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

Even hypergeometric polynomials and finite free commutators

Jacob Campbell Rafael Morales  and  Daniel Perales
Abstract.

We study in detail the class of even polynomials and their behavior with respect to finite free convolutions. To this end, we use some specific hypergeometric polynomials and a variation of the rectangular finite free convolution to understand even real-rooted polynomials in terms of positive-rooted polynomials. Then, we study some classes of even polynomials that are of interest in finite free probability, such as even hypergeometric polynomials, symmetrizations, and finite free commutators. Specifically, we provide many new examples of these objects, involving classical families of special polynomials (such as Laguerre, Hermite, and Jacobi). Finally, we relate the limiting root distributions of sequences of even polynomials with the corresponding symmetric measures that arise in free probability.

1. Introduction

In the past decade, the subject of finite free probability has grown considerably due to its connections with geometry of polynomials, combinatorics, random matrix theory, and free probability. The core objects of study in finite free probability are polynomials of a fixed degree nn, and some convolution operations on such polynomials. These convolutions, called additive and multiplicative finite free convolutions and denoted by n\boxplus_{n} and n\boxtimes_{n}, were studied a century ago [Wal22, Sze22] but were recently rediscovered [MSS22] as expected characteristic polynomials of certain random matrices. The finite free convolutions preserve various real-rootedness and interlacing properties, and when nn\to\infty, they approximate the additive and multiplicative free convolutions of Voiculescu [VDN92], which are denoted by \boxplus and \boxtimes.

Since [MSS22] first appeared in 2015, there have been several developments in finite free probability which expand on the parallel with free probability. In [Mar21], Marcus developed finite RR- and SS-transforms in analogy with the corresponding analytic functions in free probability. On the other hand, the combinatorial side of finite free probability was developed by Arizmendi, Garza-Vargas, and Perales [AP18, AGVP23] with a cumulant sequence for finite free convolutions, in analogy with the work of Nica and Speicher on free cumulants [NS06].

There are two other operations on measures in free probability which have finite analogues that are relevant to this paper. One of them is the rectangular additive convolution studied by Gribinski and Marcus [GM22], which is analogous to the rectangular free convolution defined by Benaych-Georges [BG09]. The other is the finite free commutator operation studied by Campbell [Cam22], which is analogous to the free commutator operation studied by Nica and Speicher [NS98]. A common thread between these two operations is their relation to even polynomials: these are the polynomials whose roots come in positive-negative pairs.

Thus far, the class of even polynomials has not been studied in general in the context of finite free probability. In this paper, we carry out a detailed study of the class of even polynomials and compile some basic but useful results concerning their behavior with respect to finite free convolutions. In many cases, convolutions of even real-rooted polynomials can be reinterpreted as convolutions of polynomials with all non-negative roots and some specific hypergeometric polynomials that were studied recently in [MFMP24].

The basic tool in this paper is the map that takes an even polynomial, i.e. a polynomial of the form q(x2)q(x^{2}), and returns the polynomial q(x)q(x). We call this map 𝑸m\bm{Q}_{m}, where mm is the degree of qq. An important property of this map is that q(x2)q(x^{2}) is real-rooted if and only if q(x)q(x) is positive-rooted; this is helpful because the behavior of positive-rooted polynomials with respect to finite free convolutions is well understood. However, 𝑸m\bm{Q}_{m} does not quite preserve finite free convolutions:

𝑸m(p2mq)𝑸m(p)m𝑸m(q) and 𝑸m(p2mq)𝑸m(p)m𝑸m(q).\bm{Q}_{m}(p\boxplus_{2m}q)\neq\bm{Q}_{m}(p)\boxplus_{m}\bm{Q}_{m}(q)\text{ and }\bm{Q}_{m}(p\boxtimes_{2m}q)\neq\bm{Q}_{m}(p)\boxtimes_{m}\bm{Q}_{m}(q)\text{.}

To make these into equalities, we modify the right-hand sides using some particular hypergeometric polynomials of the kind studied in [MFMP24]. In the additive case, the result is related to a variation of the rectangular convolution of [GM22]. We also study an interesting operation in finite free probability which takes a single polynomial pp and returns the even polynomial p(x)np(x)p(x)\boxplus_{n}p(-x), which we call the symmetrization of pp. This operation appeared naturally in [Cam22] in relation to the finite free commutator.

Then, we establish even versions of the algebraic results of [MFMP24], which concern the behavior of even hypergeometric polynomials with respect to finite free convolutions. In particular, we use some known product identities concerning hypergeometric series to provide many non-trivial examples of symmetrizations of hypergeometric polynomials (such as Laguerre, Hermitte, or Jacobi) which are of interest in finite free probability.

We also use our framework of even hypergeometric polynomials to provide some new insight into finite free commutators: the result of [Cam22] can be phrased in terms of even hypergeometric polynomials, and we provide some partial results concerning real-rootedness. We work out many examples of finite free commutators, and connect them asymptotically with known examples of commutators in free probability.

Finally, we study the asymptotic behavior of even polynomials in connection to free probability. In some cases we obtain new results in free probability. For instance, in the limit, the symmetrization operation tends to the analogous operation on probability measures, which we call free symmetrization: given a probability measure μ\mu, its free symmetrization is the measure μμ~\mu\boxplus\widetilde{\mu}, where μ~\widetilde{\mu} is just the pushforward of μ\mu by xxx\mapsto-x. To the best of our knowledge this type of operation has not been studied systematically in free probability, but it has appeared sporadically in different contexts (see Remark 6.5). Our machinery allows us to compute the free symmetrizations of some special distributions, including the Marchenko-Pastur, reversed Marchenko-Pastur, and free beta distributions.

Besides this introductory section, the rest of the paper is organized as follows. In Section 2, we review some preliminaries on polynomials, measures, and finite free probability. In Section 3, we establish our basic framework for even polynomials and study its behavior in relation to finite free convolutions. We specialize to the study of even hypergeometric polynomials in Section 4. In Section 5, we use even hypergeometric polynomials to study the finite free commutator operation and provide many examples. Finally, in Section 6 we study the asymptotic behavior of families of even polynomials and relate it to results in free probability.

2. Preliminaries

2.1. Polynomials and their coefficients

We start by introducing some notation. We denote by 𝒫n\mathcal{P}_{n} the set of monic polynomials (over the complex plane \mathbb{C}) of degree nn. To specify that all the roots of a polynomial belong to a specific region KK\subseteq\mathbb{C}, we use the notation 𝒫n(K)\mathcal{P}_{n}(K). For most of our results, KK is going to be either the set of real numbers \mathbb{R}, the set positive real number >0\mathbb{R}_{>0}, or the set of negative real numbers <0\mathbb{R}_{<0}.

Notation 2.1 (Roots and coefficients).

Given a polynomial p𝒫np\in\mathcal{P}_{n} we denote its roots by λ1(p),,λn(p)\lambda_{1}(p),\ldots,\lambda_{n}(p). Every polynomial p𝒫np\in\mathcal{P}_{n} can be written in the form

p(x)=k=0nxnk(1)k𝖾k(p)p(x)=\sum_{k=0}^{n}x^{n-k}(-1)^{k}\mathsf{e}_{k}(p) (1)

for some coefficients 𝖾k(p)\mathsf{e}_{k}(p). There is a specific formula for these coefficients:

𝖾k(p)=𝖾0(p)1i1<<iknλi1(p)λik(p)\mathsf{e}_{k}(p)=\mathsf{e}_{0}(p)\sum_{1\leq i_{1}<\ldots<i_{k}\leq n}\lambda_{i_{1}}(p)\cdots\lambda_{i_{k}}(p)

for 0kn0\leq k\leq n. These are the so-called elementary symmetric polynomials in the roots of pp.

Notation 2.2 (Dilation).

The dilation of a polynomial pp by a non-zero scalar α\alpha is defined as

[Dilαp](x)=αnp(α1x).[\mathrm{Dil}_{\alpha}p](x)=\alpha^{n}p(\alpha^{-1}x)\text{.}

The roots of Dilαp\mathrm{Dil}_{\alpha}p are the roots of pp scaled by α\alpha: αλ1(p),,αλn(p)\alpha\lambda_{1}(p),\dots,\alpha\lambda_{n}(p).

2.2. Measures and asymptotic empirical root distribution

We will be interested in sequences of polynomials with increasing degree, whose roots tend in the limit to a probability measure. We denote by \mathcal{M} the set of probability measures on the complex plane. Similar to our notation for polynomials, for KK\subseteq\mathbb{C}, we denote by (K)\mathcal{M}(K) the set of probability measures supported on KK. In particular, ()\mathcal{M}(\mathbb{R}) is the set of probability measures supported on the real line. We also denote by E()\mathcal{M}^{E}(\mathbb{R}) the set of symmetric probability measures on \mathbb{R}.

Notation 2.3 (Cauchy transform).

For a probability measure μ()\mu\in\mathcal{M}(\mathbb{R}), the Cauchy transform of μ\mu is defined by

Gμ(z):=1zt𝑑μ(t).G_{\mu}(z):=\int_{\mathbb{R}}\frac{1}{z-t}\,d\mu(t)\text{.}

This is an analytic function from the upper half-plane to the lower half-plane. Among other things, the Cauchy transform encodes weak convergence: if (μn)n1(\mu_{n})_{n\geq 1} is a sequence in ()\mathcal{M}(\mathbb{R}) and μ()\mu\in\mathcal{M}(\mathbb{R}), then μnμ\mu_{n}\to\mu weakly if and only if Gμn(z)Gμ(z)G_{\mu_{n}}(z)\to G_{\mu}(z) pointwise. This result can be found in e.g. [MS17, Remark 3.12, Theorem 3.13].

Notation 2.4.

For μE()\mu\in\mathcal{M}^{E}(\mathbb{R}), let Q(μ)Q(\mu) be the pushforward of μ\mu along the map 0:tt2\mathbb{R}\to\mathbb{R}_{\geq 0}:t\mapsto t^{2}. This is a probability measure supported on 0\mathbb{R}_{\geq 0}, and it has appeared before in the free probability literature [NS98, APA09]. A useful description of Q(μ)Q(\mu) was given in [APA09, Proposition 5] in terms of the Cauchy transform:

Gμ(z)=zGQ(μ)(z2).G_{\mu}(z)=zG_{Q(\mu)}(z^{2})\text{.}

In the other direction, for ν(0)\nu\in\mathcal{M}(\mathbb{R}_{\geq 0}), write

S(ν)=12(S+(ν)+S(ν)),S(\nu)=\frac{1}{2}(S_{+}(\nu)+S_{-}(\nu))\text{,}

where S+(ν)S_{+}(\nu) and S(ν)S_{-}(\nu) are the pushforwards of ν\nu along the maps

00:tt and 00:tt\mathbb{R}_{\geq 0}\to\mathbb{R}_{\geq 0}:t\mapsto\sqrt{t}\text{ and }\mathbb{R}_{\geq 0}\to\mathbb{R}_{\leq 0}:t\mapsto-\sqrt{t}

respectively.

Lemma 2.5.

Q:E()(0)Q:\mathcal{M}^{E}(\mathbb{R})\to\mathcal{M}(\mathbb{R}_{\geq 0}) is a homeomorphism with respect to weak convergence, with inverse SS.

Proof.

The map QQ is weakly continuous because it is a pushforward along a continuous map [Bil99, Theorem 2.7]. One can check that SS is the inverse of QQ by integrating against bounded continuous functions, and SS is continuous for the same reason as QQ. ∎

In this paper we will often be concerned with root distributions of polynomials. For a non-zero polynomial pp, the empirical root distribution (or zero counting measure) of pp is the measure

ρ(p):=1deg(p)α root of pδα\rho(p):=\frac{1}{\deg(p)}\sum_{\alpha\text{ root of }p}\delta_{\alpha}\in\mathcal{M}

where the roots in the sum are counted with multiplicity and δα\delta_{\alpha} is the Dirac delta (unit mass) placed at the point α\alpha.

Definition 2.6.

We say that a sequence 𝔭=(pnk)k1\mathfrak{p}=\left(p_{n_{k}}\right)_{k\geq 1} of polynomials is converging if

  • (nk)k1(n_{k})_{k\geq 1} is a strictly increasing subsequence of integers,

  • pnkp_{n_{k}} has degree nkn_{k} for k1k\geq 1, and

  • there is a measure, denoted by ρ(𝔭)\rho(\mathfrak{p}), such that ρ(pnk)ρ(𝔭)\rho(p_{n_{k}})\to\rho(\mathfrak{p}) weakly as kk\to\infty.

For our purposes, the subsequence (nk)k1(n_{k})_{k\geq 1} will usually be the full sequence of integers or the subsequence of even integers.

2.3. Free convolution

In this paper we will occasionally use some tools from free probability. In this section we will briefly review the facts we need, following the references [NS06, MS17].

Given a compactly supported measure μ()\mu\in\mathcal{M}(\mathbb{R}) with moment sequence (mk)k1(m_{k})_{k\geq 1}, the moment generating function of μ\mu is given by

Mμ(z)=k=1mkzk,M_{\mu}(z)=\sum_{k=1}^{\infty}m_{k}z^{k}\text{,}

and the Cauchy transform GμG_{\mu} has the Laurent expansion

Gμ(z)=k=0mkz(k+1)G_{\mu}(z)=\sum_{k=0}^{\infty}m_{k}z^{-(k+1)}

in a neighborhood of \infty. This yields the relation Mμ(z)=z1Gμ(z1)1M_{\mu}(z)=z^{-1}G_{\mu}(z^{-1})-1.

The RR-transform and SS-transform of μ\mu are defined by

Rμ(z)=Gμ1(z)1z,andSμ(z)=z+1zMμ1(z).R_{\mu}(z)=G_{\mu}^{\langle-1\rangle}(z)-\frac{1}{z},\qquad\text{and}\qquad S_{\mu}(z)=\frac{z+1}{z}M_{\mu}^{\langle-1\rangle}(z).

Here, the notation 1\langle-1\rangle indicates the compositional inverse.

Definition 2.7.

Let μ,ν()\mu,\nu\in\mathcal{M}(\mathbb{R}) be compactly supported. Then

  1. (1)

    their free additive convolution is the measure μν\mu\boxplus\nu defined by

    Rμν(z)=Rμ(z)+Rν(z),R_{\mu\boxplus\nu}(z)=R_{\mu}(z)+R_{\nu}(z), (2)
  2. (2)

    and, if μ,ν(0)\mu,\nu\in\mathcal{M}(\mathbb{R}_{\geq 0}), their free multiplicative convolution is the measure μν(0)\mu\boxtimes\nu\in\mathcal{M}(\mathbb{R}_{\geq 0}) defined by

    Sμν(z)=Sμ(z)Sν(z).S_{\mu\boxtimes\nu}(z)=S_{\mu}(z)S_{\nu}(z). (3)

2.4. Finite free convolution of polynomials

In this section we summarize some definitions and results on the finite free additive and multiplicative convolutions that will be used throughout this paper. First, let us establish some notation for rising and falling factorials.

Notation 2.8.

For aa\in\mathbb{C} and k0:={0}k\in\mathbb{Z}_{\geq 0}:=\mathbb{N}\cup\{0\}, the rising and falling factorials111The rising and falling factorials are both sometimes called the Pochhammer symbol. The notation (x)n(x)_{n} is sometimes used to refer to either the rising or falling factorial; we prefer the clear notation laid out in the text. are respectively defined as

(a)k¯:=a(a+1)(a+k1)\left(a\right)^{\overline{k}}:=a(a+1)\cdots(a+k-1)

and

(a)k¯:=a(a1)(ak+1)=(ak+1)k¯.\left(a\right)^{\underline{k}}:=a(a-1)\cdots(a-k+1)=\left(a-k+1\right)^{\overline{k}}\text{.}

The following relations follow from the definition and will be useful later. For aa\in\mathbb{C} and j,k0j,k\in\mathbb{Z}_{\geq 0} with jkj\leq k, we have

(a)k¯\displaystyle\left(a\right)^{\underline{k}} =(1)k(a)k¯,\displaystyle=(-1)^{k}\left(-a\right)^{\overline{k}}\text{,} (4)
(a)k¯\displaystyle\left(a\right)^{\overline{k}} =(a)kj¯(a+k1)j¯, and\displaystyle=\left(a\right)^{\overline{k-j}}\left(a+k-1\right)^{\underline{j}}\text{, and } (5)
(2a)2k¯\displaystyle\left(2a\right)^{\underline{2k}} =22k(a)k¯(a12)k¯.\displaystyle=2^{2k}\left(a\right)^{\underline{k}}\left(a-\frac{1}{2}\right)^{\underline{k}}\text{.} (6)

We are now ready to introduce the multiplicative and additive convolutions, as defined in [MSS22, Definition 1.1 and 1.4].

Definition 2.9 (Additive and multiplicative convolutions).

Consider polynomials p,q𝒫np,q\in\mathcal{P}_{n}. We define the finite free additive convolution of pp and qq as the polynomial pnq𝒫np\boxplus_{n}q\in\mathcal{P}_{n} with coefficients given by

𝖾k(pnq)=(n)k¯i+j=k𝖾i(p)𝖾j(q)(n)i¯(n)j¯for 0kn.\mathsf{e}_{k}(p\boxplus_{n}q)=\left(n\right)^{\underline{k}}\sum_{i+j=k}\frac{\mathsf{e}_{i}(p)\mathsf{e}_{j}(q)}{\left(n\right)^{\underline{i}}\left(n\right)^{\underline{j}}}\qquad\text{for }0\leq k\leq n. (7)

We define the finite free multiplicative convolution of pp and qq as the polynomial pnq𝒫np\boxtimes_{n}q\in\mathcal{P}_{n} with coefficients given by

𝖾k(pnq)=1(nk)𝖾k(p)𝖾k(q)for 0kn.\mathsf{e}_{k}(p\boxtimes_{n}q)=\frac{1}{\binom{n}{k}}\mathsf{e}_{k}(p)\mathsf{e}_{k}(q)\qquad\text{for }0\leq k\leq n. (8)
Remark 2.10.

We will also use an equivalent description of n\boxplus_{n} in terms of certain differential operators, which was also given in [MSS22]. Namely, if we can write p(x)=P[xn]p(x)=P[x^{n}] and q(x)=Q[xn]q(x)=Q[x^{n}] for some differential operators PP and QQ, then

p(x)nq(x)=P[Q[xn]]=Q[P[xn]].p(x)\boxplus_{n}q(x)=P[Q[x^{n}]]=Q[P[x^{n}]]\text{.} (9)

Notice that for a given pp, the differential operator PP is not unique. The simplest way to construct such an operator is by taking

P=k=0n(1)k𝖾k(p)(n)k¯(x)k.P=\sum_{k=0}^{n}(-1)^{k}\frac{\mathsf{e}_{k}(p)}{\left(n\right)^{\underline{k}}}\left(\tfrac{\partial}{\partial x}\right)^{k}\text{.} (10)

To construct other differential operators that yield pp one can add terms of the form (x)k\left(\frac{\partial}{\partial x}\right)^{k} with k>nk>n, which vanish when applied to xnx^{n}.

Remark 2.11 (Basic properties).

Directly from the definition, one can derive some basic properties of the binary operations n\boxplus_{n} and n\boxtimes_{n}. For example, they are bilinear, associative, and commutative. Another property of n\boxtimes_{n} is related to scaling: for α\alpha\in\mathbb{C} and p𝒫np\in\mathcal{P}_{n}, we have

p(x)n(xα)n=Dilαp.p(x)\boxtimes_{n}(x-\alpha)^{n}=\mathrm{Dil}_{\alpha}p.

The main property of finite free convolutions is that they preserve real-rootedness, in the following sense:

Theorem 2.12 ([Wal22, Sze22]).

Let p,q𝒫n()p,q\in\mathcal{P}_{n}(\mathbb{R}). Then

  1. (1)

    pnq𝒫n()p\boxplus_{n}q\in\mathcal{P}_{n}(\mathbb{R});

  2. (2)

    if either p𝒫n(0)p\in\mathcal{P}_{n}(\mathbb{R}_{\geq 0}) or q𝒫n(0)q\in\mathcal{P}_{n}(\mathbb{R}_{\geq 0}), then pnq𝒫n()p\boxtimes_{n}q\in\mathcal{P}_{n}(\mathbb{R});

  3. (3)

    if both p,q𝒫n(0)p,q\in\mathcal{P}_{n}(\mathbb{R}_{\geq 0}), then pnq𝒫n(0)p\boxtimes_{n}q\in\mathcal{P}_{n}(\mathbb{R}_{\geq 0}).

Remark 2.13.

It is easy to extend (3) in Theorem 2.12 to a “rule of signs” for the behavior of roots under the operation n\boxtimes_{n}. Specifically, since Dil1p=pn(x+1)n\mathrm{Dil}_{-1}p=p\boxtimes_{n}(x+1)^{n} by Remark 2.11, we have the following:

  • if p,q𝒫n(0)p,q\in\mathcal{P}_{n}(\mathbb{R}_{\leq 0}), then pnq𝒫n(0)p\boxtimes_{n}q\in\mathcal{P}_{n}(\mathbb{R}_{\geq 0});

  • if p𝒫n(0)p\in\mathcal{P}_{n}(\mathbb{R}_{\leq 0}) and q𝒫n(0)q\in\mathcal{P}_{n}(\mathbb{R}_{\geq 0}), then pnq𝒫n(0)p\boxtimes_{n}q\in\mathcal{P}_{n}(\mathbb{R}_{\leq 0}).

The connection between finite free probability and free probability is revealed in the asymptotic regime; this was first observed by Marcus [Mar21, Section 4] and formalized later using finite free cumulants [AP18, AGVP23].

Theorem 2.14 ([AP18, Corollary 5.5], [AGVP23, Theorem 1.4]).

Let 𝔭:=(pn)n=1\mathfrak{p}:=\left(p_{n}\right)_{n=1}^{\infty} and 𝔮:=(qn)n=1\mathfrak{q}:=\left(q_{n}\right)_{n=1}^{\infty} be two converging sequences of polynomials in the sense of Definition 2.6.

  1. (i)

    If 𝔭,𝔮𝒫()\mathfrak{p},\mathfrak{q}\subset\mathcal{P}(\mathbb{R}), then (pnnqn)n=1\left(p_{n}\boxplus_{n}q_{n}\right)_{n=1}^{\infty} weakly converges to ρ(𝔭)ρ(𝔮)\rho(\mathfrak{p})\boxplus\rho(\mathfrak{q}).

  2. (ii)

    If 𝔭𝒫()\mathfrak{p}\subset\mathcal{P}(\mathbb{R}) and 𝔮𝒫(>0)\mathfrak{q}\subset\mathcal{P}(\mathbb{R}_{>0}) then (pnnqn)n=1\left(p_{n}\boxtimes_{n}q_{n}\right)_{n=1}^{\infty} weakly converges to ρ(𝔭)ρ(𝔮)\rho(\mathfrak{p})\boxtimes\rho(\mathfrak{q}).

2.5. Hypergeometric polynomials and examples

A large class of polynomials with real roots is contained in the class of hypergeometric polynomials; these were recently studied in connection with finite free probability in [MFMP24]. This class contains several important families, such as Bessel, Laguerre and Jacobi polynomials. These polynomials – and their specializations such as Hermite polynomials – constitute a rich class of examples in the theory of finite free probability.

