\omega_{1})y_{ij}(\omega_{2})/\sqrt{n}

satisfies the conditions of Theorem 1.2.

Since μWn(ω1,ω2)μsc\mu_{W_{n}(\omega_{1},\omega_{2})}\Rightarrow\mu_{\mathrm{sc}} for 𝐏\operatorname{\mathbf{P}}-a.e. (ω1,ω2)(\omega_{1},\omega_{2}), Tonelli’s theorem implies that for 𝐏1\operatorname{\mathbf{P}}_{1}-a.e. ω1Ω1\omega_{1}\in\Omega_{1}, we have μWn(ω1,)μsc\mu_{W_{n}(\omega_{1},\cdot)}\Rightarrow\mu_{\mathrm{sc}} 𝐏2\operatorname{\mathbf{P}}_{2}-a.s. Note that the (i,j)(i,j)-entry of the random matrix Wn(ω1,)W_{n}(\omega_{1},\cdot) defined on (Ω2,2,𝒫2)(\Omega_{2},\mathcal{F}_{2},\mathcal{P}_{2}) has variance xij(ω1)2/n,x_{ij}(\omega_{1})^{2}/n, which can deviate by any amount from 1/n1/n.

A problem with this approach is that we do not know for which ω1\omega_{1} we have the a.s. convergence μWn(ω1,)μsc\mu_{W_{n}(\omega_{1},\cdot)}\Rightarrow\mu_{\mathrm{sc}}, even though we know this happens for almost all ω1\omega_{1}. For instance, the above discussion does not tell us whether μWnμsc\mu_{W_{n}}\Rightarrow\mu_{\mathrm{sc}} a.s. is true when (Wn)n𝐍(W_{n})_{n\in\mathbf{N}} is a symmetric Wigner ensemble such that

𝐏(wij=2/n)=𝐏(wij=2/n)=1/2if i+j is even\operatorname{\mathbf{P}}(w_{ij}=\sqrt{2/n})=\operatorname{\mathbf{P}}(w_{ij}=-\sqrt{2/n})=1/2\qquad\text{if $i+j$ is even}

and wij=0w_{ij}=0 if i+ji+j is odd.

Götze, Naumov, and Tikhomirov [GNT15] covered this case by proving the following:

Theorem 1.4 (semicircle law, [GNT15, Corollary 1]).

Let (Wn)n𝐍(W_{n})_{n\in\mathbf{N}} be a symmetric Wigner ensemble such that 𝐄wij=0\operatorname{\mathbf{E}}w_{ij}=0 and 𝐄|wij|2<\operatorname{\mathbf{E}}|w_{ij}|^{2}<\infty for i,j=1,,ni,j=1,\ldots,n. If the Lindeberg condition (1.2) holds, and

(1.3) 1ni=1n|j=1n𝐄|wij|21|0,\frac{1}{n}\sum_{i=1}^{n}\Bigl{|}\sum_{j=1}^{n}\operatorname{\mathbf{E}}|w_{ij}|^{2}-1\Bigr{|}\to 0,

and

(1.4) j=1n𝐄|wij|2Cfor some C< independent of n and i,\sum_{j=1}^{n}\operatorname{\mathbf{E}}|w_{ij}|^{2}\leq C\qquad\text{for some $C<\infty$ independent of $n$ and $i$,}

then μWnμsc\mu_{W_{n}}\Rightarrow\mu_{\mathrm{sc}} a.s.

From our main result (Theorem 1.6) it will follow that (1.4) is not needed in Theorem 1.4, and that (Wn)n𝐍(W_{n})_{n\in\mathbf{N}} can be assumed to be Hermitian, not necessarily real symmetric.

To illustrate that (1.3) is needed in Theorem 1.4, the authors of [GNT15] considered the random symmetric block matrix

Wn=(ABBTD)W_{n}=\begin{pmatrix}A&B\\ B^{T}&D\end{pmatrix}

where AA and DD are of size n/2×n/2\lfloor{n/2}\rfloor\times\lfloor{n/2}\rfloor and n/2×n/2\lceil{n/2}\rceil\times\lceil{n/2}\rceil, and the upper triangular entries of WnW_{n} are independent. They let all entries of WnW_{n} except the non-diagonal entries of DD be normal with mean 0 and variance 1/n1/n, and simulated the spectrum of WnW_{n} for n=2000n=2000 to see that μWn\mu_{W_{n}} does not look like a semicircle. Note that (1.3) does not hold.

Our main theorem will let us prove what was suggested by the simulation in [GNT15], namely that 𝐄μWn⇏μsc\operatorname{\mathbf{E}}\mu_{W_{n}}\not\Rightarrow\mu_{\mathrm{sc}}. More generally, we will prove for a large class of Hermitian Wigner ensembles (Wn)n𝐍(W_{n})_{n\in\mathbf{N}} that 𝐄μWnμsc\operatorname{\mathbf{E}}\mu_{W_{n}}\Rightarrow\mu_{\mathrm{sc}} (or μWnμsc\mu_{W_{n}}\Rightarrow\mu_{\mathrm{sc}} a.s., equivalently) holds if and only if (1.3) is true.

One thing we should notice is that changing o(n)o(n) rows of WnW_{n} has no effect on the limit of 𝐄μWn\operatorname{\mathbf{E}}\mu_{W_{n}} due to the following:

Lemma 1.5 (rank inequality).

Let AA and BB be n×nn\times n Hermitian matrices. If FAF_{A} and FBF_{B} are the distribution functions of μA\mu_{A} and μB\mu_{B} (defined in the same way as μWn\mu_{W_{n}}), then

supx𝐑FA(x)FB(x)rank(AB)n.\sup_{x\in\mathbf{R}}\|F_{A}(x)-F_{B}(x)\|\leq\frac{\operatorname{rank}(A-B)}{n}.
Proof.

See [BS10, Theorem A.43]. ∎

We want to say that for certain Hermitian Wigner ensembles (Wn)n𝐍(W_{n})_{n\in\mathbf{N}} with 𝐄μWnμsc\operatorname{\mathbf{E}}\mu_{W_{n}}\Rightarrow\mu_{\mathrm{sc}}, we have (1.3). However, without further restriction on (Wn)n𝐍(W_{n})_{n\in\mathbf{N}}, we can always change o(n)o(n) rows and columns of WnW_{n} so that (1.3) becomes false, while leaving the limiting distribution of 𝐄μWn\operatorname{\mathbf{E}}\mu_{W_{n}} unchanged. To avoid this problem, we assume that

(1.5) 1niJnj=1n𝐄|wij|20for any Jn{1,,n} with |Jn|/n0.\frac{1}{n}\sum_{i\in J_{n}}\sum_{j=1}^{n}\operatorname{\mathbf{E}}|w_{ij}|^{2}\to 0\qquad\text{for any $J_{n}\subset\{1,\ldots,n\}$ with $|J_{n}|/n\to 0$.}

Notice that this condition is weaker than (1.4). If WnW_{n} satisfies (1.3) and (1.5), and we change o(n)o(n) rows and columns of it to obtain Hermitian WnW_{n}^{\prime} which also satisfies (1.5), then WnW_{n}^{\prime} also satisfies (1.3). The following is our first main theorem:

Theorem 1.6 (characterization of semicircle convergence).

Let (Wn)n𝐍(W_{n})_{n\in\mathbf{N}} be a Hermitian Wigner ensemble with 𝐄wij=0\operatorname{\mathbf{E}}w_{ij}=0 and 𝐄|wij|2<\operatorname{\mathbf{E}}|w_{ij}|^{2}<\infty for i,j=1,,ni,j=1,\ldots,n satisfying (1.5) and the Lindeberg condition (1.2). Then 𝐄μWnμsc\operatorname{\mathbf{E}}\mu_{W_{n}}\Rightarrow\mu_{\mathrm{sc}} if and only if (1.3) holds.

Remark 1.7.

Since (1.3) implies

1niJnj=1n𝐄|wij|21niJn(|j=1n𝐄|wij|21|+1)0\frac{1}{n}\sum_{i\in J_{n}}\sum_{j=1}^{n}\operatorname{\mathbf{E}}|w_{ij}|^{2}\leq\frac{1}{n}\sum_{i\in J_{n}}\biggl{(}\Bigl{|}\sum_{j=1}^{n}\operatorname{\mathbf{E}}|w_{ij}|^{2}-1\Bigr{|}+1\biggr{)}\to 0

for any Jn{1,,n}J_{n}\subset\{1,\ldots,n\} with |Jn|/n0|J_{n}|/n\to 0, the sufficiency direction of Theorem 1.6 does not require (1.5). This proves the claim right after Theorem 1.4.

By Lindeberg’s universality scheme [GNT15, Theorem 2] for random matrices, it follows that we can remove the condition (1.4) (which is (5) in [GNT15]) from the semicircle law [GNT15, Theorem 1] for certain random symmetric matrices with dependent upper triangular entries.

We can actually go beyond Theorem 1.6 and allow the entries of WnW_{n} to have infinite variances, for example when wij=cijxij/nlognw_{ij}=c_{ij}x_{ij}/\sqrt{n\log n} where xijx_{ij} has a density

f(x)={1/|x|3if |x|>10if |x|1f(x)=\begin{cases}1/|x|^{3}&\text{if $|x|>1$}\\ 0&\text{if $|x|\leq 1$}\end{cases}

and cijc_{ij} is a real number close to 11. To achieve this, instead of 𝐄wij=0\operatorname{\mathbf{E}}w_{ij}=0 and the Lindeberg condition (1.2), we assume

(1.6) 1ni,j=1n(𝐄[wij;|wij|1])20\frac{1}{n}\sum_{i,j=1}^{n}\bigl{(}\operatorname{\mathbf{E}}[w_{ij};|w_{ij}|\leq 1]\bigr{)}^{2}\to 0

and

(1.7) 1ni,j=1n𝐏(|wij|>ϵ)0for all ϵ>0.\frac{1}{n}\sum_{i,j=1}^{n}\operatorname{\mathbf{P}}(|w_{ij}|>\epsilon)\to 0\qquad\text{for all $\epsilon>0$.}

If 𝐄wij=0\operatorname{\mathbf{E}}w_{ij}=0 and (1.2) hold, then (1.6) follows due to

(𝐄[wij;|wij|1])2=(𝐄[wij;|wij|>1])2𝐄[|wij|2;|wij|>1],\bigl{(}\operatorname{\mathbf{E}}[w_{ij};|w_{ij}|\leq 1]\bigr{)}^{2}=\bigl{(}\operatorname{\mathbf{E}}[w_{ij};|w_{ij}|>1]\bigr{)}^{2}\leq\operatorname{\mathbf{E}}[|w_{ij}|^{2};|w_{ij}|>1],

and (1.7) follows by

𝐏(|wij|>ϵ)ϵ2𝐄[|wij|2;|wij|>ϵ].\operatorname{\mathbf{P}}(|w_{ij}|>\epsilon)\leq\epsilon^{-2}\operatorname{\mathbf{E}}[|w_{ij}|^{2};|w_{ij}|>\epsilon].

Finally, (1.5) is replaced by

(1.8) 1niJnj=1n𝐄[|wij|2;|wij|1]0for any Jn{1,,n} with |Jn|/n0.\frac{1}{n}\sum_{i\in J_{n}}\sum_{j=1}^{n}\operatorname{\mathbf{E}}[|w_{ij}|^{2};|w_{ij}|\leq 1]\to 0\\ \text{for any $J_{n}\subset\{1,\ldots,n\}$ with $|J_{n}|/n\to 0$.}

The following is our second main theorem:

Theorem 1.8 (characterization, general version).

Let (Wn)n𝐍(W_{n})_{n\in\mathbf{N}} be a Hermitian Wigner ensemble satisfying (1.6), (1.7), and (1.8). Then 𝐄μWnμsc\operatorname{\mathbf{E}}\mu_{W_{n}}\Rightarrow\mu_{\mathrm{sc}} if and only if

(1.9) 1ni=1n|j=1n𝐄[|wij|2;|wij|1]1|0.\frac{1}{n}\sum_{i=1}^{n}\Bigl{|}\sum_{j=1}^{n}\operatorname{\mathbf{E}}[|w_{ij}|^{2};|w_{ij}|\leq 1]-1\Bigr{|}\to 0.
Remark 1.9.
  1. (1)

    As in Theorem 1.6, the sufficiency direction of Theorem 1.8 does not require 1.8. (See Remark 1.7.)

  2. (2)

    Theorem 1.6 follows easily from Theorem 1.8. See Lemma B.1 in the appendix.

  3. (3)

    In case the entries of WnW_{n} are real, the sufficiency part of Theorem 1.8 can be easily proved using Theorem 1.4 and Lemma 1.5. This is covered in Section 2.

  4. (4)

    Our full proof of the sufficiency is a careful consideration of Wigner’s moment method proof of the original semicircle law. This is arguably more elementary than the proof of Theorem 1.4 in [GNT15], which first deals with matrices with Gaussian entries using combinatorial arguments, and then generalizes the result to symmetric Wigner ensembles using Lindeberg’s universality scheme for random matrices.

The following corollary relates 𝐄μWnμsc\operatorname{\mathbf{E}}\mu_{W_{n}}\Rightarrow\mu_{\mathrm{sc}} to the convergence in distribution of the sum of a row of WnW_{n} to the standard normal random variable. The sufficiency direction when the entries of WnW_{n} are identically distributed was covered by [Jun18]. We denote the Lévy metric by LL.

