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

A Berry-Esseen type theorem for finite free convolution

Octavio Arizmendi Centro de Investigación en Matemáticas, A.C., Jalisco S/N, Col. Valenciana CP: 36023 Guanajuato, Gto, México  and  Daniel Perales
Abstract.

We prove that the rate of convergence for the central limit theorem in finite free convolution is of order n1/2n^{-1/2}.

[Uncaptioned image] This project has received funding from the European Union’s Horizon 2020 research and innovation programme under the Marie Skłodowska-Curie grant agreement No 734922.

1. Introduction

In recent years, the convolution of polynomials, first studied by Walsh [10], was revisited by Marcus, Spielman and Srivastava [8], in order to exhibit bounds for the eigenvalues of expected characteristic polynomials of certain dd-regular graphs, in their aim to construct bipartite Ramanujan graphs of all sizes [9]. The authors refer to this convolution as finite free additive convolution because of its relation to free convolution, see [1, 7, 9].

In [7], Marcus showed that the Central Limit Theorem for this convolution is given by the polynomial D1/d(Hd),D_{1/\sqrt{d}}(H_{d}), where HdH_{d} is an Hermite polynomial and may be written as

Hd(x)=d!i=0d2(1)ii!(d2i)!xd2i2i,H_{d}(x)=d!\sum_{i=0}^{\lfloor{\tfrac{d}{2}}\rfloor}{\frac{(-1)^{i}}{i!(d-2i)!}}{\frac{x^{d-2i}}{2^{i}}},

and in general for any λ>0\lambda>0 and polynomial pp of degree dd, Dλp(x):=λdp(x/λ)D_{\lambda}p(x):=\lambda^{d}p(x/\lambda) is the dilation by λ\lambda of the polynomial pp.

In this note we are interested in the rate of convergence in the Central Limit Theorem for finite free convolution. Recall that for the central limit theorem in probability, the Berry-Esseen Theorem [2, 4] states that if μ\mu is a probability measure with m1(μ)=0m_{1}(\mu)=0, m2(μ)=1m_{2}(\mu)=1, and |x|3𝑑μ<\int_{\mathbb{R}}|x|^{3}d\mu<\infty, then the distance to the standard Gaussian distribution 𝒩\mathcal{N} is bounded as follows

dkol(D1nμn,𝒩)C|x|3𝑑μn,d_{kol}(D_{\frac{1}{\sqrt{n}}}\mu^{\ast n},\mathcal{N})\leq C\frac{\int_{\mathbb{R}}|x|^{3}d\mu}{\sqrt{n}},

where dkold_{kol} denotes the Kolmogorov distance between measures, DbμD_{b}\mu denotes the dilation of a measure μ\mu by a factor b>0b>0, \ast denotes the classical convolution, and CC is an absolute constant. Our main reasult shows that as for the classical and free case [3, 5], in finite free probability we also achieve a rate of order n1/2n^{-1/2}. However, we use the Lévy distance (see Section 2.3 for more details) instead of Kolmogorov distance, the reason being that we are dealing with measures supported in dd atoms with size 1/d1/d and thus we cannot expect better.

Thus, for two polynomials of degree dd, pp and qq, let us define the distance between them to be L(p,q):=dL(μp,μq),L(p,q):=d_{L}(\mu_{p},\mu_{q}), where dLd_{L} is the Lévy distance and the measures μp\mu_{p} and μq\mu_{q} are the distributions of pp and qq, respectively.

In this language we can state our contribution as follows.

Theorem 1.1.

Let dd\in\mathbb{N} and let pp be a real polynomial of degree dd such that the first two moments of μp\mu_{p} are m1=0m_{1}=0 and m2=1m_{2}=1. Then, there exists an absolute constant CdC_{d}, only depending on dd, such that for all n>0n>0,

L(D1/n(pdn),D1/d(Hd))<Cdn,L\left(D_{1/\sqrt{n}}(p^{\boxplus_{d}n}),D_{1/\sqrt{d}}(H_{d})\right)<\frac{C_{d}}{\sqrt{n}},

The main tool to prove the above rate of convergence are the cumulants for finite free convolution, as we defined in [1]. These cumulants give a combinatorial approach to investigate this convolution and its relation to free probability. In particular we showed that finite free cumulants approach free cumulants, providing a combinatorial perspective to the fact that finite free convolution approaches free convolution in the limit. Using these cumulants we were able to show that some properties of free convolution are valid already in the finite free case. The above theorem is another instance of the fact that many properties in free probability already appear in the finite level.

