<<

arXiv:2007.11979v3 [math-ph] 2 May 2021 arxo optbesz.Wa a esi bu h itiuinof distribution the about said be can What size. product compatible a of h upr fteegnauso h sum the of eigenvalues Kie the classical Zhang, of the [17], of support versions Zhang the are and problems Forrester Horn multiplicative see randomized literature, the in recently nta,teei ra criyo xc omlsee o utt just that for Assume even follows. formulas exact orthog of scarcity with W great ensembles due Liu, is analogous are [2], For there formulas instead, Kieburg symmetry. exact and unitary of of Burda, existence e nature The Akemann, [29], fruitful therein. references Zhang a and and reg enabled Wang, limiting has Liu, numerous under formulas [28], results ense exact asymptotic Ginibre including of of part processes, presence products a expone as the as such Lyapunov and matrices, symmetry, networks, with unitary neural connection with in ensembles in gradients vanishing interest, and matr of random exploding are and [19], values networks neural singular 35], [18, systems dynamical efn pairs. Gelfand hs ievle r ie.Oernoie utpiaieHr prob matrix Horn product multiplicative the randomized of One given. are eigenvalues whose onpolm erfrterae oRf.[7 39]. [17, Refs. to reader the on refer progress we a problems For Horn matrices. unitary Haar-distributed diagonalizing where RDC ARXPOESSWT YPETCADORTHOGONAL AND SYMPLECTIC WITH PROCESSES MATRIX PRODUCT n ftepolm rsn ntecneto rdcso admma random of products of context the in arising problems the of One intersec the at seat special a enjoy matrices random of Products e od n phrases. and words Key ocrigtedsrbto ftesnua auso h product the of values singular the of distribution the Concerning T rdcso w ypetcmtie iharn etraiefac perturbative process. 1 point symm Pfaffian rank orthogonal a a form with and values matrices symplectic singular symplectic singula two to of joint matrices products the unitary perturb for truncated 1 formula rank orthog of the a Kieburg-Kuijlaars-Stivigny Haar as for acts truncated recent factor formulae a the latter by explicit the multiplied where obtain conditions matrix under we deterministic a proce connection, of Macdonald this values of a Through degenerations symplectic distributed are Corwin. Haar processes of matrix truncations product of products of values Abstract. X T a needn nre hs eladiaiayprsaeindepe are parts imaginary and real whose entries independent has hspolmi iia oternoie utpiaieHr rbesdis problems Horn multiplicative randomized the to similar is problem This ? eapysmercfnto hoyt td admpoessfo processes random study to theory function symmetric apply We NAINEVASMERCFUNCTIONS SYMMETRIC VIA INVARIANCE X C rdcso admmtie,snua au ttsis Macdonald , value singular matrices, random of Products safie arxwoesnua ausaekon n let and known, are values singular whose matrix fixed a is = A 2 1 NRWANADEGN STRAHOV EUGENE AND AHN ANDREW BA 1 2 where 1. Introduction A C and = A 1 B + r ie yterfie ievle,adtheir and eigenvalues, fixed their by given are B ftofie emta matrices Hermitian fixed two of to.Cneunl,w generalize we Consequently, ation. ssitoue yBrdnand Borodin by introduced sses epouto w matrices. two of product he dotooa arcs These matrices. orthogonal nd agl otermral special remarkably the to largely o,w hwta h squared the that show we tor, ino ueosaesincluding areas numerous of tion au est faproduct a of density value r tycass pcaiigto Specializing classes. etry cs npriua,tesquared the particular, In ices. nlo ypetcsymmetries symplectic or onal poainit hs random these into xploration h utpiaierandomized multiplicative the ms e ujar n Zhang and Kuijlaars see imes, ug n orse 3] The [39]. Forrester and burg, nlo ypetcmatrix symplectic or onal be n rnae unitary truncated and mbles X T onpolm[0 nfinding on [20] problem Horn cesse.Fravreyof variety a For system. icle itiuino singular of distribution t ndnmclsystems, dynamical in nts rcscnb omltdas formulated be can trices e st td eigenvalues study to is lem h iglrvle fthe of values singular the h ipetcs sthat is case simplest the mdb singular by rmed n,adWn [30] Wang and ang, dn n aea have and ndent ymti functions, symmetric T earandom a be A and cussed B 2 ANDREW AHN AND EUGENE STRAHOV standard normal distribution, i.e. T is a complex Ginibre matrix. In this situation the joint probability density function of the squared singular values can be explicitly computed, see Kuijlaars and Stivigny [27], Lemma 2.2, and references therein to related works. Another known case is where T is a submatrix (or truncation) of a Haar distributed . In this case the squared singular values of TX are distributed by a polynomial ensemble, see Theorem 2.1 in Kieburg, Kuijlaars, and Stivigny [24]. The results mentioned above can be interpreted as those for the Markov transition kernel, from the squared singular values of a deterministic matrix X to the squared singular values of TX, where T is a . The explicit formulae for such Markov kernels can be used to study products of random matrices, and, in particular, product matrix processes. Indeed, let us assume that X is a random matrix such that its squared singular values form a polynomial ensemble, and that T is a complex Ginibre matrix independent from X. Then the explicit formula for the relevant Markov transition kernel implies that the squared singular values of TX form a polynomial ensemble as well, see Theorem 2.1 in Kuijlaars [25]. The subsequent application of Theorem 2.1 in Kuijlaars [25] gives the distribution of the squared singular values for the product of an arbitrary number of independent Ginibre matrices (first obtained in the papers by Akemann, Kieburg, and Wei [4], and by Akemann, Ipsen, and Kieburg [5]). Now, consider sequences T T m of { l ··· 1}l=1 products of independent Ginibre matrices instead of fixed products, assume that each Tl is of size (n + νl) (n + νl 1) (where l = 1,...,m; ν0 = 0, ν1 0,...,νm 0), and for each l = 1,...,m × l − ≥ ≥ denote by yj, j =1,...,n, the squared singular values of Tl ...T1. The configuration