Corollary 1.10 (characterization, Gaussian convergence).

Under the hypotheses of Theorem 1.8, we have 𝐄μWnμsc\operatorname{\mathbf{E}}\mu_{W_{n}}\Rightarrow\mu_{\mathrm{sc}} if and only if

(1.10) 1ni=1nL(Fni,G)0,\frac{1}{n}\sum_{i=1}^{n}L(F_{ni},G)\to 0,

where FniF_{ni} and GG are the distribution functions of j=1n±|wij|\sum_{j=1}^{n}\pm|w_{ij}| and the standard normal random variable. The signs ±\pm are independent Rademacher random variables independent from WnW_{n}.

The rest of the paper is organized as follows. Section 2 is a short section that introduces Theorem 2.1. This theorem is a reduction of Theorems 1.6 and 1.8, and will ultimately imply them as shown in Appendix B. This section also proves the sufficiency part of Theorem 2.1 in the case when the entries of WnW_{n} are real.

Section 3 is the essence of the proof of the necessity part of Theorem 2.1. We add an assumption that the sixth moment of 𝐄μWn\operatorname{\mathbf{E}}\mu_{W_{n}} is bounded, but as a return we obtain a clean proof that is right to the point. The idea is to express the second and the fourth moments of 𝐄μWn\operatorname{\mathbf{E}}\mu_{W_{n}} in terms of the variances of the entries of WnW_{n}.

Section 4 shows that we can remove the additional assumption on the sixth moments, assuming that a certain lemma (Lemma 4.1) holds. The condition (1.5) is used in this section.

Section 5 proves Lemma 4.1 by a systematic computation of the moments of 𝐄μWn\operatorname{\mathbf{E}}\mu_{W_{n}}. The computation is a variant of Wigner’s original moment method, but it can handle the case when the entries have non-identical variances.

In Section 6, we prove the sufficiency part of Theorem 2.1 using the results of Section 5. The classical argument involving Dyck paths is discussed for completeness.

Finally in Section 7, we derive Corollary 1.10 from Theorem 1.8 by an elementary argument involving the Lindeberg–Feller central limit theorem.

Appendix A defines the mean of a random probability measure, and proves some of their properties that we need. Appendix B contains the fairly standard proof that Theorem 1.8 implies Theorem 1.6, and that Theorem 2.1 implies Theorem 1.8.

2. Proof of sufficiency for symmetric Wigner ensembles

It is enough to prove the following in order to prove Theorems 1.6 and 1.8. The justification for the reduction is fairly standard, and is covered by Lemmas B.1 and B.4 in the appendix.

Theorem 2.1 (characterization, reduced form).

Let (Wn)n𝐍(W_{n})_{n\in\mathbf{N}} be a Hermitian Wigner ensemble such that

(2.1) 𝐄wij=0and|wij|ϵnfor all n𝐍 and i,j=1,,nwhere1ϵ1ϵ2andϵn0,\operatorname{\mathbf{E}}w_{ij}=0\quad\text{and}\quad|w_{ij}|\leq\epsilon_{n}\qquad\text{for all $n\in\mathbf{N}$ and $i,j=1,\ldots,n$}\\ \text{where}\quad 1\geq\epsilon_{1}\geq\epsilon_{2}\geq\cdots\quad\text{and}\quad\epsilon_{n}\to 0,

and (1.5) is true. Then 𝐄μWnμsc\operatorname{\mathbf{E}}\mu_{W_{n}}\Rightarrow\mu_{\mathrm{sc}} if and only if (1.3). Note that (1.2) is automatically satisfied due to |wij|ϵn|w_{ij}|\leq\epsilon_{n}.

The following shows that we can assume (1.4) in the proof of the sufficiency part of Theorem 2.1.

Lemma 2.2.

Assume that (Wn)n𝐍(W_{n})_{n\in\mathbf{N}} satisfies (1.3) and the conditions of Theorem 2.1. Then there exists a Hermitian Wigner ensemble (Wn)n𝐍(W^{\prime}_{n})_{n\in\mathbf{N}} satisfying (1.3), (1.4), and the conditions of Theorem 2.1 such that μWnμsc\mu_{W_{n}}\Rightarrow\mu_{\mathrm{sc}} a.s. if and only if μWnμsc\mu_{W^{\prime}_{n}}\Rightarrow\mu_{\mathrm{sc}} a.s.

Proof.

If we let JnJ_{n} be the set of all i{1,,n}i\in\{1,\ldots,n\} such that j=1n𝐄|wij|2>2\sum_{j=1}^{n}\operatorname{\mathbf{E}}|w_{ij}|^{2}>2, then

|Jn|n1ni=1n|j=1n𝐄|wij|21|\frac{|J_{n}|}{n}\leq\frac{1}{n}\sum_{i=1}^{n}\Bigl{|}\sum_{j=1}^{n}\operatorname{\mathbf{E}}|w_{ij}|^{2}-1\Bigr{|}

where the right side goes to 0 by (1.3). Let Wn=(wij)i,j=1nW^{\prime}_{n}=(w_{ij}^{\prime})_{i,j=1}^{n} be given by wij=wijw_{ij}^{\prime}=w_{ij} if i,jJni,j\not\in J_{n} and wij=0w_{ij}^{\prime}=0 otherwise.

Then,

1ni=1n|j=1n𝐄|wij|21|1ni=1n|j=1n𝐄|wij|21|+2niJnj=1n𝐄|wij|20\begin{split}\frac{1}{n}\sum_{i=1}^{n}\Bigl{|}\sum_{j=1}^{n}\operatorname{\mathbf{E}}|w_{ij}^{\prime}|^{2}-1\Bigr{|}&\leq\frac{1}{n}\sum_{i=1}^{n}\Bigl{|}\sum_{j=1}^{n}\operatorname{\mathbf{E}}|w_{ij}|^{2}-1\Bigr{|}+\frac{2}{n}\sum_{i\in J_{n}}\sum_{j=1}^{n}\operatorname{\mathbf{E}}|w_{ij}|^{2}\\ &\to 0\end{split}

where the right side goes to 0 by (1.3) and (1.5). Notice that (Wn)n𝐍(W_{n}^{\prime})_{n\in\mathbf{N}} satisfies (1.4) for C=2C=2. The conditions of Theorem 2.1 can be easily shown for (Wn)n𝐍(W_{n}^{\prime})_{n\in\mathbf{N}}.

Since

rank(WnWn)n2|Jn|n0,\frac{\operatorname{rank}(W_{n}-W^{\prime}_{n})}{n}\leq\frac{2|J_{n}|}{n}\to 0,

we have μWnμsc\mu_{W_{n}}\Rightarrow\mu_{\mathrm{sc}} a.s. if and only if μWnμsc\mu_{W^{\prime}_{n}}\Rightarrow\mu_{\mathrm{sc}} a.s. by Lemma 1.5. ∎

If WnW_{n} is real symmetric, then the sufficiency part of Theorem 2.1 follows from Lemma 2.2 and Theorem 1.4. In Section 6, we will present a direct proof of the sufficiency that applies to Hermitian Wigner ensembles and does not depend on the result of [GNT15].

3. Proof of necessity under bounded sixth moments

Assume (2.1) throughout this section. In this section, we present a relatively simple proof of necessity in Theorem 2.1 under the following additional assumption:

(3.1) supn𝐍𝐑x6𝐄μWn<.\sup_{n\in\mathbf{N}}\int_{\mathbf{R}}x^{6}\,\operatorname{\mathbf{E}}\mu_{W_{n}}<\infty.

The number 66 comes out just because it is an even number greater than 44. Our proof is based on an examination of the second and the fourth moments of 𝐄μWn\operatorname{\mathbf{E}}\mu_{W_{n}}. If λ1,,λn\lambda_{1},\ldots,\lambda_{n} are the eigenvalues of WnW_{n}, then for each k𝐍k\in\mathbf{N} we have

𝐑xkμWn(dx)=1ni=1nλik=1ntrWnk,\int_{\mathbf{R}}x^{k}\,\mu_{W_{n}}(dx)=\frac{1}{n}\sum_{i=1}^{n}\lambda_{i}^{k}=\frac{1}{n}\operatorname{tr}W_{n}^{k},

and thus

(3.2) 𝐑xk𝐄μWn(dx)=1n𝐄trWnk.\int_{\mathbf{R}}x^{k}\,\operatorname{\mathbf{E}}\mu_{W_{n}}(dx)=\frac{1}{n}\operatorname{\mathbf{E}}\operatorname{tr}W_{n}^{k}.

(See Lemma A.1.)

The second moment of 𝐄μWn\operatorname{\mathbf{E}}\mu_{W_{n}} can be easily expressed in terms of the variances of wijw_{ij}.

Lemma 3.1 (computation of the second moment).

We have

𝐑x2𝐄μWn(dx)=1ni,j=1n𝐄|wij|2.\int_{\mathbf{R}}x^{2}\,\operatorname{\mathbf{E}}\mu_{W_{n}}(dx)=\frac{1}{n}\sum_{i,j=1}^{n}\operatorname{\mathbf{E}}|w_{ij}|^{2}.
Proof.

It follows from (3.2) and

trWn2=i,j=1n|wij|2.\operatorname{tr}W_{n}^{2}=\sum_{i,j=1}^{n}|w_{ij}|^{2}.\qed

Computing the fourth moment requires more effort, but is still tractable.

Lemma 3.2 (computation of the fourth moment).

If

(3.3) supn𝐍𝐑x2𝐄μWn<,\sup_{n\in\mathbf{N}}\int_{\mathbf{R}}x^{2}\,\operatorname{\mathbf{E}}\mu_{W_{n}}<\infty,

then

𝐑x4𝐄μWn(dx)2ni=1n(j=1n𝐄|wij|2)20.\int_{\mathbf{R}}x^{4}\,\operatorname{\mathbf{E}}\mu_{W_{n}}(dx)-\frac{2}{n}\sum_{i=1}^{n}\biggl{(}\sum_{j=1}^{n}\operatorname{\mathbf{E}}|w_{ij}|^{2}\biggr{)}^{2}\to 0.

We remark that (1.5) implies (3.3). The proof for that is similar to that of Lemma 4.2 below.

Proof.

Note that

𝐑x4𝐄μWn(dx)=1n𝐄trWn4=1ni1,i2,i3,i4=1n𝐄[wi1i2wi2i3wi3i4wi4i1].\int_{\mathbf{R}}x^{4}\,\operatorname{\mathbf{E}}\mu_{W_{n}}(dx)=\frac{1}{n}\operatorname{\mathbf{E}}\operatorname{tr}W_{n}^{4}=\frac{1}{n}\sum_{i_{1},i_{2},i_{3},i_{4}=1}^{n}\operatorname{\mathbf{E}}[w_{i_{1}i_{2}}w_{i_{2}i_{3}}w_{i_{3}i_{4}}w_{i_{4}i_{1}}].

Since the upper triangular entries of WnW_{n} are independent and have mean zero, in order for 𝐄[wi1i2wi2i3wi3i4wi4i1]\operatorname{\mathbf{E}}[w_{i_{1}i_{2}}w_{i_{2}i_{3}}w_{i_{3}i_{4}}w_{i_{4}i_{1}}] not to vanish,

{i1,i2},{i2,i3},{i3,i4},{i4,i1}\{i_{1},i_{2}\},\{i_{2},i_{3}\},\{i_{3},i_{4}\},\{i_{4},i_{1}\}

should either be all the same, or be partitioned into two groups, where each group consists of two identical sets. This implies either i1=i3i_{1}=i_{3} or i2=i4i_{2}=i_{4} or both. Notice that, for instance, i1=i2i3=i4i_{1}=i_{2}\neq i_{3}=i_{4} cannot happen because {i1,i2}\{i_{1},i_{2}\} would then appear only once among {i1,i2}\{i_{1},i_{2}\}, {i2,i3}\{i_{2},i_{3}\}, {i3,i4}\{i_{3},i_{4}\}, and {i4,i1}\{i_{4},i_{1}\}.

Thus, the sum on the right side equals

i,j,k=1n𝐄[|wij|2|wik|2]+i,j,k=1n𝐄[|wij|2|wjk|2]i,j=1n𝐄|wij|4,\sum_{i,j,k=1}^{n}\operatorname{\mathbf{E}}[|w_{ij}|^{2}|w_{ik}|^{2}]+\sum_{i,j,k=1}^{n}\operatorname{\mathbf{E}}[|w_{ij}|^{2}|w_{jk}|^{2}]-\sum_{i,j=1}^{n}\operatorname{\mathbf{E}}|w_{ij}|^{4},

where the last term corresponds to the case where both i1=i3i_{1}=i_{3} and i2=i4i_{2}=i_{4} are true. Since the first and the second sum both equal

i=1n(j=1n𝐄|wij|2)2+i,j=1n(𝐄|wij|4(𝐄|wij|2)2),\sum_{i=1}^{n}\biggl{(}\sum_{j=1}^{n}\operatorname{\mathbf{E}}|w_{ij}|^{2}\biggr{)}^{2}+\sum_{i,j=1}^{n}\bigl{(}\operatorname{\mathbf{E}}|w_{ij}|^{4}-(\operatorname{\mathbf{E}}|w_{ij}|^{2})^{2}\bigr{)},

we have

𝐑x4𝐄μWn(dx)=2ni=1n(j=1n𝐄|wij|2)2+1ni,j=1n(𝐄|wij|42(𝐄|wij|2)2).\int_{\mathbf{R}}x^{4}\,\operatorname{\mathbf{E}}\mu_{W_{n}}(dx)=\frac{2}{n}\sum_{i=1}^{n}\biggl{(}\sum_{j=1}^{n}\operatorname{\mathbf{E}}|w_{ij}|^{2}\biggr{)}^{2}+\frac{1}{n}\sum_{i,j=1}^{n}\bigl{(}\operatorname{\mathbf{E}}|w_{ij}|^{4}-2(\operatorname{\mathbf{E}}|w_{ij}|^{2})^{2}\bigr{)}.