Apart from this introduction this note consists of two sections. Section 2 gives the preliminaries for the theory of finite free probability and in Section 3 we give the proof of the main theorem, Theorem 1.1.

2. Preliminaries

We give very basic preliminaries on finite free convolution we refer to [1, 7] for details.

2.1. Finite Free Convolution

For two polynomials, p(x)=i=0dxdi(1)iaipp(x)=\sum_{i=0}^{d}x^{d-i}(-1)^{i}a^{p}_{i} and q(x)=i=0dxdi(1)iaiqq(x)=\sum_{i=0}^{d}x^{d-i}(-1)^{i}a^{q}_{i}, the finite free additive convolution of pp and qq is given by

p(x)dq(x)=k=0dxdr(1)ri+j=r(di)!(dj)!d!(dij)!aipajq.p(x)\boxplus_{d}q(x)=\sum_{k=0}^{d}x^{d-r}(-1)^{r}\sum_{i+j=r}\frac{(d-i)!(d-j)!}{d!(d-i-j)!}a^{p}_{i}a^{q}_{j}.

The finite RR-transform of a polynomial is defined by

(2.1) pd(s)1dsln(i=0d(d)iaip(d)isi)mod [sd],\mathcal{R}^{d}_{p}(s)\equiv-\frac{1}{d}\frac{\partial}{\partial s}\ln\left(\sum_{i=0}^{d}\frac{(-d)^{i}a^{p}_{i}}{(d)_{i}}s^{i}\right)\qquad\text{mod }[s^{d}],

when pp is the monic polynomial p(x)=i=0dxdi(1)iaipp(x)=\sum_{i=0}^{d}x^{d-i}(-1)^{i}a^{p}_{i}.

We consider the truncated RR-transform given by the sum of the first dd terms in the series expansion of pd\mathcal{R}^{d}_{p}, which will have the cumulants as coefficients.

Definition 2.1 ([1]).

Let pp be a monic polynomial of degree dd, and suppose the pd(s)\mathcal{R}^{d}_{p}(s) satisfies

pd(s)j=0d1κj+1(p)sj mod [sd].\mathcal{R}^{d}_{p}(s)\equiv\sum_{j=0}^{d-1}\kappa_{j+1}(p)s^{j}\quad\text{ mod }[s^{d}].
  1. (1)

    We call the sum of the first dd terms in the series expansion of d\mathcal{R}^{d} the truncated RR-transform and denote by ~pd(s)\mathcal{\tilde{R}}^{d}_{p}(s), i.e.

    ~pd(s):=j=0d1κj+1(p)sj.\mathcal{\tilde{R}}^{d}_{p}(s):=\sum_{j=0}^{d-1}\kappa_{j+1}(p)s^{j}.
  2. (2)

    The numbers κ1(p),κ2(p),,κd(p)\kappa_{1}(p),\kappa_{2}(p),\dots,\kappa_{d}(p) will be called the finite free cumulants. To simplify notation we will omit the dependence on pp when we deal with only one polynomial.

We want to use the combinatorial framework in terms of moments for these cumulants. Hence, for a polynomial pp with roots λ1,.,λn\lambda_{1},....,\lambda_{n} we define the moments of pp, by the formula mn(p)=1di=1dλinm_{n}(p)=\frac{1}{d}\sum^{d}_{i=1}\lambda_{i}^{n}.

These finite free cumulants satisfy the following properties which are the analog of the properties in the axiomatization of cumulants by Lehner [6], in non-commutative probability.