l l,yj l =1,...,m; j =1,...,n    R of all these singular values generates a random point process on 1,...,m >0 is called the Ginibre product process in Strahov [36]. The application of the Markov{ transition} × kernel to this l process gives the joint distribution of yj as a product of . The density formula implies that the Ginibre product matrix process is a discrete-time determinantal process, and accesses various asymptotic results, see Strahov [36] for further details. Matrix products with truncated unitary matrices can be studied in a similar way. An explicit formula for the Markov transition kernel, from the squared singular values of a deterministic matrix X to the squared singular values of TX where T is a truncated unitary matrix (see Kieburg, Kuijlaars, and Stivigny [24], Theorem 2.1) gives the joint densities of squared singular values for products of truncated unitary matrices. These densities have explicit determinantal forms in terms of Meijer G-functions, which leads to determinantal point processes formed by the squared singular values. Through the remarkable fact that the product matrix process with truncated unitary matrices can be understood as a scaling limit of the Schur process, see Borodin, Gorin, and Strahov [9], one can obtain determinantal formulas for (dynamical) correlation functions. The existence of exact formulas for matrix products mentioned above is due to the special nature of unitary symmetry. In this article we derive new formulas for the squared singular values of truncated orthogonal and symplectic matrices, and for product matrix processes formed by such truncations. By unitary, orthogonal and symplectic matrices we mean those taken from the classical unitary U(m), from the orthogonal group O(m), and from the compact Sp(2m), respectively. The orthogonal matrices can be obtained by the restricting unitary matrices from U(m) to real elements, and the symplectic matrices can be obtained by restricting unitary matrices from U(m) to real quaternion elements. If we use a 2 2 representation of the quaternions in terms of , then rank 1 pertubations for symplectic× matrices are PRODUCT MATRIX PROCESSES VIA SYMMETRIC FUNCTIONS 3 essentially rank 2 pertubations that respect the quaternion structure, and k r truncations are essentially 2k 2r truncations. × Our main result× is an explicit formula for the distribution of the squared singular values of a fixed matrix multiplied with a truncation of a Haar distributed orthogonal or symplectic matrix such that its number of rows is one less than that of the ambient Haar matrix, see Theorem 2.7 of the present paper. The condition on the number of rows effectively makes multiplication by this truncated Haar matrix behave as a rank 1 pertubation at the level of singular values, in the sense that the singular values of the initial matrix and new matrix interlace. Although this assumption on row numbers may seem limiting, we note that the Markov transition kernel for an arbitrary truncated Haar orthogonal/symplectic matrix can be obtained by a composition of Markov transition kernels of rank 1-pertubative type. Thus the rank 1 perturbations we study can be viewed as the elementary building blocks for general truncated matrix transitions. A subsequent application of Theorem 2.7 gives the joint law for the product matrix processes with truncated symplectic and orthogonal matrices, see Theorem 2.11. Theorem 2.11 is used to derive two formulas for the joint probability density of the squared singular values of matrix products constructed with truncated symplectic and orthogonal matrices, see Theorem 2.12 and Theorem 2.14. Theorem 2.12 presents the joint density as an integral where the of the integral depends only on the number of matrices in products, and Theorem 2.14 gives the joint probability density of the squared singular values in terms of a Jack symmetric function with the appropriate parameter associated to the symmetry class. We highlight a key structural result which we obtain in the symplectic case from our main result. We recast the two-product density for truncated symplectic matrices — again, where one factor is a rank 1-pertubative type — in a determinantal form in terms of Meijer G-functions. The structure of this indicates that it is a Pfaffian point process and we compute the correlation kernel via skew orthogonal polynomials. The methods we apply in the two product case are not stable under iteration, thus we are unable to see the Pfaffian structure for products of three or more matrices, if it exists. If there exists a Pfaffian structure for general products of truncated orthogonal and symplectic matrices, this may be due to special properties of the Jack functions in the parameters associated to these symmetry classes. Thus, one possible approach to find this structure in general may be exploit the Jack function representation of our density. We hope to explore this in a later work. Our entry point to these results is a connection between Macdonald processes introduced in Borodin and Corwin [8], and products of truncated orthogonal, unitary, and symplectic matri- ces. This connection is rooted in the fact that the zonal spherical functions associated to the Gelfand pairs GLn(R)/O(n), GLn(C)/U(n), GLn(H)/ Sp(2n) (where H denotes the skew field of real quaternions) are given by the Jack functions with appropriate parameters which are certain degenerations of the Macdonald symmetric functions. More concretely, the squared singular val- ues of these products can be realized as a degeneration of certain Macdonald processes. In [9], this degeneration was applied in the unitary case, and [6] considered this degeneration for the other symmetry classes. Through this connection, we produce an explicit formula for the Markov transition kernel from squared singular value of a deterministic matrix X to the squared singular values of a product TX where T is a truncated orthogonal or symplectic matrix, analogous to the result of [24, 25]. With the Markov transition kernel, we compute the density in terms of Cauchy-type and Vandermonde determinants. Using generalized Dixon-type integrals, we are able to derive an integral representation for the density where the dimension of the integral depends 4 ANDREW AHN AND EUGENE STRAHOV only on the number of products. By exploiting a variable-index symmetry in the Macdonald sym- metric functions, we derive an alternative formula for the density in terms of the Jack symmetric functions. Similar problems to those considered in this paper can be formulated for sums of random matri- T T T ces, and for the Hermitised products XM ...X2X1AX1 X2 ...XM , where each Xi is a real random matrix, and A is real antisymmetric, see Kuijlaars and Rom´an [26], Kieburg [22], Kieburg, For- rester, and Ipsen [23] and references therein for applications of the theory of spherical functions to these ensembles. An application of the theory of symmetric functions to such sums and products is a possible topic for a future research. While the focus of this article is on the structure and form of densities of the squared singular values of random matrix products, we mention a couple asymptotic applications. First, we note that the Dyson index β, where β =1, 2, 4 corresponds to the orthogonal, unitary, and symplectic symmetry classes respectively, can be generalized to arbitrary β > 0 in our model, extending the idea of β-ensembles to the setting of matrix products, see e.g. [3, Chapter 20]. This generalization is described in Section 7. By taking β , we find that the particles crystallize and the fluctuations of the particles are described→ by ∞correlated Gaussians, see Section 8. An analogous object was introduced by Gorin and Marcus [14] for the corners process and studied in the n limit by Gorin and Kleptsyn [13]. We expect that similar methods as in [13] may be employed→∞ to study the n of the β = product process. We mention→∞ one more application.∞ Recall that our Theorem 2.14 describes the density of squared singular values of matrix products in terms of the Jack function. With this density formula and integral representations for the Jack function [11], we can access the limiting behavior of the squared singular values in the regime where n is fixed but the number of products tends to . Here, we expect Lyapunov exponents to determine the asymptotic positions of the log singular∞ values, and the fluctuations to be Gaussian. This regime was considered for products of p-adic random matrices in [38] where the Lyaunov exponents and fluctuations were determined. A related regime for complex Ginibre and truncated unitary matrices was also considered in [2], [30], [6], where it was shown that if the matrix size n and the number of products p tend to infinity with n/p 0, then the largest log singular values tend to Lyapunov exponents with Gaussian fluctuations→ as well. We defer further details in this direction to a future article. The remainder of this article is organized as follows. In Section 2, we describe in more details the background behind our work and our main results. In Section 3, we prove our formula for the Markov transition and the first density formula. Section 4 derives the second density formula via generalized Dixon integration. Section 5 derives the third density formula in terms of Jack functions. We derive the density in the case of products of two truncated orthogonal matrices and the corresponding Pfaffian correlation kernel in Section 6. In Section 7, we give some details about the interpretation of our models for arbitrary β > 0. Lastly, in Section 8, we take the β limit of our model. → ∞ Acknowledgements. This work was supported by the BSF grant 2018248 “Products of random matrices via the theory of symmetric functions”. A. Ahn was partially supported by National Science Foundation Grant DMS-1664619.