Since both 𝐄|wij|4\operatorname{\mathbf{E}}|w_{ij}|^{4} and (𝐄|wij|2)2(\operatorname{\mathbf{E}}|w_{ij}|^{2})^{2} are bounded above by ϵn2𝐄|wij|2\epsilon_{n}^{2}\operatorname{\mathbf{E}}|w_{ij}|^{2}, we have

|1ni,j=1n(𝐄|wij|42(𝐄|wij|2)2)|3ϵn2ni,j=1n𝐄|wij|20\biggl{|}\frac{1}{n}\sum_{i,j=1}^{n}\bigl{(}\operatorname{\mathbf{E}}|w_{ij}|^{4}-2(\operatorname{\mathbf{E}}|w_{ij}|^{2})^{2}\bigr{)}\biggr{|}\leq\frac{3\epsilon_{n}^{2}}{n}\sum_{i,j=1}^{n}\operatorname{\mathbf{E}}|w_{ij}|^{2}\to 0

by (3.3) and Lemma 3.1. ∎

Now we are ready to prove the necessity part of Theorem 2.1 assuming (3.1). Assume 𝐄μWnμsc\operatorname{\mathbf{E}}\mu_{W_{n}}\Rightarrow\mu_{\mathrm{sc}}. By Skorokhod’s representation theorem [Bil12, Theorem 25.6], we can take real-valued random variables X,X1,X2,X,X_{1},X_{2},\ldots on a common probability space such that 𝐄μWn\operatorname{\mathbf{E}}\mu_{W_{n}} is the distribution of XnX_{n}, μsc\mu_{\mathrm{sc}} is the distribution of XX, and XnXX_{n}\to X a.s.

Since

(3.4) supn𝐍𝐄Xn6=supn𝐍𝐑x6𝐄μWn(dx)<,\sup_{n\in\mathbf{N}}\operatorname{\mathbf{E}}X_{n}^{6}=\sup_{n\in\mathbf{N}}\int_{\mathbf{R}}x^{6}\,\operatorname{\mathbf{E}}\mu_{W_{n}}(dx)<\infty,

(Xn2)n𝐍(X_{n}^{2})_{n\in\mathbf{N}} and (Xn4)n𝐍(X_{n}^{4})_{n\in\mathbf{N}} are uniformly integrable. Thus, XnXX_{n}\to X a.s. implies 𝐄Xn2𝐄X2\operatorname{\mathbf{E}}X_{n}^{2}\to\operatorname{\mathbf{E}}X^{2} and 𝐄Xn4𝐄X4\operatorname{\mathbf{E}}X_{n}^{4}\to\operatorname{\mathbf{E}}X^{4}. By Lemma 3.1 and Lemma 3.2, we have

1ni,j=1n𝐄|wij|2𝐄X2=𝐑x2μsc(dx)=1\frac{1}{n}\sum_{i,j=1}^{n}\operatorname{\mathbf{E}}|w_{ij}|^{2}\to\operatorname{\mathbf{E}}X^{2}=\int_{\mathbf{R}}x^{2}\,\mu_{\mathrm{sc}}(dx)=1

and

2ni=1n(j=1n𝐄|wij|2)2𝐄X4=𝐑x4μsc(dx)=2.\frac{2}{n}\sum_{i=1}^{n}\biggl{(}\sum_{j=1}^{n}\operatorname{\mathbf{E}}|w_{ij}|^{2}\biggr{)}^{2}\to\operatorname{\mathbf{E}}X^{4}=\int_{\mathbf{R}}x^{4}\,\mu_{\mathrm{sc}}(dx)=2.

See [AGZ10, 2.1.1] for a computation of the moments of μsc\mu_{\mathrm{sc}}.

Here is the punchline: using (𝐄Y)2𝐄Y2(\operatorname{\mathbf{E}}Y)^{2}\leq\operatorname{\mathbf{E}}Y^{2} in the first line, and then applying the two convergences we have just established, we have

(1ni=1n|j=1n𝐄|wij|21|)21ni=1n(j=1n𝐄|wij|21)2=1ni=1n(j=1n𝐄|wij|2)22ni,j=1n𝐄|wij|2+112+1=0.\begin{split}\biggl{(}\frac{1}{n}\sum_{i=1}^{n}\Bigl{|}\sum_{j=1}^{n}\operatorname{\mathbf{E}}|w_{ij}|^{2}-1\Bigr{|}\biggr{)}^{2}&\leq\frac{1}{n}\sum_{i=1}^{n}\biggl{(}\sum_{j=1}^{n}\operatorname{\mathbf{E}}|w_{ij}|^{2}-1\biggr{)}^{2}\\ &=\frac{1}{n}\sum_{i=1}^{n}\biggl{(}\sum_{j=1}^{n}\operatorname{\mathbf{E}}|w_{ij}|^{2}\biggr{)}^{2}-\frac{2}{n}\sum_{i,j=1}^{n}\operatorname{\mathbf{E}}|w_{ij}|^{2}+1\\ &\to 1-2+1=0.\end{split}

Thus, the necessity in Theorem 2.1 is proved under the assumption (3.1).

4. Lifting the bounded sixth moment condition

In this section, we prove the necessity part of Theorem 2.1 without assuming the bounded sixth moment condition (3.1). We rely on the following lemma, which will be proved in the next section.

Lemma 4.1 (bounded eighth moments).

If (1.4) and (2.1) hold, then

supn𝐍𝐑x8𝐄μWn(dx)<.\sup_{n\in\mathbf{N}}\int_{\mathbf{R}}x^{8}\,\operatorname{\mathbf{E}}\mu_{W_{n}}(dx)<\infty.

The number 88 is here just because it is even and greater than 66. In fact, our proof easily extends to any even natural number. Given Kn{1,,n}K_{n}\subset\{1,\ldots,n\} for all n𝐍n\in\mathbf{N}, let WnKW_{n}^{K} be the matrix obtained from WnW_{n} by replacing wijw_{ij} with 0 for all (i,j)Kn×Kn(i,j)\not\in K_{n}\times K_{n}.

Lemma 4.2.

Assume (1.5). For any ϵ>0\epsilon>0, there exist Kn{1,,n}K_{n}\subset\{1,\ldots,n\} with |Kn|(1ϵ)n|K_{n}|\geq(1-\epsilon)n such that

j=1n𝐄|wij|2Cfor all n𝐍 and iKn\sum_{j=1}^{n}\operatorname{\mathbf{E}}|w_{ij}|^{2}\leq C\qquad\text{for all $n\in\mathbf{N}$ and $i\in K_{n}$}

for some C<C<\infty.

Proof.

We may assume ϵ<1\epsilon<1. Suppose that the claim is false, and let KnK_{n} be the set of size (1ϵ)n\lceil{(1-\epsilon)n}\rceil consisting of i{1,,n}i\in\{1,\ldots,n\} with (1ϵ)n\lceil{(1-\epsilon)n}\rceil smallest j=1n𝐄|wij|2\sum_{j=1}^{n}\operatorname{\mathbf{E}}|w_{ij}|^{2}. Notice that

miniKncj=1n𝐄|wij|2maxiKnj=1n𝐄|wij|2\min_{i\in K_{n}^{c}}\sum_{j=1}^{n}\operatorname{\mathbf{E}}|w_{ij}|^{2}\geq\max_{i\in K_{n}}\sum_{j=1}^{n}\operatorname{\mathbf{E}}|w_{ij}|^{2}\to\infty

along some subsequence (nk)k𝐍(n_{k})_{k\in\mathbf{N}}, where Knc:={1,,n}KnK_{n}^{c}:=\{1,\ldots,n\}\setminus K_{n}. For all n=n1,n2,n=n_{1},n_{2},\ldots such that the left side in the previous display is at least 11, let JnJ_{n} be any subset of KncK_{n}^{c} such that

|Jn|=ϵnminiKncj=1n𝐄|wij|2.|J_{n}|=\biggl{\lceil}\frac{\epsilon n}{\min_{i\in K_{n}^{c}}\sum_{j=1}^{n}\operatorname{\mathbf{E}}|w_{ij}|^{2}}\biggr{\rceil}.

Then |Jnk|/nk0|J_{n_{k}}|/n_{k}\to 0 follows from xx+1\lceil{x}\rceil\leq x+1. However,

1niJnj=1n𝐄|wij|2|Jn|nminiKncj=1n𝐄|wij|2ϵ\frac{1}{n}\sum_{i\in J_{n}}\sum_{j=1}^{n}\operatorname{\mathbf{E}}|w_{ij}|^{2}\geq\frac{|J_{n}|}{n}\min_{i\in K_{n}^{c}}\sum_{j=1}^{n}\operatorname{\mathbf{E}}|w_{ij}|^{2}\geq\epsilon

for n=n1,n2,n=n_{1},n_{2},\ldots for which JnJ_{n} is defined. If we let Jn:=J_{n}:=\emptyset for all nn for which JnJ_{n} is undefined, then JnJ_{n} contradicts (1.5). ∎

Given Kn{1,,n}K_{n}\subset\{1,\ldots,n\} for each n𝐍n\in\mathbf{N}, let WnKW_{n}^{K} be the matrix obtained from WnW_{n} by replacing wijw_{ij} with 0 for all (i,j)Kn×Kn(i,j)\not\in K_{n}\times K_{n}.

Lemma 4.3.

If (1.5) and (2.1) hold, then there are Kn{1,,n}K_{n}\subset\{1,\ldots,n\} with |Kn|/n1|K_{n}|/n\to 1 such that

(4.1) supn𝐍𝐑x6𝐄μWnK(dx)<.\sup_{n\in\mathbf{N}}\int_{\mathbf{R}}x^{6}\,\operatorname{\mathbf{E}}\mu_{W_{n}^{K}}(dx)<\infty.
Proof.

Let ϵ(0,1)\epsilon\in(0,1), and KnK_{n} and CC be as in the preceding lemma. Suppose that

𝐑x6𝐄μWnkK(dx)c>36as k\int_{\mathbf{R}}x^{6}\,\operatorname{\mathbf{E}}\mu_{W_{n_{k}}^{K}}(dx)\to c>3^{6}\qquad\text{as $k\to\infty$}

for some n1<n2<n_{1}<n_{2}<\cdots. By Lemmas 4.1 and 4.2,

supn𝐍𝐑x8𝐄μWnK(dx)<.\sup_{n\in\mathbf{N}}\int_{\mathbf{R}}x^{8}\,\operatorname{\mathbf{E}}\mu_{W_{n}^{K}}(dx)<\infty.

This implies that (𝐄μWnkK)k𝐍(\operatorname{\mathbf{E}}\mu_{W_{n_{k}}^{K}})_{k\in\mathbf{N}} is tight, thus it has a subsequence weakly convergent to some μ\mu, which we still, by abuse of notation, denote by (𝐄μWnkK)k𝐍(\operatorname{\mathbf{E}}\mu_{W_{n_{k}}^{K}})_{k\in\mathbf{N}}. By Skorokhod’s theorem and the uniform integrability argument that followed (3.4), we have

𝐑x6μ(dx)=limk𝐑x6𝐄μWnkK(dx)>36,\int_{\mathbf{R}}x^{6}\,\mu(dx)=\lim_{k\to\infty}\int_{\mathbf{R}}x^{6}\,\operatorname{\mathbf{E}}\mu_{W_{n_{k}}^{K}}(dx)>3^{6},

and thus μ([3,3]c)>0\mu([-3,3]^{c})>0.

If WnKW_{n}^{K} has kk eigenvalues outside [3,3][-3,3], then the Cauchy interlacing law [Tao12, Exercise 1.3.14] implies that WnW_{n} has at least kk eigenvalues outside [3,3][-3,3]. Thus

μWn([3,3]c)(1ϵ)μWnK([3,3]c),\mu_{W_{n}}([-3,3]^{c})\geq(1-\epsilon)\mu_{W_{n}^{K}}([-3,3]^{c}),

and therefore

𝐄μWn([3,3]c)(1ϵ)𝐄μWnK([3,3]c)\operatorname{\mathbf{E}}\mu_{W_{n}}([-3,3]^{c})\geq(1-\epsilon)\operatorname{\mathbf{E}}\mu_{W_{n}^{K}}([-3,3]^{c})

by Lemma A.1. Since 𝐄μWnkμsc\operatorname{\mathbf{E}}\mu_{W_{n_{k}}}\Rightarrow\mu_{\mathrm{sc}}, the portmanteau theorem implies

μsc((3,3)c)lim supk𝐄μWnk((3,3)c)(1ϵ)lim supk𝐄μWnkK((3,3)c)(1ϵ)lim infk𝐄μWnkK([3,3]c)(1ϵ)μ([3,3]c)>0,\begin{split}\mu_{\mathrm{sc}}((-3,3)^{c})&\geq\limsup_{k\to\infty}\operatorname{\mathbf{E}}\mu_{W_{n_{k}}}((-3,3)^{c})\\ &\geq(1-\epsilon)\limsup_{k\to\infty}\operatorname{\mathbf{E}}\mu_{W_{n_{k}}^{K}}((-3,3)^{c})\\ &\geq(1-\epsilon)\liminf_{k\to\infty}\operatorname{\mathbf{E}}\mu_{W_{n_{k}}^{K}}([-3,3]^{c})\\ &\geq(1-\epsilon)\mu([-3,3]^{c})>0,\end{split}

but this contradicts the fact that μsc\mu_{\mathrm{sc}} is supported on [2,2][-2,2]. Thus, we have

lim supn𝐑x6𝐄μWnK(dx)36.\limsup_{n\to\infty}\int_{\mathbf{R}}x^{6}\,\operatorname{\mathbf{E}}\mu_{W_{n}^{K}}(dx)\leq 3^{6}.