Definition 2.15.

For i,j0i,j\in\mathbb{Z}_{\geq 0} and n1n\geq 1, pick some parameters

a1,,ai{0,1n,2n,,n1n} and b1,,bjja_{1},\ldots,a_{i}\in\mathbb{R}\setminus\left\{0,\tfrac{1}{n},\tfrac{2}{n},\ldots,\tfrac{n-1}{n}\right\}\text{ and }b_{1},\ldots,b_{j}\in\mathbb{R}^{j} (11)

Define the polynomial n[.b1,,bja1,,ai.]\mathcal{H}_{n}{\left[\genfrac{.}{.}{0.0pt}{1}{b_{1},\ldots,b_{j}}{a_{1},\ldots,a_{i}}\right]} to be the unique monic polynomial of degree nn with coefficients in representation (1) given by

𝖾k(n[.b1,,bja1,,ai.]):=(nk)(nb1)k¯(nbj)k¯(na1)k¯(nai)k¯\mathsf{e}_{k}\left(\mathcal{H}_{n}{\left[\genfrac{.}{.}{0.0pt}{1}{b_{1},\ldots,b_{j}}{a_{1},\ldots,a_{i}}\right]}\right):=\binom{n}{k}\frac{\left(nb_{1}\right)^{\underline{k}}\cdots\left(nb_{j}\right)^{\underline{k}}}{\left(na_{1}\right)^{\underline{k}}\cdots\left(na_{i}\right)^{\underline{k}}}

for 1kn1\leq k\leq n.

To simplify notation, for a tuple 𝒂=(a1,,ai)\bm{a}=(a_{1},\ldots,a_{i}) we will write

(𝒂)k¯:=s=1i(as)k¯.\left(\bm{a}\right)^{\underline{k}}:=\prod_{s=1}^{i}\left(a_{s}\right)^{\overline{k}}\text{.}

Then, for tuples 𝒂=(a1,,ai)\bm{a}=(a_{1},\ldots,a_{i}) and 𝒃=(b1,,bj)\bm{b}=(b_{1},\ldots,b_{j}) satisfying Eq. 11, the hypergeometric polynomial n[.𝒃𝒂.]\mathcal{H}_{n}{\left[\genfrac{.}{.}{0.0pt}{1}{\bm{b}}{\bm{a}}\right]} has coefficients given by

𝖾k(n[.𝒃𝒂.])=(nk)(𝒃n)k¯(𝒂n)k¯\mathsf{e}_{k}\left(\mathcal{H}_{n}{\left[\genfrac{.}{.}{0.0pt}{1}{\bm{b}}{\bm{a}}\right]}\right)=\binom{n}{k}\frac{\left(\bm{b}n\right)^{\underline{k}}}{\left(\bm{a}n\right)^{\underline{k}}}

for 1kn1\leq k\leq n.

Remark 2.16.

The reason we call n[.𝒃𝒂.]\mathcal{H}_{n}{\left[\genfrac{.}{.}{0.0pt}{1}{\bm{b}}{\bm{a}}\right]} “hypergeometric” is that it can be identified as a terminating generalized hypergeometric series [KLS10, OLBC10]:

n[.𝒃𝒂.](x)=(1)n(𝒃n)n¯(𝒂n)n¯i+1Fj(.n,𝒂nn+1𝒃nn+1.;x)\mathcal{H}_{n}{\left[\genfrac{.}{.}{0.0pt}{1}{\bm{b}}{\bm{a}}\right]}(x)=\frac{(-1)^{n}\left(\bm{b}n\right)^{\underline{n}}}{\left(\bm{a}n\right)^{\underline{n}}}{\ }_{i+1}F_{j}{\left(\genfrac{.}{.}{0.0pt}{}{-n,\bm{a}n-n+1}{\bm{b}n-n+1};x\right)} (12)

where 𝒄nn+1\bm{c}n-n+1 means that we multiply each entry of 𝒄\bm{c} by nn and then add n+1-n+1, and we use the standard notation Fji{}_{i}F_{j} for a generalized hypergeometric series. Namely, for tuples 𝒂=(a1,,ai)i\bm{a}=(a_{1},\ldots,a_{i})\in\mathbb{R}^{i} and 𝒃=(b1,,bj)j\bm{b}=(b_{1},\ldots,b_{j})\in\mathbb{R}^{j}, we write

iFj(.𝒂𝒃.;x):=k=0(𝒂)k¯(𝒃)k¯xkk!.{\ }_{i}F_{j}{\left(\genfrac{.}{.}{0.0pt}{}{\bm{a}}{\bm{b}};x\right)}:=\sum_{k=0}^{\infty}\frac{\left(\bm{a}\right)^{\overline{k}}}{\left(\bm{b}\right)^{\overline{k}}}\frac{x^{k}}{k!}\text{.} (13)

In [MFMP24] it was noticed that these hypergeometric polynomials behave well with respect to finite free convolution.

Theorem 2.17 ([MFMP24, Equations (82)-(84)]).

Consider tuples 𝐚1\bm{a}_{1}, 𝐚2\bm{a}_{2}, 𝐚3\bm{a}_{3}, 𝐛1\bm{b}_{1}, 𝐛2\bm{b}_{2}, and 𝐛3\bm{b}_{3} with lengths i1i_{1}, i2i_{2}, i3i_{3}, j1j_{1}, j2j_{2}, and j3j_{3} respectively. Then

  1. (1)

    the multiplicative convolution from Definition 2.9 is given by

    n[.𝒃1𝒂1.]nn[.𝒃2𝒂2.]=n[.𝒃1,𝒃2𝒂1,𝒂2.];\mathcal{H}_{n}{\left[\genfrac{.}{.}{0.0pt}{1}{\bm{b}_{1}}{\bm{a}_{1}}\right]}\boxtimes_{n}\mathcal{H}_{n}{\left[\genfrac{.}{.}{0.0pt}{1}{\bm{b}_{2}}{\bm{a}_{2}}\right]}=\mathcal{H}_{n}{\left[\genfrac{.}{.}{0.0pt}{1}{\bm{b}_{1},\ \bm{b}_{2}}{\bm{a}_{1},\ \bm{a}_{2}}\right]}\text{;} (14)
  2. (2)

    if we have the factorization

    j1Fi1(.n𝒃1n𝒂1.;x)j2Fi2(.n𝒃2n𝒂2.;x)=j3Fi3(.n𝒃3n𝒂3.;x){\ }_{j_{1}}F_{i_{1}}{\left(\genfrac{.}{.}{0.0pt}{}{-n\bm{b}_{1}}{-n\bm{a}_{1}};x\right)}{\ }_{j_{2}}F_{i_{2}}{\left(\genfrac{.}{.}{0.0pt}{}{-n\bm{b}_{2}}{-n\bm{a}_{2}};x\right)}={\ }_{j_{3}}F_{i_{3}}{\left(\genfrac{.}{.}{0.0pt}{}{-n\bm{b}_{3}}{-n\bm{a}_{3}};x\right)} (15)

    and write sl=(1)il+jl+1s_{l}=(-1)^{i_{l}+j_{l}+1} for l=1,2,3l=1,2,3, then the additive convolution from Definition 2.9 is given by

    n[.𝒃1𝒂1.](s1x)nn[.𝒃2𝒂2.](s2x)=n[.𝒃3𝒂3.](s3x).\mathcal{H}_{n}{\left[\genfrac{.}{.}{0.0pt}{1}{\bm{b}_{1}}{\bm{a}_{1}}\right]}(s_{1}x)\boxplus_{n}\mathcal{H}_{n}{\left[\genfrac{.}{.}{0.0pt}{1}{\bm{b}_{2}}{\bm{a}_{2}}\right]}(s_{2}x)=\mathcal{H}_{n}{\left[\genfrac{.}{.}{0.0pt}{1}{\bm{b}_{3}}{\bm{a}_{3}}\right]}(s_{3}x)\text{.} (16)

The limiting distribution of hypergeometric polynomials can be expressed concretely in terms of the SS-transform.

Theorem 2.18 ([MFMP25, Theorem 3.9], [AFPU24, Corollary 10.8]).

For integers i,j0i,j\geq 0, consider tuples 𝐀=(A1,,Ai)([0,1))i\bm{A}=(A_{1},\dots,A_{i})\in(\mathbb{R}\setminus[0,1))^{i} and 𝐁=(B1,,Bj)({0})j\bm{B}=(B_{1},\dots,B_{j})\in(\mathbb{R}\setminus\{0\})^{j}. Assume that 𝔭=(pn)n0\mathfrak{p}=(p_{n})_{n\geq 0} is a sequence of polynomials such that

pn=Dilnijn[.𝒃n𝒂n.]𝒫n(>0),p_{n}=\mathrm{Dil}_{n^{i-j}}\mathcal{H}_{n}{\left[\genfrac{.}{.}{0.0pt}{1}{\bm{b}_{n}}{\bm{a}_{n}}\right]}\in\mathcal{P}_{n}(\mathbb{R}_{>0}), (17)

where the tuples of parameters 𝐚ni\bm{a}_{n}\in\mathbb{R}^{i} and 𝐛nj\bm{b}_{n}\in\mathbb{R}^{j} have a limit given by

limn𝒂n=𝑨,andlimn𝒃n=𝑩.\lim_{n\to\infty}\bm{a}_{n}=\bm{A},\qquad\text{and}\qquad\lim_{n\to\infty}\bm{b}_{n}=\bm{B}. (18)

Then 𝔭\mathfrak{p} is a converging sequences in the sense of Definition 2.6. Moreover, ρ(𝔭)(0)\rho(\mathfrak{p})\in\mathcal{M}(\mathbb{R}_{\geq 0}) has SS-transform given by

Sρ(𝔭)(z)=r=1i(z+Ar)s=1j(z+Bs)S_{\rho(\mathfrak{p})}(z)=\frac{\prod_{r=1}^{i}(z+A_{r})}{\prod_{s=1}^{j}(z+B_{s})} (19)

Recall that since ρ(𝔭)\rho(\mathfrak{p}) is supported in (0)\mathcal{M}(\mathbb{R}_{\geq 0}), then the measure is determined by the SS-transform. In view of the previous result, we will use the following notation.

Notation 2.19.

If μ(0)\mu\in\mathcal{M}(\mathbb{R}_{\geq 0}) is a measure with SS-transform of the form

Sμ(z)=r=1i(z+ar)s=1j(z+bs),S_{\mu}(z)=\frac{\prod_{r=1}^{i}(z+a_{r})}{\prod_{s=1}^{j}(z+b_{s})}\text{,}

for some parameters a1,,aia_{1},\dots,a_{i} and b1,,bjb_{1},\dots,b_{j}, then we say that μ\mu is SS-rational measure and denote it by ρ[.b1,,bja1,,ai.]\rho{\left[\genfrac{.}{.}{0.0pt}{1}{b_{1},\dots,b_{j}}{a_{1},\dots,a_{i}}\right]}.

Moreover, we denote the square root of an SS-rational function by

ρE[.b1,,bja1,,ai.]:=S(ρ[.b1,,bja1,,ai.])\rho^{E}{\left[\genfrac{.}{.}{0.0pt}{1}{b_{1},\dots,b_{j}}{a_{1},\dots,a_{i}}\right]}:=S\left(\rho{\left[\genfrac{.}{.}{0.0pt}{1}{b_{1},\dots,b_{j}}{a_{1},\dots,a_{i}}\right]}\right)

where SS is the bijection from 2.4.

Notice that, if ar=bsa_{r}=b_{s} for some rir\leq i and sjs\leq j, then

ρ[.b1,,bja1,,ai.]=ρ[.b1,,bs1,bs+1,,bja1,,ar1,ar+1,,ai.].\rho{\left[\genfrac{.}{.}{0.0pt}{1}{b_{1},\dots,b_{j}}{a_{1},\dots,a_{i}}\right]}=\rho{\left[\genfrac{.}{.}{0.0pt}{1}{b_{1},\cdots,b_{s-1},b_{s+1},\cdots,b_{j}}{a_{1},\cdots,a_{r-1},a_{r+1},\cdots,a_{i}}\right]}. (20)

As particular cases, one has that Laguerre tends to Marchenko-Pastur, Bessel tends to reversed Marchenko-Pastur, and Jacobi tends to Free beta.

Example 2.20 (Identities).

The simplest cases of Definition 2.15 are the following:

n[.𝒂𝒂.](x)=n[..](x)=(x1)n and n[.0𝒂.](x)=xn.\mathcal{H}_{n}{\left[\genfrac{.}{.}{0.0pt}{1}{\bm{a}}{\bm{a}}\right]}(x)=\mathcal{H}_{n}{\left[\genfrac{.}{.}{0.0pt}{1}{\cdot}{\cdot}\right]}(x)=(x-1)^{n}\text{ and }\mathcal{H}_{n}{\left[\genfrac{.}{.}{0.0pt}{1}{0}{\bm{a}}\right]}(x)=x^{n}\text{.}

These polynomials are identities for the operations n\boxtimes_{n} and n\boxplus_{n} respectively.

Example 2.21 (Laguerre and Hermite polynomials).

The analogues of the semicircular and free Poisson distributions in finite free probability are the Hermite and Laguerre polynomials respectively. These appear in finite free analogues of the central limit theorem and Poisson limit theorem [Mar21].

The Laguerre polynomials of interest in this paper can be written in terms of the hypergeometric polynomials n[.b.]\mathcal{H}_{n}{\left[\genfrac{.}{.}{0.0pt}{1}{b}{\cdot}\right]}. There are some ranges of the parameter bb which yield real-rooted polynomials:

  • if b{1n,,n1n}b\in\{\frac{1}{n},\ldots,\frac{n-1}{n}\}, then n[.b.]𝒫(0)\mathcal{H}_{n}{\left[\genfrac{.}{.}{0.0pt}{1}{b}{\cdot}\right]}\in\mathcal{P}(\mathbb{R}_{\geq 0}), and 0 is a root with multiplicity (1b)n(1-b)n;

  • if b(n2n,n1n)b\in(\frac{n-2}{n},\frac{n-1}{n}), then n[.b.]𝒫()\mathcal{H}_{n}{\left[\genfrac{.}{.}{0.0pt}{1}{b}{\cdot}\right]}\in\mathcal{P}(\mathbb{R});

  • if b>11nb>1-\frac{1}{n}, then n[.b.]𝒫(>0)\mathcal{H}_{n}{\left[\genfrac{.}{.}{0.0pt}{1}{b}{\cdot}\right]}\in\mathcal{P}(\mathbb{R}_{>0}).

See [MFMP24, Table 1, (56)] for more details.

We will be particularly interested in the following scaled version of a Laguerre polynomial, to which we give a special name: for λ1\lambda\geq 1, write

Ln(λ)=Dil1nn[.λ.].L_{n}^{(\lambda)}=\mathrm{Dil}_{\frac{1}{n}}\mathcal{H}_{n}{\left[\genfrac{.}{.}{0.0pt}{1}{\lambda}{\cdot}\right]}\text{.}

This polynomial can be understood as the finite analogue of the free Poisson (also known as Marchenko-Pastur) distribution in free probability. Indeed, it appears as the limiting polynomial in the finite free Poisson limit theorem [Mar21] and it converges in ERD to the free Poisson distribution with rate λ\lambda:

ρ(Ln(λ))12π(r+x)(xr)xδx[r,r+]dx\rho(L_{n}^{(\lambda)})\to\frac{1}{2\pi}\frac{\sqrt{(r_{+}-x)(x-r_{-})}}{x}\delta_{x\in[r_{-},r_{+}]}\,dx

weakly as nn\to\infty, where r±:=λ+1±2λr_{\pm}:=\lambda+1\pm 2\sqrt{\lambda}. This is a classical result; see [MFMP24, Section 5.3] and its references.

The Hermite polynomials used in this paper are defined as follows:

H2m(x):=mmm[.112m.](mx2)=Dil1mmE[.112m.].H_{2m}(x):=m^{-m}\mathcal{H}_{m}{\left[\genfrac{.}{.}{0.0pt}{1}{1-\frac{1}{2m}}{\cdot}\right]}(mx^{2})=\mathrm{Dil}_{\sqrt{\frac{1}{m}}}\mathcal{H}^{E}_{m}{\left[\genfrac{.}{.}{0.0pt}{1}{1-\frac{1}{2m}}{\cdot}\right]}\text{.}

This polynomial is known to be the appropriate finite analogue of the semicircular distribution in free probability. It appears as the limiting polynomial in the finite free CLT [Mar21] and it converges in ERD to the semicircular distribution with radius 22:

ρ(H2m)12π4x2δx[2,2]dx\rho(H_{2m})\to\frac{1}{2\pi}\sqrt{4-x^{2}}\delta_{x\in[-2,2]}\,dx

weakly as nn\to\infty. This is also a well-known classical result, see e.g. [KM16].

The resemblance between the definitions of HH and L(λ)L^{(\lambda)} is meaningful, and will be elaborated in Example 6.3.

Example 2.22 (Bessel polynomials).

The Bessel polynomials of interest in this paper can be written in terms of the hypergeometric polynomials n[.a.]\mathcal{H}_{n}{\left[\genfrac{.}{.}{0.0pt}{1}{\cdot}{a}\right]}. Some known results on their roots are the following:

  • if a(0,1n)a\in(0,\frac{1}{n}), then n[.a.]𝒫()\mathcal{H}_{n}{\left[\genfrac{.}{.}{0.0pt}{1}{\cdot}{a}\right]}\in\mathcal{P}(\mathbb{R});

  • if a<0a<0, then n[.a.]𝒫(<0)\mathcal{H}_{n}{\left[\genfrac{.}{.}{0.0pt}{1}{\cdot}{a}\right]}\in\mathcal{P}(\mathbb{R}_{<0}).

For a<0a<0, let us make the notation

Cn(a):=Dilnn[.a.]𝒫(<0).C_{n}^{(a)}:=\mathrm{Dil}_{-n}\mathcal{H}_{n}{\left[\genfrac{.}{.}{0.0pt}{1}{\cdot}{a}\right]}\in\mathcal{P}(\mathbb{R}_{<0})\text{.}

The asymptotics of these Bessel polynomials are also known:

ρ(Cn(a))a2π(r+x)(xr)x2δx[r,r+]dx\rho(C_{n}^{(a)})\to\frac{-a}{2\pi}\frac{\sqrt{(r_{+}-x)(x-r_{-})}}{x^{2}}\delta_{x\in[r_{-},r_{+}]}\,dx

weakly as nn\to\infty, where r±:=1a2±21ar_{\pm}:=-\frac{1}{a-2\pm 2\sqrt{1-a}}. See [MFMP24, Section 5.3].

Example 2.23 (Jacobi polynomials).

The Jacobi polynomials of interest in this paper can be written in terms of the hypergeometric polynomials n[.ba.]\mathcal{H}_{n}{\left[\genfrac{.}{.}{0.0pt}{1}{b}{a}\right]}. Let us recall some simple combinations of parameters which produce real-rooted Jacobi polynomials:

  • n[.ba.]𝒫n([0,1])\mathcal{H}_{n}{\left[\genfrac{.}{.}{0.0pt}{1}{b}{a}\right]}\in\mathcal{P}_{n}([0,1]) when b>11nb>1-\frac{1}{n} and a>b+11na>b+1-\frac{1}{n};

  • n[.ba.]𝒫n(<0)\mathcal{H}_{n}{\left[\genfrac{.}{.}{0.0pt}{1}{b}{a}\right]}\in\mathcal{P}_{n}(\mathbb{R}_{<0}) when b>11nb>1-\frac{1}{n} and a<0a<0;

  • n[.ba.]𝒫n(>0)\mathcal{H}_{n}{\left[\genfrac{.}{.}{0.0pt}{1}{b}{a}\right]}\in\mathcal{P}_{n}(\mathbb{R}_{>0}) when a<0a<0 and b<a1+1nb<a-1+\frac{1}{n}.

See [MFMP24] for further information. A more particular case that we will use in this paper is the polynomial on the left-hand side of the following factorization ([KLS10, Equation (1.7.1)], [MFMP24, Equation (30)]):

n[.bb+rn.]=(x1)nrr[.nr(b1)+1nrb+1.]\mathcal{H}_{n}{\left[\genfrac{.}{.}{0.0pt}{1}{b}{b+\frac{r}{n}}\right]}=(x-1)^{n-r}\mathcal{H}_{r}{\left[\genfrac{.}{.}{0.0pt}{1}{\frac{n}{r}(b-1)+1}{\frac{n}{r}b+1}\right]} (21)

for 0rn0\leq r\leq n. For b>1rn2b>1-\frac{r}{n^{2}}, by the first bullet point above, the hypergeometric polynomial on the right-hand side of Eq. 21 is in 𝒫n([0,1])\mathcal{P}_{n}([0,1]), so the polynomial on the left-hand side is in 𝒫n([0,1])\mathcal{P}_{n}([0,1]) as well.

Similarly, one can pick parameters to produce Jacobi polynomials with roots at only 0 and 11:

Rn(r)(x):=n[.r/n1.](x)=(x1)rxnr for 0kn.R_{n}^{(r)}(x):=\mathcal{H}_{n}{\left[\genfrac{.}{.}{0.0pt}{1}{r/n}{1}\right]}(x)=(x-1)^{r}x^{n-r}\text{ for }0\leq k\leq n\text{.} (22)

This is the characteristic polynomial of an orthogonal projection on n\mathbb{C}^{n} with rank rr. As such, it plays a special role in relation to free probability, where projections are an important type of noncommutative random variable.

Some asymptotic results concerning Jacobi polynomials can be found in [MFMP24] and its references.

3. Even polynomials in finite free probability

In this section we will study in detail the basic properties of even polynomials, with special emphasis on their behavior under finite free convolutions.

Definition 3.1.

We say that a polynomial p𝒫np\in\mathcal{P}_{n} is even if one of the following equivalent statements holds:

  1. (1)

    for every root α\alpha of pp, there is a root β\beta of pp with β=α\beta=-\alpha and mult(β)=mult(α)\mathrm{mult}(\beta)=\mathrm{mult}(\alpha);

  2. (2)

    pp is an even function if nn is even, or pp is an odd function if nn is odd;

  3. (3)

    𝖾k(p)=0\mathsf{e}_{k}(p)=0 for all odd 0kn0\leq k\leq n;

  4. (4)

    p=Dil1(p)p=\mathrm{Dil}_{-1}(p).

We will denote by 𝒫2mE\mathcal{P}_{2m}^{E} the set of all even polynomials of even degree 2m2m.

The equivalence between (1), (2) and (3) above is well known. The equivalence of (3) and (4) follows from noticing that Dil1\mathrm{Dil}_{-1} simply changes the sign of the roots.

Remark 3.2 (Even polynomials of odd degree).

To simplify the presentation, throughout this paper we will focus on studying even polynomials of even degree. However, the reader should keep in mind that the case of odd degree is completely analogous to the case of even degree, except that we need to add a root at 0. Indeed, notice that if pp is an even polynomial of odd degree, then pp must have a root at 0, implying that it is of the form p(x)=xq(x)p(x)=xq(x) where qq is an even polynomial of even degree.

Remark 3.3.