2. Formulation of the problem and the main results The starting point of the present research is the well-known fact that the distribution of the squared singular values of a truncation of a Haar distributed matrix taken from unitary, symplectic, or orthogonal group is given by the Jacobi ensemble from Random Matrix Theory. PRODUCT MATRIX PROCESSES VIA SYMMETRIC FUNCTIONS 5

2.1. The Jacobi ensembles related to truncated unitary, orthogonal, and symplectic matrices. In this paper we are dealing with random matrix ensembles related to classical compact groups U(m), O(m), and Sp(2m). Here U(m) denotes the group of m m unitary matrices over C, O(m) denotes the group of m m unitary matrices over R, and Sp×(2m) denotes the group of m m unitary matrices over the field× of real quaternions H. In what follows we refer to U(m) as to the× unitary group, to O(m) as to the orthogonal group, and to Sp(2m) as to the symplectic group. Let S be a Haar distributed matrix taken from the unitary group U(m), or from the orthogonal group O(m), or from the symplectic group Sp(2m). Let the integers k, r be chosen such that the condition 1 k,r m is satisfied. The submatrix T of S defined by ≤ ≤ S1,1 ... S1,r T = .  .  Sk,1 ... Sk,r   is called a k r truncation of S. × Proposition 2.1. Let ν be a positive integer, and assume that T is a (n + ν) n-truncation of a Haar distributed matrix S taken from the unitary group U(m), or from the orthogonal× group O(m), or from the symplectic group Sp(2m). In addition, assume that the condition m 2n + ν ≥ is satisfied. Then the distribution of the eigenvalues (x1,...,xn) of T ∗T is given by the probability measure on [0, 1]n defined by

(θ) Pn,Jacobi (x1,...,xn) dx1 ...dxn n (2.1) 1 2θ θ(ν+1) 1 θ(m 2n ν+1) 1 = x x (x ) − (1 x ) − − − dx ...dx . (θ) | j − k| j − j 1 n Zn,Jacobi 1 j

m1 n n − Γ(θj) Γ(θ (m 2n ν + j))Γ(θj) (2.2) Z(θ) = n!(Γ(θ))n − − , n,Jacobi Γ(θ(j + n)) (Γ(θ))2 j=ν1+1 j=1 Y Y θ = 1 in case T is a (n + ν) n-truncation of a Haar distributed matrix taken from the unitary 1 × group U(m), θ = 2 in case T is a (n + ν) n-truncation of a Haar distributed matrix taken from the orthogonal group O(m), and θ =2 in case× T is a (n + ν) n-truncation of a Haar distributed matrix taken from the symplectic group Sp(2m). × Proof. See Forrester [16], Section 3.8.3.  Proposition 2.1 is a fundamental fact of Random Matrix Theory. In particular, Proposition 2.1 implies that the eigenvalues (x1,...,xn) of T ∗T form a determinantal point process in case T is a truncation of a unitary matrix, or Pfaffian point process in case T is a truncation of an , or a matrix from Sp(2m). The correlation functions of such point processes can be found explicitly in terms of special functions, and different scaling limits can be obtained. The question arises whether it is possible to extend the results of Proposition 2.1 to products of truncated matrices. The paper by Kieburg, Kuijlaars, and Stivigny [24] gives an answer to this question in the case of truncation of matrices taken from U(m). 6 ANDREW AHN AND EUGENE STRAHOV

2.2. The Kieburg-Kuijlaars-Stivigny theorem, and some of its consequences. In Ref. [24] Kieburg, Kuijlaars, and Stivigny proved that the squared singular values of a fixed matrix multiplied with a truncation of a Haar distributed unitary matrix form a polynomial ensemble (in the sense of Kuijlaars [25]). Namely, the following Theorem holds true Theorem 2.2. Assume that S is a Haar distributed unitary matrix of size m m, and let T be a (n + ν) l truncation of S. In addition, let X be a nonrandom matrix of size×l n such that the conditions× × (2.3) 1 n l m, m n + ν +1 ≤ ≤ ≤ ≥ are satisfied, and such that the eigenvalues (x1,...,xn) of X∗X are pairwise distinct and nonzero. Then the vector (y1,...,yn) of eigenvalues of (TX)∗ (TX) has density n n m+n ν m n ν 1 n (y) const x− y det (x y ) − − − △ , j j k − j + j,k=1 (x) j=1 ! j=1 ! △ Y Y   where (x) = 1 i0 called the product matrix process associated with the truncated { } × unitary matrices. Proposition 2.1 (with θ = 1) and Theorem 2.2 lead to representation of the joint probability distribution of xk,...,xk (where 1 k p) in terms of the product of determinants. 1 n ≤ ≤ Theorem 2.4. Consider the product matrix process associated with truncated unitary matrices. k k The joint probability distribution of x1,...,xn is given by 1 (xp)  Zn,p △ p 1 − n νr+1 mr+1 n νr+1 1 n m +1 (2.6) r+1 r r+1 − − − r r det xj xk xj (xk) − × − + k,j=1 r=1 Y h n  i (1) 1 1 n det wk xj dx ...dx , × k,j=1 h i PRODUCT MATRIX PROCESSES VIA SYMMETRIC FUNCTIONS 7

p p where (x y)+ = max(0, x y), the Vandermonde determinant (x ) is defined by (x ) = −p p − l l l △ △ xj xi , for 1 l p we write dx = dx1 ...dxn, and 1 i

Theorem 2.5. Let U1, ..., Up be independent Haar distributed unitary matrices. Assume that the size of each U , 2 j p, is equal to m m , and denote by T the truncation of U of j ≤ ≤ j × j j j size (n + νj) (n + νj 1). If conditions (2.4), (2.5) are satisfied, then the squared singular values × − (y1,...,yn) of Y = Tp ...T1 have the joint probability density n 1 (p) (2.9) (yk yj) det wk (yj) , Zn,p − k,j=1 1 j

(j) νj mj n νj 1 (j 1) y dx (2.10) w (y)= x (1 x) − − − w − k − k x x Z0   for j =2,...,p.