Since ϵ>0\epsilon>0 is arbitrary, we have actually proved that for each ϵ>0\epsilon>0 we can choose Knϵ{1,,n}K_{n}^{\epsilon}\subset\{1,\ldots,n\} such that |Knϵ|(1ϵ)n|K^{\epsilon}_{n}|\geq(1-\epsilon)n and

lim supn𝐑x6𝐄μWnKϵ(dx)36.\limsup_{n\to\infty}\int_{\mathbf{R}}x^{6}\,\operatorname{\mathbf{E}}\mu_{W_{n}^{K^{\epsilon}}}(dx)\leq 3^{6}.

Choose positive integers m1<m2<m_{1}<m_{2}<\cdots so that

𝐑x6𝐄μWnK1/k36+1for all nmk,\int_{\mathbf{R}}x^{6}\,\operatorname{\mathbf{E}}\mu_{W_{n}^{K^{1/k}}}\leq 3^{6}+1\qquad\text{for all $n\geq m_{k}$,}

and let Kn:=Kn1/kK_{n}:=K_{n}^{1/k} for n=mk,,mk+11n=m_{k},\ldots,m_{k+1}-1 and Kn:=K_{n}:=\emptyset for n=1,,m11n=1,\ldots,m_{1}-1. (We are redefining KnK_{n} by abuse of notation.) Then we have |Kn|/n1|K_{n}|/n\to 1 and (4.1). ∎

We are ready to prove the necessity part of Theorem 2.1. Assume (1.5), (2.1), and 𝐄μWnμsc\operatorname{\mathbf{E}}\mu_{W_{n}}\Rightarrow\mu_{\mathrm{sc}}. By Lemma 1.3, we have μWnμsc\mu_{W_{n}}\Rightarrow\mu_{\mathrm{sc}} a.s. If K1,K2,K_{1},K_{2},\ldots are as in Lemma 4.3, then

rank(WnWnK)n2(n|Kn|)n0,\frac{\operatorname{rank}(W_{n}-W_{n}^{K})}{n}\leq\frac{2(n-|K_{n}|)}{n}\to 0,

and thus μWnKμsc\mu_{W_{n}^{K}}\Rightarrow\mu_{\mathrm{sc}} a.s. by Lemma 1.5. By another application of Lemma 1.3, 𝐄μWnKμsc\operatorname{\mathbf{E}}\mu_{W_{n}^{K}}\Rightarrow\mu_{\mathrm{sc}}. As we have (4.1), the previous section tells us that

1niKn|jKn𝐄|wij|21|0.\frac{1}{n}\sum_{i\in K_{n}}\Bigl{|}\sum_{j\in K_{n}}\operatorname{\mathbf{E}}|w_{ij}|^{2}-1\Bigr{|}\to 0.

Since |Knc|/n0|K_{n}^{c}|/n\to 0, the assumption (1.5) implies (1.3). Thus, the necessity part of Theorem 2.1 is proved assuming that Lemma 4.1 holds.

5. Computation of moments

The goal of this section is to prove Lemma 4.1 and also establish some arguments needed in the next section. We use a variant of Wigner’s original moment method that can handle entries with non-identical variances. Those that are very familiar with these arguments may want to jump ahead to the proof of Lemma 4.1.

Assume (1.4) and (2.1) throughout this section. Recall that

(3.2) 𝐑xk𝐄μWn(dx)=1n𝐄trWnk\int_{\mathbf{R}}x^{k}\,\operatorname{\mathbf{E}}\mu_{W_{n}}(dx)=\frac{1}{n}\operatorname{\mathbf{E}}\operatorname{tr}W_{n}^{k}

for all k𝐍k\in\mathbf{N}. In this section, we compute the asymptotics of n1𝐄trWnkn^{-1}\operatorname{\mathbf{E}}\operatorname{tr}W_{n}^{k} as nn\to\infty.

Fix k𝐍k\in\mathbf{N}. The boldface lower case letters 𝐢,𝐣,\mathbf{i},\mathbf{j},\ldots will denote (i0,,ik)(i_{0},\ldots,i_{k}), (j0,,jk)(j_{0},\ldots,j_{k}), and so on. Let us call a (k+1)(k+1)-tuple 𝐢\mathbf{i} with i0=iki_{0}=i_{k} a closed walk of length kk. For any closed walk 𝐢\mathbf{i} with i0,,ik{1,,n}i_{0},\ldots,i_{k}\in\{1,\ldots,n\}, let

w𝐢:=s=1kwis1is.w_{\mathbf{i}}:=\prod_{s=1}^{k}w_{i_{s-1}i_{s}}.

Notice that

1n𝐄trWnk=1n𝐢𝐄w𝐢\frac{1}{n}\operatorname{\mathbf{E}}\operatorname{tr}W_{n}^{k}=\frac{1}{n}\sum_{\mathbf{i}}\operatorname{\mathbf{E}}w_{\mathbf{i}}

where 𝐢\mathbf{i} ranges over all closed walks 𝐢\mathbf{i} (of length kk) with i0,,ik{1,,n}i_{0},\ldots,i_{k}\in\{1,\ldots,n\}.

Now we gather together the closed walks which have the same “shape.” Let us say that two closed walks 𝐢\mathbf{i} and 𝐣\mathbf{j} are isomorphic if for any s,t=0,,ks,t=0,\ldots,k we have is=iti_{s}=i_{t} if and only if js=jtj_{s}=j_{t}. A canonical closed walk of length kk on t𝐍t\in\mathbf{N} vertices is a closed walk 𝐜\mathbf{c} such that

  1. (1)

    c0=ck=1c_{0}=c_{k}=1,

  2. (2)

    {c0,,ck}={1,,t}\{c_{0},\ldots,c_{k}\}=\{1,\ldots,t\}, and

  3. (3)

    cimax{c0,,ci1}+1c_{i}\leq\max\{c_{0},\ldots,c_{i-1}\}+1 for each i=1,,ki=1,\ldots,k.

Let γ(k,t)\gamma(k,t) denote the set of such walks. It is straightforward to show that any closed walk is isomorphic to exactly one canonical closed walk. For each 𝐜γ(k,t)\mathbf{c}\in\gamma(k,t), let L(n,𝐜)L(n,\mathbf{c}) denote the set of all closed walks 𝐢\mathbf{i} with i0,,ik{1,,n}i_{0},\ldots,i_{k}\in\{1,\ldots,n\} which are isomorphic to 𝐜\mathbf{c}. Then we have

(5.1) 1n𝐄trWnk=1ni0,,ik=1n𝐄w𝐢=1nt=1k𝐜γ(k,t)𝐢L(n,𝐜)𝐄w𝐢,\frac{1}{n}\operatorname{\mathbf{E}}\operatorname{tr}W_{n}^{k}=\frac{1}{n}\sum_{i_{0},\ldots,i_{k}=1}^{n}\operatorname{\mathbf{E}}w_{\mathbf{i}}=\frac{1}{n}\sum_{t=1}^{k}\sum_{\mathbf{c}\in\gamma(k,t)}\sum_{\mathbf{i}\in L(n,\mathbf{c})}\operatorname{\mathbf{E}}w_{\mathbf{i}},

where the upper bound of tt is (rather arbitrarily) set to kk since γ(k,t)\gamma(k,t) is empty for all t>kt>k.

We will fix t𝐍t\in\mathbf{N} and 𝐜γ(k,t)\mathbf{c}\in\gamma(k,t), and compute n1𝐢L(n,𝐜)𝐄w𝐢n^{-1}\sum_{\mathbf{i}\in L(n,\mathbf{c})}\operatorname{\mathbf{E}}w_{\mathbf{i}}. As a first step, we get an easy case out of the way.

Lemma 5.1 (zeroed out terms).

If 𝐜\mathbf{c} crosses some edge {i,j}\{i,j\} exactly once, i.e., {cs1,cs}={i,j}\{c_{s-1},c_{s}\}=\{i,j\} for exactly one s{1,,k}s\in\{1,\ldots,k\}, then 𝐄w𝐢=0\operatorname{\mathbf{E}}w_{\mathbf{i}}=0 for any n𝐍n\in\mathbf{N} and 𝐢L(n,𝐜)\mathbf{i}\in L(n,\mathbf{c}).

Proof.

Since we have (2.1) and the upper triangular entries of WnW_{n} are independent, w𝐢w_{\mathbf{i}} is the product of wijw_{ij} (or wjiw_{ji}) and a bounded random variable which is independent from wijw_{ij}. Since 𝐄wij=0\operatorname{\mathbf{E}}w_{ij}=0, we have 𝐄w𝐢=0\operatorname{\mathbf{E}}w_{\mathbf{i}}=0. ∎

Now assume that 𝐜\mathbf{c} does not cross any edge exactly once, i.e., for each s{1,,k}s\in\{1,\ldots,k\} there is some r{1,,k}r\in\{1,\ldots,k\} distinct from ss such that {cs1,cs}={cr1,cr}\{c_{s-1},c_{s}\}=\{c_{r-1},c_{r}\}. To compute n1iL(n,𝐜)𝐄w𝐢n^{-1}\sum_{i\in L(n,\mathbf{c})}\operatorname{\mathbf{E}}w_{\mathbf{i}}, we introduce some notation. Let G(𝐜)G(\mathbf{c}) be the graph (possibly having loops but no multiple edges) with the vertex set

{i0,,ik}\{i_{0},\ldots,i_{k}\}

and the edge set

{{it1,it}|t=1,,k}.\bigl{\{}\{i_{t-1},i_{t}\}\bigm{|}t=1,\ldots,k\bigr{\}}.

For a tree TT, let V(T)V(T) and E(T)E(T) denote the vertex set and the edge set of TT. Given a finite tree TT and n𝐍n\in\mathbf{N}, we let I(T,n)I(T,n) denote the set of injections from the vertex set V(T)V(T) of TT to {1,,n}\{1,\ldots,n\}. For each FI(T,n)F\in I(T,n), let us write

Π(F):=eE(T)𝐄|wF(xe)F(ye)|2\Pi(F):=\prod_{e\in E(T)}\operatorname{\mathbf{E}}|w_{F(x_{e})F(y_{e})}|^{2}

where xe,yeV(T)x_{e},y_{e}\in V(T) denote the endpoints of eE(T)e\in E(T). We are omitting the dependence of Π(F)\Pi(F) on TT, but there should be no confusion. Note that the value Π(F)\Pi(F) is well-defined because WnW_{n} is Hermitian.

Now we get back to the problem of computing n1iL(n,𝐜)𝐄w𝐢n^{-1}\sum_{i\in L(n,\mathbf{c})}\operatorname{\mathbf{E}}w_{\mathbf{i}}. As each edge of G(𝐜)G(\mathbf{c}) is crossed at least twice by 𝐜\mathbf{c}, there are at most k/2k/2 edges in G(𝐜)G(\mathbf{c}). As G(𝐜)G(\mathbf{c}) is a connected graph with tt vertices, we have tk/2+1t\leq k/2+1, and G(𝐜)G(\mathbf{c}) has a spanning tree SS with t1t-1 edges. Choose some SS. For each 𝐢L(n,𝐜)\mathbf{i}\in L(n,\mathbf{c}), consider the injection F𝐢:{1,,t}{1,,n}F_{\mathbf{i}}\colon\{1,\ldots,t\}\to\{1,\ldots,n\} given by F𝐢(cs):=isF_{\mathbf{i}}(c_{s}):=i_{s} for each s=0,,ks=0,\ldots,k.

First assume t=k/2+1t=k/2+1. Since SS has k/2k/2 edges and each edge of SS is crossed twice by 𝐜\mathbf{c}, we have S=G(𝐜)S=G(\mathbf{c}). As each edge of G(𝐜)G(\mathbf{c}) is traversed exactly once in each direction, and the map L(n,𝐜)I(S,n)L(n,\mathbf{c})\to I(S,n) given by 𝐢F𝐢\mathbf{i}\mapsto F_{\mathbf{i}} is a bijection, we have

(5.2) 1n𝐢L(n,𝐜)𝐄w𝐢=1n𝐢L(n,𝐜)Π(F𝐢)=1nFI(S,n)Π(F).\frac{1}{n}\sum_{\mathbf{i}\in L(n,\mathbf{c})}\operatorname{\mathbf{E}}w_{\mathbf{i}}=\frac{1}{n}\sum_{\mathbf{i}\in L(n,\mathbf{c})}\Pi(F_{\mathbf{i}})=\frac{1}{n}\sum_{F\in I(S,n)}\Pi(F).