From Definition 2.9, it is immediate that the set 𝒫2mE\mathcal{P}_{2m}^{E} is closed under 2m\boxplus_{2m} and 2m\boxtimes_{2m}. Furthermore, 𝒫2mE\mathcal{P}_{2m}^{E} is absorbing with respect to multiplicative convolution. These claims can be nicely summarized as follows:

𝒫2mE2m𝒫2mE=𝒫2mEand𝒫2mE2m𝒫2m=𝒫2mE.\mathcal{P}_{2m}^{E}\boxplus_{2m}\mathcal{P}^{E}_{2m}=\mathcal{P}_{2m}^{E}\qquad\text{and}\qquad\mathcal{P}_{2m}^{E}\boxtimes_{2m}\mathcal{P}_{2m}=\mathcal{P}_{2m}^{E}.

3.1. Degree doubling operation

A very simple way to construct even polynomials is by squaring the dummy variable. This is a very natural operation, and has appeared in the context of finite free probability [MSS22, GM22]. Since we will extensively use this operation and its inverse we will fix some notation.

Notation 3.4 (Degree doubling operation).

Define 𝑺m:𝒫m𝒫2mE\bm{S}_{m}:\mathcal{P}_{m}\to\mathcal{P}_{2m}^{E} by

𝑺m(p)=p(x2)\bm{S}_{m}(p)=p(x^{2})

for p𝒫mp\in\mathcal{P}_{m}.

Notation 3.5 (Even and odd parts).

Define 𝑸m:𝒫2m𝒫m\bm{Q}_{m}:\mathcal{P}_{2m}\to\mathcal{P}_{m} as follows: for p𝒫2mp\in\mathcal{P}_{2m}, define 𝑸m(p)\bm{Q}_{m}(p) by the coefficients

𝖾k(𝑸m(p))=(1)k𝖾2k(p).\mathsf{e}_{k}(\bm{Q}_{m}(p))=(-1)^{k}\mathsf{e}_{2k}(p)\text{.} (23)

Notice that the operations 𝑺m\bm{S}_{m} and 𝑸m\bm{Q}_{m} are linear, and for p𝒫mp\in\mathcal{P}_{m}, the roots of 𝑺m(p)\bm{S}_{m}(p) are

λ1(p),λ1(p),λ2(p),λ2(p),,λm(p),λm(p).-\sqrt{\lambda_{1}(p)},\sqrt{\lambda_{1}(p)},-\sqrt{\lambda_{2}(p)},\sqrt{\lambda_{2}(p)},\ldots,-\sqrt{\lambda_{m}(p)},\sqrt{\lambda_{m}(p)}. (24)

Another simple observation is that 𝑸m𝑺m:𝒫m𝒫m\bm{Q}_{m}\circ\bm{S}_{m}:\mathcal{P}_{m}\to\mathcal{P}_{m} is just the identity map, whereas the map 𝑺m𝑸m:𝒫2m𝒫2mE\bm{S}_{m}\circ\bm{Q}_{m}:\mathcal{P}_{2m}\to\mathcal{P}_{2m}^{E} yields an even polynomial that has the same even coefficients as the original polynomial. These observations provide a bijection that we will use constantly throughout this paper:

Fact 3.6.

𝑺m\bm{S}_{m} restricts to a bijection 𝒫m(0)𝒫2mE()\mathcal{P}_{m}(\mathbb{R}_{\geq 0})\to\mathcal{P}_{2m}^{E}(\mathbb{R}), and the inverse of 𝐒m\bm{S}_{m} is 𝐐m\bm{Q}_{m}.

Notice that is a finite free analogue of Lemma 2.5. In Proposition 6.1 we will check that this bijection behaves well with respect to limits of empirical root distributions.

3.2. Symmetrization

Another way to construct even polynomials is by taking the additive convolution of p(x)p(x) with p(x)p(-x), yielding an even polynomial of the same degree. This operation appeared naturally in [Cam22] when studying commutators in the context of finite free probability.

Notation 3.7.

For p𝒫np\in\mathcal{P}_{n}, the symmetrization of pp is the polynomial

Sym(p):=pn(Dil1p).\mathrm{Sym}\left(p\right):=p\boxplus_{n}\left(\mathrm{Dil}_{-1}p\right)\text{.} (25)

It follows directly from the definition that Sym(p)=Sym(Dil1p)\mathrm{Sym}\left(p\right)=\mathrm{Sym}\left(\mathrm{Dil}_{-1}p\right). Notice also that if p𝒫nEp\in\mathcal{P}_{n}^{E}, then Sym(p)=pnp\mathrm{Sym}\left(p\right)=p\boxplus_{n}p. In the following lemma we collect more properties of the symmetrization:

Lemma 3.8.

Let p,q𝒫np,q\in\mathcal{P}_{n} and α\alpha\in\mathbb{C}, and write δα:=(xα)n\delta_{\alpha}:=(x-\alpha)^{n}. Then

  1. (1)

    Sym(p)𝒫nE\mathrm{Sym}\left(p\right)\in\mathcal{P}_{n}^{E};

  2. (2)

    Sym(pnq)=Sym(p)nSym(q)\mathrm{Sym}\left(p\boxplus_{n}q\right)=\mathrm{Sym}\left(p\right)\boxplus_{n}\mathrm{Sym}\left(q\right);

  3. (3)

    Sym(Dilαp)=DilαSym(p)\mathrm{Sym}\left(\mathrm{Dil}_{\alpha}p\right)=\mathrm{Dil}_{\alpha}\mathrm{Sym}\left(p\right);

  4. (4)

    Sym(pnδα)=Sym(p)\mathrm{Sym}\left(p\boxplus_{n}\delta_{\alpha}\right)=\mathrm{Sym}\left(p\right);

  5. (5)

    if p𝒫2m()p\in\mathcal{P}_{2m}(\mathbb{R}), then 𝑸m(Sym(p))𝒫m(0)\bm{Q}_{m}(\mathrm{Sym}\left(p\right))\in\mathcal{P}_{m}(\mathbb{R}_{\geq 0}).

Proof.

The proof of (1) follows from the observation that

Dil1Sym(p)=Dil1(pnDil1p)=(Dil1p)np=Sym(p).\mathrm{Dil}_{-1}\mathrm{Sym}\left(p\right)=\mathrm{Dil}_{-1}\left(p\boxplus_{n}\mathrm{Dil}_{-1}p\right)=\left(\mathrm{Dil}_{-1}p\right)\boxplus_{n}p=\mathrm{Sym}\left(p\right)\text{.}

Alternatively, one can use the formula for the coefficients of the additive convolution and notice that the negative signs of Dil1p\mathrm{Dil}_{-1}p will generate cancellations, causing the odd coefficients of Sym(p)\mathrm{Sym}\left(p\right) to vanish.

For (2), observe that both sides are equal to

pnDil1pnqnDil1q.p\boxplus_{n}\mathrm{Dil}_{-1}p\boxplus_{n}q\boxplus_{n}\mathrm{Dil}_{-1}q\text{.}

Part (3) is a direct consequence of the fact that dilation operation distributes over additive convolutions. For the proof of (4), first notice that

Sym(δα)=δαnδα=δ0.\mathrm{Sym}\left(\delta_{\alpha}\right)=\delta_{\alpha}\boxplus_{n}\delta_{-\alpha}=\delta_{0}\text{.}

So by (2), we have Sym(pnδα)=Sym(p)\mathrm{Sym}\left(p\boxplus_{n}\delta_{\alpha}\right)=\mathrm{Sym}\left(p\right). Finally, (5) follows from (1) and 3.6. ∎

3.3. Multiplicative convolution

The degree doubling operation behaves well with respect to multiplicative convolution:

Proposition 3.9.

For p,q𝒫2mp,q\in\mathcal{P}_{2m}, we have

𝑸m(p2mq)=𝑸m(p)m𝑸m(q)mm[.12m112m.].\bm{Q}_{m}(p\boxtimes_{2m}q)=\bm{Q}_{m}(p)\boxtimes_{m}\bm{Q}_{m}(q)\boxtimes_{m}\mathcal{H}_{m}{\left[\genfrac{.}{.}{0.0pt}{1}{-\frac{1}{2m}}{1-\frac{1}{2m}}\right]}\text{.} (26)

Equivalently, we can express this in terms of the degree doubling operation: for p,q𝒫mp,q\in\mathcal{P}_{m}, we have

𝑺m(pmq)=𝑺m(p)2m𝑺m(q)2m𝑺m(m[.112m, 112m12m,12m.]).\bm{S}_{m}(p\boxtimes_{m}q)=\bm{S}_{m}(p)\boxtimes_{2m}\bm{S}_{m}(q)\boxtimes_{2m}\bm{S}_{m}\left(\mathcal{H}_{m}{\left[\genfrac{.}{.}{0.0pt}{1}{1-\frac{1}{2m},\,1-\frac{1}{2m}}{-\frac{1}{2m},\,-\frac{1}{2m}}\right]}\right)\text{.} (27)
Proof.

To prove Eq. 26, we check the equality for every coefficient: for 0km0\leq k\leq m the coefficient 𝖾k(𝑸m(p2mq))\mathsf{e}_{k}(\bm{Q}_{m}(p\boxtimes_{2m}q)) of the left hand side polynomial is given by

(1)k𝖾2k(p2mq)=(1)k(2m2k)𝖾2k(p)𝖾2k(q)\displaystyle(-1)^{k}\mathsf{e}_{2k}(p\boxtimes_{2m}q)=\frac{(-1)^{k}}{\binom{2m}{2k}}\mathsf{e}_{2k}(p)\mathsf{e}_{2k}(q) (Definition 2.9)
=(1)k(1)k𝖾k(𝑸m(p))(1)k𝖾k(𝑸m(q))(2k)!(2m)2k¯\displaystyle=(-1)^{k}(-1)^{k}\mathsf{e}_{k}(\bm{Q}_{m}(p))(-1)^{k}\mathsf{e}_{k}(\bm{Q}_{m}(q))\frac{(2k)!}{\left(2m\right)^{\underline{2k}}} (Eq. 23)
=𝖾k(𝑸m(p))𝖾k(𝑸m(q))(1)kk!(k12)k¯(m)k¯(m12)k¯\displaystyle=\mathsf{e}_{k}(\bm{Q}_{m}(p))\mathsf{e}_{k}(\bm{Q}_{m}(q))\frac{(-1)^{k}k!\left(k-\frac{1}{2}\right)^{\underline{k}}}{\left(m\right)^{\underline{k}}\left(m-\frac{1}{2}\right)^{\underline{k}}} (Eq. 6)
=𝖾k(𝑸m(p))𝖾k(𝑸m(q))(mk)(mk)(mk)(12)k¯(m12)k¯\displaystyle=\frac{\mathsf{e}_{k}(\bm{Q}_{m}(p))\mathsf{e}_{k}(\bm{Q}_{m}(q))}{\binom{m}{k}\binom{m}{k}}\binom{m}{k}\frac{\left(-\frac{1}{2}\right)^{\underline{k}}}{\left(m-\frac{1}{2}\right)^{\underline{k}}}
=𝖾k(𝑸m(p)m𝑸m(q)mm[.12m112m.]).\displaystyle=\mathsf{e}_{k}\left(\bm{Q}_{m}(p)\boxtimes_{m}\bm{Q}_{m}(q)\boxtimes_{m}\mathcal{H}_{m}{\left[\genfrac{.}{.}{0.0pt}{1}{-\frac{1}{2m}}{1-\frac{1}{2m}}\right]}\right)\text{.} (Definition 2.9)

The proof of Eq. 27 is analogous. ∎

3.4. Additive and rectangular convolution

The effect of taking even parts of finite free convolutions is somewhat more complicated in the additive case. Here, one finds that a variation of the rectangular convolution of Gribinski and Marcus [GM22] plays a key role.

Definition 3.10.

For p,q𝒫mp,q\in\mathcal{P}_{m}, define pm1/2q𝒫mp\boxplus_{m}^{-1/2}q\in\mathcal{P}_{m} by the coefficients

𝖾k(pm1/2q)=(m)k¯(m12)k¯i+j=k𝖾i(p)(m)i¯(m12)i¯𝖾j(q)(m)j¯(m12)j¯.\mathsf{e}_{k}(p\boxplus_{m}^{-1/2}q)=\left(m\right)^{\underline{k}}\left(m-\tfrac{1}{2}\right)^{\underline{k}}\sum_{i+j=k}\frac{\mathsf{e}_{i}(p)}{\left(m\right)^{\underline{i}}\left(m-\frac{1}{2}\right)^{\underline{i}}}\frac{\mathsf{e}_{j}(q)}{\left(m\right)^{\underline{j}}\left(m-\frac{1}{2}\right)^{\underline{j}}}\text{.}
Remark 3.11.

Definition 3.10 can be rephrased in terms of hypergeometric polynomials:

m[.112m.]m(pm1/2q)=(m[.112m.]mp)m(m[.112m.]mq).\mathcal{H}_{m}{\left[\genfrac{.}{.}{0.0pt}{1}{\cdot}{1-\frac{1}{2m}}\right]}\boxtimes_{m}\left(p\boxplus_{m}^{-1/2}q\right)=\left(\mathcal{H}_{m}{\left[\genfrac{.}{.}{0.0pt}{1}{\cdot}{1-\frac{1}{2m}}\right]}\boxtimes_{m}p\right)\boxplus_{m}\left(\mathcal{H}_{m}{\left[\genfrac{.}{.}{0.0pt}{1}{\cdot}{1-\frac{1}{2m}}\right]}\boxtimes_{m}q\right)\text{.}

With this definition in hand we can prove the following relation222We learn this relation, in an online talk by Gribinski at UC Berkeley Probabilistic Operator Algebra Seminar on April 24, 2023. The slides can be found at https://math.berkeley.edu/~jgarzav/Slides_Gribinski.pdf:

Proposition 3.12.

For p,q𝒫2mEp,q\in\mathcal{P}_{2m}^{E}, we have

𝑸m(p2mq)=𝑸m(p)m1/2𝑸m(q).\bm{Q}_{m}(p\boxplus_{2m}q)=\bm{Q}_{m}(p)\boxplus_{m}^{-1/2}\bm{Q}_{m}(q)\text{.} (28)
Proof.

For 0km0\leq k\leq m, we have

𝖾k(m)(𝑸m(p2mq))=(1)k𝖾2k(2m)(p2mq)\displaystyle\mathsf{e}_{k}^{(m)}(\bm{Q}_{m}(p\boxplus_{2m}q))=(-1)^{k}\mathsf{e}_{2k}^{(2m)}(p\boxplus_{2m}q)
=(1)k(2m)2k¯i+j=k𝖾2i(p)(2m)2i¯𝖾2j(q)(2m)2j¯\displaystyle=(-1)^{k}\left(2m\right)^{\underline{2k}}\sum_{i+j=k}\frac{\mathsf{e}_{2i}(p)}{\left(2m\right)^{\underline{2i}}}\frac{\mathsf{e}_{2j}(q)}{\left(2m\right)^{\underline{2j}}}
=(1)k4k(m)k¯(m12)k¯i+j=k(1)i𝖾i(𝑸m(p))4i(m)i¯(m12)i¯(1)j𝖾j(𝑸m(q))4j(m)j¯(m12)j¯\displaystyle=(-1)^{k}4^{k}\left(m\right)^{\underline{k}}\left(m-\tfrac{1}{2}\right)^{\underline{k}}\sum_{i+j=k}\frac{(-1)^{i}\mathsf{e}_{i}(\bm{Q}_{m}(p))}{4^{i}\left(m\right)^{\underline{i}}\left(m-\frac{1}{2}\right)^{\underline{i}}}\frac{(-1)^{j}\mathsf{e}_{j}(\bm{Q}_{m}(q))}{4^{j}\left(m\right)^{\underline{j}}\left(m-\frac{1}{2}\right)^{\underline{j}}}
=𝖾k(𝑸m(p)m1/2𝑸m(q)).\displaystyle=\mathsf{e}_{k}\left(\bm{Q}_{m}(p)\boxplus_{m}^{-1/2}\bm{Q}_{m}(q)\right)\text{.}

Since the coefficients match, the polynomials are the same. ∎

The previous relation implies that the set 𝒫m(0)\mathcal{P}_{m}(\mathbb{R}_{\geq 0}) is closed under the (m,12)(m,-\frac{1}{2}) rectangular convolution from Definition 3.10.

Corollary 3.13.

If p,q𝒫m(0)p,q\in\mathcal{P}_{m}(\mathbb{R}_{\geq 0}), then pm1/2q𝒫m(0)p\boxplus_{m}^{-1/2}q\in\mathcal{P}_{m}(\mathbb{R}_{\geq 0}).

Proof.

Using (28), we can write

pm1/2q=𝑸m(𝑺m(p)2m𝑺m(q)).p\boxplus_{m}^{-1/2}q=\bm{Q}_{m}(\bm{S}_{m}(p)\boxplus_{2m}\bm{S}_{m}(q))\text{.}

Since p,q𝒫m(0)p,q\in\mathcal{P}_{m}(\mathbb{R}_{\geq 0}), we have 𝑺m(p),𝑺m(q)𝒫2mE()\bm{S}_{m}(p),\bm{S}_{m}(q)\in\mathcal{P}_{2m}^{E}(\mathbb{R}) and 𝑺m(p)2m𝑺m(q)𝒫2mE()\bm{S}_{m}(p)\boxplus_{2m}\bm{S}_{m}(q)\in\mathcal{P}_{2m}^{E}(\mathbb{R}). Therefore, after applying the degree halving operation we conclude that pm1/2q=𝑸m(𝑺m(p)2m𝑺m(q))𝒫m(0).p\boxplus_{m}^{-1/2}q=\bm{Q}_{m}(\bm{S}_{m}(p)\boxplus_{2m}\bm{S}_{m}(q))\in\mathcal{P}_{m}(\mathbb{R}_{\geq 0}).

4. Even hypergeometric polynomials

The purpose of this section is to study the specifics of how even hypergeometric polynomials interact with finite free convolution. Since this large class of polynomials contains several regions of parameters where the polynomials have all real, positive, or negative roots, understanding the multiplicative and additive convolutions in these cases will provide us with a large sample of even polynomials.

Our approach resembles that of [MFMP24], with the difference that we want to study hypergeometric polynomials with the variable x2x^{2} rather than xx. This requires some adjustment to the convolution formulas, because of the dependence on the degree, which is now 2m2m instead of mm.

Notation 4.1 (Even hypergeometric polynomials).

Given i,j,mi,j,m\in\mathbb{N}, 𝒂=(a1,,ai)i\bm{a}=(a_{1},\dots,a_{i})\in\mathbb{R}^{i} and 𝒃=(b1,,bj)j\bm{b}=(b_{1},\dots,b_{j})\in\mathbb{R}^{j}, write

mE[.𝒃𝒂.]:=𝑺m(m[.𝒃𝒂.])\mathcal{H}^{E}_{m}{\left[\genfrac{.}{.}{0.0pt}{1}{\bm{b}}{\bm{a}}\right]}:=\bm{S}_{m}\left(\mathcal{H}_{m}{\left[\genfrac{.}{.}{0.0pt}{1}{\bm{b}}{\bm{a}}\right]}\right)

where 𝑺m\bm{S}_{m} is the bijection from 3.5. In terms of coefficients, we have

𝖾2k(mE[.𝒃𝒂.])\displaystyle\mathsf{e}_{2k}\left(\mathcal{H}^{E}_{m}{\left[\genfrac{.}{.}{0.0pt}{1}{\bm{b}}{\bm{a}}\right]}\right) =(1)k𝖾k(m[.𝒃𝒂.])\displaystyle=(-1)^{k}\mathsf{e}_{k}\left(\mathcal{H}_{m}{\left[\genfrac{.}{.}{0.0pt}{1}{\bm{b}}{\bm{a}}\right]}\right)
=(1)k(mk)(𝒃m)k¯(𝒂m)k¯\displaystyle=(-1)^{k}\binom{m}{k}\frac{\left(\bm{b}m\right)^{\underline{k}}}{\left(\bm{a}m\right)^{\underline{k}}}
=(1)k(2m2k)(k12)k¯(𝒃m)k¯(m12)k¯(𝒂m)k¯\displaystyle=(-1)^{k}\binom{2m}{2k}\frac{\left(k-\frac{1}{2}\right)^{\underline{k}}\left(\bm{b}m\right)^{\underline{k}}}{\left(m-\frac{1}{2}\right)^{\underline{k}}\left(\bm{a}m\right)^{\underline{k}}}

for 0km0\leq k\leq m.

Example 4.2 (Bernoulli polynomials).

The simplest even hypergeometric polynomial is

B2m(x)=mE[..](x)=(x21)m.B_{2m}(x)=\mathcal{H}^{E}_{m}{\left[\genfrac{.}{.}{0.0pt}{1}{\cdot}{\cdot}\right]}(x)=(x^{2}-1)^{m}\text{.}

We will call this a Bernoulli polynomial because its empirical root distribution is a Bernoulli distribution with equal weights at 11 and 1-1.

Example 4.3 (Hermite polynomials).

Another important sequence of even hypergeometric polynomials is the one we encountered in Example 2.21:

H2m=Dil1mmE[.112m.].H_{2m}=\mathrm{Dil}_{\sqrt{\frac{1}{m}}}\mathcal{H}^{E}_{m}{\left[\genfrac{.}{.}{0.0pt}{1}{1-\frac{1}{2m}}{\cdot}\right]}\text{.}

4.1. Preliminary results

To study the convolution of even hypergeometric polynomials we first need to understand how the polynomials look in differential form.

A formula to write polynomials m[.𝒃𝒂.]\mathcal{H}_{m}{\left[\genfrac{.}{.}{0.0pt}{1}{\bm{b}}{\bm{a}}\right]} as differential operators was implicitly found in [MFMP24]. The formula to write mE[.𝒃𝒂.]\mathcal{H}^{E}_{m}{\left[\genfrac{.}{.}{0.0pt}{1}{\bm{b}}{\bm{a}}\right]} as a differential operator can be derived in a similar way and it can be generalized for larger powers. We will first prove a general lemma for an arbitrary power and then we specialize to the cases we are concerned.

Lemma 4.4.

Given a constant cc\in\mathbb{R}, integers i,j,m,li,j,m,l\in\mathbb{N}, and tuples of parameters 𝐚=(a1,,ai)i\bm{a}=(a_{1},\dots,a_{i})\in\mathbb{R}^{i} and 𝐛=(b1,,bj)j\bm{b}=(b_{1},\dots,b_{j})\in\mathbb{R}^{j}, if a polynomial has the following differential form:

p(x)=jFi(.m𝒃m𝒂.;c(x)l)xlm,p(x)={\ }_{j}F_{i}{\left(\genfrac{.}{.}{0.0pt}{}{-m\bm{b}}{-m\bm{a}};c\left(\tfrac{\partial}{\partial x}\right)^{l}\right)}x^{lm}\text{,}

then we can express it as the following hypergeometric polynomial:

p(x)=((1)i+j+1llc)mm[.𝒃, 11lm, 12lm,, 1l1lm𝒂.]((1)i+j+1llcxl).p(x)=\left((-1)^{i+j+1}l^{l}c\right)^{m}\mathcal{H}_{m}{\left[\genfrac{.}{.}{0.0pt}{1}{\bm{b},\,1-\tfrac{1}{lm},\,1-\tfrac{2}{lm},\,\ldots,\,1-\tfrac{l-1}{lm}}{\bm{a}}\right]}\left(\frac{(-1)^{i+j+1}}{l^{l}c}x^{l}\right)\text{.}
Proof.