(l) Remark 2.6. The function wk (x) can be expressed as a Meijer G-function, m n, ..., m n, m 2n + k w(l)(x)= c Gl,0 l 2 1 x k l l,l ν−, ..., ν−, ν −+ k 1  l 2 1 −  (2.11) l cl Γ(ν1 + k 1+ s) j=2 Γ(νj + s) s = − x− ds, 0

Then the joint distribution of the ordered eigenvalues (y1,...,yn) of (TX)∗(TX) is given by the probability measure

θ n 1 θ n θ(ν+1) 1 1 Γ(θ (ν + n + 1)) (y) 1 − y − (2.14) △ det i dy (Γ(θ))n Γ(θ(ν + 1)) (x) x y θ(ν+2) 1 i i j i,j=1! i=1 xi − △   −  Y supported in 0 y x ... y x < 1. ≤ 1 ≤ 1 ≤ ≤ n ≤ n Here θ =1 corresponds to the multiplication with a truncation of a Haar distributed unitary matrix, 1 θ = 2 corresponds to the multiplication with a truncation of a Haar distributed orthogonal matrix, and θ = 2 corresponds to the multiplication with a truncation of a Haar distributed symplectic matrix. Remark 2.8. For the case of the orthogonal group many ensembles were studied in multivari- ate statistics, see Muirhead [34]. However, to the best of our knowledge products of truncated orthogonal matrices were not considered previously, and our Theorem 2.7 is a new result. Remark 2.9. The condition m = n + ν + 1 means that the number of rows of T is one less than that of the ambient unitary matrix U. As a result, T behaves as a “rank-one” perturbation which is reflected by the support satisfying the interlacing condition.

Remark 2.10. (a) The joint distribution of the eigenvalues (y1,...,yn) can be also given by the probability measure (2.15) n n n θ(ν+1) 1 1 Γ(θ (ν + n + 1)) θ 1 (y) 0 yi − n xi yj − △ det (xk yj)+ dyi (Γ(θ)) Γ(θ(ν + 1)) | − | 2θ 1 − j,k=1 θ(ν+2) 1 i,j=1 ( (x)) − i=1 xi − Y △ h  i Y PRODUCT MATRIX PROCESSES VIA SYMMETRIC FUNCTIONS 9

n supported in [0, 1] . Indeed, for 0

n 0 1, if 0 n + ν + 1, then the methods of the present paper cannot be applied, and the extension of the Kieburg-Kuijlaars-Stivigny theorem to truncations of orthogonal or symplectic matrices is an open question.

In the same way as in the case of unitary matrices we introduce the product matrix processes associated with truncated orthogonal and truncated symplectic matrices. Let S1, ..., Sp be inde- pendent Haar distributed symplectic or orthogonal matrices. As in the case of unitary matrices (see Section 2.2) assume that the size of each matrix Sj,1 j p, is equal to mj mj, and denote ≤k ≤ k × by Tj the (n + νj) (n + νj 1) truncation of Sj. Let x1,...,xn be the vector of the squared singular values of T× ...T .− We will refer to the point process on 1,...,p R formed by k 1  { } × >0 k configurations k, xj k =1,...,p; j =1,...,n as to the product matrix process with truncated   symplectic/orthogonal  matrices .

Theorem 2.11. Consider the product matrix process with truncated unitary matrices (θ = 1), 1 or with truncated orthogonal matrices θ = 2 , or with truncated symplectic matrices (θ = 2), and suppose that the parameters n, ν , ..., ν satisfy the conditions 1 p  (2.17) m 2n + ν , 1 ≥ 1 and

(2.18) m = n + ν +1, 2 k p. k k ≤ ≤ k k Let us agree that ν0 =0. Then the joint probability distribution of x1,...,xn (where 1 k p) is given by the probability measure ≤ ≤  1 θ p n − n θ(νk+1) 1 1 1 xk − ( (xp))θ det i k 1 k θ(ν +2) 1 Zn,p,θ △  x − x  k 1 k k=2 i j ! i=1 x − − (2.19) Y − i,j=1 Y i  n   p θ θ(ν1+1) 1 θ(m1 2n ν1+1) 1  x1 x1 − 1 x1 − − − dxk × △ i − i i=1 k=1  Y   Y supported in point configurations satisfying the condition

k k 1 k k 1 (2.20) 0 x x − ... x x − 1 ≤ 1 ≤ 1 ≤ ≤ n ≤ n ≤ 10 ANDREW AHN AND EUGENE STRAHOV

k k k for 2 k p. Here for 1 k p we write dx = dx1 ...dxn, and Zn,p,θ denote the normalization constant≤ given≤ by the formula≤ ≤ (2.21) p m1 n n np Γ(θ (νk + 1)) − Γ(θj) Γ(θ (m1 2n ν1 + j))Γ(θj) Zn,p,θ = (Γ(θ)) − − 2 . Γ(θ (νk + n + 1)) Γ(θ(j + n)) Γ(θ) j=ν1+1 j=1 kY=2 Y Y Theorem 2.11 will be proved in Section 3.4. Theorem 2.11 can be used to derive the distribution of squared singular values for a product matrix formed by truncated orthogonal or symplectic p p matrices. Indeed, in order to obtain the distribution of (x1,...,xn) in Theorem 2.11 it is enough p p to integrate the joint probability distribution (2.19) over all configurations (x1,...,xn) with 1 k p 1. Such an integration can be performed due to certain identities equivalent to those≤ derived≤ − by Dixon [12]. The result is the following Theorem which will be proved in Section 4.