Now assume t<k/2+1t<k/2+1. By |wij|ϵn|w_{ij}|\leq\epsilon_{n} and the fact that 𝐜\mathbf{c} crosses any edge of G(𝐜)G(\mathbf{c}) at least twice, we have

|𝐄w𝐢|ϵnk2(t1)Π(F𝐢).|\operatorname{\mathbf{E}}w_{\mathbf{i}}|\leq\epsilon_{n}^{k-2(t-1)}\Pi(F_{\mathbf{i}}).

Note that ϵnk2(t1)0\epsilon_{n}^{k-2(t-1)}\to 0 since t<k/2+1t<k/2+1. By using the bijection L(n,𝐜)I(S,n)L(n,\mathbf{c})\to I(S,n) again, we have

(5.3) 1n𝐢L(n,𝐜)|𝐄w𝐢|ϵnk2(t1)n𝐢L(n,𝐜)Π(F𝐢)=ϵnk2(t1)nFI(S,n)Π(F).\begin{split}\frac{1}{n}\sum_{\mathbf{i}\in L(n,\mathbf{c})}|\operatorname{\mathbf{E}}w_{\mathbf{i}}|&\leq\frac{\epsilon_{n}^{k-2(t-1)}}{n}\sum_{\mathbf{i}\in L(n,\mathbf{c})}\Pi(F_{\mathbf{i}})\\ &=\frac{\epsilon_{n}^{k-2(t-1)}}{n}\sum_{F\in I(S,n)}\Pi(F).\end{split}

The right side tends to 0 by the following.

Lemma 5.2 (contribution of a tree).

If TT is a finite tree with mm edges, xV(T)x\in V(T), n𝐍n\in\mathbf{N}, and i{1,,n}i\in\{1,\ldots,n\}, then

(5.4) FI(T,n)F(x)=iΠ(F)Cm\sum_{\begin{subarray}{c}F\in I(T,n)\\ F(x)=i\end{subarray}}\Pi(F)\leq C^{m}

where CC is as in (1.4). Note that it follows that

1nFI(T,n)Π(F)Cm.\frac{1}{n}\sum_{F\in I(T,n)}\Pi(F)\leq C^{m}.
Proof.

There is nothing to prove if m=0m=0. To proceed by induction, assume that (5.4) holds, and let TT be a tree with m+1m+1 edges. Pick any leaf yV(T)y\in V(T) distinct from xx, and let zV(T)z\in V(T) be the only vertex that is adjacent to yy. Since

FI(T,n)F(x)=i,F(z)=jΠ(F)HI(Ty,n)H(x)=i,H(z)=j(Π(H)l=1n𝐄|wjl|2)CHI(Ty,n)H(x)=i,H(z)=jΠ(H)\begin{split}\sum_{\begin{subarray}{c}F\in I(T,n)\\ F(x)=i,F(z)=j\end{subarray}}\Pi(F)&\leq\sum_{\begin{subarray}{c}H\in I(T\setminus y,n)\\ H(x)=i,H(z)=j\end{subarray}}\biggl{(}\Pi(H)\sum_{l=1}^{n}\operatorname{\mathbf{E}}|w_{jl}|^{2}\biggr{)}\\ &\leq C\sum_{\begin{subarray}{c}H\in I(T\setminus y,n)\\ H(x)=i,H(z)=j\end{subarray}}\Pi(H)\end{split}

by (1.4) for all j=1,,nj=1,\ldots,n, we have

FI(T,n)F(x)=iΠ(F)=j=1nFI(T,n)F(x)=i,F(z)=jΠ(F)Cj=1nHI(Ty,n)H(x)=i,H(z)=jΠ(H)=CHI(Ty,n)H(x)=iΠ(H)Cm+1\begin{split}\sum_{\begin{subarray}{c}F\in I(T,n)\\ F(x)=i\end{subarray}}\Pi(F)&=\sum_{j=1}^{n}\sum_{\begin{subarray}{c}F\in I(T,n)\\ F(x)=i,F(z)=j\end{subarray}}\Pi(F)\\ &\leq C\sum_{j=1}^{n}\sum_{\begin{subarray}{c}H\in I(T\setminus y,n)\\ H(x)=i,H(z)=j\end{subarray}}\Pi(H)=C\sum_{\begin{subarray}{c}H\in I(T\setminus y,n)\\ H(x)=i\end{subarray}}\Pi(H)\leq C^{m+1}\end{split}

by the induction hypothesis. ∎

We have now shown the following; see (5.2) and (5.3).

Lemma 5.3 (contribution of a canonical walk).

Let t𝐍t\in\mathbf{N} and 𝐜γ(k,t)\mathbf{c}\in\gamma(k,t). Assume that 𝐜\mathbf{c} does not cross any edge exactly one, i.e., for each s{1,,k}s\in\{1,\ldots,k\} there is some r{1,,k}r\in\{1,\ldots,k\} distinct from ss such that {cs1,cs}={cr1,cr}\{c_{s-1},c_{s}\}=\{c_{r-1},c_{r}\}. Then we have tk/2+1t\leq k/2+1. If t<k/2+1t<k/2+1, we have

1n𝐢L(n,𝐜)𝐄w𝐢0.\frac{1}{n}\sum_{\mathbf{i}\in L(n,\mathbf{c})}\operatorname{\mathbf{E}}w_{\mathbf{i}}\to 0.

If t=k/2+1t=k/2+1, then G(𝐜)G(\mathbf{c}) is a tree, and we have

1n𝐢L(n,𝐜)𝐄w𝐢=1nFI(G(𝐜),n)Π(F).\frac{1}{n}\sum_{\mathbf{i}\in L(n,\mathbf{c})}\operatorname{\mathbf{E}}w_{\mathbf{i}}=\frac{1}{n}\sum_{F\in I(G(\mathbf{c}),n)}\Pi(F).

Combining (3.2), (5.1), Lemma 5.1, and Lemma 5.3, we obtain the following approximation to the moments of 𝐄μWn\operatorname{\mathbf{E}}\mu_{W_{n}}.

Lemma 5.4 (computation of moments).

Let Γk\Gamma_{k} be the set of all 𝐜γ(k,k/2+1)\mathbf{c}\in\gamma(k,k/2+1) which crosses each edge of G(𝐜)G(\mathbf{c}) twice. (Note that G(𝐜)G(\mathbf{c}) should be a tree, and that Γk\Gamma_{k} is finite.) Then,

𝐑xk𝐄μWn(dx)1n𝐜ΓkFI(G(𝐜),n)Π(F)0.\int_{\mathbf{R}}x^{k}\,\operatorname{\mathbf{E}}\mu_{W_{n}}(dx)-\frac{1}{n}\sum_{\mathbf{c}\in\Gamma_{k}}\sum_{F\in I(G(\mathbf{c}),n)}\Pi(F)\to 0.

We now prove Lemma 4.1 as promised.

Proof of Lemma 4.1.

Let Γk\Gamma_{k} be as in Lemma 5.4. By Lemma 5.2, we have

1nFI(G(𝐜),n)Π(F)C4for all 𝐜Γ8.\frac{1}{n}\sum_{F\in I(G(\mathbf{c}),n)}\Pi(F)\leq C^{4}\qquad\text{for all $\mathbf{c}\in\Gamma_{8}$.}

Since Γ8\Gamma_{8} is finite, we have

supn𝐍(1n𝐜Γ8FI(G(𝐜),n)Π(F))<,\sup_{n\in\mathbf{N}}\biggl{(}\frac{1}{n}\sum_{\mathbf{c}\in\Gamma_{8}}\sum_{F\in I(G(\mathbf{c}),n)}\Pi(F)\biggr{)}<\infty,

and thus

supn𝐍𝐑x8𝐄μWn(dx)<\sup_{n\in\mathbf{N}}\int_{\mathbf{R}}x^{8}\,\operatorname{\mathbf{E}}\mu_{W_{n}}(dx)<\infty

by Lemma 5.4. ∎

6. Proof of sufficiency

We continue to use the notation introduced in Section 5. On top of (1.3) and (2.1), we assume (1.4), which is possible by Lemma 2.2. We are one lemma away from proving the sufficiency part of Theorem 2.1. Our proof is essentially a manifestation of Wigner’s original idea.

Lemma 6.1 (each tree contributes one).

For any finite tree TT we have

(6.1) limn1nFI(T,n)Π(F)=1.\lim_{n\to\infty}\frac{1}{n}\sum_{F\in I(T,n)}\Pi(F)=1.

(Compare it to Lemma 5.2.)

Proof.

There is nothing to prove if TT has no edges. To proceed by induction, assume that (6.1) holds if TT has mm edges, and let TT be a tree with m+1m+1 edges. Let xV(T)x\in V(T) be a leaf of TT, and yy be the only vertex of TT that is adjacent to xx. Note that

|1nFI(T,n)Π(F)1nHI(Tx,n)(Π(H)i=1n𝐄|wH(y)i|2)|(m+1)ϵn2nHI(Tx,n)Π(H)0\begin{split}\biggl{|}\frac{1}{n}\sum_{F\in I(T,n)}\Pi(F)-\frac{1}{n}\sum_{H\in I(T\setminus x,n)}&\biggl{(}\Pi(H)\sum_{i=1}^{n}\operatorname{\mathbf{E}}|w_{H(y)i}|^{2}\biggr{)}\biggr{|}\\ &\leq\frac{(m+1)\epsilon_{n}^{2}}{n}\sum_{H\in I(T\setminus x,n)}\Pi(H)\to 0\end{split}

by |wij|ϵn|w_{ij}|\leq\epsilon_{n} and Lemma 5.2 (or the induction hypothesis). Applying Lemma 5.2 once again, we have

|1nHI(Tx,n)(Π(H)i=1n(𝐄|wH(y)i|21n))|=1n|j=1n[i=1n(𝐄|wji|21n)HI(Tx,n)H(y)=jΠ(H)]|Cmnj=1n|i=1n(𝐄|wji|21n)|0\begin{split}\biggl{|}\frac{1}{n}\sum_{H\in I(T\setminus x,n)}&\biggl{(}\Pi(H)\sum_{i=1}^{n}\biggl{(}\operatorname{\mathbf{E}}|w_{H(y)i}|^{2}-\frac{1}{n}\biggr{)}\biggr{)}\biggr{|}\\ &=\frac{1}{n}\biggl{|}\sum_{j=1}^{n}\biggl{[}\sum_{i=1}^{n}\biggl{(}\operatorname{\mathbf{E}}|w_{ji}|^{2}-\frac{1}{n}\biggr{)}\cdot\sum_{\begin{subarray}{c}H\in I(T\setminus x,n)\\ H(y)=j\end{subarray}}\Pi(H)\biggr{]}\biggr{|}\\ &\leq\frac{C^{m}}{n}\sum_{j=1}^{n}\biggl{|}\sum_{i=1}^{n}\biggl{(}\operatorname{\mathbf{E}}|w_{ji}|^{2}-\frac{1}{n}\biggr{)}\biggr{|}\to 0\end{split}

by (1.3). The claimed result follows from the previous two displays and

limn1nHI(Tx,n)Π(H)=1.\lim_{n\to\infty}\frac{1}{n}\sum_{H\in I(T\setminus x,n)}\Pi(H)=1.\qed

By Lemma 5.4 and Lemma 6.1, we have

𝐑xk𝐄μWn(dx)|Γk|.\int_{\mathbf{R}}x^{k}\,\operatorname{\mathbf{E}}\mu_{W_{n}}(dx)\to|\Gamma_{k}|.

If kk is odd, then |Γk|=0|\Gamma_{k}|=0 because k/2+1k/2+1 is not an integer.

Assume that kk is even. A Dyck path of length kk is a finite sequence (x0,,xk)(x_{0},\ldots,x_{k}) satisfying

  1. (1)

    x0=xk=0x_{0}=x_{k}=0,

  2. (2)

    |xsxs1|=1|x_{s}-x_{s-1}|=1 for all s=1,,ks=1,\ldots,k, and

  3. (3)

    xs0x_{s}\geq 0 for all s=0,,ks=0,\ldots,k.

Given 𝐜Γk\mathbf{c}\in\Gamma_{k}, let D(𝐜):=(x0,,xk)D(\mathbf{c}):=(x_{0},\ldots,x_{k}) where xsx_{s} is the distance between c0c_{0} and csc_{s} in G(𝐜)G(\mathbf{c}). Then it is clear that D(𝐜)D(\mathbf{c}) is indeed a Dyck path, and it is not difficult to see that DD is a bijection from Γk\Gamma_{k} to the set of all Dyck paths of length kk. It is well-known that there are exactly 1k/2+1(kk/2)\frac{1}{k/2+1}\binom{k}{k/2} Dyck paths of length kk; see [vLW01, Example 14.8]. Thus, we have |Γk|=1k/2+1(kk/2)|\Gamma_{k}|=\frac{1}{k/2+1}\binom{k}{k/2}.