We have

jFi(.𝒃m𝒂m.;c(x)l)xlm\displaystyle{\ }_{j}F_{i}{\left(\genfrac{.}{.}{0.0pt}{}{-\bm{b}m}{-\bm{a}m};c\left(\tfrac{\partial}{\partial x}\right)^{l}\right)}x^{lm} =k=0(m𝒃)k¯(m𝒂)k¯ckk!(x)lkxlm\displaystyle=\sum_{k=0}^{\infty}\frac{\left(-m\bm{b}\right)^{\overline{k}}}{\left(-m\bm{a}\right)^{\overline{k}}}\frac{c^{k}}{k!}\left(\tfrac{\partial}{\partial x}\right)^{lk}x^{lm}
=k=0m(1)(i+j)k(m𝒃)k¯(lm)lk¯(m𝒂)k¯ckk!xlmlk.\displaystyle=\sum_{k=0}^{m}(-1)^{(i+j)k}\frac{\left(m\bm{b}\right)^{\underline{k}}\left(lm\right)^{\underline{lk}}}{\left(m\bm{a}\right)^{\underline{k}}}\frac{c^{k}}{k!}x^{lm-lk}\text{.}

If ss is not a multiple of ll, then 𝖾s(p)=0\mathsf{e}_{s}(p)=0. Otherwise, we have

(1)lk𝖾lk(p)\displaystyle\quad(-1)^{lk}\mathsf{e}_{lk}(p)
=(m𝒃)k¯(m𝒂)k¯(1)(i+j)kckk!(lm)lk¯\displaystyle=\frac{\left(m\bm{b}\right)^{\underline{k}}}{\left(m\bm{a}\right)^{\underline{k}}}(-1)^{(i+j)k}\frac{c^{k}}{k!}\left(lm\right)^{\underline{lk}}
=(m𝒃)k¯(m𝒂)k¯ckk!(1)(i+j)kllk(m)k¯(m1l)k¯(ml1l)k¯\displaystyle=\frac{\left(m\bm{b}\right)^{\underline{k}}}{\left(m\bm{a}\right)^{\underline{k}}}\frac{c^{k}}{k!}(-1)^{(i+j)k}l^{lk}\left(m\right)^{\underline{k}}\left(m-\tfrac{1}{l}\right)^{\underline{k}}\cdots\left(m-\tfrac{l-1}{l}\right)^{\underline{k}}
=((1)i+jllc)m(mk)(m𝒃)k¯(m1l)k¯(ml1l)k¯(m𝒂)k¯((1)i+jllc)mk.\displaystyle=\left((-1)^{i+j}l^{l}c\right)^{m}\binom{m}{k}\frac{\left(m\bm{b}\right)^{\underline{k}}\left(m-\tfrac{1}{l}\right)^{\underline{k}}\cdots\left(m-\tfrac{l-1}{l}\right)^{\underline{k}}}{\left(m\bm{a}\right)^{\underline{k}}}\left(\frac{(-1)^{i+j}}{l^{l}c}\right)^{m-k}\text{.}

So we have proven the claim:

p(x)=((1)i+j+1llc)mm[.𝒃, 11lm, 12lm,, 1l1lm𝒂.]((1)i+j+1llcxl).p(x)=\left((-1)^{i+j+1}l^{l}c\right)^{m}\mathcal{H}_{m}{\left[\genfrac{.}{.}{0.0pt}{1}{\bm{b},\,1-\tfrac{1}{lm},\,1-\tfrac{2}{lm},\,\ldots,\,1-\tfrac{l-1}{lm}}{\bm{a}}\right]}\left(\frac{(-1)^{i+j+1}}{l^{l}c}x^{l}\right)\text{.}

As particular cases, we obtain the following:

Corollary 4.5.

With the assumptions of Lemma 4.4, in the case l=1l=1 we have

jFi(.m𝒃m𝒂.;cx)xm=Dilc(1)i+j+1m[.𝒃𝒂.](x).{\ }_{j}F_{i}{\left(\genfrac{.}{.}{0.0pt}{}{-m\bm{b}}{-m\bm{a}};c\tfrac{\partial}{\partial x}\right)}x^{m}=\mathrm{Dil}_{c(-1)^{i+j+1}}\mathcal{H}_{m}{\left[\genfrac{.}{.}{0.0pt}{1}{\bm{b}}{\bm{a}}\right]}(x)\text{.} (29)

In the case l=2l=2 we have

jFi(.m𝒃m𝒂.;c(x)2)x2m=Dil2c(1)i+j+1mE[.𝒃, 112m𝒂.](x).{\ }_{j}F_{i}{\left(\genfrac{.}{.}{0.0pt}{}{-m\bm{b}}{-m\bm{a}};c\left(\tfrac{\partial}{\partial x}\right)^{2}\right)}x^{2m}=\mathrm{Dil}_{2\sqrt{c(-1)^{i+j+1}}}\mathcal{H}^{E}_{m}{\left[\genfrac{.}{.}{0.0pt}{1}{\bm{b},\ 1-\tfrac{1}{2m}}{\bm{a}}\right]}(x)\text{.} (30)

It is worth emphasizing that in the right hand side of (30), we first double the degree of the hypergeometric polynomial and then dilate it. Notice that the same polynomial can be obtained by first dilating (by the square of the constant) and then doubling the degree.

With these formulas in hand we can now readily generalize the last part of Theorem 2.17 and relate the product of hypergeometric series (evaluated in any power of xx) to the additive convolution of hypergeometric polynomials (evaluated on the corresponding powers of xx). The idea is to use the definition of n\boxplus_{n} in terms of differential operators and the fact that we just proved that differential operators on hypergeometric series applied to xnx^{n} yield hypergeometric polynomials. We first provide the result in its more general form, and then specialize to the cases that we are more interested in.

Theorem 4.6 (Additive convolution of hypergeometric polynomials).

Let c1,c2,c3c_{1},c_{2},c_{3}\in\mathbb{R} be constants, and let l1,l2,l3,nl_{1},l_{2},l_{3},n\in\mathbb{N} be numbers such that lkl_{k} divides nn for k=1,2,3k=1,2,3. Consider tuples 𝐚1,𝐚2,𝐚3,𝐛1,𝐛2,𝐛3\bm{a}_{1},\bm{a}_{2},\bm{a}_{3},\bm{b}_{1},\bm{b}_{2},\bm{b}_{3} of sizes i1,i2,i3,j1,j2,j3i_{1},i_{2},i_{3},j_{1},j_{2},j_{3}\in\mathbb{N}, and assume that

j1Fi1(.n𝒃1n𝒂1.;c1xl1)j2Fi2(.n𝒃2n𝒂2.;c2xl2)=j3Fi3(.n𝒃3n𝒂3.;c3xl3).{\ }_{j_{1}}F_{i_{1}}{\left(\genfrac{.}{.}{0.0pt}{}{-n\bm{b}_{1}}{-n\bm{a}_{1}};c_{1}x^{l_{1}}\right)}{\ }_{j_{2}}F_{i_{2}}{\left(\genfrac{.}{.}{0.0pt}{}{-n\bm{b}_{2}}{-n\bm{a}_{2}};c_{2}x^{l_{2}}\right)}={\ }_{j_{3}}F_{i_{3}}{\left(\genfrac{.}{.}{0.0pt}{}{-n\bm{b}_{3}}{-n\bm{a}_{3}};c_{3}x^{l_{3}}\right)}\text{.}

Then, if for k=1,2,3k=1,2,3 we consider the polynomials

pk(x)=((1)ik+jk+1lklkck)nlknlk[.lk𝒃, 11n,, 1lk1nlk𝒂.]((1)ik+jk+1lklkckxlk)p_{k}(x)=\left((-1)^{i_{k}+j_{k}+1}l_{k}^{l_{k}}c_{k}\right)^{\frac{n}{l_{k}}}\mathcal{H}_{\frac{n}{l_{k}}}{\left[\genfrac{.}{.}{0.0pt}{1}{l_{k}\bm{b},\,1-\tfrac{1}{n},\,\ldots,\,1-\tfrac{l_{k}-1}{n}}{l_{k}\bm{a}}\right]}\left(\frac{(-1)^{i_{k}+j_{k}+1}}{l_{k}^{l_{k}}c_{k}}x^{l_{k}}\right)

we get that p1np2=p3p_{1}\boxplus_{n}p_{2}=p_{3}.

Proof.

Fix k{1,2,3}k\in\{1,2,3\}. From Lemma 4.4 applied to ckc_{k}\in\mathbb{R}, integer values ik,jk,nlk,lki_{k},j_{k},\frac{n}{l_{k}},l_{k}\in\mathbb{N}, and tuples of parameters lk𝒂kil_{k}\bm{a}_{k}\in\mathbb{R}^{i} and lk𝒃kjl_{k}\bm{b}_{k}\in\mathbb{R}^{j}, we know that a polynomial written differential form

jkFik(.nlklk𝒃knlklk𝒂k.;ck(x)lk)xn{\ }_{j_{k}}F_{i_{k}}{\left(\genfrac{.}{.}{0.0pt}{}{-\frac{n}{l_{k}}l_{k}\bm{b}_{k}}{-\frac{n}{l_{k}}l_{k}\bm{a}_{k}};c_{k}\left(\tfrac{\partial}{\partial x}\right)^{l_{k}}\right)}x^{n}

is precisely the polynomial

pk(x)=((1)ik+jk+1lklkc)nlknlk[.lk𝒃k, 11n,, 1lk1nlk𝒂k.]((1)ik+jk+1lklkckxlk).p_{k}(x)=\left((-1)^{i_{k}+j_{k}+1}l_{k}^{l_{k}}c\right)^{\frac{n}{l_{k}}}\mathcal{H}_{\frac{n}{l_{k}}}{\left[\genfrac{.}{.}{0.0pt}{1}{l_{k}\bm{b}_{k},\ 1-\tfrac{1}{n},\ \ldots\ ,\ 1-\tfrac{l_{k}-1}{n}}{l_{k}\bm{a}_{k}}\right]}\left(\frac{(-1)^{i_{k}+j_{k}+1}}{l_{k}^{l_{k}}c_{k}}x^{l_{k}}\right)\text{.}

Then the result follows from the definition of additive convolution using differential operators. ∎

4.2. Symmetrization

With the results from last section in hand, we are ready to study the symmetrization of some hypergeometric polynomials using some well-known results of products of hypergeometric functions.

Lemma 4.7.

Consider tuples 𝐚,𝐚,𝐛,𝐛\bm{a},\bm{a}^{\prime},\bm{b},\bm{b}^{\prime} of sizes i,i,j,ji,i^{\prime},j,j^{\prime}, and assume that

jFi(.2m𝒃2m𝒂.;x)jFi(.2m𝒃2m𝒂.;x)=jFi(.m𝒃m𝒂.;cx2).{\ }_{j}F_{i}{\left(\genfrac{.}{.}{0.0pt}{}{-2m\bm{b}}{-2m\bm{a}};x\right)}{\ }_{j}F_{i}{\left(\genfrac{.}{.}{0.0pt}{}{-2m\bm{b}}{-2m\bm{a}};-x\right)}={\ }_{j^{\prime}}F_{i^{\prime}}{\left(\genfrac{.}{.}{0.0pt}{}{-m\bm{b}^{\prime}}{-m\bm{a}^{\prime}};cx^{2}\right)}\text{.} (31)

Then we have

Sym(2m[.𝒃𝒂.])=Dil2c(1)i+j+1mE[.𝒃, 112m𝒂.].\mathrm{Sym}\left(\mathcal{H}_{2m}{\left[\genfrac{.}{.}{0.0pt}{1}{\bm{b}}{\bm{a}}\right]}\right)=\mathrm{Dil}_{2\sqrt{c(-1)^{i^{\prime}+j^{\prime}+1}}}\mathcal{H}^{E}_{m}{\left[\genfrac{.}{.}{0.0pt}{1}{\bm{b}^{\prime},\,1-\tfrac{1}{2m}}{\bm{a}^{\prime}}\right]}\text{.} (32)
Proof.

The result follows from applying Theorem 4.6 to the particular case where n=2mn=2m, l1=l2=1l_{1}=l_{2}=1, l3=2l_{3}=2, c1=1c_{1}=1, c2=1c_{2}=-1, c3=cc_{3}=c, 𝒂1=𝒂2=𝒂\bm{a}_{1}=\bm{a}_{2}=\bm{a}, and 𝒃1=𝒃2=𝒃\bm{b}_{1}=\bm{b}_{2}=\bm{b}. ∎

Using this result, we can compute the symmetrization of certain hypergeometric polynomials using product identities for hypergeometric series of the form (31). Some of these formulas are elementary, like the product of binomial functions or the product of Bessel functions, while more involved ones can be found in works of Ramanujan, Preece, and Bailey. We use Grinshpan’s survey [Gri13] as a convenient reference. Specifically, in Proposition 4.8 we reproduce Equations (9), (19), (21), (20), and (23) from [Gri13] as Eq. 33Eq. 37 respectively.

Proposition 4.8 (Product of hypergeometric series).

Given real parameters m,a,b,c,dm,a,b,c,d, the following identities hold:

1F0(.2mb.;x)1F0(.2mb.;x)=1F0(.2mb.;x2);{\ }_{1}F_{0}{\left(\genfrac{.}{.}{0.0pt}{}{-2mb}{\cdot};x\right)}{\ }_{1}F_{0}{\left(\genfrac{.}{.}{0.0pt}{}{-2mb}{\cdot};-x\right)}={\ }_{1}F_{0}{\left(\genfrac{.}{.}{0.0pt}{}{-2mb}{\cdot};x^{2}\right)}\text{;} (33)
0F1(.2ma.;x)0F1(.2ma.;x){\ }_{0}F_{1}{\left(\genfrac{.}{.}{0.0pt}{}{\cdot}{-2ma};x\right)}{\ }_{0}F_{1}{\left(\genfrac{.}{.}{0.0pt}{}{\cdot}{-2ma};-x\right)}
=0F3(.2ma,ma,ma+12.;x24);={\ }_{0}F_{3}{\left(\genfrac{.}{.}{0.0pt}{}{\cdot}{-2ma,\ -ma,\ -ma+\frac{1}{2}};-\frac{x^{2}}{4}\right)}\text{;} (34)
1F1(.2mb2ma.;x)1F1(.2mb2ma.;x){\ }_{1}F_{1}{\left(\genfrac{.}{.}{0.0pt}{}{-2mb}{-2ma};x\right)}{\ }_{1}F_{1}{\left(\genfrac{.}{.}{0.0pt}{}{-2mb}{-2ma};-x\right)}
=2F3(.2mb,2ma+2mb2ma,ma,ma+12.;x24);={\ }_{2}F_{3}{\left(\genfrac{.}{.}{0.0pt}{}{-2mb,\ -2ma+2mb}{-2ma,\ -ma,\ -ma+\frac{1}{2}};\frac{x^{2}}{4}\right)}\text{;} (35)
2F0(.2mb,2md.;x)2F0(.2mb,2md.;x){\ }_{2}F_{0}{\left(\genfrac{.}{.}{0.0pt}{}{-2mb,\ -2md}{\cdot};x\right)}{\ }_{2}F_{0}{\left(\genfrac{.}{.}{0.0pt}{}{-2mb,\ -2md}{\cdot};-x\right)}
=4F1(.2mb,2md,m(b+d),m(b+d)+122mb2md.;4x2);={\ }_{4}F_{1}{\left(\genfrac{.}{.}{0.0pt}{}{-2mb,\ -2md,\ -m(b+d),\ -m(b+d)+\frac{1}{2}}{-2mb-2md};4x^{2}\right)}\text{;} (36)
0F2(.a,c.;x)0F2(.a,c.;x){\ }_{0}F_{2}{\left(\genfrac{.}{.}{0.0pt}{}{\cdot}{a,\ c};x\right)}{\ }_{0}F_{2}{\left(\genfrac{.}{.}{0.0pt}{}{\cdot}{a,\ c};-x\right)}
=3F8(.a+c13,a+c3,a+c+13a,c,a2,c2,a+12,c+12,a+c12,a+c2.;27x264).={\ }_{3}F_{8}{\left(\genfrac{.}{.}{0.0pt}{}{\frac{a+c-1}{3},\ \frac{a+c}{3},\ \frac{a+c+1}{3}}{a,\ c,\ \frac{a}{2},\ \frac{c}{2},\ \frac{a+1}{2},\ \frac{c+1}{2},\ \frac{a+c-1}{2},\ \frac{a+c}{2}};-\frac{27x^{2}}{64}\right)}\text{.} (37)

Using Proposition 4.8 and Lemma 4.7 we can compute the symmetrizations of various classical polynomials, such as Laguerre and Bessel polynomials, as well as their multiplicative convolutions. We collect these results in Table 1.

Polynomial pp Sym(p)\mathrm{Sym}\left(p\right)
Laguerre 2m[.b.]\mathcal{H}_{2m}{\left[\genfrac{.}{.}{0.0pt}{1}{b}{\cdot}\right]} Dil2mE[.2b, 112m.]\mathrm{Dil}_{2}\mathcal{H}^{E}_{m}{\left[\genfrac{.}{.}{0.0pt}{1}{2b,\,1-\frac{1}{2m}}{\cdot}\right]}
Bessel 2m[.a.]\mathcal{H}_{2m}{\left[\genfrac{.}{.}{0.0pt}{1}{\cdot}{a}\right]} DilimE[.112m2a,a,a12m.]\mathrm{Dil}_{i}\mathcal{H}^{E}_{m}{\left[\genfrac{.}{.}{0.0pt}{1}{1-\frac{1}{2m}}{2a,\,a,\,a-\frac{1}{2m}}\right]}
Jacobi 2m[.ba.]\mathcal{H}_{2m}{\left[\genfrac{.}{.}{0.0pt}{1}{b}{a}\right]} mE[.112m, 2b, 2a2b2a,a,a12m.]\mathcal{H}^{E}_{m}{\left[\genfrac{.}{.}{0.0pt}{1}{1-\frac{1}{2m},\,2b,\,2a-2b}{2a,\,a,\,a-\frac{1}{2m}}\right]}
Lag 2m\boxtimes_{2m} Lag 2m[.b,d.]\mathcal{H}_{2m}{\left[\genfrac{.}{.}{0.0pt}{1}{b,\,d}{\cdot}\right]} Dil4mE[.112m, 2b, 2d,b+d,b+d12m2b+2d.]\mathrm{Dil}_{4}\mathcal{H}^{E}_{m}{\left[\genfrac{.}{.}{0.0pt}{1}{1-\frac{1}{2m},\,2b,\,2d,\,b+d,\,b+d-\frac{1}{2m}}{2b+2d}\right]}
Bes 2m\boxtimes_{2m} Bes 2m[.a,c.]\mathcal{H}_{2m}{\left[\genfrac{.}{.}{0.0pt}{1}{\cdot}{a,\,c}\right]} Dil827mE[.23(a+c+12m),23(a+c),23(a+c12m), 112m2a, 2c,a,c,a12m,c12m,a+c+12m,a+c.]\mathrm{Dil}_{\frac{8}{\sqrt{-27}}}\mathcal{H}^{E}_{m}{\left[\genfrac{.}{.}{0.0pt}{1}{\frac{2}{3}\left(a+c+\frac{1}{2m}\right),\,\frac{2}{3}\left(a+c\right),\,\frac{2}{3}\left(a+c-\frac{1}{2m}\right),\,1-\frac{1}{2m}}{2a,\,2c,\,a,\,c,\,a-\frac{1}{2m},\,c-\frac{1}{2m},\,a+c+\frac{1}{2m},\,a+c}\right]}
Table 1. Symmetrizations of hypergeometric polynomials

We should emphasize that in the last row of Table 1, the case corresponding to the multiplicative convolution of two Bessel polynomials, one must do the change of variable a=2maa^{\prime}=-2ma and b=2bmb^{\prime}=-2bm in Eq. 37 before applying Lemma 4.7. Notice also, that in this case, if we let a+c=32ma+c=\frac{-3}{2m}, the formula simplifies to

Sym(2m[.a,c.])=Dil827mE[.43m, 112m2a, 2c,a,c,a12m,c12m.].\mathrm{Sym}\left(\mathcal{H}_{2m}{\left[\genfrac{.}{.}{0.0pt}{1}{\cdot}{a,\,c}\right]}\right)=\mathrm{Dil}_{\frac{8}{\sqrt{-27}}}\mathcal{H}^{E}_{m}{\left[\genfrac{.}{.}{0.0pt}{1}{\frac{-4}{3m},\,1-\frac{1}{2m}}{2a,\,2c,\,a,\,c,\,a-\frac{1}{2m},\,c-\frac{1}{2m}}\right]}\text{.} (38)

4.3. Multiplicative convolution

The multiplicative convolution of two even hypergeometric polynomials has a very nice expression that follows from Proposition 3.9 and Theorem 2.17.

Proposition 4.9 (Multiplicative convolution of even hypergeometric polynomials).

Consider tuples 𝐚1,𝐚2,𝐛1,𝐛2\bm{a}_{1},\bm{a}_{2},\bm{b}_{1},\bm{b}_{2} of sizes i1,i2,j1,j2i_{1},i_{2},j_{1},j_{2}. Then

mE[.𝒃1𝒂1.]2mmE[.𝒃2𝒂2.]=mE[.12m,𝒃1,𝒃2112m,𝒂1,𝒂2.].\mathcal{H}^{E}_{m}{\left[\genfrac{.}{.}{0.0pt}{1}{\bm{b}_{1}}{\bm{a}_{1}}\right]}\boxtimes_{2m}\mathcal{H}^{E}_{m}{\left[\genfrac{.}{.}{0.0pt}{1}{\bm{b}_{2}}{\bm{a}_{2}}\right]}=\mathcal{H}^{E}_{m}{\left[\genfrac{.}{.}{0.0pt}{1}{-\frac{1}{2m},\,\bm{b}_{1},\,\bm{b}_{2}}{1-\frac{1}{2m},\,\bm{a}_{1},\,\bm{a}_{2}}\right]}\text{.}
Example 4.10 (Multiplicative convolution of two Bernoulli polynomialk).

By Proposition 4.9, the multiplicative convolution of two Bernoulli polynomials B2mB_{2m} from Example 4.2 is given by

B2m2mB2m=mE[.12m112m.].B_{2m}\boxtimes_{2m}B_{2m}=\mathcal{H}^{E}_{m}{\left[\genfrac{.}{.}{0.0pt}{1}{-\frac{1}{2m}}{1-\frac{1}{2m}}\right]}\text{.}

For m>1m>1, this polynomial is not necessarily real-rooted; instead we can conclude that all the roots lie in the unit circle 𝕋\mathbb{T}. Indeed, B2m𝒫2m(𝕋)B_{2m}\in\mathcal{P}_{2m}(\mathbb{T}) and 𝒫2m(𝕋)\mathcal{P}_{2m}(\mathbb{T}) is closed under multiplicative convolution (see [Mar66, Theorem 16.1 and Corollary 16.1a]), hence

B2m2mB2m=mE[.12m112m.]𝒫2m(𝕋).B_{2m}\boxtimes_{2m}B_{2m}=\mathcal{H}^{E}_{m}{\left[\genfrac{.}{.}{0.0pt}{1}{-\frac{1}{2m}}{1-\frac{1}{2m}}\right]}\in\mathcal{P}_{2m}(\mathbb{T})\text{.}
Example 4.11 (Multiplicative convolution of two Hermite polynomials).