  1. (1)

    Polynomial in the first nn moments: kn(p)k_{n}(p) is a polynomial in the first nn moments of pp with leading term

    dn(d)nmn(p).\frac{d^{n}}{(d)_{n}}m_{n}(p).
  2. (2)

    Homogeneity: for all monic polynomials pp and λ0\lambda\neq 0 we have

    κn(Dλ(p))=λnκn(p).\kappa_{n}(D_{\lambda}(p))=\lambda^{n}\kappa_{n}(p).
  3. (3)

    Additivity: for all monic polynomials pp and qq, we have

    κn(pdq)=κn(p)+κn(q).\kappa_{n}(p\boxplus_{d}q)=\kappa_{n}(p)+\kappa_{n}(q).

2.2. Moment-cumulant formula

Moment-cumulant formulas involve summing over partitions on the set [n].[n]. Let us introduce this definition and some notation.

Definition 2.2.

We call π={V1,,Vr}\pi=\{V_{1},...,V_{r}\} a partition of the set [n]:={1,2,,n}[n]:=\{1,2,\dots,n\} if ViV_{i} (1ir)(1\leq i\leq r) are pairwise disjoint, non-void subsets of [n][n], such that V1V2Vr={1,2,,n}V_{1}\cup V_{2}...\cup V_{r}=\{1,2,\dots,n\}. We call V1,V2,,VrV_{1},V_{2},\dots,V_{r} the blocks of π\pi. The number of blocks of π\pi is denoted by |π|\left|\pi\right|. We will denote the set of partitions of [n][n] by 𝒫(n)\mathcal{P}(n).

The set 𝒫(n)\mathcal{P}(n) can be equipped with the partial order \leq of reverse refinement (πσ\pi\leq\sigma if and only if every block of π\pi is completely contained in a block of σ\sigma). With this order the minimum is given by the partition with nn blocks, 0n={{1},{2},,{n}}0_{n}=\{\{1\},\{2\},\cdots,\{n\}\}, and the maximum is given by the partition with 11 block, 1n={{1,2,,n}}1_{n}=\{\{1,2,\cdots,n\}\}.

Thus one can consider the incidence algebra of 𝒫(n)\mathcal{P}(n). For two partitions σ,ρ\sigma,\rho in the set of partitions 𝒫(n)\mathcal{P}(n) the Möbius function is given by

μ(σ,ρ)=(1)|σ||ρ|(2!)r3(3!)r4((n1)!)rn,\mu(\sigma,\rho)=(-1)^{{|\sigma|-|\rho|}}(2!)^{{r_{3}}}(3!)^{{r_{4}}}\cdots((n-1)!)^{{r_{n}}},

where rir_{i} is the number of blocks of ρ\rho that contain exactly ii blocks of σ\sigma. In particular, for σ=0n\sigma=0_{n} we have

μ(0n,ρ)=(1)n|ρ|(2!)t3(3!)t4((n1)!)tn,\mu(0_{n},\rho)=(-1)^{n-{|\rho|}}(2!)^{{t_{3}}}(3!)^{{t_{4}}}\cdots((n-1)!)^{{t_{n}}},

where tit_{i} is the number of blocks of ρ\rho of size ii.

Given a sequence of complex numbers f={fn}nf=\{f_{n}\}_{n\in\mathbb{N}} we may extend ff to partitions in a multiplicative way by the formula

fπ=f|V1|f|V2|f|Vn|,f_{\pi}=f_{|V_{1}|}f_{|V_{2}|}\cdots f_{|V_{n}|},

where V1,,VnV_{1},\dots,V_{n} are the blocks of π.\pi. In this note we will frequently use the multiplicative extensions of the Pochhammer sequence (d)n=(d)(d1)(dn+1)(d)_{n}=(d)(d-1)\cdots(d-n+1) and the factorial sequence n!n!, whose extensions will be denoted by (d)π(d)_{\pi} and N!πN!_{\pi}, respectively.

In [1], we gave formulas that relate the moments and coefficients of a polynomial with its finite free cumulants. First, we have a formula that writes coefficients in terms of cumulants.

Proposition 2.3 (Coefficient-cumulant formula).

Let p(x)=i=0dxdi(1)iaip(x)=\sum_{i=0}^{d}x^{d-i}(-1)^{i}a_{i} be a polynomial of degree dd and let (κn)n=1d(\kappa_{n})^{d}_{n=1} be its finite free cumulants. The following formulas hold.