A direct computation (see [AGZ10, 2.1.1]) yields

𝐑xkμsc(dx)=1k/2+1(kk/2)for all even k𝐍,\int_{\mathbf{R}}x^{k}\,\mu_{\mathrm{sc}}(dx)=\frac{1}{k/2+1}\binom{k}{k/2}\qquad\text{for all even $k\in\mathbf{N}$,}

where the odd moments of μsc\mu_{\mathrm{sc}} are all zero due to the symmetry. Thus,

𝐑xk𝐄μWn(dx)𝐑xkμsc(dx)for all k𝐍.\int_{\mathbf{R}}x^{k}\,\operatorname{\mathbf{E}}\mu_{W_{n}}(dx)\to\int_{\mathbf{R}}x^{k}\,\mu_{\mathrm{sc}}(dx)\qquad\text{for all $k\in\mathbf{N}$.}

Since

|k=11k!𝐑xkμsc(dx)rk|k=1|2r|kk!<\biggl{|}\sum_{k=1}^{\infty}\frac{1}{k!}\int_{\mathbf{R}}x^{k}\,\mu_{\mathrm{sc}}(dx)\,r^{k}\biggr{|}\leq\sum_{k=1}^{\infty}\frac{|2r|^{k}}{k!}<\infty

by the ratio test for all r𝐑r\in\mathbf{R}, the probability measure μsc\mu_{\mathrm{sc}} is determined by its moments by [Bil12, Theorem 30.1]. Therefore, the moment convergence theorem [Bil12, Theorem 30.2] tells us that 𝐄μWnμsc\operatorname{\mathbf{E}}\mu_{W_{n}}\Rightarrow\mu_{\mathrm{sc}}.

7. Gaussian convergence

Assume that (Wn)n𝐍(W_{n})_{n\in\mathbf{N}} satisfies the conditions of Theorem 1.8. In this section, we prove Corollary 1.10 by showing that (1.9) and (1.10) are equivalent. We need the following two simple facts.

Lemma 7.1 (converging averages).

For each n𝐍n\in\mathbf{N}, let x1(n),,xn(n)0x_{1}^{(n)},\ldots,x_{n}^{(n)}\geq 0. If

1ni=1nxi(n)0,\frac{1}{n}\sum_{i=1}^{n}x_{i}^{(n)}\to 0,

then we can take nonempty Kn{1,,n}K_{n}\subset\{1,\ldots,n\} for each n𝐍n\in\mathbf{N} so that

|Kn|n1andmaxiKn|xi(n)|0.\frac{|K_{n}|}{n}\to 1\qquad\text{and}\qquad\max_{i\in K_{n}}|x_{i}^{(n)}|\to 0.
Proof.

For each ϵ>0\epsilon>0, we have

|{i:xi(n)>ϵ}|n1nϵi=1nxi(n)0.\frac{|\{i:x_{i}^{(n)}>\epsilon\}|}{n}\leq\frac{1}{n\epsilon}\sum_{i=1}^{n}x_{i}^{(n)}\to 0.

We can take positive ϵ1,ϵ2,\epsilon_{1},\epsilon_{2},\ldots with ϵn0\epsilon_{n}\to 0 such that

|{i:xi(n)>ϵn}|n0.\frac{|\{i:x_{i}^{(n)}>\epsilon_{n}\}|}{n}\to 0.

Let KnK_{n} be {i:xi(n)ϵn}\{i:x_{i}^{(n)}\leq\epsilon_{n}\} if it is nonempty, and let Kn:={1}K_{n}:=\{1\} otherwise. ∎

Lemma 7.2 (uniform convergence).

Let (E,d)(E,d) be a metric space, A1,A2,EA_{1},A_{2},\ldots\subset E, and xEx\in E. Then the following are equivalent:

  1. (1)

    xnxx_{n}\to x for any choice of x1A1,x2A2,x_{1}\in A_{1},x_{2}\in A_{2},\ldots\,.

  2. (2)

    supyAnd(x,y)0\sup_{y\in A_{n}}d(x,y)\to 0.

Proof.

We omit the easy proof. ∎

By (1.7) and Lemma 7.1, for each ϵ>0\epsilon>0 we have nonempty Knϵ{1,,n}K_{n}^{\epsilon}\subset\{1,\ldots,n\} with

|Knϵ|n1andmaxiKnϵj=1n𝐏(|wij|>ϵ)0.\frac{|K_{n}^{\epsilon}|}{n}\to 1\qquad\text{and}\qquad\max_{i\in K_{n}^{\epsilon}}\sum_{j=1}^{n}\operatorname{\mathbf{P}}(|w_{ij}|>\epsilon)\to 0.

We can take ϵ1ϵ2\epsilon_{1}\geq\epsilon_{2}\geq\cdots with ϵn0\epsilon_{n}\to 0 such that

(7.1) |Knϵn|n1andmaxiKnϵnj=1n𝐏(|wij|>ϵn)0.\frac{|K_{n}^{\epsilon_{n}}|}{n}\to 1\qquad\text{and}\qquad\max_{i\in K_{n}^{\epsilon_{n}}}\sum_{j=1}^{n}\operatorname{\mathbf{P}}(|w_{ij}|>\epsilon_{n})\to 0.

First assume (1.9). By Lemma 7.1, we have nonempty Kn{1,,n}K_{n}\subset\{1,\ldots,n\} such that

|Kn|n1andmaxiKn|j=1n𝐄[|wij|2;|wij|1]1|0.\frac{|K_{n}|}{n}\to 1\qquad\text{and}\qquad\max_{i\in K_{n}}\Bigl{|}\sum_{j=1}^{n}\operatorname{\mathbf{E}}[|w_{ij}|^{2};|w_{ij}|\leq 1]-1\Bigr{|}\to 0.

By (7.1), we can make KnK_{n} smaller so that

maxiKnj=1n𝐏(|wij|>ϵ)0\max_{i\in K_{n}}\sum_{j=1}^{n}\operatorname{\mathbf{P}}(|w_{ij}|>\epsilon)\to 0

also holds while retaining |Kn|/n1|K_{n}|/n\to 1.

Let inKni_{n}\in K_{n} for each n𝐍n\in\mathbf{N}. By Lemma 7.2, we have

j=1n𝐄[|winj|2;|winj|1]1andj=1n𝐏(|winj|>ϵn)0.\sum_{j=1}^{n}\operatorname{\mathbf{E}}[|w_{i_{n}j}|^{2};|w_{i_{n}j}|\leq 1]\to 1\qquad\text{and}\qquad\sum_{j=1}^{n}\operatorname{\mathbf{P}}(|w_{i_{n}j}|>\epsilon_{n})\to 0.

Since

j=1n𝐄[|winj|2;ϵ<|winj|1]j=1n𝐏(|winj|>ϵ)0\sum_{j=1}^{n}\operatorname{\mathbf{E}}[|w_{i_{n}j}|^{2};\epsilon<|w_{i_{n}j}|\leq 1]\leq\sum_{j=1}^{n}\operatorname{\mathbf{P}}(|w_{i_{n}j}|>\epsilon)\to 0

for all ϵ>0\epsilon>0, the Lindeberg–Feller central limit theorem [Kal02, Theorem 5.12] implies

j=1n±|winj|𝟏{|winj|1}Z\sum_{j=1}^{n}\pm|w_{i_{n}j}|\operatorname{\mathbf{1}}\{|w_{i_{n}j}|\leq 1\}\Rightarrow Z

where ZZ is standard normal. As j=1n𝐏(|winj|>1)0\sum_{j=1}^{n}\operatorname{\mathbf{P}}(|w_{i_{n}j}|>1)\to 0, it follows that j=1n±|winj|Z\sum_{j=1}^{n}\pm|w_{i_{n}j}|\Rightarrow Z.

Since i1,i2,i_{1},i_{2},\ldots are arbitrary, Lemma 7.2 implies

maxiKnL(Fnin,G)0.\max_{i\in K_{n}}L(F_{ni_{n}},G)\to 0.

As the Lévy distance is bounded above by 11, we conclude that

1ni=1nL(Fni,G)maxiKnL(Fnin,G)+n|Kn|n0.\frac{1}{n}\sum_{i=1}^{n}L(F_{ni},G)\leq\max_{i\in K_{n}}L(F_{ni_{n}},G)+\frac{n-|K_{n}|}{n}\to 0.

Now we assume (1.10). By Lemma 7.1 and (7.1), we have nonempty Kn{1,,n}K_{n}\subset\{1,\ldots,n\} with |Kn|/n1|K_{n}|/n\to 1,

maxiKnL(Fni,G)0,andmaxiKnj=1n𝐏(|wij|>ϵn)0.\max_{i\in K_{n}}L(F_{ni},G)\to 0,\qquad\text{and}\qquad\max_{i\in K_{n}}\sum_{j=1}^{n}\operatorname{\mathbf{P}}(|w_{ij}|>\epsilon_{n})\to 0.

Let inKni_{n}\in K_{n} for each n𝐍n\in\mathbf{N}. By Lemma 7.2, we have

j=1n±|winj|Zandj=1n𝐏(|winj|>ϵn)0.\sum_{j=1}^{n}\pm|w_{i_{n}j}|\Rightarrow Z\qquad\text{and}\qquad\sum_{j=1}^{n}\operatorname{\mathbf{P}}(|w_{i_{n}j}|>\epsilon_{n})\to 0.

Since j=1n𝐏(|winj|>1)0\sum_{j=1}^{n}\operatorname{\mathbf{P}}(|w_{i_{n}j}|>1)\to 0, we have

(7.2) j=1n±|winj|𝟏{|winj|1}Z.\sum_{j=1}^{n}\pm|w_{i_{n}j}|\operatorname{\mathbf{1}}\{|w_{i_{n}j}|\leq 1\}\Rightarrow Z.

Let

cn:=j=1n𝐄[|winj|2;|winj|1].c_{n}:=\sum_{j=1}^{n}\operatorname{\mathbf{E}}[|w_{i_{n}j}|^{2};|w_{i_{n}j}|\leq 1].

If cn0c_{n}\to 0 along some subsequence, then

𝐄(j=1n±|winj|𝟏{|winj|1})20\operatorname{\mathbf{E}}\Bigl{(}\sum_{j=1}^{n}\pm|w_{i_{n}j}|\operatorname{\mathbf{1}}\{|w_{i_{n}j}|\leq 1\}\Bigr{)}^{2}\to 0

along that subsequence, but it contradicts (7.2). If cnc(0,]c_{n}\to c\in(0,\infty] along some subsequence, then

1cnj=1n±|winj|𝟏{|winj|1}Z\frac{1}{\sqrt{c_{n}}}\sum_{j=1}^{n}\pm|w_{i_{n}j}|\operatorname{\mathbf{1}}\{|w_{i_{n}j}|\leq 1\}\Rightarrow Z

along that subsequence by the Lindeberg–Feller central limit theorem, and so c=1c=1. Thus, we have cn1c_{n}\to 1.

Since i1,i2,i_{1},i_{2},\ldots are arbitrary, Lemma 7.2 and (1.8) imply

1ni=1n|j=1n𝐄[|wij|2;|wij|1]1|maxiKn|j=1n𝐄[|wij|2;|wij|1]1|+1niKncj=1n𝐄[|wij|2;|wij|1]+n|Kn|n0.\begin{split}\frac{1}{n}\sum_{i=1}^{n}\Bigl{|}\sum_{j=1}^{n}\operatorname{\mathbf{E}}[|w_{ij}|^{2}&;|w_{ij}|\leq 1]-1\Bigr{|}\\ &\leq\max_{i\in K_{n}}\Bigl{|}\sum_{j=1}^{n}\operatorname{\mathbf{E}}[|w_{ij}|^{2};|w_{ij}|\leq 1]-1\Bigr{|}\\ &\quad+\frac{1}{n}\sum_{i\in K_{n}^{c}}\sum_{j=1}^{n}\operatorname{\mathbf{E}}[|w_{ij}|^{2};|w_{ij}|\leq 1]+\frac{n-|K_{n}|}{n}\to 0.\end{split}

Appendix A Mean probability measures

In this section, we clarify what we mean by 𝐄μWn\operatorname{\mathbf{E}}\mu_{W_{n}}, and prove that 𝐄μWnμsc\operatorname{\mathbf{E}}\mu_{W_{n}}\Rightarrow\mu_{\mathrm{sc}} is equivalent to μWnμsc\mu_{W_{n}}\Rightarrow\mu_{\mathrm{sc}} a.s. if (Wn)n𝐍(W_{n})_{n\in\mathbf{N}} is a Hermitian Wigner ensemble.

Let Pr(𝐑)\Pr(\mathbf{R}) be the set of all Borel probability measures on 𝐑\mathbf{R}. Equip Pr(𝐑)\Pr(\mathbf{R}) with the smallest σ\sigma-field that makes Pr(𝐑)μμ(,x]\Pr(\mathbf{R})\ni\mu\mapsto\mu(-\infty,x] measurable for all x𝐑x\in\mathbf{R}.

For any random element μ\mu of Pr(𝐑)\Pr(\mathbf{R}), it is straightforward to show that 𝐑x𝐄μ(,x]\mathbf{R}\ni x\mapsto\operatorname{\mathbf{E}}\mu(-\infty,x] is a distribution function of some Borel probability measure on 𝐑\mathbf{R}. Let 𝐄μ\operatorname{\mathbf{E}}\mu denote that measure. Then 𝐄μ\operatorname{\mathbf{E}}\mu has the following property:

Lemma A.1 (change of order).

Let μ\mu be a random element of Pr(𝐑)\Pr(\mathbf{R}), and f:𝐑𝐑f\colon\mathbf{R}\to\mathbf{R} be (Borel) measurable.