The multiplicative convolution of two Hermite polynomials H2mH_{2m} from Example 4.3 is given by

H2m2mH2m=Dil1mmE[.12m, 112m.].H_{2m}\boxtimes_{2m}H_{2m}=\mathrm{Dil}_{\frac{1}{m}}\mathcal{H}^{E}_{m}{\left[\genfrac{.}{.}{0.0pt}{1}{-\frac{1}{2m},\,1-\frac{1}{2m}}{\cdot}\right]}\text{.}

This polynomial is not real-rooted in general.

4.4. Additive convolution

The additive convolution of even hypergeometric polynomials can now be described as a particular instance of Theorem 4.6, where we let all the powers to be squares.

Proposition 4.12 (Additive convolution of even hypergeometric polynomials).

Consider tuples 𝐚1,𝐚2,𝐚3,𝐛1,𝐛2,𝐛3\bm{a}_{1},\bm{a}_{2},\bm{a}_{3},\bm{b}_{1},\bm{b}_{2},\bm{b}_{3} of sizes i1,i2,i3,j1,j2,j3i_{1},i_{2},i_{3},j_{1},j_{2},j_{3}, and assume that

j1Fi1(.m𝒃1m𝒂1.;c1x)j2Fi2(.m𝒃2m𝒂2.;c2x)=j3Fi3(.m𝒃3m𝒂3.;c3x).{\ }_{j_{1}}F_{i_{1}}{\left(\genfrac{.}{.}{0.0pt}{}{-m\bm{b}_{1}}{-m\bm{a}_{1}};c_{1}x\right)}{\ }_{j_{2}}F_{i_{2}}{\left(\genfrac{.}{.}{0.0pt}{}{-m\bm{b}_{2}}{-m\bm{a}_{2}};c_{2}x\right)}={\ }_{j_{3}}F_{i_{3}}{\left(\genfrac{.}{.}{0.0pt}{}{-m\bm{b}_{3}}{-m\bm{a}_{3}};c_{3}x\right)}\text{.} (39)

Then, if we let sk=ck(1)ik+jk+1s_{k}=\sqrt{c_{k}(-1)^{i_{k}+j_{k}+1}} for k=1,2,3k=1,2,3, we have

Dils1mE[.𝒃1, 112m𝒂1.]2mDils2mE[.𝒃2, 112m𝒂2.]=Dils3mE[.𝒃3, 112m𝒂3.].\mathrm{Dil}_{s_{1}}\mathcal{H}^{E}_{m}{\left[\genfrac{.}{.}{0.0pt}{1}{\bm{b}_{1},\,1-\frac{1}{2m}}{\bm{a}_{1}}\right]}\boxplus_{2m}\mathrm{Dil}_{s_{2}}\mathcal{H}^{E}_{m}{\left[\genfrac{.}{.}{0.0pt}{1}{\bm{b}_{2},\,1-\frac{1}{2m}}{\bm{a}_{2}}\right]}=\mathrm{Dil}_{s_{3}}\mathcal{H}^{E}_{m}{\left[\genfrac{.}{.}{0.0pt}{1}{\bm{b}_{3},\,1-\frac{1}{2m}}{\bm{a}_{3}}\right]}\text{.} (40)
Proof.

If we evaluate the assumption (39) in x24\frac{x^{2}}{4} we get

j1Fi1(.m𝒃1m𝒂1.;c1x24)j2Fi2(.m𝒃2m𝒂2.;c2x24)=j3Fi3(.m𝒃3m𝒂3.;c3x24).{\ }_{j_{1}}F_{i_{1}}{\left(\genfrac{.}{.}{0.0pt}{}{-m\bm{b}_{1}}{-m\bm{a}_{1}};c_{1}\frac{x^{2}}{4}\right)}{\ }_{j_{2}}F_{i_{2}}{\left(\genfrac{.}{.}{0.0pt}{}{-m\bm{b}_{2}}{-m\bm{a}_{2}};c_{2}\frac{x^{2}}{4}\right)}={\ }_{j_{3}}F_{i_{3}}{\left(\genfrac{.}{.}{0.0pt}{}{-m\bm{b}_{3}}{-m\bm{a}_{3}};c_{3}\frac{x^{2}}{4}\right)}\text{.}

Then we can use Theorem 4.6 with constants c14,c24,c34\frac{c_{1}}{4},\frac{c_{2}}{4},\frac{c_{3}}{4}\in\mathbb{R}, integers l1=l2=l3=ml_{1}=l_{2}=l_{3}=m and n=2mn=2m, and parameters 12𝒂k,12𝒃k\frac{1}{2}\bm{a}_{k},\frac{1}{2}\bm{b}_{k} for k=1,2,3k=1,2,3. This gives p12mp2=p3p_{1}\boxplus_{2m}p_{2}=p_{3}, where

pk=DilskmE[.𝒃k, 112m𝒂k.].p_{k}=\mathrm{Dil}_{s_{k}}\mathcal{H}^{E}_{m}{\left[\genfrac{.}{.}{0.0pt}{1}{\bm{b}_{k},\,1-\frac{1}{2m}}{\bm{a}_{k}}\right]}\text{.}

Then Eq. 40 follows from scaling the polynomials. ∎

Remark 4.13.

Notice from Proposition 3.12 that the even parts of these polynomials can be related using the rectangular convolution m1/2\boxplus_{m}^{-1/2}.

Similar to Section 4.2, we can compute the additive convolutions of certain hypergeometric polynomials using product identities of hypergeometric series that fit into the form of Eq. 39. We again use Grinshpan’s survey [Gri13]. Specifically, in Proposition 4.14 we reproduce Equations (18), (7), (8), (10), and (11) from [Gri13] as Eq. 41Eq. 45 respectively.

Proposition 4.14 (Product of hypergeometric series).

Given real parameters m,a,a1,a2,b1,b2,c1,c2m,a,a_{1},a_{2},b_{1},b_{2},c_{1},c_{2}, the following identities hold:

0F1(.ma1.;x)0F1(.ma2.;x){\ }_{0}F_{1}{\left(\genfrac{.}{.}{0.0pt}{}{\cdot}{-ma_{1}};x\right)}{\ }_{0}F_{1}{\left(\genfrac{.}{.}{0.0pt}{}{\cdot}{-ma_{2}};x\right)}
=2F3(.m(a1+a2)2,m(a1+a2)12ma1,ma2,m(a1+a2)1.;4x);={\ }_{2}F_{3}{\left(\genfrac{.}{.}{0.0pt}{}{\frac{-m(a_{1}+a_{2})}{2},\frac{-m(a_{1}+a_{2})-1}{2}}{-ma_{1},-ma_{2},-m(a_{1}+a_{2})-1};4x\right)}\text{;} (41)
0F0(..;c1x)0F0(..;c2x)=0F0(..;(c1+c2)x);{\ }_{0}F_{0}{\left(\genfrac{.}{.}{0.0pt}{}{\cdot}{\cdot};-c_{1}x\right)}{\ }_{0}F_{0}{\left(\genfrac{.}{.}{0.0pt}{}{\cdot}{\cdot};-c_{2}x\right)}={\ }_{0}F_{0}{\left(\genfrac{.}{.}{0.0pt}{}{\cdot}{\cdot};-(c_{1}+c_{2})x\right)}\text{;} (42)
1F0(.mb1.;x)1F0(.mb2.;x)=1F0(.m(b1+b2).;x);{\ }_{1}F_{0}{\left(\genfrac{.}{.}{0.0pt}{}{-mb_{1}}{\cdot};x\right)}{\ }_{1}F_{0}{\left(\genfrac{.}{.}{0.0pt}{}{-mb_{2}}{\cdot};x\right)}={\ }_{1}F_{0}{\left(\genfrac{.}{.}{0.0pt}{}{-m(b_{1}+b_{2})}{\cdot};x\right)}\text{;} (43)
1F0(.m(b1+b2a).;x)2F1(.m(ab1),m(ab2)ma.;x){\ }_{1}F_{0}{\left(\genfrac{.}{.}{0.0pt}{}{-m(b_{1}+b_{2}-a)}{\cdot};x\right)}{\ }_{2}F_{1}{\left(\genfrac{.}{.}{0.0pt}{}{-m(a-b_{1}),-m(a-b_{2})}{-ma};x\right)}
=2F1(.mb1,mb2ma.;x);={\ }_{2}F_{1}{\left(\genfrac{.}{.}{0.0pt}{}{-mb_{1},-mb_{2}}{-ma};x\right)}\text{;} (44)
[2F1(.mb1,mb2m(b1+b2)+1/2.;x)]2\left[{\ }_{2}F_{1}{\left(\genfrac{.}{.}{0.0pt}{}{-mb_{1},-mb_{2}}{-m(b_{1}+b_{2})+1/2};x\right)}\right]^{2}
=3F2(.2mb1,m(b1+b2),2mb2m(b1+b2)+12,2m(b1+b2).;x).={\ }_{3}F_{2}{\left(\genfrac{.}{.}{0.0pt}{}{-2mb_{1},-m(b_{1}+b_{2}),-2mb_{2}}{-m(b_{1}+b_{2})+\frac{1}{2},-2m(b_{1}+b_{2})};x\right)}. (45)

Using Proposition 4.14 and Proposition 4.12 we can compute the additive convolutions of some even hypergeometric polynomials; we collect the results in Table 2.

pp qq p2mqp\boxplus_{2m}q
mE[.112ma1.]\mathcal{H}^{E}_{m}{\left[\genfrac{.}{.}{0.0pt}{1}{1-\frac{1}{2m}}{a_{1}}\right]} mE[.112ma2.]\mathcal{H}^{E}_{m}{\left[\genfrac{.}{.}{0.0pt}{1}{1-\frac{1}{2m}}{a_{2}}\right]} Dil2mE[.a1+a22,a1+a22+12m, 112ma1,a2,a1+a2+1m.]\mathrm{Dil}_{2}\mathcal{H}^{E}_{m}{\left[\genfrac{.}{.}{0.0pt}{1}{\frac{a_{1}+a_{2}}{2},\,\frac{a_{1}+a_{2}}{2}+\frac{1}{2m},\,1-\frac{1}{2m}}{a_{1},\,a_{2},\,a_{1}+a_{2}+\frac{1}{m}}\right]}
Dilc1mE[.112m.]\mathrm{Dil}_{\sqrt{c_{1}}}\mathcal{H}^{E}_{m}{\left[\genfrac{.}{.}{0.0pt}{1}{1-\frac{1}{2m}}{\cdot}\right]} Dilc2mE[.112m.]\mathrm{Dil}_{\sqrt{c_{2}}}\mathcal{H}^{E}_{m}{\left[\genfrac{.}{.}{0.0pt}{1}{1-\frac{1}{2m}}{\cdot}\right]} Dilc1+c2mE[.112m.]\mathrm{Dil}_{\sqrt{c_{1}+c_{2}}}\mathcal{H}^{E}_{m}{\left[\genfrac{.}{.}{0.0pt}{1}{1-\frac{1}{2m}}{\cdot}\right]}
mE[.b1, 112m.]\mathcal{H}^{E}_{m}{\left[\genfrac{.}{.}{0.0pt}{1}{b_{1},\,1-\frac{1}{2m}}{\cdot}\right]} mE[.b2, 112m.]\mathcal{H}^{E}_{m}{\left[\genfrac{.}{.}{0.0pt}{1}{b_{2},\,1-\frac{1}{2m}}{\cdot}\right]} mE[.b1+b2, 112m.]\mathcal{H}^{E}_{m}{\left[\genfrac{.}{.}{0.0pt}{1}{b_{1}+b_{2},\,1-\frac{1}{2m}}{\cdot}\right]}
mE[.b1+b2a, 112m.]\mathcal{H}^{E}_{m}{\left[\genfrac{.}{.}{0.0pt}{1}{b_{1}+b_{2}-a,\,1-\frac{1}{2m}}{\cdot}\right]} mE[.ab1,ab2, 112ma.]\mathcal{H}^{E}_{m}{\left[\genfrac{.}{.}{0.0pt}{1}{a-b_{1},\,a-b_{2},\,1-\frac{1}{2m}}{a}\right]} mE[.b1,b2, 112ma.]\mathcal{H}^{E}_{m}{\left[\genfrac{.}{.}{0.0pt}{1}{b_{1},\,b_{2},\,1-\frac{1}{2m}}{a}\right]}
mE[.b1,b2, 112mb1+b212m.]\mathcal{H}^{E}_{m}{\left[\genfrac{.}{.}{0.0pt}{1}{b_{1},\,b_{2},\,1-\frac{1}{2m}}{b_{1}+b_{2}-\frac{1}{2m}}\right]} mE[.b1,b2, 112mb1+b212m.]\mathcal{H}^{E}_{m}{\left[\genfrac{.}{.}{0.0pt}{1}{b_{1},\,b_{2},\,1-\frac{1}{2m}}{b_{1}+b_{2}-\frac{1}{2m}}\right]} mE[.2b1,b1+b2, 2b2, 112mb1+b212m, 2(b1+b2).]\mathcal{H}^{E}_{m}{\left[\genfrac{.}{.}{0.0pt}{1}{2b_{1},\,b_{1}+b_{2},\,2b_{2},\,1-\frac{1}{2m}}{b_{1}+b_{2}-\frac{1}{2m},\,2(b_{1}+b_{2})}\right]}
Table 2. Sum of even hypergeometric polynomials
Example 4.15 (Additive convolution of Hermite polynomials).

Notice that Row 2 of Table 2 asserts that the additive convolution of two Hermite polynomials yields another Hermite polynomial, recovering a result which also follows from the machinery of [Mar21, AP18].

Example 4.16 (Additive convolution of Bernoulli polynomials).

To compute the additive convolution of two Bernoulli polynomials, we can take a1=a2=112ma_{1}=a_{2}=1-\frac{1}{2m} in Row 1 of Table 2. After a cancellation of the parameter 12m1-\frac{2}{m} appearing downstairs and upstairs in each polynomial, we obtain

B2m2mB2m=mE[..]2mmE[..]=Dil2mE[.12.].B_{2m}\boxplus_{2m}B_{2m}=\mathcal{H}^{E}_{m}{\left[\genfrac{.}{.}{0.0pt}{1}{\cdot}{\cdot}\right]}\boxplus_{2m}\mathcal{H}^{E}_{m}{\left[\genfrac{.}{.}{0.0pt}{1}{\cdot}{\cdot}\right]}=\mathrm{Dil}_{2}\mathcal{H}^{E}_{m}{\left[\genfrac{.}{.}{0.0pt}{1}{1}{2}\right]}\text{.}

Notice that the even part of the right hand side is a Jacobi polynomial.

The special significance of this example, for us, is that it mirrors a basic example of free convolution which can be found in [NS06, Example 12.8]. Namely, if μ=12(δ1+δ1)\mu=\frac{1}{2}(\delta_{1}+\delta_{-1}) is the measure with atoms at ±1\pm 1 and mass 1/21/2 each, the free convolution μμ\mu\boxplus\mu is a so-called arcsine distribution (centered at 0 and supported on [2,2][-2,2]). So the analogue of the arcsine distribution in finite free probability should be the dilation of the squared Jacobi polynomial Dil2mE[.12.]\mathrm{Dil}_{2}\mathcal{H}^{E}_{m}{\left[\genfrac{.}{.}{0.0pt}{1}{1}{2}\right]}.

5. Finite free commutators

One of the main insights of [MSS22] is that the operations n\boxplus_{n} and n\boxtimes_{n}, with their peculiar algebraic descriptions reviewed in Definition 2.9, actually have very natural interpretations involving random matrices. Specifically, pnqp\boxplus_{n}q is the expected characteristic polynomial of A+UBUA+UBU^{*}, where p(x)=det(xInA)p(x)=\det(xI_{n}-A) and q(x)=det(xInB)q(x)=\det(xI_{n}-B) are the characteristic polynomials of some diagonal n×nn\times n matrices AA and BB, and UU is a random n×nn\times n unitary matrix. Similarly, pnqp\boxtimes_{n}q is the expected characteristic polynomial of AUBUAUBU^{*}.

A natural next step is to look at other polynomials in AA and UBUUBU^{*}, and try to extract algebraic descriptions of their expected characteristic polynomials. In particular, knowing the historical development of free probability, one might gravitate towards the self-adjoint commutator i(AUBUUBUA)i(AUBU^{*}-UBU^{*}A). The algebraic description of this commutator operation, due to [Cam22], can be set up in purely algebraic terms involving n\boxplus_{n}, n\boxtimes_{n}, and a particular special polynomial:

Notation 5.1.

Let

zn(x):=k=0n/2xn2k(1)k(n2k)(n)k¯k!(2k)!n+1kn+1z_{n}(x):=\sum_{k=0}^{\lfloor n/2\rfloor}x^{n-2k}(-1)^{k}\binom{n}{2k}\left(n\right)^{\underline{k}}\frac{k!}{(2k)!}\frac{n+1-k}{n+1}

and for polynomials p(x)p(x) and q(x)q(x) with degree nn, write

pnq:=Sym(p)nSym(q)nzn.p\,\square_{n}\,q:=\mathrm{Sym}\left(p\right)\boxtimes_{n}\mathrm{Sym}\left(q\right)\boxtimes_{n}z_{n}\text{.} (46)

The use of the symbol \square is inspired by its use in [NS98].

With n=2mn=2m, one can write

z2m=Dil12mE[.2, 2, 112m12m,12m, 2+1m.].z_{2m}=\mathrm{Dil}_{\frac{1}{2}}\mathcal{H}^{E}_{m}{\left[\genfrac{.}{.}{0.0pt}{1}{2,\,2,\,1-\frac{1}{2m}}{-\frac{1}{2m},\,-\frac{1}{2m},\,2+\frac{1}{m}}\right]}\text{.}

By Theorem 2.12, it follows that z2m𝒫n(>0)z_{2m}\in\mathcal{P}_{n}(\mathbb{R}_{>0}).

The point of the operation n\square_{n} is that it encodes the expected characteristic polynomials of randomly rotated matrices:

Theorem 5.2 ([Cam22]).

Let AA and BB be normal n×nn\times n matrices with characteristic polynomials p(x)=χ(A):=det[xIA]p(x)=\chi(A):=\det[xI-A] and q(x)=χ(B):=det[xIB]q(x)=\chi(B):=\det[xI-B]. Then

𝔼Uχ[i(AUBUUBUA)]=p(x)nq(x)\mathbb{E}_{U}\chi[i(AUBU^{*}-UBU^{*}A)]=p(x)\,\square_{n}\,q(x)

where UU is a random n×nn\times n unitary matrix.

The realization of znz_{n} as a hypergeometric polynomial suggests a connection between special polynomials and analytic questions about finite free commutators. Our conjecture is that the commutator preserves real-rootedness in all cases:

Conjecture 5.3.

For p,q𝒫2m()p,q\in\mathcal{P}_{2m}(\mathbb{R}), we have p2mq𝒫2m()p\,\square_{2m}\,q\in\mathcal{P}_{2m}(\mathbb{R}).

A general proof of this conjecture has turned out to be elusive, but we can provide some partial results. First, we can rephrase the result of [Cam22] in terms of our framework for even polynomials:

Proposition 5.4 (Even part of commutator).

For p,q𝒫2m()p,q\in\mathcal{P}_{2m}(\mathbb{R}), we have

𝑸m(p2mq)=𝑸m(Sym(p))m𝑸m(Sym(q))mDil14m[.2, 2112m, 2+1m.].\bm{Q}_{m}(p\,\square_{2m}\,q)=\bm{Q}_{m}\left(\mathrm{Sym}\left(p\right)\right)\boxtimes_{m}\bm{Q}_{m}\left(\mathrm{Sym}\left(q\right)\right)\boxtimes_{m}\mathrm{Dil}_{\frac{1}{4}}\mathcal{H}_{m}{\left[\genfrac{.}{.}{0.0pt}{1}{2,\,2}{1-\frac{1}{2m},\,2+\frac{1}{m}}\right]}\text{.}
Proof.

When one applies 𝑸m\bm{Q}_{m} to the right-hand side of Eq. 46 and uses Proposition 3.9, one obtains the expression

𝑸m(Sym(p))m𝑸m(Sym(q))mDil14m[.2, 2, 112m12m,12m, 2+1m.]\displaystyle\quad\bm{Q}_{m}(\mathrm{Sym}\left(p\right))\boxtimes_{m}\bm{Q}_{m}(\mathrm{Sym}\left(q\right))\boxtimes_{m}\mathrm{Dil}_{\frac{1}{4}}\mathcal{H}_{m}{\left[\genfrac{.}{.}{0.0pt}{1}{2,\,2,\,1-\frac{1}{2m}}{-\frac{1}{2m},\,-\frac{1}{2m},\,2+\frac{1}{m}}\right]}
mm[.12m112m.]mm[.12m112m.].\displaystyle\qquad\boxtimes_{m}\mathcal{H}_{m}{\left[\genfrac{.}{.}{0.0pt}{1}{-\frac{1}{2m}}{1-\frac{1}{2m}}\right]}\boxtimes_{m}\mathcal{H}_{m}{\left[\genfrac{.}{.}{0.0pt}{1}{-\frac{1}{2m}}{1-\frac{1}{2m}}\right]}\text{.}

By Theorem 2.17, the hypergeometric polynomials can be combined, and cancellation of parameters leaves

m[.2, 2112m, 2+1m.],\mathcal{H}_{m}{\left[\genfrac{.}{.}{0.0pt}{1}{2,\,2}{1-\frac{1}{2m},\,2+\frac{1}{m}}\right]}\text{,}

hence the claim. ∎

Remark 5.5.

It is very important to notice that the polynomial

m[.2, 2112m, 2+1m.],\mathcal{H}_{m}{\left[\genfrac{.}{.}{0.0pt}{1}{2,\,2}{1-\frac{1}{2m},\,2+\frac{1}{m}}\right]}\text{,}

appearing in Proposition 5.4, does not belong to 𝒫m(0)\mathcal{P}_{m}(\mathbb{R}_{\geq 0}).

Actually, if this polynomial were to be in 𝒫m(0)\mathcal{P}_{m}(\mathbb{R}_{\geq 0}), then 5.3 would follow from part (3) of Theorem 2.12, after noticing that

𝑸m(Sym(p))𝒫m(0) and 𝑸m(Sym(q))𝒫m(0)\bm{Q}_{m}\left(\mathrm{Sym}\left(p\right)\right)\in\mathcal{P}_{m}(\mathbb{R}_{\geq 0})\text{ and }\bm{Q}_{m}\left(\mathrm{Sym}\left(q\right)\right)\in\mathcal{P}_{m}(\mathbb{R}_{\geq 0})

by part (5) of Lemma 3.8.