(2.2) an=(d)ndnn!π𝒫(n)d|π|μ(0n,π)κπ,n.a_{n}=\frac{(d)_{n}}{d^{n}n!}\sum_{\pi\in\mathcal{P}(n)}d^{|\pi|}\mu(0_{n},\pi)\kappa_{\pi},\qquad n\in\mathbb{N}.

We also have a moment-cumulant formula for finite free cumulants:

Proposition 2.4.

Let pp be a monic polynomial of degree dd and let (mn)n=1(m_{n})^{\infty}_{n=1} and (κn)n=1d(\kappa_{n})^{d}_{n=1}, be the moments and cumulants of pp, respectively. Then

κn\displaystyle\kappa_{n} =(d)n1(n1)!σ𝒫(n)d|σ|μ(0,σ)mσπσμ(π,1n)(d)π,\displaystyle=\frac{(-d)^{n-1}}{(n-1)!}\sum_{\sigma\in\mathcal{P}(n)}d^{|\sigma|}\mu(0,\sigma)m_{\sigma}\sum_{\pi\geq\sigma}\frac{\mu(\pi,1_{n})}{(d)_{\pi}},

for n=1,,dn=1,\ldots,d and

mn\displaystyle m_{n} =(1)ndn+1(n1)!σ𝒫(n)d|σ|μ(0,σ)κσπσμ(π,1n)(d)π,\displaystyle=\frac{(-1)^{n}}{d^{n+1}(n-1)!}\sum_{\sigma\in\mathcal{P}(n)}d^{|\sigma|}\mu(0,\sigma)\kappa_{\sigma}\sum_{\pi\geq\sigma}-\mu(\pi,1_{n})(d)_{\pi},

for nn\in\mathbb{N}.

Remark 2.5.

The explicit moment-cumulant formulas of the first three finite cumulants are

κ1\displaystyle\kappa_{1} =\displaystyle= m1,κ2=dd1(m2m12),\displaystyle m_{1},\qquad\kappa_{2}=\frac{d}{d-1}(m_{2}-m_{1}^{2}),
κ3\displaystyle\kappa_{3} =\displaystyle= d2(d1)(d2)(2m133m1m2+m3),\displaystyle\frac{d^{2}}{(d-1)(d-2)}(2m_{1}^{3}-3m_{1}m_{2}+m_{3}),

and the explicit moment-cumulant formulas of the first three finite moments are

m1\displaystyle m_{1} =\displaystyle= κ1,m2=d1dκ2+κ12,\displaystyle\kappa_{1},\qquad m_{2}=\frac{d-1}{d}\kappa_{2}+\kappa_{1}^{2},
m3\displaystyle m_{3} =\displaystyle= (d1)(d2)d2κ3+3(d1)dκ2κ1+κ13.\displaystyle\frac{(d-1)(d-2)}{d^{2}}\kappa_{3}+\frac{3(d-1)}{d}\kappa_{2}\kappa_{1}+\kappa_{1}^{3}.

2.3. Convergence of polynomials and Lévy distance

In this setting of [1, 7] convergence of polynomials is pointwise convergence of the coefficients. We prefer to consider the weak convergence of the induced measures since it is common with the free probability setting. Thus, for a polynomial pp, with roots λ1,λ2,,λn\lambda_{1},\lambda_{2},\dots,\lambda_{n}, we define its distribution μp\mu_{p} as the uniform measure on the roots of pp, μp=1diδλi\mu_{p}=\tfrac{1}{d}\sum_{i}\delta_{\lambda_{i}}.

To quantify this convergence we use the Lévy distance

dL(μ,ν):=inf{ϵ>0|F(xϵ)ϵG(x)F(x+ϵ)+ϵfor all x},d_{L}(\mu,\nu):=\inf\{\epsilon>0\>|\>F(x-\epsilon)-\epsilon\leq G(x)\leq F(x+\epsilon)+\epsilon\>\>\>\text{for all }x\in\mathbb{R}\},

where FF and GG are the cumulative distribution functions of μ\mu and ν\nu respectively.