Theorem 2.12. Let X = Tp ...T1, where each Tj is the (n + νj) (n + νj 1) truncation of a Haar × − distributed unitary, orthogonal, or symplectic matrix Sj of size mj mj, 1 j p. Assume that conditions (2.17) and (2.18) are satisfied. In addition, assume that×ν ≤ν +1≤ > 0. Then the p+1 − 1 squared singular values (x1,...,xn) of X have the joint density

n 1 2θ θ,n,p θ(m1 2n ν1+p) 1 θ(ν1+1) 1 (2.22) ( (x ,...,x )) I (x ,...,x ) (1 x ) − − − x − , Z W △ 1 n m1;ν1,...,νp 1 n − i i n,p,θ n,p i=1 Y 1 where θ =1 in case T1,..., Tp are truncated unitary matrices, θ = 2 in case T1,..., Tp are truncated orthogonal matrices, or θ = 2 in case T1,..., Tp are truncated symplectic matrices. The constant Zn,p,θ is defined by equation (2.21), and the constant Wn,p is given by

p Γ(θ (m n ν )Γ(θ (ν ν + 1))) (2.23) W = 1 − − r r − 1 . n,p Γ(θ)n r+2Γ(θ (m 2n ν + r 1)) r=2 − 1 1 Y − − − θ,n,p The function Im1;ν1,...,νp (x1,...,xn) in equation (2.22) has the following integral representation

θ,n,p Im1;ν1,...,νp (x1,...,xn)

= dv1,1 dv2,1dv2,2 ...... dvp 1,1 ...dvp 1,p 1 (vp 1,1,...,vp 1,p 1) − − − △ − − − ΩZ1 ZΩ2 Z Z Ωp−1Z p 2 r+1 r r − (2.24) 2(1 θ) θ 1 θ(νr νr+1 1) ( (v ,...,v )) − v v − (v ) − − × △ r,1 r,r | r+1,i − r,j| r,i r=1 i=1 j=1 i=1 Y Y Y Y  p 1 n − θ(m1 2n ν1+p 1) θ(νp ν1+1) 1 θ (1 vp 1,i)− − − − vp 1,i− − xj vp 1,i − , × − − − | − − | i=1 j=1 Y Y  where the integration region Ω1 is defined by (2.25) Ω = < v < 0 , 1 {−∞ 1,1 } and the integration regions Ω , 2 r p 1, are defined by r ≤ ≤ − (2.26) Ωr = < vr,1 < vr 1,1 < vr,2 < vr 1,2 < vr,3 <...

Remark 2.13. Assume that p = 2 (the case corresponding to the product of two matrices). Then θ,n,p the function Im1;ν1,...,νp (x1,...,xn) takes the form

0 n θ,n,p=2 θ(m1 2n ν1+1) θ(ν2 ν1+1) 1 θ (2.27) I (x ,...,x )= dv(1 v)− − − v − − x v − , m1;ν1,ν2 1 n − | j − | Z j=1 −∞ Y and the density of the distribution of the squared singular values (x1,...,xn) of X = T2T1 is proportional to n 2θ θ(m1 2n ν1+2) 1 θ(ν1+1) ( (x ,...,x )) (1 x ) − − − x △ 1 n − i i i=1 (2.28) + Y ∞ n θ(m1 2n ν1+1) θ(ν2 ν1+1) 1 θ dv(1 + v)− − − v − − x + v − . × | i | i=1 Z0 Y We have found an alternative representation for the distribution of the squared singular values of the total product matrix X = Tp ...T1. Namely, under certain additional restrictions the joint singular value density of a product of truncated unitary, symplectic, or orthogonal matrices can be expressed using the Jack symmetric functions. This will be done in Section 5. The result is the following

Theorem 2.14. Let X = Tp ...T1, where each Tj is the (n + νj) (n + νj 1) truncation of a Haar × − distributed unitary, orthogonal, or symplectic matrix Sj of size mj mj, 1 j p. Assume that conditions (2.17) and (2.18) are satisfied. Set × ≤ ≤

p M = (m n ν ) , i − − i i=1 X and assume there exists a partition µ with l(µ) M such that the numbers ≤ θ (ν + 1) ,...,θ (m n) ,...,θ (ν + 1) ,...,θ (m n) 1 1 − p p − is a rearrangement of µ + θ(M 1) >µ + θ(M 2) >...>µ . 1 − 2 − M Then the squared singular values (x1 <...

M n 1 Jµ x1,...,xn, 1 − ; θ ˆ J (1M ; θ) Zn,p,θ µ  (2.29) n 2θ θ(M n)+θ 1 dxi (x x ) (1 x ) − − , × j − i − i x 1 i

the constant Zˆ is defined by • n,p,θ

p

p mi n n Γ θ (mi n νi) n + i Γ(θi) − Γ(θj) − − − (2.30) Zˆ =  i=1  . n,p,θ Γ(θ(j + n)) P Γ(θ) i=1 j=ν +1 i=1 Y Yi Y Remark 2.15. Assume that θ = p = 1. In this case M = m n ν , µ is a rectangular Young 1 − − 1 diagram with m1 n ν1 rows, and with ν1 +1 boxes in each row. Theorem 2.14 gives the following expression for the− distribution− of the squared singular values of the truncated unitary matrix

n m1 2n ν1 2 m1 2n ν1 dxi (2.31) const s x ,...,x ;1 − − (x x ) (1 x ) − − . µ 1 n j − i − i x 1 i

n m1 2n ν1 ν1+1 sµ x1,...,xn;1 − − = xj . j=1  Y Thus we obtain formula (2.1) (with θ = 1) from Theorem 2.14.