This means that to prove 5.3, one must find some kind of “extra” real-rootedness in the symmetrizations 𝑸m(Sym(p))\bm{Q}_{m}\left(\mathrm{Sym}\left(p\right)\right). It is unclear if this extra positivity holds in general. With this idea in mind, however, we can explicitly formulate a theorem that is similar to 5.3, but requires an extra assumption.

Theorem 5.6.

Let p,q𝒫2m()p,q\in\mathcal{P}_{2m}(\mathbb{R}) and suppose that

𝑸m(Sym(q))=m[.112m.]mr\bm{Q}_{m}(\mathrm{Sym}\left(q\right))=\mathcal{H}_{m}{\left[\genfrac{.}{.}{0.0pt}{1}{1-\frac{1}{2m}}{\cdot}\right]}\boxtimes_{m}r

for some r𝒫m(0)r\in\mathcal{P}_{m}(\mathbb{R}_{\geq 0}). Then p2mq𝒫2m()p\,\square_{2m}\,q\in\mathcal{P}_{2m}(\mathbb{R}).

Proof.

Starting with the expression

𝑸m(p2mq)=𝑸m(Sym(p))m𝑸m(Sym(q))Dil14m[.2, 2112m, 2+1m.]\bm{Q}_{m}(p\,\square_{2m}q)=\bm{Q}_{m}(\mathrm{Sym}\left(p\right))\boxtimes_{m}\bm{Q}_{m}(\mathrm{Sym}\left(q\right))\boxtimes\mathrm{Dil}_{\frac{1}{4}}\mathcal{H}_{m}{\left[\genfrac{.}{.}{0.0pt}{1}{2,\,2}{1-\frac{1}{2m},\,2+\frac{1}{m}}\right]}

from Proposition 5.4, it suffices to show the polynomial on the right-hand side has all its roots in 0\mathbb{R}_{\geq 0}. We can compute

𝑸m(p2mq)\displaystyle\quad\bm{Q}_{m}(p\,\square_{2m}q)
=𝑸2m(Sym(p))mrmm[.112m.]mDil14m[.2, 2112m, 2+1m.]\displaystyle=\bm{Q}_{2m}(\mathrm{Sym}\left(p\right))\boxtimes_{m}r\boxtimes_{m}\mathcal{H}_{m}{\left[\genfrac{.}{.}{0.0pt}{1}{1-\frac{1}{2m}}{\cdot}\right]}\boxtimes_{m}\mathrm{Dil}_{\frac{1}{4}}\mathcal{H}_{m}{\left[\genfrac{.}{.}{0.0pt}{1}{2,\,2}{1-\frac{1}{2m},\,2+\frac{1}{m}}\right]}
=𝑸m(Sym(p))mrmDil14m[.2, 22+1m.]\displaystyle=\bm{Q}_{m}(\mathrm{Sym}\left(p\right))\boxtimes_{m}r\boxtimes_{m}\mathrm{Dil}_{\frac{1}{4}}\mathcal{H}_{m}{\left[\genfrac{.}{.}{0.0pt}{1}{2,\,2}{2+\frac{1}{m}}\right]}

and all three polynomials above are in 𝒫(0)\mathcal{P}(\mathbb{R}_{\geq 0}). By Theorem 2.12, this shows 𝑸2m(p2mq)𝒫m(0)\bm{Q}_{2m}(p\,\square_{2m}\,q)\in\mathcal{P}_{m}(\mathbb{R}_{\geq 0}), and in turn that p2mq𝒫2m()p\,\square_{2m}\,q\in\mathcal{P}_{2m}(\mathbb{R}). ∎

Remark 5.7.

The assumption in Theorem 5.6 is rather restrictive: it is satisfied by the Hermite, Laguerre, and Bessel polynomials that are of interest in finite free probability, but it fails for many other polynomials. However, it does all the work of guaranteeing real-rootedness of p2mqp\,\square_{2m}\,q and allows one to put any real-rooted polynomial in the other argument.

5.1. Examples

First, let us work out some examples which are outside the scope of Theorem 5.6. For these, we will need a particular result from the special function literature:

Remark 5.8.

The polynomial

m[.1, 1112m, 2+1m.]\mathcal{H}_{m}{\left[\genfrac{.}{.}{0.0pt}{1}{1,\,1}{1-\frac{1}{2m},\,2+\frac{1}{m}}\right]}

is positive-rooted. To see this, we can refer to [DJ02, Theorem 3.6] for the fact that

m[.1, 1112m, 2.]𝒫m(0,1)\mathcal{H}_{m}{\left[\genfrac{.}{.}{0.0pt}{1}{1,\,1}{1-\frac{1}{2m},\,2}\right]}\in\mathcal{P}_{m}(0,1)

and use Theorem 2.17 to write

m[.1, 1112m, 2+1m.]=m[.1, 1112m, 2.]mm[.22+1m.].\mathcal{H}_{m}{\left[\genfrac{.}{.}{0.0pt}{1}{1,\,1}{1-\frac{1}{2m},\,2+\frac{1}{m}}\right]}=\mathcal{H}_{m}{\left[\genfrac{.}{.}{0.0pt}{1}{1,\,1}{1-\frac{1}{2m},\,2}\right]}\boxtimes_{m}\mathcal{H}_{m}{\left[\genfrac{.}{.}{0.0pt}{1}{2}{2+\frac{1}{m}}\right]}\text{.}

The latter polynomial is positive-rooted, as explained in (21), so the left-hand side is positive-rooted by Theorem 2.12.

Example 5.9 (Projections and Bernoulli).

Recall the special Jacobi polynomial R2m(r)(x)=x2mr(x1)rR_{2m}^{(r)}(x)=x^{2m-r}(x-1)^{r} from Example 2.23. To compute the finite free commutator of two such polynomials, let 0r,sm0\leq r,s\leq m. By Row 3 of Table 1, with n=2mn=2m, we have

Sym(R2m(r))=mE[.rm, 2rm2, 1.] and Sym(R2m(s))=mE[.sm, 2sm2, 1.]\mathrm{Sym}\left(R_{2m}^{(r)}\right)=\mathcal{H}^{E}_{m}{\left[\genfrac{.}{.}{0.0pt}{1}{\frac{r}{m},\,2-\frac{r}{m}}{2,\,1}\right]}\text{ and }\mathrm{Sym}\left(R_{2m}^{(s)}\right)=\mathcal{H}^{E}_{m}{\left[\genfrac{.}{.}{0.0pt}{1}{\frac{s}{m},\,2-\frac{s}{m}}{2,\,1}\right]}

so by Proposition 5.4, we have

𝑸m(R2m(r)2mR2m(s))\displaystyle\bm{Q}_{m}(R_{2m}^{(r)}\,\square_{2m}\,R_{2m}^{(s)})
=𝑸m(Sym(R2m(r)))m𝑸m(Sym(R2m(s)))mDil14m[.2, 2112m, 2+1m.]\displaystyle=\bm{Q}_{m}\left(\mathrm{Sym}\left(R_{2m}^{(r)}\right)\right)\boxtimes_{m}\bm{Q}_{m}\left(\mathrm{Sym}\left(R_{2m}^{(s)}\right)\right)\boxtimes_{m}\mathrm{Dil}_{\frac{1}{4}}\mathcal{H}_{m}{\left[\genfrac{.}{.}{0.0pt}{1}{2,\,2}{1-\frac{1}{2m},\,2+\frac{1}{m}}\right]}
=m[.rm, 2rm2, 1.]mm[.sm, 2sm2, 1.]mDil14m[.2, 2112m, 2+1m.]\displaystyle=\mathcal{H}_{m}{\left[\genfrac{.}{.}{0.0pt}{1}{\frac{r}{m},\,2-\frac{r}{m}}{2,\,1}\right]}\boxtimes_{m}\mathcal{H}_{m}{\left[\genfrac{.}{.}{0.0pt}{1}{\frac{s}{m},\,2-\frac{s}{m}}{2,\,1}\right]}\boxtimes_{m}\mathrm{Dil}_{\frac{1}{4}}\mathcal{H}_{m}{\left[\genfrac{.}{.}{0.0pt}{1}{2,\,2}{1-\frac{1}{2m},\,2+\frac{1}{m}}\right]}
=Dil14m[.rm, 2rm,sm, 2sm1, 1, 112m, 2+1m.]\displaystyle=\mathrm{Dil}_{\frac{1}{4}}\mathcal{H}_{m}{\left[\genfrac{.}{.}{0.0pt}{1}{\frac{r}{m},\,2-\frac{r}{m},\,\frac{s}{m},\,2-\frac{s}{m}}{1,\,1,\,1-\frac{1}{2m},\,2+\frac{1}{m}}\right]}

and

R2m(r)2mR2m(s)=Dil12mE[.rm, 2rm,sm, 2sm1, 1, 112m, 2+1m.].R_{2m}^{(r)}\,\square_{2m}\,R_{2m}^{(s)}=\mathrm{Dil}_{\frac{1}{2}}\mathcal{H}^{E}_{m}{\left[\genfrac{.}{.}{0.0pt}{1}{\frac{r}{m},\,2-\frac{r}{m},\,\frac{s}{m},\,2-\frac{s}{m}}{1,\,1,\,1-\frac{1}{2m},\,2+\frac{1}{m}}\right]}\text{.}

With r=s=mr=s=m, i.e. with the roots evenly split between 0 and 11, the above reads as

R2m(m)2mR2m(m)=Dil12mE[.1, 1112m, 2+1m.].R_{2m}^{(m)}\,\square_{2m}\,R_{2m}^{(m)}=\mathrm{Dil}_{\frac{1}{2}}\mathcal{H}^{E}_{m}{\left[\genfrac{.}{.}{0.0pt}{1}{1,\,1}{1-\frac{1}{2m},\,2+\frac{1}{m}}\right]}\text{.}

Notice that the polynomial R2m(m)(x)=xm(x1)mR_{2m}^{(m)}(x)=x^{m}(x-1)^{m} is very similar to the Bernoulli polynomial B2m(x)=(x1)m(x+1)mB_{2m}(x)=(x-1)^{m}(x+1)^{m}, introduced in Example 4.2. Indeed, we can get one from the other by performing a shift and dilation:

B2m=(Dil2R2m(m))2mδ1,B_{2m}=(\mathrm{Dil}_{2}R_{2m}^{(m)})\boxplus_{2m}\delta_{1}\text{,} (47)

where δ1(x)=(x+1)2m\delta_{1}(x)=(x+1)^{2m}. This means that symmetrizations of these two polynomials are the same up to a dilation by 2, and the commutators coincide up to a dilation by 4. Specifically, by Lemma 3.8, one can check that

Sym(B2m)\displaystyle\mathrm{Sym}\left(B_{2m}\right) =Sym((Dil2R2m(m))2mδ1)\displaystyle=\mathrm{Sym}\left((\mathrm{Dil}_{2}R_{2m}^{(m)})\boxplus_{2m}\delta_{1}\right)
=Sym(Dil2R2m(m))=Dil2Sym(R2m(m)).\displaystyle=\mathrm{Sym}\left(\mathrm{Dil}_{2}R_{2m}^{(m)}\right)=\mathrm{Dil}_{2}\mathrm{Sym}\left(R_{2m}^{(m)}\right)\text{.}

Thus, by Eq. 46 we have

B2m2mB2m\displaystyle B_{2m}\,\square_{2m}\,B_{2m} =Sym(B2m)2mSym(B2m)2mz2m\displaystyle=\mathrm{Sym}\left(B_{2m}\right)\boxtimes_{2m}\mathrm{Sym}\left(B_{2m}\right)\boxtimes_{2m}z_{2m}
=Dil2Sym(R2m(m))2mDil2Sym(R2m(m))2mz2m\displaystyle=\mathrm{Dil}_{2}\mathrm{Sym}\left(R_{2m}^{(m)}\right)\boxtimes_{2m}\mathrm{Dil}_{2}\mathrm{Sym}\left(R_{2m}^{(m)}\right)\boxtimes_{2m}z_{2m}
=Dil4(R2m(m)2mR2m(m)),\displaystyle=\mathrm{Dil}_{4}\left(R_{2m}^{(m)}\,\square_{2m}\,R_{2m}^{(m)}\right),

so we conlude that

B2m2mB2m=Dil2mE[.1, 1112m, 2+1m.].B_{2m}\,\square_{2m}\,B_{2m}=\mathrm{Dil}_{2}\mathcal{H}^{E}_{m}{\left[\genfrac{.}{.}{0.0pt}{1}{1,\,1}{1-\frac{1}{2m},\,2+\frac{1}{m}}\right]}\text{.}

As explained in Remark 5.8, the polynomial appearing as the finite free commutator in these examples is real-rooted. The limit ERDs of these finite free commutators will be described in Example 6.8 and Example 6.9.

It is also worthwhile to work out some detailed examples involving polynomials which are covered by Theorem 5.6:

Example 5.10 (Hermite polynomials).

Let H2mH_{2m} be the Hermite polynomial defined in Example 2.21. By Example 4.15, we have

Sym(H2m)=Dil2mmE[.112m.]\mathrm{Sym}\left(H_{2m}\right)=\mathrm{Dil}_{\sqrt{\frac{2}{m}}}\mathcal{H}^{E}_{m}{\left[\genfrac{.}{.}{0.0pt}{1}{1-\frac{1}{2m}}{\cdot}\right]}

so with p=q=H2mp=q=H_{2m}, we have

𝑸m(H2m2mH2m)\displaystyle\bm{Q}_{m}(H_{2m}\,\square_{2m}\,H_{2m}) =Dil2mm[.112m.]mDil2mm[.112m.]\displaystyle=\mathrm{Dil}_{\frac{2}{m}}\mathcal{H}_{m}{\left[\genfrac{.}{.}{0.0pt}{1}{1-\frac{1}{2m}}{\cdot}\right]}\boxtimes_{m}\mathrm{Dil}_{\frac{2}{m}}\mathcal{H}_{m}{\left[\genfrac{.}{.}{0.0pt}{1}{1-\frac{1}{2m}}{\cdot}\right]}
mDil14m[.2, 2112m, 2+1m.]\displaystyle\qquad\boxtimes_{m}\mathrm{Dil}_{\frac{1}{4}}\mathcal{H}_{m}{\left[\genfrac{.}{.}{0.0pt}{1}{2,\,2}{1-\frac{1}{2m},\,2+\frac{1}{m}}\right]}
=Dil1m2m[.112m, 2, 22+1m.]\displaystyle=\mathrm{Dil}_{\frac{1}{m^{2}}}\mathcal{H}_{m}{\left[\genfrac{.}{.}{0.0pt}{1}{1-\frac{1}{2m},\,2,\,2}{2+\frac{1}{m}}\right]}

This polynomial is positive-rooted because

m[.112m, 2, 22+1m.]=m[.112m2+1m.]mm[.2.]mm[.2.];\mathcal{H}_{m}{\left[\genfrac{.}{.}{0.0pt}{1}{1-\frac{1}{2m},\,2,\,2}{2+\frac{1}{m}}\right]}=\mathcal{H}_{m}{\left[\genfrac{.}{.}{0.0pt}{1}{1-\frac{1}{2m}}{2+\frac{1}{m}}\right]}\boxtimes_{m}\mathcal{H}_{m}{\left[\genfrac{.}{.}{0.0pt}{1}{2}{\cdot}\right]}\boxtimes_{m}\mathcal{H}_{m}{\left[\genfrac{.}{.}{0.0pt}{1}{2}{\cdot}\right]}\text{;}

the three polynomials on the right-hand side are well-known examples of positive-rooted Jacobi and Laguerre polynomials, reviewed in Example 2.23 and Example 2.21 respectively. This shows that

H2m2mH2m=Dil1mmE[.112m, 2, 22+1m.]𝒫2m().H_{2m}\,\square_{2m}\,H_{2m}=\mathrm{Dil}_{\frac{1}{m}}\mathcal{H}^{E}_{m}{\left[\genfrac{.}{.}{0.0pt}{1}{1-\frac{1}{2m},\,2,\,2}{2+\frac{1}{m}}\right]}\in\mathcal{P}_{2m}(\mathbb{R})\text{.}

The limit in ERD of this finite free commutator will be described in Example 6.10.

Example 5.11 (Hermite and projection).

Now let us consider the commutator of a Hermite polynomial H2mH_{2m} with a projection-like polynomial R2m(m)R_{2m}^{(m)}, whose roots are evenly split between 0 and 11. As computed above, we have

Sym(H2m)=Dil2/mmE[.112m.] and Sym(R2m(m))=mE[.12.]\mathrm{Sym}\left(H_{2m}\right)=\mathrm{Dil}_{\sqrt{2/m}}\mathcal{H}^{E}_{m}{\left[\genfrac{.}{.}{0.0pt}{1}{1-\frac{1}{2m}}{\cdot}\right]}\text{ and }\mathrm{Sym}\left(R_{2m}^{(m)}\right)=\mathcal{H}^{E}_{m}{\left[\genfrac{.}{.}{0.0pt}{1}{1}{2}\right]}

so

𝑸m(H2m2mR2m(m))\displaystyle\bm{Q}_{m}(H_{2m}\,\square_{2m}\,R_{2m}^{(m)}) =Dil2mm[.112m.]mm[.12.]\displaystyle=\mathrm{Dil}_{\frac{2}{m}}\mathcal{H}_{m}{\left[\genfrac{.}{.}{0.0pt}{1}{1-\frac{1}{2m}}{\cdot}\right]}\boxtimes_{m}\mathcal{H}_{m}{\left[\genfrac{.}{.}{0.0pt}{1}{1}{2}\right]}
mDil14m[.2, 2112m, 2+1m.]\displaystyle\qquad\boxtimes_{m}\mathrm{Dil}_{\frac{1}{4}}\mathcal{H}_{m}{\left[\genfrac{.}{.}{0.0pt}{1}{2,\,2}{1-\frac{1}{2m},\,2+\frac{1}{m}}\right]}
=Dil12mm[.1, 22+1m.].\displaystyle=\mathrm{Dil}_{\frac{1}{2m}}\mathcal{H}_{m}{\left[\genfrac{.}{.}{0.0pt}{1}{1,\,2}{2+\frac{1}{m}}\right]}\text{.}

This polynomial is positive-rooted because

m[.1, 22+1m.]=m[.12+1m.]mm[.2.]\mathcal{H}_{m}{\left[\genfrac{.}{.}{0.0pt}{1}{1,\,2}{2+\frac{1}{m}}\right]}=\mathcal{H}_{m}{\left[\genfrac{.}{.}{0.0pt}{1}{1}{2+\frac{1}{m}}\right]}\boxtimes_{m}\mathcal{H}_{m}{\left[\genfrac{.}{.}{0.0pt}{1}{2}{\cdot}\right]}

and the two polynomials on the right-hand side are again clear examples of positive-rooted Jacobi and Laguerre polynomials. This shows that

H2m2mR2m(m)=Dil12mmE[.1, 22+1m.]𝒫2m().H_{2m}\,\square_{2m}\,R_{2m}^{(m)}=\mathrm{Dil}_{\frac{1}{\sqrt{2m}}}\mathcal{H}^{E}_{m}{\left[\genfrac{.}{.}{0.0pt}{1}{1,\,2}{2+\frac{1}{m}}\right]}\in\mathcal{P}_{2m}(\mathbb{R})\text{.}

The limit in ERD of this finite free commutator will be described in Example 6.11.

Example 5.12 (Laguerre polynomials).

Recall the polynomial

Ln(λ)=Dil1/nn[.λ.],L_{n}^{(\lambda)}=\mathrm{Dil}_{1/n}\mathcal{H}_{n}{\left[\genfrac{.}{.}{0.0pt}{1}{\lambda}{\cdot}\right]}\text{,}

from Example 2.21. By Row 1 of Table 1, with n=2mn=2m, we have

Sym(Ln(λ))=Dil1mmE[.2λ, 112m.]\mathrm{Sym}\left(L_{n}^{(\lambda)}\right)=\mathrm{Dil}_{\frac{1}{m}}\mathcal{H}^{E}_{m}{\left[\genfrac{.}{.}{0.0pt}{1}{2\lambda,\,1-\frac{1}{2m}}{\cdot}\right]}

so we have

𝑸m(Ln(λ)2mLn(μ))\displaystyle\bm{Q}_{m}(L_{n}^{(\lambda)}\,\square_{2m}\,L_{n}^{(\mu)}) =Dil1m2m[.2λ, 112m.]Dil1m2m[.2μ, 112m.]\displaystyle=\mathrm{Dil}_{\frac{1}{m^{2}}}\mathcal{H}_{m}{\left[\genfrac{.}{.}{0.0pt}{1}{2\lambda,\,1-\frac{1}{2m}}{\cdot}\right]}\boxtimes\mathrm{Dil}_{\frac{1}{m^{2}}}\mathcal{H}_{m}{\left[\genfrac{.}{.}{0.0pt}{1}{2\mu,\,1-\frac{1}{2m}}{\cdot}\right]}
mDil14m[.2, 2112m, 2+1m.]\displaystyle\qquad\boxtimes_{m}\mathrm{Dil}_{\frac{1}{4}}\mathcal{H}_{m}{\left[\genfrac{.}{.}{0.0pt}{1}{2,\,2}{1-\frac{1}{2m},\,2+\frac{1}{m}}\right]}
=Dil14m4m[.2λ, 2μ, 2, 2, 112m2+1m.]\displaystyle=\mathrm{Dil}_{\frac{1}{4m^{4}}}\mathcal{H}_{m}{\left[\genfrac{.}{.}{0.0pt}{1}{2\lambda,\,2\mu,\,2,\,2,\,1-\frac{1}{2m}}{2+\frac{1}{m}}\right]}

This polynomial is positive-rooted for λ,μ12\lambda,\mu\geq\frac{1}{2} because

m[.2λ, 2μ, 2, 2, 112m2+1m.]\displaystyle\mathcal{H}_{m}{\left[\genfrac{.}{.}{0.0pt}{1}{2\lambda,\,2\mu,\,2,\,2,\,1-\frac{1}{2m}}{2+\frac{1}{m}}\right]} =m[.2λ.]mm[.2μ.]mm[.2.]m[.2.]\displaystyle=\mathcal{H}_{m}{\left[\genfrac{.}{.}{0.0pt}{1}{2\lambda}{\cdot}\right]}\boxtimes_{m}\mathcal{H}_{m}{\left[\genfrac{.}{.}{0.0pt}{1}{2\mu}{\cdot}\right]}\boxtimes_{m}\mathcal{H}_{m}{\left[\genfrac{.}{.}{0.0pt}{1}{2}{\cdot}\right]}\boxtimes\mathcal{H}_{m}{\left[\genfrac{.}{.}{0.0pt}{1}{2}{\cdot}\right]}
mm[.112m2+1m.];\displaystyle\qquad\boxtimes_{m}\mathcal{H}_{m}{\left[\genfrac{.}{.}{0.0pt}{1}{1-\frac{1}{2m}}{2+\frac{1}{m}}\right]}\text{;}

the first two are positive-rooted when λ,μ12\lambda,\mu\geq\frac{1}{2}, and the rest are positive-rooted as reviewed in Example 2.21 and Example 2.23. This shows that

L2m(λ)2mL2m(μ)=Dil12m2mE[.2λ, 2μ, 2, 2, 112m2+1m.]𝒫2m().L_{2m}^{(\lambda)}\,\square_{2m}\,L_{2m}^{(\mu)}=\mathrm{Dil}_{\frac{1}{2m^{2}}}\mathcal{H}^{E}_{m}{\left[\genfrac{.}{.}{0.0pt}{1}{2\lambda,\,2\mu,\,2,\,2,\,1-\frac{1}{2m}}{2+\frac{1}{m}}\right]}\in\mathcal{P}_{2m}(\mathbb{R})\text{.}

The limit in ERD of this finite free commutator will be described in Example 6.12.