  1. (1)

    If ff is nonnegative, then 𝐑f𝑑μ\int_{\mathbf{R}}f\,d\mu is measurable, and we have

    (A.1) 𝐑fd𝐄μ=𝐄[𝐑f𝑑μ].\int_{\mathbf{R}}f\,d\operatorname{\mathbf{E}}\mu=\operatorname{\mathbf{E}}\biggl{[}\int_{\mathbf{R}}f\,d\mu\biggr{]}.
  2. (2)

    If 𝐑|f|𝐄μ<\int_{\mathbf{R}}|f|\,\operatorname{\mathbf{E}}\mu<\infty, then 𝐑f𝑑μ\int_{\mathbf{R}}f\,d\mu is a.s. finite and measurable, and we have (A.1).

Proof.

Since (2) follows immediately from (1), we will prove (1) only. As the statement of (1) holds for f=1(,x]f=1_{(-\infty,x]} for all x𝐑x\in\mathbf{R}, Dynkin’s π\pi-λ\lambda theorem implies that the statement holds for all measurable A𝐑A\subset\mathbf{R}. By the simple function approximation argument, the statement extends to all nonnegative measurable ff. ∎

In order to talk about 𝐄μWn\operatorname{\mathbf{E}}\mu_{W_{n}}, we first need to establish the measurability of μWn\mu_{W_{n}} for each x𝐑x\in\mathbf{R}.

Lemma A.2 (measurability).

If WW is a random n×nn\times n Hermitian matrix, then μW\mu_{W} is measurable.

Proof.

Let f(x):=det(xIW)f(x):=\det(xI-W). Given an interval [a,b][a,b] where a<ba<b, the event of having an eigenvalue of WW with multiplicity at least kk (where knk\leq n) in [a,b][a,b] is equal to

{infq[a,b]𝐐((f(q))2+(f(q))2++(f(k1)(q))2)=0},\Bigl{\{}\inf_{q\in[a,b]\cap\mathbf{Q}}\bigl{(}(f(q))^{2}+(f^{\prime}(q))^{2}+\cdots+(f^{(k-1)}(q))^{2}\bigr{)}=0\Bigr{\}},

which is indeed measurable. Using this and by partitioning 𝐑\mathbf{R} into many small intervals, one can show that {λj(W)x}\{\lambda_{j}(W)\leq x\} is measurable for each j=1,,nj=1,\ldots,n and x𝐑x\in\mathbf{R}. Since

μW(,x]=1nj=1n𝟏{λj(W)x},\mu_{W}(-\infty,x]=\frac{1}{n}\sum_{j=1}^{n}\operatorname{\mathbf{1}}\{\lambda_{j}(W)\leq x\},

μW\mu_{W} is measurable. ∎

Now we turn to the proof of Lemma 1.3. We need the following inequality, which was found independently by Guntuboyina and Leeb [GL09], and Bordenave, Caputo, and Chafaï [BCC11].

Lemma A.3 (concentration for spectral measures).

Let (Wn)n𝐍(W_{n})_{n\in\mathbf{N}} be a Hermitian Wigner ensemble. If the total variation of f:𝐑𝐑f\colon\mathbf{R}\to\mathbf{R} is less than or equal to 11, then

𝐏(|𝐑f𝑑μWn𝐄𝐑f𝑑μWn|t)2exp(nt2/2).\operatorname{\mathbf{P}}\biggl{(}\Bigl{|}\int_{\mathbf{R}}f\,d\mu_{W_{n}}-\operatorname{\mathbf{E}}\int_{\mathbf{R}}f\,d\mu_{W_{n}}\Bigr{|}\geq t\biggr{)}\leq 2\exp(-nt^{2}/2).
Proof.

See [BCC11, Lemma C.2]. ∎

Proof of Lemma 1.3.

Assume 𝐄μWnμsc\operatorname{\mathbf{E}}\mu_{W_{n}}\Rightarrow\mu_{\mathrm{sc}}. For each p,q𝐐p,q\in\mathbf{Q} with p<qp<q, let fp,q:𝐑𝐑f_{p,q}\colon\mathbf{R}\to\mathbf{R} be 11 on (,p](-\infty,p], 0 on [q,)[q,\infty), and linear and continuous on [p,q][p,q]. Then

𝐄[𝐑fp,qdμWn]𝐑fp,qdμsc.for all rational p<q.\operatorname{\mathbf{E}}\biggl{[}\int_{\mathbf{R}}f_{p,q}\,d\mu_{W_{n}}\biggr{]}\to\int_{\mathbf{R}}f_{p,q}\,d\mu_{\mathrm{sc}}.\qquad\text{for all rational $p<q$.}

By Lemma A.3 and the Borel-Cantelli lemma,

𝐑fp,q𝑑μWn𝐑fp,q𝑑μscfor all rational p<q, a.s.\int_{\mathbf{R}}f_{p,q}\,d\mu_{W_{n}}\to\int_{\mathbf{R}}f_{p,q}\,d\mu_{\mathrm{sc}}\qquad\text{for all rational $p<q$, a.s.}

This proves μWnμsc\mu_{W_{n}}\Rightarrow\mu_{\mathrm{sc}} a.s.

To show the converse, assume that μWnμsc\mu_{W_{n}}\Rightarrow\mu_{\mathrm{sc}} a.s., and let f:𝐑𝐑f\colon\mathbf{R}\to\mathbf{R} be continuous and bounded. Since ff is bounded, we can apply the dominated convergence theorem to obtain

𝐄[𝐑f𝑑μWn]𝐑f𝑑μsc.\operatorname{\mathbf{E}}\biggl{[}\int_{\mathbf{R}}f\,d\mu_{W_{n}}\biggr{]}\to\int_{\mathbf{R}}f\,d\mu_{\mathrm{sc}}.

This shows 𝐄μWnμsc\operatorname{\mathbf{E}}\mu_{W_{n}}\Rightarrow\mu_{\mathrm{sc}}. ∎

Appendix B Reductions

In this section, we prove that Theorem 1.6 follows from Theorem 1.8 (Lemma B.1), and that Theorem 1.8 follows from Theorem 2.1 (Lemma B.4).

Lemma B.1.

Theorem 1.6 follows from Theorem 1.8.

To prove this, we use the following two lemmas.

Lemma B.2 (perturbation inequality).

If AA and BB are n×nn\times n Hermitian matrices, and FAF_{A} and FBF_{B} are the distribution functions of μA\mu_{A} and μB\mu_{B}, then

(L(FA,FB))3tr((AB)2)\bigl{(}L(F_{A},F_{B})\bigr{)}^{3}\leq\operatorname{tr}\bigl{(}(A-B)^{2}\bigr{)}

where LL is the Lévy metric.

Proof.

See [BS10, Theorem A.41]. ∎

Lemma B.3.

If (1.2) holds, then the following are true:

  1. (1)

    (1.5) holds if and only if (1.8) holds.

  2. (2)

    (1.3) holds if and only if (1.9) holds.

Proof.

From (1.2) it follows that

1ni,j=1n𝐄[|wij|2;|wij|>1]0.\frac{1}{n}\sum_{i,j=1}^{n}\operatorname{\mathbf{E}}[|w_{ij}|^{2};|w_{ij}|>1]\to 0.

(1) For any Jn{1,,n}J_{n}\subset\{1,\ldots,n\} with |Jn|/n0|J_{n}|/n\to 0, we have

(B.1) |1niJnj=1n𝐄|wij|21niJnj=1n𝐄[|wij|2;|wij|1]|1ni,j=1n𝐄[|wij|2;|wij|>1]0.\biggl{|}\frac{1}{n}\sum_{i\in J_{n}}\sum_{j=1}^{n}\operatorname{\mathbf{E}}|w_{ij}|^{2}-\frac{1}{n}\sum_{i\in J_{n}}\sum_{j=1}^{n}\operatorname{\mathbf{E}}[|w_{ij}|^{2};|w_{ij}|\leq 1]\biggr{|}\\ \leq\frac{1}{n}\sum_{i,j=1}^{n}\operatorname{\mathbf{E}}[|w_{ij}|^{2};|w_{ij}|>1]\to 0.

Thus, (1.5) holds if and only if (1.8) holds.

(2) Since

(B.2) |1ni=1n|j=1n𝐄|wij|21|1ni=1n|j=1n𝐄[|wij|2;|wij|1]1||1ni,j=1n𝐄[|wij|2;|wij|>1]0,\biggl{|}\frac{1}{n}\sum_{i=1}^{n}\Bigl{|}\sum_{j=1}^{n}\operatorname{\mathbf{E}}|w_{ij}|^{2}-1\Bigr{|}-\frac{1}{n}\sum_{i=1}^{n}\Bigl{|}\sum_{j=1}^{n}\operatorname{\mathbf{E}}[|w_{ij}|^{2};|w_{ij}|\leq 1]-1\Bigr{|}\biggr{|}\\ \leq\frac{1}{n}\sum_{i,j=1}^{n}\operatorname{\mathbf{E}}[|w_{ij}|^{2};|w_{ij}|>1]\to 0,

(1.3) holds if and only if (1.9) holds. ∎

Proof of Lemma B.1.

We first show the sufficiency direction of Theorem 1.6. Recall that we proved that (1.6) and (1.7) follows from (1.2) and 𝐄wij=0\operatorname{\mathbf{E}}w_{ij}=0. By Lemma B.3, we have (1.8) and (1.9). Thus, 𝐄μWnμsc\operatorname{\mathbf{E}}\mu_{W_{n}}\Rightarrow\mu_{\mathrm{sc}} follows from Theorem 1.8.

Now let us show the necessity. On top of (1.6) and (1.7), we have (1.8) by Lemma B.3. Thus, (1.9) follows from Theorem 1.8, and therefore we have (1.3) by Lemma B.3. ∎

Next we reduce Theorem 1.8 to Theorem 2.1.

Lemma B.4.
  1. (1)

    The sufficiency part of Theorem 2.1 implies the sufficiency part of Theorem 1.8.

  2. (2)

    The necessity part of Theorem 2.1 implies the necessity part of Theorem 1.8.

We need the following lemma.

Lemma B.5 (reduction to vanishing bounds).

Let (Wn)n𝐍(W_{n})_{n\in\mathbf{N}} be a Hermitian Wigner ensemble satisfying (1.6) and (1.7). Then there exist 1/2η1η21/2\geq\eta_{1}\geq\eta_{2}\geq\cdots with ηn0\eta_{n}\to 0 such that if we let

Wn(wij)i,j=1n:=(wij𝟏{|wij|ηn}𝐄[wij;|wij|ηn])i,j=1n,W_{n}^{\prime}\equiv(w_{ij}^{\prime})_{i,j=1}^{n}:=\bigl{(}w_{ij}\operatorname{\mathbf{1}}\{|w_{ij}|\leq\eta_{n}\}-\operatorname{\mathbf{E}}[w_{ij};|w_{ij}|\leq\eta_{n}]\bigr{)}_{i,j=1}^{n},

then the following are true:

  1. (1)

    𝐄μWnμsc\operatorname{\mathbf{E}}\mu_{W_{n}}\Rightarrow\mu_{\mathrm{sc}} if and only if 𝐄μWnμsc\operatorname{\mathbf{E}}\mu_{W_{n}^{\prime}}\Rightarrow\mu_{\mathrm{sc}}.

  2. (2)

    (1.9) if and only if

    (B.3) 1ni=1n|j=1n𝐄|wij|21|0.\frac{1}{n}\sum_{i=1}^{n}\Bigl{|}\sum_{j=1}^{n}\operatorname{\mathbf{E}}|w_{ij}^{\prime}|^{2}-1\Bigr{|}\to 0.
  3. (3)

    (1.8) if and only if

    (B.4) 1niJnj=1n𝐄|wij|20for any Jn{1,,n} with |Jn|/n0.\frac{1}{n}\sum_{i\in J_{n}}\sum_{j=1}^{n}\operatorname{\mathbf{E}}|w_{ij}^{\prime}|^{2}\to 0\quad\text{for any $J_{n}\subset\{1,\ldots,n\}$ with $|J_{n}|/n\to 0$.}

The proof of (1) will be based on the following lemma.

Lemma B.6 (Bernstein’s inequality).

Suppose that X1,,XnX_{1},\ldots,X_{n} are independent real-valued random variables with |Xi|1|X_{i}|\leq 1 and 𝐄Xi=0\operatorname{\mathbf{E}}X_{i}=0 for i=1,,ni=1,\ldots,n. If S:=X1++XnS:=X_{1}+\cdots+X_{n}, then

𝐏(Sx)exp(x22(𝐄S2+x))for all x>0.\operatorname{\mathbf{P}}(S\geq x)\leq\exp\biggl{(}\frac{-x^{2}}{2(\operatorname{\mathbf{E}}S^{2}+x)}\biggr{)}\qquad\text{for all $x>0$.}
Proof.

The proof of [Bil99, M20] with a slight modification works. ∎

Proof of Lemma B.5.

Choose 1/2η1η21/2\geq\eta_{1}\geq\eta_{2}\cdots with ηn0\eta_{n}\to 0 such that

(B.5) 1ni,j=1n𝐏(|wij|>ηn)0.\frac{1}{n}\sum_{i,j=1}^{n}\operatorname{\mathbf{P}}(|w_{ij}|>\eta_{n})\to 0.