Comparing the results of Theorem 2.14 with those of Theorem 2.5 we obtain a representation of the Schur symmetric function (with a suitable choice of variables) in terms of Meijer G-functions. Namely, the following Corollary holds true

Corollary 2.16. Suppose µ is a partition such that l(µ) p 1 M and M p +1 n. Then ≤ − ≤ − ≥

M n s x ,...,x , 1 − 1 1 (2.32) µ 1 n = det [w (x )]n , s (1M ) n M n (x x ) j i i,j=1 µ  i=1 (1 xi) − j i − 1 i

Γ(M n p + 2)Γ(M n + j) w (x)= − − − j Γ(M n p +1+ j) (2.33) − − p,0 µ1 + M, ..., µp 1 + M p +2, M p n + j +1 Gp,p − − − − x . × µ1 + M 1, ..., µp 1 + M p +1, j 1  − − − − 

Remark 2.17. Consider the case where p = N k + 1. In this case l(λ) N k, and the − ≤ − functions wj(x) can be written as contour integrals,

N k s ( 1) − Γ(N k + j) x ds wj(x)= − − , 2πiΓ(j) [s (λ1 + N 1)] ... [s (λN k + N (N k))] [s j + 1] IΣ − − − − − − − PRODUCT MATRIX PROCESSES VIA SYMMETRIC FUNCTIONS 13 where 1 j k, and the contour Σ encloses all the singularities of the integrand. Taking this into account≤ we≤ see that equation (2.32) gives

N k N k k sλ x1,...,xk, 1 − ( 1) − Γ(N k + j) = − j=1 − s (1N ) k N k k λ  i=1 (1 xi) − (xj xi) Q j=1 Γ(j) − 1 i

The determinant in the integrand can be computed explicitly using the formula for the Cauchy determinant,

1 1 ... 1 s1 s1 1 s1 k+1 (s s ) (j i) 1 1− ... −1 j − i − s2 s2 1 s2 k+1 1 i

k k k Since (j i)= Γ(j), and j=1 Γ(N k + j)= j=1(N j)!, we obtain 1 i

N k k s x ,...,x , 1 − (N j)! λ 1 k = j=1 − s (1N ) k N k λ  i=1 (xi Q1) − (xj xi) − 1 i

2.4. The correlation functions for singular values of the product of two truncated symplectic matrices. Let X = T T , where T is the (n + ν ) n truncation of a Haar distributed 2 1 1 1 × symplectic matrix S1 of size m1 m1, and T2 is the (n + ν2) (n + ν1) truncation of a Haar distributed symplectic matrix S of× size m m . In this particular× case we find that the density 2 2 × 2 of squared singular values (x1,...,xn) of X can be written as a determinant. 14 ANDREW AHN AND EUGENE STRAHOV

(2) Proposition 2.18. The density Pn,Product (x1,...,xn) of squared singular values (x1,...,xn) of X = T2T1 is equal to 2n 1 1 ... x1 − . .  . .  2n 1 (2) 1 1 ... xn − (2.36) Pn,Product (x1,...,xn)= (2) det  (2) (2)  ,  W1 (x1) ... W2n (x1)  Zn,Product    . .   . .   (2) (2)   W (xn) ... W (xn)   1 2n  (2)   where Zn,Product is the normalization constant,

m1 n m (2) Γ(2(ν2 + 1)) − Γ(2j) (2.37) Zn,Product = Γ(2(m1 2n ν1 + j)) Γ(2j), Γ(2(ν2 + n + 1)) Γ(2(j + n)) − − j=ν1+1 j=1 Y Y and 2ν +2 2(m 2n +1)+ j 1 (2.38) W (2)(x)= G2,0 2 1 x . j 2,2 2ν +1− 2ν + j 1 −  2 1 − 

This density is over 0 < x1 <...

The proof of Proposition 2.18 will be given in Section 6.1. Assume that we have a probability measure on Rn which can be written as 1 (2.39) Pn (x1,...,xn) dx1 ...dxn = det (φj (xk) ψj (xk))1 k n, 0 j 2n 1 dx1 ...dxn, Zn ≤ ≤ ≤ ≤ − where φj, ψj are certain functions, and Zn is the normalizing constant. We will refer to (2.39) as to a symplectic-type ensemble. Define the correlation kernel, Kn(x, y), of the symplectic-type ensemble (2.39) as a 2 2 matrix valued kernel of the operator Kn for which the following condition is satisfied × n 2

(2.40) ... Pn (x1,...,xn) (1 + f (xi)) dx1 ...dxn = det (I + Knf) . i=1 ! Z Z Y 2 Here Kn denotes the operator on L and f is the operator of multiplication by that function. Equation (2.36) implies that the squared singular values of the product of two truncated symplectic matrices form a symplectic type ensemble. The standard methods of Random Matrix Theory enable us to obtain in Section 6.2 the following

(2) Proposition 2.19. The correlation kernel for the density Pn,Product (x1,...,xn) defined by equation (2.36) can be written as

(1,1) (1,2) K Kn,Product(x, y) Kn,Product(x, y) (2.41) n,Product(x, y)= (2,1) (2,2) , Kn,Product(x, y) Kn,Product(x, y) ! where 2n 1 − (1,1) 2,0 2ν2 +2 2(m1 2n +1)+ k Product l (2.42) Kn,Product(x, y)= G2,2 − x qk,l y , 2ν2 +1 2ν1 + k k,l=0   X