3. Proof of Theorem 1.1

Before going in to the proof of the main theorem we prove a couple of lemmas about the support and cumulants of polynomials with mean 0 and variance 11.

Lemma 3.1.

Let pp be a real polynomial of degree dd with κ1=0\kappa_{1}=0 and κ2=1\kappa_{2}=1. Then the support of pp is contained in (d1,d1)(-\sqrt{d-1},\sqrt{d-1}).

Proof.

If κ1=0\kappa_{1}=0 and κ2=1\kappa_{2}=1 then

1=κ2=dd1m2=1d1i=1dλi2.1=\kappa_{2}=\frac{d}{d-1}m_{2}=\frac{1}{d-1}\sum_{i=1}^{d}\lambda_{i}^{2}.

This means that λi2<d1\lambda_{i}^{2}<d-1 (strict because there is at least another non-zero λ\lambda) and thus |λi|<d1|\lambda_{i}|<\sqrt{d-1} for all i=1,,di=1,\dots,d. ∎

Lemma 3.2.

Let pp be a real polynomial of degree dd with κ1=0\kappa_{1}=0 and κ2=1\kappa_{2}=1. Then there exists a constant cdc_{d}, depending only on dd, such that max2sd|κs(p)|<cd.\max_{2\leq s\leq d}|\kappa_{s}(p)|<c_{d}.

Proof.

By the previous lemma mn(d1)nm_{n}\leq(d-1)^{n} and then max2sd|ms(p)|<(d1)d\max_{2\leq s\leq d}|m_{s}(p)|<(d-1)^{d}, so we can bound uniformly κn\kappa_{n} by the moment-cumulant formulas. ∎

Now we are able to prove the main theorem which we state again for convenience of the reader.

Proposition 3.3.

Let pp be a real polynomial with κ1=0\kappa_{1}=0 and κ2=1\kappa_{2}=1. Then, there exists CdC_{d} such that for all n>0n>0

L(D1/n(pdn),D1/d(Hd))<Cdn.L\left(D_{1/\sqrt{n}}(p^{\boxplus_{d}n}),D_{1/\sqrt{d}}(H_{d})\right)<\frac{C_{d}}{\sqrt{n}}.
Proof.

Let us denote h=D1/d(Hd)h=D_{1/\sqrt{d}}(H_{d}), pn=pdnp_{n}=p^{\boxplus_{d}n} and qn=D1/n(pn)q_{n}=D_{1/\sqrt{n}}(p_{n}). By the coefficient-cumulant formula, we know that

ajqn\displaystyle a_{j}^{q_{n}} =\displaystyle= (d)jdjj!π𝒫(j)d|π|μ(0j,π)κπ(qn)\displaystyle\frac{(d)_{j}}{d^{j}j!}\sum_{\pi\in\mathcal{P}(j)}d^{|\pi|}\mu(0_{j},\pi)\kappa_{\pi}(q_{n})
=\displaystyle= ajh+(d)jdjj!π𝒫(j)\𝒫12(j)d|π|μ(0j,π)κπ(qn),\displaystyle a_{j}^{h}+\frac{(d)_{j}}{d^{j}j!}\sum_{\pi\in\mathcal{P}(j)\backslash\mathcal{P}_{12}(j)}d^{|\pi|}\mu(0_{j},\pi)\kappa_{\pi}(q_{n}),

where 𝒫12(j)\mathcal{P}_{12}(j) is the set of partitions π=(V1,,Vr)𝒫(j)\pi=(V_{1},\ldots,V_{r})\in\mathcal{P}(j) such that |Vi|2|V_{i}|\leq 2 for all i{1,,r}i\in\{1,\ldots,r\} (i.e., π=(V1,,Vr)𝒫(j)\𝒫12(j)\pi=(V_{1},\ldots,V_{r})\in\mathcal{P}(j)\backslash\mathcal{P}_{12}(j), if |Vi|>2|V_{i}|>2 for some i{1,,r}i\in\{1,\ldots,r\}).