Example 5.13 (Bessel polynomials).

Let a,b<0a,b<0 and let p=Cn(a)p=C_{n}^{(a)} and q=Cn(b)q=C_{n}^{(b)}, using the notation of Example 2.22. Then with n=2mn=2m, we have

Sym(Cn(a))=Dil2mimE[.112m2a,a,a12m.]\mathrm{Sym}\left(C_{n}^{(a)}\right)=\mathrm{Dil}_{2mi}\mathcal{H}^{E}_{m}{\left[\genfrac{.}{.}{0.0pt}{1}{1-\frac{1}{2m}}{2a,\,a,\,a-\frac{1}{2m}}\right]}

and

𝑸m(C2m(a)2mC2m(b))\displaystyle\bm{Q}_{m}(C_{2m}^{(a)}\,\square_{2m}\,C_{2m}^{(b)}) =Dil4m2m[.112m2a,a,a12m.]mDil4m2m[.112m2b,b,b12m.]\displaystyle=\mathrm{Dil}_{-4m^{2}}\mathcal{H}_{m}{\left[\genfrac{.}{.}{0.0pt}{1}{1-\frac{1}{2m}}{2a,\,a,\,a-\frac{1}{2m}}\right]}\boxtimes_{m}\mathrm{Dil}_{-4m^{2}}\mathcal{H}_{m}{\left[\genfrac{.}{.}{0.0pt}{1}{1-\frac{1}{2m}}{2b,\,b,\,b-\frac{1}{2m}}\right]}
mDil14m[.2, 2112m, 2+1m.]\displaystyle\qquad\boxtimes_{m}\mathrm{Dil}_{\frac{1}{4}}\mathcal{H}_{m}{\left[\genfrac{.}{.}{0.0pt}{1}{2,\,2}{1-\frac{1}{2m},\,2+\frac{1}{m}}\right]}
=Dil4m4m[.112m, 2, 22a,a,a12m, 2b,b,b12m, 2+1m.].\displaystyle=\mathrm{Dil}_{4m^{4}}\mathcal{H}_{m}{\left[\genfrac{.}{.}{0.0pt}{1}{1-\frac{1}{2m},\,2,\,2}{2a,\,a,\,a-\frac{1}{2m},\,2b,\,b,\,b-\frac{1}{2m},\,2+\frac{1}{m}}\right]}\text{.}

This polynomial is positive-rooted because

m[.112m, 2, 22a,a,a12m, 2b,b,b12m, 2+1m.]\displaystyle\mathcal{H}_{m}{\left[\genfrac{.}{.}{0.0pt}{1}{1-\frac{1}{2m},\,2,\,2}{2a,\,a,\,a-\frac{1}{2m},\,2b,\,b,\,b-\frac{1}{2m},\,2+\frac{1}{m}}\right]} =m[.2.]mm[.2.]mm[.112m2+1m.]\displaystyle=\mathcal{H}_{m}{\left[\genfrac{.}{.}{0.0pt}{1}{2}{\cdot}\right]}\boxtimes_{m}\mathcal{H}_{m}{\left[\genfrac{.}{.}{0.0pt}{1}{2}{\cdot}\right]}\boxtimes_{m}\mathcal{H}_{m}{\left[\genfrac{.}{.}{0.0pt}{1}{1-\frac{1}{2m}}{2+\frac{1}{m}}\right]}
mm[.2a.]mm[.a.]mm[.a12m.]\displaystyle\qquad\boxtimes_{m}\mathcal{H}_{m}{\left[\genfrac{.}{.}{0.0pt}{1}{\cdot}{2a}\right]}\boxtimes_{m}\mathcal{H}_{m}{\left[\genfrac{.}{.}{0.0pt}{1}{\cdot}{a}\right]}\boxtimes_{m}\mathcal{H}_{m}{\left[\genfrac{.}{.}{0.0pt}{1}{\cdot}{a-\frac{1}{2m}}\right]}
mm[.2b.]mm[.b.]mm[.b12m.].\displaystyle\qquad\boxtimes_{m}\mathcal{H}_{m}{\left[\genfrac{.}{.}{0.0pt}{1}{\cdot}{2b}\right]}\boxtimes_{m}\mathcal{H}_{m}{\left[\genfrac{.}{.}{0.0pt}{1}{\cdot}{b}\right]}\boxtimes_{m}\mathcal{H}_{m}{\left[\genfrac{.}{.}{0.0pt}{1}{\cdot}{b-\frac{1}{2m}}\right]}\text{.}

The Laguerre and Jacobi polynomials on the right-hand side are positive-rooted as explained in previous examples. The Bessel polynomials are all negative-rooted as explained in Example 2.22, and the “rule of signs” from Remark 2.13 shows that the multiplicative convolution of six negative-rooted polynomials is positive-rooted. All told, this shows that

Cn(a)2mCn(b)=Dil2m2mE[.112m, 2, 22a,a,a12m, 2b,b,b12m, 2+1m.]𝒫2m().C_{n}^{(a)}\,\square_{2m}\,C_{n}^{(b)}=\mathrm{Dil}_{2m^{2}}\mathcal{H}^{E}_{m}{\left[\genfrac{.}{.}{0.0pt}{1}{1-\frac{1}{2m},\,2,\,2}{2a,\,a,\,a-\frac{1}{2m},\,2b,\,b,\,b-\frac{1}{2m},\,2+\frac{1}{m}}\right]}\in\mathcal{P}_{2m}(\mathbb{R})\text{.}
Example 5.14 (Jacobi polynomials).

For this example, we omit details, since they are cumbersome and identical to previous examples. If p=2m[.ba.]p=\mathcal{H}_{2m}{\left[\genfrac{.}{.}{0.0pt}{1}{b}{a}\right]} and q=2m[.dc.]q=\mathcal{H}_{2m}{\left[\genfrac{.}{.}{0.0pt}{1}{d}{c}\right]}, then

p2mq=Dil1/2mE[.2b, 2a2b, 2d, 2c2d, 2, 2, 112m2a,a,a12m, 2c,c,c12m, 2+1m.].p\,\square_{2m}\,q=\mathrm{Dil}_{1/2}\mathcal{H}^{E}_{m}{\left[\genfrac{.}{.}{0.0pt}{1}{2b,\,2a-2b,\,2d,\,2c-2d,\,2,\,2,\,1-\frac{1}{2m}}{2a,\,a,\,a-\frac{1}{2m},\,2c,\,c,\,c-\frac{1}{2m},\,2+\frac{1}{m}}\right]}\text{.}

For example, with b=d=1b=d=1 and a=c=2a=c=2, the above is

Dil1/2mE[.2, 2, 2, 2, 112m4, 4, 212m, 212m, 2+1m.].\mathrm{Dil}_{1/2}\mathcal{H}^{E}_{m}{\left[\genfrac{.}{.}{0.0pt}{1}{2,\,2,\,2,\,2,\,1-\frac{1}{2m}}{4,\,4,\,2-\frac{1}{2m},\,2-\frac{1}{2m},\,2+\frac{1}{m}}\right]}\text{.}

We collect these examples in Table 3.

pp qq p2mqp\,\square_{2m}\,q
2m[.121.]\mathcal{H}_{2m}{\left[\genfrac{.}{.}{0.0pt}{1}{\frac{1}{2}}{1}\right]} 2m[.121.]\mathcal{H}_{2m}{\left[\genfrac{.}{.}{0.0pt}{1}{\frac{1}{2}}{1}\right]} Dil12mE[.1, 1112m, 2+1m.]\mathrm{Dil}_{\frac{1}{2}}\mathcal{H}^{E}_{m}{\left[\genfrac{.}{.}{0.0pt}{1}{1,\,1}{1-\frac{1}{2m},\,2+\frac{1}{m}}\right]}
mE[.112m.]\mathcal{H}^{E}_{m}{\left[\genfrac{.}{.}{0.0pt}{1}{1-\frac{1}{2m}}{\cdot}\right]} mE[.112m.]\mathcal{H}^{E}_{m}{\left[\genfrac{.}{.}{0.0pt}{1}{1-\frac{1}{2m}}{\cdot}\right]} Dil12mE[.112m, 2, 22+1m.]\mathrm{Dil}_{\frac{1}{2}}\mathcal{H}^{E}_{m}{\left[\genfrac{.}{.}{0.0pt}{1}{1-\frac{1}{2m},\,2,\,2}{2+\frac{1}{m}}\right]}
mE[.112m.]\mathcal{H}^{E}_{m}{\left[\genfrac{.}{.}{0.0pt}{1}{1-\frac{1}{2m}}{\cdot}\right]} 2m[.121.]\mathcal{H}_{2m}{\left[\genfrac{.}{.}{0.0pt}{1}{\frac{1}{2}}{1}\right]} Dil12mE[.1, 22+1m.]\mathrm{Dil}_{\frac{1}{2}}\mathcal{H}^{E}_{m}{\left[\genfrac{.}{.}{0.0pt}{1}{1,\,2}{2+\frac{1}{m}}\right]}
2m[.b.]\mathcal{H}_{2m}{\left[\genfrac{.}{.}{0.0pt}{1}{b}{\cdot}\right]} 2m[.d.]\mathcal{H}_{2m}{\left[\genfrac{.}{.}{0.0pt}{1}{d}{\cdot}\right]} Dil12mE[.2b, 2d, 2, 2, 112m2+1m.]\mathrm{Dil}_{\frac{1}{2}}\mathcal{H}^{E}_{m}{\left[\genfrac{.}{.}{0.0pt}{1}{2b,\,2d,\,2,\,2,\,1-\frac{1}{2m}}{2+\frac{1}{m}}\right]}
2m[.12.]\mathcal{H}_{2m}{\left[\genfrac{.}{.}{0.0pt}{1}{1}{2}\right]} 2m[.12.]\mathcal{H}_{2m}{\left[\genfrac{.}{.}{0.0pt}{1}{1}{2}\right]} Dil12mE[.2, 2, 2, 2, 112m4, 4, 212m, 212m, 2+1m.]\mathrm{Dil}_{\frac{1}{2}}\mathcal{H}^{E}_{m}{\left[\genfrac{.}{.}{0.0pt}{1}{2,\,2,\,2,\,2,\,1-\frac{1}{2m}}{4,\,4,\,2-\frac{1}{2m},\,2-\frac{1}{2m},\,2+\frac{1}{m}}\right]}
Table 3. Examples of finite free commutators

6. Asymptotics and connection to free probability

Most of the results from previous sections have counterparts in free probability, if we let the degree nn tend to infinity while the empirical root distribution tends to a compact measure. Some of these results are well known in free probability, and in fact served as original motivation for the results in finite free probability. However, to the best of our knowledge, many examples are new, and might shed some light on future directions in this area.

We begin by explicitly proving the intuitive fact that when we let mm\to\infty, the map 𝑺m\bm{S}_{m} from 3.4 tends to the map SS from 2.4. Thus, the bijection between 𝒫m(0)\mathcal{P}_{m}(\mathbb{R}_{\geq 0}) and 𝒫2mE()\mathcal{P}_{2m}^{E}(\mathbb{R}) from 3.6, in the limit becomes the bijection between (0)\mathcal{M}(\mathbb{R}_{\geq 0}) and E()\mathcal{M}^{E}(\mathbb{R}) from Lemma 2.5.

Proposition 6.1 (Even polynomials approximate symmetric measures).

Let 𝔭=(pm)m1\mathfrak{p}=(p_{m})_{m\geq 1} be a converging sequence of positive-rooted polynomials. Then 𝔭:=(𝐒m(p2m))m1\sqrt{\mathfrak{p}}:=(\bm{S}_{m}(p_{2m}))_{m\geq 1} is a converging sequence of even real-rooted polynomials, with ρ(𝔭)=S(ρ(𝔭))\rho(\sqrt{\mathfrak{p}})=S(\rho(\mathfrak{p})).

Similarly, if 𝔮=(q2m)m1\mathfrak{q}=(q_{2m})_{m\geq 1} is a converging sequence of even real-rooted polynomials, then 𝔮2:=(𝐐m(q2m))m1\mathfrak{q}^{2}:=(\bm{Q}_{m}(q_{2m}))_{m\geq 1} is a converging sequence of positive-rooted polynomials with ρ(𝔮2)=Q(ρ(𝔮))\rho(\mathfrak{q}^{2})=Q(\rho(\mathfrak{q})).

Proof.

The basic tool involved in the proof of this proposition is the fact [Bil99, Theorem 2.7] that pushforwards along continuous functions preserve weak limits.

For the first claim, the assumptions amount to limmρ(pm)=ρ(𝔭)\lim_{m\to\infty}\rho(p_{m})=\rho(\mathfrak{p}). Then

limmS+(ρ(pm))=S+(ρ(𝔭)) and limmS(ρ(pm))=S(ρ(𝔭))\lim_{m\to\infty}S_{+}(\rho(p_{m}))=S_{+}(\rho(\mathfrak{p}))\text{ and }\lim_{m\to\infty}S_{-}(\rho(p_{m}))=S_{-}(\rho(\mathfrak{p}))

since 00:tt\mathbb{R}_{\geq 0}\to\mathbb{R}_{\geq 0}:t\mapsto\sqrt{t} and 00:tt\mathbb{R}_{\geq 0}\to\mathbb{R}_{\leq 0}:t\mapsto-\sqrt{t} are continuous. This shows that

limmS(ρ(pm))=S(ρ(𝔭)).\lim_{m\to\infty}S(\rho(p_{m}))=S(\rho(\mathfrak{p}))\text{.}

Since S(ρ(pm))=ρ(𝑺m(pm))S(\rho(p_{m}))=\rho(\bm{S}_{m}(p_{m})), we have

limmρ(𝑺m(pm))=S(ρ(𝔭)),\lim_{m\to\infty}\rho(\bm{S}_{m}(p_{m}))=S(\rho(\mathfrak{p}))\text{,}

hence 𝔭=(𝑺m(pm)m1\sqrt{\mathfrak{p}}=(\bm{S}_{m}(p_{m})_{m\geq 1} is a converging sequence with ρ(𝔭)=S(ρ(𝔭))\rho(\sqrt{\mathfrak{p}})=S(\rho(\mathfrak{p})).

For the second claim, suppose that 𝔮=(q2m)m1\mathfrak{q}=(q_{2m})_{m\geq 1} is a converging sequence of even real-rooted polynomials, i.e. limmρ(q2m)=ρ(𝔮)\lim_{m\to\infty}\rho(q_{2m})=\rho(\mathfrak{q}). Then since 0:tt2\mathbb{R}\to\mathbb{R}_{\geq 0}:t\mapsto t^{2} is continuous, we have

limmQ(ρ(q2m))=Q(limmρ(q2m))=Q(ρ(𝔮)).\lim_{m\to\infty}Q(\rho(q_{2m}))=Q\left(\lim_{m\to\infty}\rho(q_{2m})\right)=Q(\rho(\mathfrak{q}))\text{.}

Since Q(ρ(q2m)=ρ(𝑸m(q2m))Q(\rho(q_{2m})=\rho(\bm{Q}_{m}(q_{2m})), this shows that

limmρ(𝑸m(q2m))=Q(ρ(𝔮)),\lim_{m\to\infty}\rho(\bm{Q}_{m}(q_{2m}))=Q(\rho(\mathfrak{q}))\text{,}

i.e. 𝔮2=(𝑸m(q2m))m1\mathfrak{q}^{2}=(\bm{Q}_{m}(q_{2m}))_{m\geq 1} is a converging sequence with ρ(𝔮2)=Q(ρ(𝔮))\rho(\mathfrak{q}^{2})=Q(\rho(\mathfrak{q})). ∎

Using this result, we can study the limiting root distributions of even polynomial. Specifically, the examples from Section 4 involving hypergeometric polynomials can be studied asymptotically, using free convolutions and even measures. Recall from Theorem 2.18 and 2.19 that the asymptotic root distribution of hypergeometric polynomials is given by a SS-rational measure ρ[.b1,,bja1,,ai.]\rho{\left[\genfrac{.}{.}{0.0pt}{1}{b_{1},\dots,b_{j}}{a_{1},\dots,a_{i}}\right]}, determined by having an SS-transform of the form (z+a1)(z+ai)(z+b1)(z+bj)\frac{(z+a_{1})\cdots(z+a_{i})}{(z+b_{1})\cdots(z+b_{j})}. Recall also that we denote the square root of an SS-rational function by

ρE[.b1,,bja1,,ai.].\rho^{E}{\left[\genfrac{.}{.}{0.0pt}{1}{b_{1},\dots,b_{j}}{a_{1},\dots,a_{i}}\right]}.

We now state an analogue of Theorem 2.18 that holds for even hypergeometric polynomials:

Theorem 6.2.

For non-negative integers i,ji,j consider tuples of parameters 𝐀=(A1,,Ai)([0,1))i\bm{A}=(A_{1},\dots,A_{i})\in(\mathbb{R}\setminus[0,1))^{i} and 𝐁=(B1,,Bj)({0})j\bm{B}=(B_{1},\dots,B_{j})\in(\mathbb{R}\setminus\{0\})^{j}. Assume that 𝔭=(pm)m0\mathfrak{p}=(p_{m})_{m\geq 0} is a sequence of polynomials given by

pm=DilmijmE[.𝒃m𝒂m.]𝒫2mE(),p_{m}=\mathrm{Dil}_{\sqrt{m^{i-j}}}\mathcal{H}^{E}_{m}{\left[\genfrac{.}{.}{0.0pt}{1}{\bm{b}_{m}}{\bm{a}_{m}}\right]}\in\mathcal{P}^{E}_{2m}(\mathbb{R}), (48)

where the tuples of parameters 𝐚mi\bm{a}_{m}\in\mathbb{R}^{i} and 𝐛mj\bm{b}_{m}\in\mathbb{R}^{j} have a limit given by

limm𝒂m=𝑨,andlimm𝒃m=𝑩.\lim_{m\to\infty}\bm{a}_{m}=\bm{A},\qquad\text{and}\qquad\lim_{m\to\infty}\bm{b}_{m}=\bm{B}. (49)

Then 𝔭\mathfrak{p} is a converging sequences (in the sense of Definition 2.6) that converges to the measure

ρE[.B1,,BjA1,,Ai.]E().\rho^{E}{\left[\genfrac{.}{.}{0.0pt}{1}{B_{1},\dots,B_{j}}{A_{1},\dots,A_{i}}\right]}\in\mathcal{M}^{E}(\mathbb{R}).
Proof.

Follows from Proposition 6.1 and Theorem 2.18. ∎

Example 6.3 (Hermite polynomials and Laguerre polynomials).

In free probability, it is known [NS06, Proposition 12.13] that the square of a semicircular element has a free Poisson distribution with rate 11. We can observe a similar phenomenon by applying our operations 𝑸m\bm{Q}_{m} and 𝑺m\bm{S}_{m} to Hermite and Laguerre polynomials; namely, it corresponds to the observation in Example 2.21 that Hermite polynomials are related to certain Laguerre polynomials with squared variables.

More precisely, recall that the respective finite analogues of the free Poisson distribution with rate 11 and the semicircular distribution with rate 22 are the polynomials

Ln(1)(x)=nnn[.1.](nx)andH2m(x)=mmm[.112m.](mx2).L_{n}^{(1)}(x)=n^{-n}\mathcal{H}_{n}{\left[\genfrac{.}{.}{0.0pt}{1}{1}{\cdot}\right]}(nx)\quad\text{and}\quad H_{2m}(x)=m^{-m}\mathcal{H}_{m}{\left[\genfrac{.}{.}{0.0pt}{1}{1-\frac{1}{2m}}{\cdot}\right]}(mx^{2})\text{.}

We have

𝑸m(H2m)=mmm[.112m.](mx)\bm{Q}_{m}(H_{2m})=m^{-m}\mathcal{H}_{m}{\left[\genfrac{.}{.}{0.0pt}{1}{1-\frac{1}{2m}}{\cdot}\right]}(mx)

and since limm(112m)=1\lim_{m\to\infty}(1-\frac{1}{2m})=1, by Theorem 2.18 we have

limmρ(𝑸m(H2m))=ρ[.1.]=limnρ(Ln(1)).\lim_{m\to\infty}\rho(\bm{Q}_{m}(H_{2m}))=\rho{\left[\genfrac{.}{.}{0.0pt}{1}{1}{\cdot}\right]}=\lim_{n\to\infty}\rho(L_{n}^{(1)})\text{.}

We thus recover the aforementioned result concerning the square of a semicircular variable.

6.1. Examples of free symmetrization

The symmetrization operation from 3.7 tends to the corresponding symmetrization operation in free probability, defined as follows:

Notation 6.4.

For μ()\mu\in\mathcal{M}(\mathbb{R}), define the free symmetrization of μ\mu by

Sym(μ):=μ(Dil1μ)E().\mathrm{Sym}\left(\mu\right):=\mu\boxplus(\mathrm{Dil}_{-1}\mu)\in\mathcal{M}^{E}(\mathbb{R})\text{.}
Remark 6.5.

Despite being a simple operation, to the best of our knowledge Sym(μ)\mathrm{Sym}\left(\mu\right) has not been studied systematically in free probability. Some particular instances where this operation has appeared include [HM10, Final remarks], [MS11, page 274], [PAS12, Example 25], and [AHS13, Example 7.9 (3)].

Notice also that the term of free symmetrization has appeared before in the literature, but denoting a different notion, like in [HM10, Section 4]. It is unclear there is a connection between both notions.

It would be interesting to study Sym(μ)\mathrm{Sym}\left(\mu\right) in detail, due to its potential connections to the commutator. Below we present some basic properties, that are analogous to those of the of the symmetrization of polynomials.

Lemma 6.6.

Let μ,ν()\mu,\nu\in\mathcal{M}(\mathbb{R}) and α\alpha\in\mathbb{R}, and let δα\delta_{\alpha} be the Dirac measure with and atom of mass one in α\alpha. Then

  1. (1)

    Sym(μ)E\mathrm{Sym}\left(\mu\right)\in\mathcal{M}^{E};

  2. (2)

    Sym(μν)=Sym(μ)Sym(ν)\mathrm{Sym}\left(\mu\boxplus\nu\right)=\mathrm{Sym}\left(\mu\right)\boxplus\mathrm{Sym}\left(\nu\right);

  3. (3)

    Sym(Dilαμ)=DilαSym(μ)\mathrm{Sym}\left(\mathrm{Dil}_{\alpha}\mu\right)=\mathrm{Dil}_{\alpha}\mathrm{Sym}\left(\mu\right);

  4. (4)

    Sym(μδα)=Sym(μ)\mathrm{Sym}\left(\mu\boxplus\delta_{\alpha}\right)=\mathrm{Sym}\left(\mu\right);

  5. (5)

    𝑸(Sym(μ))(0)\bm{Q}(\mathrm{Sym}\left(\mu\right))\in\mathcal{M}(\mathbb{R}_{\geq 0}).