PRODUCT MATRIX PROCESSES VIA SYMMETRIC FUNCTIONS 15 (2.43) (1,2) Kn,Product(x, y) 2n 1 − 2,0 2ν2 +2 2(m1 2n +1)+ k Product 2,0 2ν2 +2 2(m1 2n +1)+ l = G2,2 − x qk,l G2,2 − y , − 2ν2 +1 2ν1 + k 2ν2 +1 2ν1 + l k,l=0     X 2n 1 − (2,1) k Product l (2.44) Kn,Product(x, y)= x qk,l y , k,lX=0 and 2n 1 − (2,2) k Product 2,0 2ν2 +2 2(m1 2n +1)+ l (2.45) Kn,Product(x, y)= x qk,l G2,2 − y . − 2ν2 +1 2ν1 + l k,l=0   X 2n 1 2n 1 Here QProduct = qProduct − is the inverse of CProduct = cProduct − defined by i,j i,j=0 i,j i,j=0  (j i) Γ( i + j +2 ν + 1) (2.46) cProduct = − 1 . i,j (2ν + i + 2)(2ν + j + 2) Γ(2(m 2n +1)+ i + j + 1) 2 2 1 − We see that in order to obtain explicit formulae for the matrix entries of the kernel Kn,Product(x, y) 2n 1 we need to find the inverse of the matrix CProduct = cProduct − defined by equation (2.46). This i,j i,j=0 can be done using the following result that will be proved in Section 6.3.  Proposition 2.20. The inverse of the matrix 2n 1 Γ(a + i + j) − C = (j i) − Γ(a + b + i + j + 1)  i,j=0 2n 1 is the matrix Q =(Qi,j)i,j=0− , where (2.47) n 1 k ( 1)i+jΓ(b + 1) − 24k (a + b +4l 1)(a + b +4k + 1)Γ(a +2l)Γ(a +1+2k) Qi,j = − 4l − a b a+1 b+1 Γ(a + i)Γ(a + j) 2 Γ(l + 1)Γ 2 + 2 + l Γ 2 + l Γ 2 + l Xk=0 Xl=0 a +1 b +1 a b Θ(k +1)Θ + k Θ + k Θ  + + k  × 2 2 2 2       2l 2k +1 Γ(a + b +2l + i 1)Γ(a + b +2k + j) × − i j      2l 2k +1 Γ(a + b +2l + j 1)Γ(a + b +2k + i) , − − j i     and where Θ(x)=Γ(x)/Γ(2x). Proposition (2.19) together with Proposition (2.20) enable us to give explicit formulae for the matrix entries of the kernel Kn,Product(x, y). Theorem 2.21. Set a =2ν +1, a =2ν +1, b =2(m 2n ν )+1, and define 1 1 2 2 1 1 − − 1 m m (a2 + i + 1)Γ(a1 + b1 + m + i 1) (2.48) P Product(x)= ( 1)i − xi, m − i Γ(a + i) i=0 1 X   16 ANDREW AHN AND EUGENE STRAHOV and (2.49) p Product j p (a2 + j + 1)Γ(a1 + b1 + p + j 1) 2,0 a2 +1 b1 + j + a1 Qp (y)= ( 1) − G2,2 y . − j Γ(a + j) a2 a1 + j 1 j=0   1  −  X K With these notation the matrix entries of the kernel n,Product(x, y) can be written as (2.50) n 1 k 4k (1,1) − 2 (a1 + b1 +4l 1)(a1 + b1 +4k + 1)Γ(a1 +2l)Γ(a1 +1+2k) K (x, y) =Γ(b1 + 1) − n,Product 24l a1 b1 a1+1 b1+1 Γ(l + 1)Γ 2 + 2 + l Γ 2 + l Γ 2 + l Xk=0 Xl=0 a +1 b +1 a b Θ(k +1)Θ 1 + k Θ 1 + k Θ  1 + 1 + k  × 2 2 2 2       QProduct(x)P Product(y) QProductP (y)Product , × 2l 2k+1 − 2k+1 2l (2.51)  n 1 k 4k (1,2) − 2 (a1 + b1 +4l 1)(a1 + b1 +4k + 1)Γ(a1 +2l)Γ(a1 +1+2k) K (x, y)= Γ(b1 + 1) − n,Product 24l a1 b1 a1+1 b1+1 − Γ(l + 1)Γ 2 + 2 + l Γ 2 + l Γ 2 + l Xk=0 Xl=0 a +1 b +1 a b Θ(k +1)Θ 1 + k Θ 1 + k Θ 1+ 1 + k   × 2 2 2 2       QProduct(x)QProduct(y) QProductQ (y)Product , × 2l 2k+1 − 2k+1 2l (2.52)  n 1 k − 24k (a + b +4l 1)(a + b +4k + 1)Γ(a +2l)Γ(a +1+2k) K(2,1) (x, y) =Γ(b + 1) 1 1 − 1 1 1 1 n,Product 1 4l a1 b1 a1+1 b1+1 2 Γ(l + 1)Γ 2 + 2 + l Γ 2 + l Γ 2 + l Xk=0 Xl=0 a +1 b +1 a b Θ(k +1)Θ 1 + k Θ 1 + k Θ  1 + 1 + k  × 2 2 2 2       P Product(x)P Product(y) P ProductP (y)Product , × 2l 2k+1 − 2k+1 2l and  (2.53) n 1 k − 24k (a + b +4l 1)(a + b +4k + 1)Γ(a +2l)Γ(a +1+2k) K(2,2) (x, y)= Γ(b + 1) 1 1 − 1 1 1 1 n,Product 1 4l a1 b1 a1+1 b1+1 − 2 Γ(l + 1)Γ 2 + 2 + l Γ 2 + l Γ 2 + l Xk=0 Xl=0 a +1 b +1 a b Θ(k +1)Θ 1 + k Θ 1 + k Θ 1+ 1 + k   × 2 2 2 2       P Product(x)QProduct(y) P ProductQ (y)Product . × 2l 2k+1 − 2k+1 2l The proof of this Theorem will be given in Section 6.4. 

3. Proofs of Theorem 2.7 and Theorem 2.11 3.1. The Markov kernel associated with the Macdonald measure on Young diagrams. In this section we use the notation of Macdonald [32]. Let Λ be the algebra of symmetric functions over the field of complex numbers C. Let Pλ(x; q, t), Qλ(x; q, t) where q, t (0, 1) denote the ordinary and dual (q, t)-Macdonald symmetric functions respectively indexed∈ by Young diagrams PRODUCT MATRIX PROCESSES VIA SYMMETRIC FUNCTIONS 17

λ. We will generally suppress the q and t and write Pλ(x), Qλ(x) or Pλ, Qλ unless the presence of the q and t is pertinent. The Macdonald symmetric functions form a for Λ. Suppose we have two sets of vari- ables A = (a1, a2,...) and B = (b1, b2,...). From the Cauchy identity for Macdonald symmetric functions we obtain that

(3.1) MMacdonald (λ; A, B)=1, λ Y X∈ where MMacdonald (λ; A, B) is defined by 1 1 (3.2) M (λ; A, B)= P (A) Q (B)= Q (A) P (B) , Macdonald Π(A; B) λ λ Π(A; B) λ λ and

∞ (taibj; q) i 1 (3.3) Π (A; B)= ∞ , (u; q) = 1 q − u . (a b ; q) ∞ − i,j i j i=1 Y ∞ Y  If both sets of variables A = (a1, a2,...) and B = (b1, b2,...) give positive specializations of the algebra Λ of symmetric functions then MMacdonald (λ; A, B) can be understood as a probability measure on the set of all Young diagrams Y. We emphasize that MMacdonald (λ; A, B) depends on the Macdonald parameters q and t. In what follows we will refer to MMacdonald (λ; A, B) as to the Macdonald measure on Young diagrams. The skew Macdonald symmetric functions Pλ/µ, Qλ/µ are defined by

(3.4) Pλ (A, B)= Pλ/µ (A) Pµ (B) , Qλ (A, B)= Qλ/µ (A) Qµ (B) , µ Y µ Y X∈ X∈ as identities on Λ Λ. Equations (3.4) suggest to introduce the Markov kernel K (λ,µ; A, B) ⊗ Markov for the Macdonald measure MMacdonald (λ; A, B) by the formula