Recall that

|κs(qn)|=|κs(D1/n(pn))|=nns/2|κs(p)|n1s/2c,|\kappa_{s}(q_{n})|=|\kappa_{s}(D_{1/\sqrt{n}}(p_{n}))|=\frac{n}{n^{s/2}}|\kappa_{s}(p)|\leq n^{1-s/2}c,

for s=3,,ds=3,\ldots,d, where c:=cdc:=c_{d} from Lemma 3.2. Thus, for any 3jd3\leq j\leq d and π=(V1,,Vr)𝒫(j)\𝒫12(j)\pi=(V_{1},\ldots,V_{r})\in\mathcal{P}(j)\backslash\mathcal{P}_{12}(j) we get

(3.1) |κπ(qn)|crnrn|V1|++|Vr|2=crnrj2crnj3j2=crnj6cdn12.|\kappa_{\pi}(q_{n})|\leq c^{r}\cdot n^{r}\cdot n^{-\frac{|V_{1}|+\cdots+|V_{r}|}{2}}=c^{r}n^{r-\frac{j}{2}}\leq c^{r}n^{\frac{j}{3}-\frac{j}{2}}=c^{r}n^{-\frac{j}{6}}\leq c^{d}n^{-\frac{1}{2}}.

Then,

|ajqnajh|cdK1(d)n,j{1,,d}|a_{j}^{q_{n}}-a_{j}^{h}|\leq\frac{c^{d}K_{1}(d)}{\sqrt{n}},\qquad\forall j\in\{1,\ldots,d\}

where

K1(d)=max1jd(d)jdjj!π𝒫(j)\𝒫12(j)d|π||μ(0j,π)|.K_{1}(d)=\max_{1\leq j\leq d}\frac{(d)_{j}}{d^{j}j!}\sum_{\pi\in\mathcal{P}(j)\backslash\mathcal{P}_{12}(j)}d^{|\pi|}|\mu(0_{j},\pi)|.

Let’s denote z1,z2,,zdz_{1},z_{2},\cdots,z_{d} the dd distinct roots of hh and δ=12min1i<jd|zizj|\delta=\frac{1}{2}\min_{1\leq i<j\leq d}|z_{i}-z_{j}|. For 0<ε<δ0<\varepsilon<\delta we define Bi={z:|zzi|ε}B_{i}=\{z\in\mathbb{C}:|z-z_{i}|\leq\varepsilon\} and Bi={z:|zzi|=ε}\partial B_{i}=\{z\in\mathbb{C}:|z-z_{i}|=\varepsilon\}. For a fixed root ii, using the previous bound we can see that for any zBiz\in\partial B_{i} we have that

|qn(z)h(z)||j=0dzdj(1)j(ajqnajh)|j=1d|z|dm|ajqnajh||q_{n}(z)-h(z)|\leq\left|\sum_{j=0}^{d}z^{d-j}(-1)^{j}(a^{q_{n}}_{j}-a_{j}^{h})\right|\leq\sum_{j=1}^{d}|z|^{d-m}|a^{q_{n}}_{j}-a_{j}^{h}|
cdK1(d)nj=1d(|zi|+|ε|)djcdK1(d)K2(d)n\leq\frac{c^{d}K_{1}(d)}{\sqrt{n}}\sum_{j=1}^{d}(|z_{i}|+|\varepsilon|)^{d-j}\leq\frac{c^{d}K_{1}(d)K_{2}(d)}{\sqrt{n}}

where

K2(d)=max1idj=1d(|zi|+|ε|)dj.K_{2}(d)=\max_{1\leq i\leq d}\sum_{j=1}^{d}(|z_{i}|+|\varepsilon|)^{d-j}.

On the other hand, if zBiz\in\partial B_{i}, we know that

|h(z)|=|(zz0)(zzn1)|=|zz1||zzn||zzi|δd1=εδd1.|h(z)|=|(z-z_{0})\cdots(z-z_{n-1})|=|z-z_{1}|\cdots|z-z_{n}|\geq|z-z_{i}|\delta^{d-1}=\varepsilon\delta^{d-1}.