Proof.

The proof is analogous to that of Lemma 3.8; one just needs to substitute the polynomials for measures, and change n\boxplus_{n} to \boxplus. ∎

Alternatively one could prove Lemma 6.6 by letting nn\to\infty in Lemma 3.8, and noticing that the symmetrization of polynomials tends to the free symmetrization. Since this last fact will we useful later, let us state it explicitly.

Lemma 6.7.

Let 𝔭=(pm)m1𝒫(0)\mathfrak{p}=(p_{m})_{m\geq 1}\subset\mathcal{P}(\mathbb{R}_{\geq 0}) be a converging sequence of polynomials. Then Sym(𝔭):=(Sym(pm))m1𝒫E()\mathrm{Sym}\left(\mathfrak{p}\right):=(\mathrm{Sym}\left(p_{m}\right))_{m\geq 1}\subset\mathcal{P}^{E}(\mathbb{R}) is also a converging sequence of polynomials, and ρ(Sym(𝔭))=Sym(ρ(𝔭))\rho(\mathrm{Sym}\left(\mathfrak{p}\right))=\mathrm{Sym}\left(\rho(\mathfrak{p})\right).

Proof.

The claim follows from Theorem 2.14 and the fact that (Dil1pm)m1(\mathrm{Dil}_{-1}p_{m})_{m\geq 1} is a converging sequence of polynomials with limit Dil1ρ(𝔭)\mathrm{Dil}_{-1}\rho(\mathfrak{p}). ∎

As a direct application of this lemma we can take limits in Table 1 to obtain their analogues in free probability. We present these results in Table 4.

Measure μ\mu Sym(μ)\mathrm{Sym}\left(\mu\right)
MP ρ[.b.]\rho{\left[\genfrac{.}{.}{0.0pt}{1}{b}{\cdot}\right]} Dil2ρE[.2b, 1.]\mathrm{Dil}_{2}\rho^{E}{\left[\genfrac{.}{.}{0.0pt}{1}{2b,\,1}{\cdot}\right]}
RMP ρ[.a.]\rho{\left[\genfrac{.}{.}{0.0pt}{1}{\cdot}{a}\right]} DiliρE[.12a,a,a.]\mathrm{Dil}_{i}\rho^{E}{\left[\genfrac{.}{.}{0.0pt}{1}{1}{2a,\,a,\,a}\right]}
Free beta ρ[.ba.]\rho{\left[\genfrac{.}{.}{0.0pt}{1}{b}{a}\right]} ρE[.1, 2b, 2a2b2a,a,a.]\rho^{E}{\left[\genfrac{.}{.}{0.0pt}{1}{1,\,2b,\,2a-2b}{2a,\,a,\,a}\right]}
MP 2m\boxtimes_{2m} MP ρ[.b,d.]\rho{\left[\genfrac{.}{.}{0.0pt}{1}{b,\,d}{\cdot}\right]} Dil4ρE[.1, 2b, 2d,b+d,b+d2b+2d.]\mathrm{Dil}_{4}\rho^{E}{\left[\genfrac{.}{.}{0.0pt}{1}{1,\,2b,\,2d,\,b+d,\,b+d}{2b+2d}\right]}
RMP 2m\boxtimes_{2m} RMP ρ[.a,c.]\rho{\left[\genfrac{.}{.}{0.0pt}{1}{\cdot}{a,\,c}\right]} Dil827ρE[.23(a+c),23(a+c),23(a+c), 12a, 2c,a,c,a,c,a+c,a+c.]\mathrm{Dil}_{\frac{8}{\sqrt{-27}}}\rho^{E}{\left[\genfrac{.}{.}{0.0pt}{1}{\frac{2}{3}\left(a+c\right),\,\frac{2}{3}\left(a+c\right),\,\frac{2}{3}\left(a+c\right),\,1}{2a,\,2c,\,a,\,c,\,a,\,c,\,a+c,\,a+c}\right]}
Table 4. Symmetrization of SS-rational measures

Similarly, in Table 5 we compute the free additive convolution of some symmetric measures. These results follow from letting mm\to\infty in Table 2.

μ\mu ν\nu μν\mu\boxplus\nu
ρE[.1a1.]\rho^{E}{\left[\genfrac{.}{.}{0.0pt}{1}{1}{a_{1}}\right]} ρE[.1a2.]\rho^{E}{\left[\genfrac{.}{.}{0.0pt}{1}{1}{a_{2}}\right]} Dil2ρE[.a1+a22,a1+a22, 1a1,a2,a1+a2.]\mathrm{Dil}_{2}\rho^{E}{\left[\genfrac{.}{.}{0.0pt}{1}{\frac{a_{1}+a_{2}}{2},\,\frac{a_{1}+a_{2}}{2},\,1}{a_{1},\,a_{2},\,a_{1}+a_{2}}\right]}
Dilc1ρE[.1.]\mathrm{Dil}_{\sqrt{c_{1}}}\rho^{E}{\left[\genfrac{.}{.}{0.0pt}{1}{1}{\cdot}\right]} Dilc2ρE[.1.]\mathrm{Dil}_{\sqrt{c_{2}}}\rho^{E}{\left[\genfrac{.}{.}{0.0pt}{1}{1}{\cdot}\right]} Dilc1+c2ρE[.1.]\mathrm{Dil}_{\sqrt{c_{1}+c_{2}}}\rho^{E}{\left[\genfrac{.}{.}{0.0pt}{1}{1}{\cdot}\right]}
ρE[.b1, 1.]\rho^{E}{\left[\genfrac{.}{.}{0.0pt}{1}{b_{1},\,1}{\cdot}\right]} ρE[.b2, 1.]\rho^{E}{\left[\genfrac{.}{.}{0.0pt}{1}{b_{2},\,1}{\cdot}\right]} ρE[.b1+b2, 1.]\rho^{E}{\left[\genfrac{.}{.}{0.0pt}{1}{b_{1}+b_{2},\,1}{\cdot}\right]}
ρE[.b1+b2a, 1.]\rho^{E}{\left[\genfrac{.}{.}{0.0pt}{1}{b_{1}+b_{2}-a,\,1}{\cdot}\right]} ρE[.ab1,ab2, 1a.]\rho^{E}{\left[\genfrac{.}{.}{0.0pt}{1}{a-b_{1},\,a-b_{2},\,1}{a}\right]} ρE[.b1,b2, 1a.]\rho^{E}{\left[\genfrac{.}{.}{0.0pt}{1}{b_{1},\,b_{2},\,1}{a}\right]}
ρE[.b1,b2, 1b1+b2.]\rho^{E}{\left[\genfrac{.}{.}{0.0pt}{1}{b_{1},\,b_{2},\,1}{b_{1}+b_{2}}\right]} ρE[.b1,b2, 1b1+b2.]\rho^{E}{\left[\genfrac{.}{.}{0.0pt}{1}{b_{1},\,b_{2},\,1}{b_{1}+b_{2}}\right]} ρE[.2b1, 2b2, 12(b1+b2).]\rho^{E}{\left[\genfrac{.}{.}{0.0pt}{1}{2b_{1},\,2b_{2},\,1}{2(b_{1}+b_{2})}\right]}
Table 5. Free additive convolutions of even measures

Notice that Row 2 in Table 5 is just the well-known fact that the free convolution of semicircular measures yields another semicircular measure.

Also, recall that if μE()\mu\in\mathcal{M}^{E}(\mathbb{R}) than Sym(μ)=μμ\mathrm{Sym}\left(\mu\right)=\mu\boxplus\mu. Thus, when we let μ=ν\mu=\nu in Table 5 we obtain results on the symmetrization of even measures.

A particular case, that will be useful later, is when we let a1=a2=1a_{1}=a_{2}=1 in Row 1 of Table 2. This corresponds to studying the limit of Example 4.16, and recovers the well-known fact that the additive convolution of Bernoulli distributions gives an arcsine distribution.

6.2. Examples of commutators

Example 6.8 (Continuation of Example 5.9).

It was established in Example 5.9 that with B2m(x):=(x21)mB_{2m}(x):=(x^{2}-1)^{m}, we have

B2m2mB2m=Dil2mE[.1, 1112m, 2+1m.]𝒫2m().B_{2m}\,\square_{2m}\,B_{2m}=\mathrm{Dil}_{2}\mathcal{H}^{E}_{m}{\left[\genfrac{.}{.}{0.0pt}{1}{1,\,1}{1-\frac{1}{2m},\,2+\frac{1}{m}}\right]}\in\mathcal{P}_{2m}(\mathbb{R})\text{.}

From [NS98], it is known that with μ=12(δ1+δ1)\mu=\frac{1}{2}(\delta_{1}+\delta_{-1}), one has μμ=μμ\mu\,\square\,\mu=\mu\boxplus\mu. The computation of the latter [NS06, Example 12.8] is a standard example in free probability: μμ\mu\boxplus\mu is identified as the so-called arcsine distribution.

We can test our finite result by taking the limit of ERDs:

limmρ(B2m2mB2m)=Dil2ρE[.12.]=limmρ(Sym(B2m)).\lim_{m\to\infty}\rho(B_{2m}\,\square_{2m}\,B_{2m})=\mathrm{Dil}_{2}\rho^{E}{\left[\genfrac{.}{.}{0.0pt}{1}{1}{2}\right]}=\lim_{m\to\infty}\rho(\mathrm{Sym}\left(B_{2m}\right))\text{.}

Since B2mB_{2m} is even, Sym(B2m)=B2m2mB2m\mathrm{Sym}\left(B_{2m}\right)=B_{2m}\boxplus_{2m}B_{2m}, so we recover the corresponding result from free probability.

Example 6.9 (Continuation of Example 5.9).

Recall from Example 5.9 that

R2m(m)2mR2m(m)=Dil12mE[.1, 1112m, 2+1m.]R_{2m}^{(m)}\,\square_{2m}R_{2m}^{(m)}=\mathrm{Dil}_{\frac{1}{2}}\mathcal{H}^{E}_{m}{\left[\genfrac{.}{.}{0.0pt}{1}{1,\,1}{1-\frac{1}{2m},\,2+\frac{1}{m}}\right]}

In [NS98], it is shown that the commutator of free projections with trace 12\frac{1}{2} has the arcsine distribution on the interval [12,12][-\frac{1}{2},\frac{1}{2}]. This distribution may be described as a scaled version of the free additive convolution of the measure 12(δ1+δ1)\frac{1}{2}(\delta_{-1}+\delta_{1}) with itself. From the formula above, we have

limmρ(R2m(m)2mR2m(m))\displaystyle\lim_{m\to\infty}\rho\left(R_{2m}^{(m)}\,\square_{2m}R_{2m}^{(m)}\right) =limmρ(Dil12mE[.1, 1112m, 2+1m.])=Dil12ρE[.12.].\displaystyle=\lim_{m\to\infty}\rho\left(\mathrm{Dil}_{\frac{1}{2}}\mathcal{H}^{E}_{m}{\left[\genfrac{.}{.}{0.0pt}{1}{1,\,1}{1-\frac{1}{2m},\,2+\frac{1}{m}}\right]}\right)=\mathrm{Dil}_{\frac{1}{2}}\rho^{E}{\left[\genfrac{.}{.}{0.0pt}{1}{1}{2}\right]}\text{.}

Since this is the arcsine distribution, we recover the result from [NS98].

Example 6.10 (Continuation of Example 5.10).

In Example 5.10, we found that

H2m2mH2m=Dil1mmE[.2, 2, 112m2+1m.]𝒫2m().H_{2m}\,\square_{2m}\,H_{2m}=\mathrm{Dil}_{\frac{1}{m}}\mathcal{H}^{E}_{m}{\left[\genfrac{.}{.}{0.0pt}{1}{2,\,2,\,1-\frac{1}{2m}}{2+\frac{1}{m}}\right]}\in\mathcal{P}_{2m}(\mathbb{R})\text{.}

In [NS98], it is shown that the commutator of freely independent semicircular elements has the same distribution as the difference of two freely independent free Poisson elements. Our computation matches this in the limit: from Table 1, we have

Sym(L2m(1))=Dil1mmE[.2, 112m.]\mathrm{Sym}\left(L_{2m}^{(1)}\right)=\mathrm{Dil}_{\frac{1}{m}}\mathcal{H}^{E}_{m}{\left[\genfrac{.}{.}{0.0pt}{1}{2,\,1-\frac{1}{2m}}{\cdot}\right]}

so

limmρ(H2m2mH2m)=Dil1mρE[.2, 1.]=limmρ(Sym(L2m(1))).\lim_{m\to\infty}\rho(H_{2m}\,\square_{2m}\,H_{2m})=\mathrm{Dil}_{\frac{1}{m}}\rho^{E}{\left[\genfrac{.}{.}{0.0pt}{1}{2,\,1}{\cdot}\right]}=\lim_{m\to\infty}\rho\left(\mathrm{Sym}\left(L_{2m}^{(1)}\right)\right)\text{.}
Example 6.11 (Continuation of Example 5.11).

In Example 5.11, we found that

H2m2mR2m(m)=Dil12mmE[.1, 22+1m.]𝒫2m().H_{2m}\,\square_{2m}\,R_{2m}^{(m)}=\mathrm{Dil}_{\frac{1}{\sqrt{2m}}}\mathcal{H}^{E}_{m}{\left[\genfrac{.}{.}{0.0pt}{1}{1,\,2}{2+\frac{1}{m}}\right]}\in\mathcal{P}_{2m}(\mathbb{R})\text{.}

In [NS98], it is shown that the free commutator of a semicircular element and a projection with trace 1/21/2 has a semicircular distribution with radius 2\sqrt{2}. In the limit, the formula above gives

limmρ(H2m2mR2m(m))\displaystyle\lim_{m\to\infty}\rho(H_{2m}\,\square_{2m}\,R_{2m}^{(m)}) =limmρ(Dil12mmE[.1, 22+1m.])\displaystyle=\lim_{m\to\infty}\rho\left(\mathrm{Dil}_{\frac{1}{\sqrt{2m}}}\mathcal{H}^{E}_{m}{\left[\genfrac{.}{.}{0.0pt}{1}{1,\,2}{2+\frac{1}{m}}\right]}\right)
=limmρ(Dil12mmE[.1.])\displaystyle=\lim_{m\to\infty}\rho\left(\mathrm{Dil}_{\frac{1}{\sqrt{2m}}}\mathcal{H}^{E}_{m}{\left[\genfrac{.}{.}{0.0pt}{1}{1}{\cdot}\right]}\right)
=limmρ(Dil12mmE[.112m.])\displaystyle=\lim_{m\to\infty}\rho\left(\mathrm{Dil}_{\frac{1}{\sqrt{2m}}}\mathcal{H}^{E}_{m}{\left[\genfrac{.}{.}{0.0pt}{1}{1-\frac{1}{2m}}{\cdot}\right]}\right)
=limmρ(Dil12H2m)\displaystyle=\lim_{m\to\infty}\rho\left(\mathrm{Dil}_{\frac{1}{\sqrt{2}}}H_{2m}\right)

which is semicircular with radius 2\sqrt{2}.

Example 6.12 (Continuation of Example 5.12).

In Example 5.12, we found that

L2m(λ)2mL2m(μ)=Dil12m2mE[.2λ, 2μ, 2, 2, 112m2+1m.]𝒫2m().L_{2m}^{(\lambda)}\,\square_{2m}\,L_{2m}^{(\mu)}=\mathrm{Dil}_{\frac{1}{2m^{2}}}\mathcal{H}^{E}_{m}{\left[\genfrac{.}{.}{0.0pt}{1}{2\lambda,\,2\mu,\,2,\,2,\,1-\frac{1}{2m}}{2+\frac{1}{m}}\right]}\in\mathcal{P}_{2m}(\mathbb{R})\text{.}

Asymptotically, we then have

limmρ(Ln(λ)2mLn(μ))\displaystyle\lim_{m\to\infty}\rho\left(L_{n}^{(\lambda)}\,\square_{2m}\,L_{n}^{(\mu)}\right) =limmρ(Dil12m2mE[.2λ, 2μ, 2, 2, 112m2+1m.])\displaystyle=\lim_{m\to\infty}\rho\left(\mathrm{Dil}_{\frac{1}{2m^{2}}}\mathcal{H}^{E}_{m}{\left[\genfrac{.}{.}{0.0pt}{1}{2\lambda,\,2\mu,\,2,\,2,\,1-\frac{1}{2m}}{2+\frac{1}{m}}\right]}\right)
=limmρ(Dil12m2mE[.2λ, 2μ, 2, 1.]).\displaystyle=\lim_{m\to\infty}\rho\left(\mathrm{Dil}_{\frac{1}{2m^{2}}}\mathcal{H}^{E}_{m}{\left[\genfrac{.}{.}{0.0pt}{1}{2\lambda,\,2\mu,\,2,\,1}{\cdot}\right]}\right)\text{.}

The free commutator of free Poisson variables with parameters λ\lambda and μ\mu is not explicitly computed in [NS98], but the pieces are all there: the SS-transform of the square of this free commutator comes out to

4(z+2λ)(z+2μ)(z+2)(z+1)\frac{4}{(z+2\lambda)(z+2\mu)(z+2)(z+1)}

which matches the limiting SS-transform above.

Acknowledgements

This project originated during the 2023 Workshop in Analysis and Probability at Texas A&M University. Two of the authors had fruitful discussions during the conferences YMC*A and IWOTA in August 2024, and thank all the organizers.

D.P. was partially supported by the AMS-Simons Travel Grant, and appreciates the hospitality of University of Virginia during April 2024.

References

  • [AFPU24] Octavio Arizmendi, Katsuori Fujie, Daniel Perales, and Yuki Ueda. S-transform in finite free probability. Preprint arXiv:2408.09337, 2024.
  • [AGVP23] Octavio Arizmendi, Jorge Garza-Vargas, and Daniel Perales. Finite free cumulants: Multiplicative convolutions, genus expansion and infinitesimal distributions. Transactions of the American Mathematical Society, 376(06):4383–4420, 2023.
  • [AHS13] Octavio Arizmendi, Takahiro Hasebe, and Noriyoshi Sakuma. On the law of free subordinators. ALEA: Latin American Journal of Probability and Mathematical Statistics, 10(1):271–291, 2013.
  • [AP18] Octavio Arizmendi and Daniel Perales. Cumulants for finite free convolution. Journal of Combinatorial Theory, Series A, 155:244–266, 2018.
  • [APA09] Octavio Arizmendi and Victor Pérez-Abreu. The SS-transform of symmetric probability measures with unbounded supports. Proceedings of the American Mathematical Society, 137(9):3057–3066, 2009.
  • [BG09] Florent Benaych-Georges. Rectangular random matrices, related convolution. Probability Theory and Related Fields, 144:471–515, 2009.
  • [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.
  • [Cam22] Jacob Campbell. Commutators in finite free probability, I, 2022. arXiv:2209.00523 [math.CO].
  • [DJ02] K. Driver and K. Jordaan. Zeros of F23(n,b,cd,e;z){}_{3}F_{2}\left(\begin{array}[]{c}-n,b,c\\ d,e\end{array};z\right) polynomials. Numer. Algorithms, 30:323–333, 2002.
  • [GM22] Aurelien Gribinski and Adam W. Marcus. A rectangular additive convolution for polynomials. Combinatorial Theory, 2 (1), 2022.
  • [Gri13] Arcadii Z. Grinshpan. Generalized hypergeometric functions: product identities and weighted norm inequalities. The Ramanujan Journal, 31:53–66, 2013.
  • [HM10] Melanie Hinz and Wojciech Młotkowski. Multiplicative free square of the free poisson measure and examples of free symmetrization. In Colloquium Mathematicum, volume 119, pages 127–136. Instytut Matematyczny Polskiej Akademii Nauk, 2010.
  • [KLS10] Roelof Koekoek, Peter A. Lesky, and René F. Swarttouw. Hypergeometric orthogonal polynomials and their qq-analogues. Springer Monographs in Mathematics. Springer-Verlag, Berlin, 2010. With a foreword by Tom H. Koornwinder.
  • [KM16] Miklós Kornyik and György Michaletzky. Wigner matrices, the moments of roots of Hermite polynomials and the semicircle law. Journal of Approximation Theory, 211:29–41, 2016.
  • [Mar66] Morris Marden. Geometry of polynomials, volume No. 3 of Mathematical Surveys. American Mathematical Society, Providence, RI, second edition, 1966.
  • [Mar21] Adam W. Marcus. Polynomial convolutions and (finite) free probability, 2021. arXiv:2108.07054 [math.CO].
  • [MFMP24] Andrei Martínez-Finkelshtein, Rafael Morales, and Daniel Perales. Real roots of hypergeometric polynomials via finite free convolution. International Mathematics Research Notices, 06 2024.
  • [MFMP25] Andrei Martínez-Finkelshtein, Rafael Morales, and Daniel Perales. Zeros of generalized hypergeometric polynomials via finite free convolution: Applications to multiple orthogonality. Constructive Approximation, pages 1–70, 2025.
  • [MS11] Wojciech Młotkowski and Noriyoshi Sakuma. Symmetrization of probability measures, pushforward of order 2 and the boolean convolution. Banach Center Publications, 96:271–276, 2011.
  • [MS17] James A. Mingo and Roland Speicher. Free probability and random matrices, volume 35 of Fields Institute Monographs. Springer, New York; Fields Institute for Research in Mathematical Sciences, Toronto, ON, 2017.
  • [MSS22] Adam W. Marcus, Daniel A. Spielman, and Nikhil Srivastava. Finite free convolutions of polynomials. Probability Theory and Related Fields, 182(3-4):807–848, 2022.
  • [NS98] Alexandru Nica and Roland Speicher. Commutators of free random variables. Duke Mathematical Journal, 92(3):553–592, 1998.
  • [NS06] Alexandru Nica and Roland Speicher. Lectures on the combinatorics of free probability, volume 335 of London Mathematical Society Lecture Note Series. Cambridge University Press, Cambridge, 2006.
  • [OLBC10] Frank W. J. Olver, Daniel W. Lozier, Ronald F. Boisvert, and Charles W. Clark, editors. NIST handbook of mathematical functions. U.S. Department of Commerce, National Institute of Standards and Technology, Washington, DC; Cambridge University Press, Cambridge, 2010. With 1 CD-ROM (Windows, Macintosh and UNIX).
  • [PAS12] Victor Pérez-Abreu and Noriyoshi Sakuma. Free infinite divisibility of free multiplicative mixtures of the wigner distribution. Journal of Theoretical Probability, 25:100–121, 2012.
  • [Sze22] G. Szegő. Bemerkungen zu einem Satz von J. H. Grace über die Wurzeln algebraischer Gleichungen. Mathematische Zeitschrift, 13(1):28–55, 1922.
  • [VDN92] Dan V. Voiculescu, Ken J. Dykema, and Alexandru Nica. Free random variables, volume 1 of CRM Monograph Series. American Mathematical Society, Providence, RI, 1992.
  • [Wal22] Joseph L. Walsh. On the location of the roots of certain types of polynomials. Transactions of the American Mathematical Society, 24(3):163–180, 1922.