Proof of (1). By Lemma 1.3, it is enough to show that μWnμsc\mu_{W_{n}}\Rightarrow\mu_{\mathrm{sc}} a.s. if and only if μWnμsc\mu_{W_{n}^{\prime}}\Rightarrow\mu_{\mathrm{sc}} a.s. Let W~:=(wij𝟏{|wij|ηn})i,j=1n\widetilde{W}:=(w_{ij}\operatorname{\mathbf{1}}\{|w_{ij}|\leq\eta_{n}\})_{i,j=1}^{n}. We will first show that μWnμsc\mu_{W_{n}}\Rightarrow\mu_{\mathrm{sc}} a.s. if and only if μW~nμsc\mu_{\widetilde{W}_{n}}\Rightarrow\mu_{\mathrm{sc}} a.s. Note that

rank(WnW~n)n2n1ijn𝟏{|wij|>ηn}.\frac{\operatorname{rank}(W_{n}-\widetilde{W}_{n})}{n}\leq\frac{2}{n}\sum_{1\leq i\leq j\leq n}\operatorname{\mathbf{1}}\{|w_{ij}|>\eta_{n}\}.

By Lemma 1.5, it is enough to show that the right side tends to 0 a.s.

Let ϵ>0\epsilon>0 be given. By (B.5), we have some n0𝐍n_{0}\in\mathbf{N} such that

1ijn𝐏(|wij|>ηn)ϵn/2for all nn0.\sum_{1\leq i\leq j\leq n}\operatorname{\mathbf{P}}(|w_{ij}|>\eta_{n})\leq\epsilon n/2\qquad\text{for all $n\geq n_{0}$.}

Since 𝟏{|wij|>ηn}\operatorname{\mathbf{1}}\{|w_{ij}|>\eta_{n}\}, 1ijn1\leq i\leq j\leq n, are independent, Bernstein’s inequality (Lemma B.6) implies

𝐏(1ijn𝟏{|wij|>ηn}ϵn)𝐏(1ijn(𝟏{|wij|>ηn}𝐏(|wij|>ηn))ϵn/2)exp(ϵ2n2/81ijn𝐏(|wij|>ηn)+ϵn/2)exp(ϵ2n2/8ϵn)=exp(ϵn/8).\begin{split}\operatorname{\mathbf{P}}\biggl{(}\sum_{1\leq i\leq j\leq n}&\operatorname{\mathbf{1}}\{|w_{ij}|>\eta_{n}\}\geq\epsilon n\biggr{)}\\ &\leq\operatorname{\mathbf{P}}\biggl{(}\sum_{1\leq i\leq j\leq n}\bigl{(}\operatorname{\mathbf{1}}\{|w_{ij}|>\eta_{n}\}-\operatorname{\mathbf{P}}(|w_{ij}|>\eta_{n})\bigr{)}\geq\epsilon n/2\biggr{)}\\ &\leq\exp\biggl{(}\frac{-\epsilon^{2}n^{2}/8}{\sum_{1\leq i\leq j\leq n}\operatorname{\mathbf{P}}(|w_{ij}|>\eta_{n})+\epsilon n/2}\biggr{)}\\ &\leq\exp\biggl{(}\frac{-\epsilon^{2}n^{2}/8}{\epsilon n}\biggr{)}=\exp(-\epsilon n/8).\end{split}

As

n=1exp(ϵn/8)<for all ϵ>0,\sum_{n=1}^{\infty}\exp(-\epsilon n/8)<\infty\qquad\text{for all $\epsilon>0$,}

the Borel-Cantelli lemma implies

1n1ijn𝟏{|wij|>ηn}0a.s.\frac{1}{n}\sum_{1\leq i\leq j\leq n}\operatorname{\mathbf{1}}\{|w_{ij}|>\eta_{n}\}\to 0\qquad\text{a.s.}

This implies that μWnμsc\mu_{W_{n}}\Rightarrow\mu_{\mathrm{sc}} a.s. if and only if μW~nμsc\mu_{\widetilde{W}_{n}}\Rightarrow\mu_{\mathrm{sc}} a.s. as explained above.

To show that μW~nμsc\mu_{\widetilde{W}_{n}}\Rightarrow\mu_{\mathrm{sc}} a.s. if and only if μWnμsc\mu_{W_{n}^{\prime}}\Rightarrow\mu_{\mathrm{sc}} a.s., use Lemma B.2 to note that

(L(μW~n,μWn))31ni,j=1n(𝐄[wij;|wij|ηn])2.\bigl{(}L(\mu_{\widetilde{W}_{n}},\mu_{W_{n}^{\prime}})\bigr{)}^{3}\leq\frac{1}{n}\sum_{i,j=1}^{n}\bigl{(}\operatorname{\mathbf{E}}[w_{ij};|w_{ij}|\leq\eta_{n}]\bigr{)}^{2}.

Since |a2b2|=|ab||a+b||a^{2}-b^{2}|=|a-b||a+b|, the difference between the right side and

1ni,j=1n(𝐄[wij;|wij|1])2\frac{1}{n}\sum_{i,j=1}^{n}\bigl{(}\operatorname{\mathbf{E}}[w_{ij};|w_{ij}|\leq 1]\bigr{)}^{2}

is bounded above by

1ni,j=1n|𝐄[wij;ηn<|wij|1]||𝐄[wij;|wij|1]+𝐄[wij;|wij|ηn]|2ni,j=1n𝐏(|wij|>ηn)0.\begin{split}\frac{1}{n}\sum_{i,j=1}^{n}\bigl{|}\operatorname{\mathbf{E}}[w_{ij};\eta_{n}<|w_{ij}|\leq 1]\bigr{|}&\cdot\bigl{|}\operatorname{\mathbf{E}}[w_{ij};|w_{ij}|\leq 1]+\operatorname{\mathbf{E}}[w_{ij};|w_{ij}|\leq\eta_{n}]\bigr{|}\\ &\leq\frac{2}{n}\sum_{i,j=1}^{n}\operatorname{\mathbf{P}}(|w_{ij}|>\eta_{n})\to 0.\end{split}

Thus, by (1.6), we have L(μW~n,μWn)0L(\mu_{\widetilde{W}_{n}},\mu_{W_{n}^{\prime}})\to 0 a.s.

Proof of (2). Since ||a||b|||ab|\bigl{|}|a|-|b|\bigr{|}\leq|a-b|, we have

|1ni=1n|j=1n𝐄[|wij|2;|wij|1]1|1ni=1n|j=1n𝐄|wij|21||1ni,j=1n|𝐄[|wij|2;|wij|1]𝐄|wij|2|=1ni,j=1n𝐄[|wij|2;ηn<|wij|1]+1ni,j=1n(𝐄[wij;|wij|ηn])2.\begin{split}\biggl{|}\frac{1}{n}\sum_{i=1}^{n}&\Bigl{|}\sum_{j=1}^{n}\operatorname{\mathbf{E}}[|w_{ij}|^{2};|w_{ij}|\leq 1]-1\Bigr{|}-\frac{1}{n}\sum_{i=1}^{n}\Bigl{|}\sum_{j=1}^{n}\operatorname{\mathbf{E}}|w_{ij}^{\prime}|^{2}-1\Bigr{|}\biggr{|}\\ &\leq\frac{1}{n}\sum_{i,j=1}^{n}\Bigl{|}\operatorname{\mathbf{E}}[|w_{ij}|^{2};|w_{ij}|\leq 1]-\operatorname{\mathbf{E}}|w^{\prime}_{ij}|^{2}\Bigr{|}\\ &=\frac{1}{n}\sum_{i,j=1}^{n}\operatorname{\mathbf{E}}[|w_{ij}|^{2};\eta_{n}<|w_{ij}|\leq 1]+\frac{1}{n}\sum_{i,j=1}^{n}\bigl{(}\operatorname{\mathbf{E}}[w_{ij};|w_{ij}|\leq\eta_{n}]\bigr{)}^{2}.\end{split}

The first term on the right side is bounded above by

1ni,j=1n𝐏(|wij|>ηn)0,\frac{1}{n}\sum_{i,j=1}^{n}\operatorname{\mathbf{P}}(|w_{ij}|>\eta_{n})\to 0,

and we have shown just above that the second term also tends to 0.

Proof of (3). The difference between the left sides of (1.8) and (B.4) is also bounded above by

1ni,j=1n𝐄[|wij|2;ηn<|wij|1]+1ni,j=1n(𝐄[wij;|wij|ηn])2,\frac{1}{n}\sum_{i,j=1}^{n}\operatorname{\mathbf{E}}[|w_{ij}|^{2};\eta_{n}<|w_{ij}|\leq 1]+\frac{1}{n}\sum_{i,j=1}^{n}\bigl{(}\operatorname{\mathbf{E}}[w_{ij};|w_{ij}|\leq\eta_{n}]\bigr{)}^{2},

which tends to 0 as nn\to\infty. ∎

Proof of Lemma B.4.

Assume that WnW_{n} is given as in Theorem 1.8, and define WnW^{\prime}_{n} as in Lemma B.5. If we let ϵn:=2ηn\epsilon_{n}:=2\eta_{n}, then (Wn)n𝐍(W_{n}^{\prime})_{n\in\mathbf{N}} satisfies the conditions of Theorem 2.1. In particular, (1.5) for WnW_{n}^{\prime} follows from (3) of Lemma B.5.

Proof of (1). Assume that the sufficiency part of Theorem 2.1 holds. If (1.9) holds, then (B.3) holds by (2) of Lemma B.5, and thus 𝐄μWnμsc\operatorname{\mathbf{E}}\mu_{W_{n}^{\prime}}\Rightarrow\mu_{\mathrm{sc}} by the sufficiency part of Theorem 2.1. Then (1) of Lemma B.5 tells us that 𝐄μWnμsc\operatorname{\mathbf{E}}\mu_{W_{n}}\Rightarrow\mu_{\mathrm{sc}}.

Proof of (2). Assume that the necessity part of Theorem 2.1 holds. If 𝐄μWnμsc\operatorname{\mathbf{E}}\mu_{W_{n}}\Rightarrow\mu_{\mathrm{sc}}, then 𝐄μWnμsc\operatorname{\mathbf{E}}\mu_{W_{n}^{\prime}}\Rightarrow\mu_{\mathrm{sc}} by (1) of Lemma B.5, and thus (B.3) holds by Theorem 2.1. This in turn implies (1.9) by (2) of Lemma B.5. ∎

References

  • [AGZ10] Greg W. Anderson, Alice Guionnet, and Ofer Zeitouni. An introduction to random matrices, volume 118 of Cambridge Studies in Advanced Mathematics. Cambridge University Press, Cambridge, 2010.
  • [Arn71] Ludwig Arnold. On wigner’s semicircle law for the eigenvalues of random matrices. Zeitschrift für Wahrscheinlichkeitstheorie und verwandte Gebiete, 19(3):191–198, 1971.
  • [BCC11] Charles Bordenave, Pietro Caputo, and Djalil Chafaï. Spectrum of non-Hermitian heavy tailed random matrices. Comm. Math. Phys., 307(2):513–560, 2011.
  • [Bil99] Patrick Billingsley. Convergence of probability measures. Wiley Series in Probability and Statistics: Probability and Statistics. John Wiley & Sons, Inc., New York, second edition, 1999. A Wiley-Interscience Publication.
  • [Bil12] Patrick Billingsley. Probability and measure. Wiley Series in Probability and Statistics. John Wiley & Sons, Inc., Hoboken, NJ, 2012. Anniversary edition [of MR1324786], With a foreword by Steve Lalley and a brief biography of Billingsley by Steve Koppes.
  • [BS10] Zhidong Bai and Jack W. Silverstein. Spectral analysis of large dimensional random matrices. Springer Series in Statistics. Springer, New York, second edition, 2010.
  • [Gir90] V. L. Girko. Theory of random determinants, volume 45 of Mathematics and its Applications (Soviet Series). Kluwer Academic Publishers Group, Dordrecht, 1990. Translated from the Russian.
  • [GL09] Adityanand Guntuboyina and Hannes Leeb. Concentration of the spectral measure of large Wishart matrices with dependent entries. Electron. Commun. Probab., 14:334–342, 2009.
  • [GNT15] F. Götze, A. A. Naumov, and A. N. Tikhomirov. Limit theorems for two classes of random matrices with dependent entries. Theory Probab. Appl., 59(1):23–39, 2015.
  • [Jun18] Paul Jung. Lévy-Khintchine random matrices and the Poisson weighted infinite skeleton tree. Trans. Amer. Math. Soc., 370(1):641–668, 2018.
  • [Kal02] Olav Kallenberg. Foundations of modern probability. Probability and its Applications (New York). Springer-Verlag, New York, second edition, 2002.
  • [Pas73] Leonid A Pastur. Spectra of random self adjoint operators. Russian mathematical surveys, 28(1):1, 1973.
  • [Tao12] Terence Tao. Topics in random matrix theory, volume 132 of Graduate Studies in Mathematics. American Mathematical Society, Providence, RI, 2012.
  • [vLW01] J. H. van Lint and R. M. Wilson. A course in combinatorics. Cambridge University Press, Cambridge, second edition, 2001.
  • [Wig55] Eugene P. Wigner. Characteristic vectors of bordered matrices with infinite dimensions. Ann. of Math. (2), 62:548–564, 1955.
  • [Wig58] Eugene P Wigner. On the distribution of the roots of certain symmetric matrices. Annals of Mathematics, pages 325–327, 1958.