Finally, if we take

n>c2dK(d)ε2,n>\frac{c^{2d}K(d)}{\varepsilon^{2}},

where K(d)=K12(d)K22(d)δ2d2K(d)=\frac{K^{2}_{1}(d)K^{2}_{2}(d)}{\delta^{2d-2}}. Since c2dK(d)c^{2d}K(d) does not depend on ii, we get that for any i=1,,ni=1,\ldots,n, if zBiz\in\partial B_{i}, then

|qn(z)h(z)|cdK1(d)K2(d)n<εδd1|h(z)||h(z)|+|qn(z)|.|q_{n}(z)-h(z)|\leq\frac{c^{d}K_{1}(d)K_{2}(d)}{\sqrt{n}}<\varepsilon\delta^{d-1}\leq|h(z)|\leq|h(z)|+|q_{n}(z)|.

Thus, Rouché’s theorem implies that qnq_{n} and hh have the same number of roots (counting multiplicity) in BiB_{i} for i=1,,ni=1,\ldots,n. By the definition of the BiB_{i} we know that they are pairwise disjoint and each one contains exactly one of the dd roots of hh. Thus, each BiB_{i} contains exactly one of the dd roots of qnq_{n} implying that distance between the roots of qnq_{n} and hh, (and therefore the Lévy distance) is less than ε\varepsilon. ∎

Observe that Theorem 1.1 directly gives a bound for TT in the next proposition.

Proposition 3.4 ([1]).

Let pxdp\neq x^{d} be a real polynomial, then there exists T>0T>0 such that for all t>Tt>T the polynomial pdtp^{\boxplus_{d}t} has dd different real roots.

Finally, we show that one cannot do better than O(n)O(\sqrt{n}) as long as m3(p)0m_{3}(p)\neq 0.

Proposition 3.5.

Let pp be a real polynomial with κ1=0\kappa_{1}=0 and κ2=1\kappa_{2}=1 and |m3|=α0|m_{3}|=\alpha\neq 0. Then, for all n>0n>0

L(D1/n(pdn),D1/d(Hd))α3dn.L\left(D_{1/\sqrt{n}}(p^{\boxplus_{d}n}),D_{1/\sqrt{d}}(H_{d})\right)\geq\frac{\alpha}{3d\sqrt{n}}.
Proof.

We use again the notation h=D1/d(Hd)h=D_{1/\sqrt{d}}(H_{d}), pn=pdnp_{n}=p^{\boxplus_{d}n} and qn=D1/n(pn)q_{n}=D_{1/\sqrt{n}}(p_{n}) and suppose that L(qn,h)<α3dn.L(q_{n},h)<\frac{\alpha}{3d\sqrt{n}}. Since κ1(qn)=0\kappa_{1}(q_{n})=0, from the moment cumulant formulas we have m3(qn)=(d1)(d2)d2κ3(qn)m_{3}(q_{n})=\frac{(d-1)(d-2)}{d^{2}}\kappa_{3}(q_{n}) and then

|m3(qn)|=(d1)(d2)d2|κ3(qn)|=(d1)(d2)d2nn3/2|κ3(p)|=|m3(p)|n=αn.|m_{3}(q_{n})|=\tfrac{(d-1)(d-2)}{d^{2}}|\kappa_{3}(q_{n})|=\tfrac{(d-1)(d-2)}{d^{2}}\frac{n}{n^{3/2}}|\kappa_{3}(p)|=\frac{|m_{3}(p)|}{\sqrt{n}}=\frac{\alpha}{\sqrt{n}}.

Since m3(h)=0m_{3}(h)=0, we can compute

m3(qn)=1di=1dλi3(qn)=1di=1dλi3(qn)1di=1dλi3(h),m_{3}(q_{n})=\frac{1}{d}\sum_{i=1}^{d}\lambda^{3}_{i}(q_{n})=\frac{1}{d}\sum_{i=1}^{d}\lambda^{3}_{i}(q_{n})-\frac{1}{d}\sum_{i=1}^{d}\lambda^{3}_{i}(h),

and thus

|m3(qn)|\displaystyle|m_{3}(q_{n})| \displaystyle\leq 1di|λi3(qn)λi3(h)|\displaystyle\frac{1}{d}\sum_{i}|\lambda^{3}_{i}(q_{n})-\lambda^{3}_{i}(h)|
=\displaystyle= 1di=1d|λi(qn)λi(h)||λi2(qn)+λi(qn)λi(h)+λi2(h)|\displaystyle\frac{1}{d}\sum_{i=1}^{d}|\lambda_{i}(q_{n})-\lambda_{i}(h)||\lambda_{i}^{2}(q_{n})+\lambda_{i}(q_{n})\lambda_{i}(h)+\lambda_{i}^{2}(h)|
<\displaystyle< 1di=1d(α3dn)(d+d+d)=αn.\displaystyle\frac{1}{d}\sum_{i=1}^{d}\left(\frac{\alpha}{3d\sqrt{n}}\right)(d+d+d)=\frac{\alpha}{\sqrt{n}.}

Where we used in the last inequality the assumption that L(qn,h)<α3dn.L(q_{n},h)<\frac{\alpha}{3d\sqrt{n}}. This is a contradiction since the inequality is strict. ∎

Remark 3.6.

A specific example with κ30\kappa_{3}\neq 0 is the finite free Poisson distribution which has cumulants κn=α\kappa_{n}=\alpha for all nn. If αd\alpha d is a positive integer we obtain a valid polynomial. This is a modification of a Laguerre polynomial, thus we obtain a precise estimate for the distance between the roots of certain Laguerre polynomials and the Hermite polynomials.

Remark 3.7.

A closer look at (3.1) shows that if m3(p)=0m_{3}(p)=0 then the convergence rate is of order 1/n.1/n. Indeed, m3(p)=0m_{3}(p)=0 implies κ3(qn)=κ3(p)=0\kappa_{3}(q_{n})=\kappa_{3}(p)=0. So in (3.1) we only need to consider partitions with |Vi|4|V_{i}|\geq 4. In this case, for any 4jd4\leq j\leq d we have

|κπ(p)|crnrj2crnj4j2=crnj4cdn1.|\kappa_{\pi}(p)|\leq c^{r}n^{r-\frac{j}{2}}\leq c^{r}n^{\frac{j}{4}-\frac{j}{2}}=c^{r}n^{-\frac{j}{4}}\leq c^{d}n^{-1}.

Finally, let us mention that while it is very tempting to let dd go to infinity, possibly together with nn, to obtain a Berry-Esseen bound in free probability, there are two problems. First, the quantity CdC_{d}, as we obtained it, increases broadly as dd\to\infty, and second, there is, for the moment not a good bound between finite free and free convolutions.

References

  • [1] O. Arizmendi and D. Perales. Cumulants for finite free convolution. Journal of Combinatorial Theory, Series A, 155:244–266, 2018.
  • [2] A. C. Berry. The accuracy of the gaussian approximation to the sum of independent variates. Transactions of the American Mathematical Society, 49(1):122–136, 1941.
  • [3] G. P. Chistyakov and F. Götze. Limit theorems in free probability theory. i. The Annals of Probability, pages 54–90, 2008.
  • [4] C.-G. Esseen. On the Liapounoff limit of error in the theory of probability. Almqvist & Wiksell, 1942.
  • [5] V. Kargin. Berry–esseen for free random variables. Journal of Theoretical Probability, 20(2):381–395, 2007.
  • [6] F. Lehner. Free cumulants and enumeration of connected partitions. European Journal of Combinatorics, 23(8):1025–1031, 2002.
  • [7] A. Marcus. Polynomial convolutions and (finite) free probability. Preprint, 2015.
  • [8] A. Marcus, D. A. Spielman, and N. Srivastava. Finite free convolutions of polynomials. arXiv preprint arXiv:1504.00350, 2015.
  • [9] A. Marcus, D. A. Spielman, and N. Srivastava. Interlacing families IV: Bipartite ramanujan graphs of all sizes. In 2015 IEEE 56th Annual Symposium on Foundations of Computer Science (FOCS), pages 1358–1377. IEEE, 2015.
  • [10] J. L. Walsh. On the location of the roots of certain types of polynomials. Transactions of the American Mathematical Society, 24(3):163–180, 1922.