1 Pλ (B) 1 Qλ (B) (3.5) KMarkov (λ,µ; A, B)= Qλ/µ (A)= Pλ/µ (A) , Π(A; B) Pµ (B) Π(A; B) Qµ (B) where in the last equality we have exploited the relationship P ,P Q = h µ µiP . λ/µ P ,P λ/µ h λ λi Here ., . is the scalar product in the algebra of symmetric functions defined as in Macdonald [32], Chapterh i VI, equation (1.5). Using equations (3.4) we obtain (3.6) K (λ,µ; C, B) M (µ; A, B)= M (λ; C A, B) , Markov Macdonald Macdonald ⊔ µ Y X∈ where C =(c1,c2,...) is an additional sequence of variables, and C A denotes the union of two collections C and A of independent variables. In addition, ⊔

(3.7) KMarkov (λ,µ; A, B)=1, λ Y X∈ as it follows from the summation formula

(3.8) Pλ/µ (B) Qλ/ν (A)=Π(A; B) Qµ/τ (A) Pν/τ (B) , λ Y τ Y X∈ X∈ 18 ANDREW AHN AND EUGENE STRAHOV see Macdonald [32], VI. 7. If A and B give positive specializations of the algebra Λ of symmetric functions then KMarkov (λ,µ; A, B) can be understood as a probability measure on the set of all Young diagrams Y parameterized by A, B, and µ. We note a remarkable identity which relates the variables and indices of Macdonald symmetric functions which will be used in the sequel. For any partitions λ, ν Y of length n, we have ∈ ≤ ν1 n 1 ν2 n 2 νn λ1 n 1 λ2 n 2 λn Pλ(q t − , q t − ,...,q ) Pν(q t − , q t − ,...,q ) (3.9) n 1 n 2 = n 1 n 2 , Pλ(q − , q − ,..., 1) Pν(q − , q − ,..., 1) see [32, Chapter VI, (6.6)]. 3.2. The convergence of the Markov kernel to the distribution of the squared singular values of TX. Assume that S is a Haar distributed matrix taken from the unitary group U(m) 1 (in this case we say that the Jack parameter θ is equal to 1), from the orthogonal group θ = 2 , or from the symplectic group (θ = 2). Let T bea(n + ν) l truncation of S, and assume that the conditions m 2n + ν and m n + ν + 1 are satisfied.× Let X be a non-random matrix of size l n with squared≥ singular values≥ (x ,...,x ) such that × 1 n 0 < x1 <...

If A is given by equation (3.10), then KMarkov (λ,µ ; A, B) is concentrated on the Young diagrams with n rows or less, so we can assume that λ has n rows at most. The next Proposition says that the distribution of the eigenvalues of (TX)∗(TX) can be obtained by a limiting procedure from the Markov kernel KMarkov (λ,µ; A, B).

Proposition 3.1. Let λ be a random Young diagram whose distribution is defined by KMarkov (λ,µ; A, B). Assume that the Macdonald parameters q and t depend on ǫ> 0, and are given by ǫ θ θǫ (3.11) q = q(ǫ)= e− , t = t(ǫ)=(q(ǫ)) = e− .

Let µ(ǫ)=(µ1(ǫ),...,µn(ǫ)) be a Young diagram with n rows, and suppose that the length of each row of µ depends on ǫ in such a way that the limits

µ1(ǫ) µn(ǫ) (3.12) lim (q(ǫ)) = x1,..., lim (q(ǫ)) = xn ǫ 0+ ǫ 0+ →   →   exist. In addition, let y(ǫ)=(y1(ǫ),...,yn(ǫ)) be a random configuration associated with the Young diagram λ as ǫλ1 ǫλn y1(ǫ)= e− ,...,yn(ǫ)= e− .

As ǫ 0+, the distribution of y1(ǫ), ..., yn(ǫ) will coincide with that of squared singular values of TX→. Proposition 3.1 follows from Proposition 3.9 in the previous work of the first author [6]. Since this is our crucial link with symmetric function theory, we provide an alternative proof here. We recall the Jack symmetric functions which are the workhorse behind this proposition. In the limit q, t 1 such that t = qθ, we define → Jλ(x1,...,xn; θ) = lim Pλ(x1,...,xn) q 1 → for λ Y, see [32, Chapter VI, Section 10]. The Jack functions for ℓ(λ) n form a basis for the space∈ of symmetric polynomials in n-variables. ≤ PRODUCT MATRIX PROCESSES VIA SYMMETRIC FUNCTIONS 19

There are three key ingredients. The first is a connection between the Jack functions and products of matrices. Lemma 3.2. Suppose V is a Haar distributed O(n), U(n), Sp(2n) for θ = 1/2, 1, 2 respectively. Let W and X be deterministic n n matrices. Then × Jκ (X∗X; θ) Jκ (W ∗W ; θ) E ∗ ∗ ∗ (3.13) Jκ (X V W WVX; θ)= n , Jκ (1 ; θ) where κ Y, ℓ(κ) n, and Jκ (A; θ) denotes the value of the Jack polynomial on the eigenvalues of A. ∈ ≤ Proof. Equation (3.13) follows immediately from the functional relation for zonal spherical func- tions, and from the fact that the zonal spherical functions for the relevant Gelfand pairs can be written in terms of symmetric functions. Indeed, consider first the case corresponding to θ = 1. Let G = GL (n; C), and let K be the unitary group U(n). It is known that (G,K) is a Gelfand pair, see [32, Chapter VII (§5)]. Denote by Ωλ(x), x G, the zonal spherical function associated with (G,K). Then we have ∈

sλ (x∗x) (3.14) Ωλ(x)= n , x G, sλ (1 ) ∈ where by sλ (x∗x) we mean the Schur polynomial sλ evaluated on eigenvalues of x∗x. Since Ωλ(x) is the zonal spherical function it satisfies the functional relation

(3.15) Ωλ(xky)dk = Ωλ(x)Ωλ(y),

UZ(n) for all x, y GL(n, C). This relation can be rewritten more explicitly as ∈ sλ (x∗x) sλ (y∗y) (3.16) sλ ((xky)∗(xky)) dk = n , sλ (1 ) UZ(n) which is equivalent to the statement of the Lemma for θ =1. If G = GL (n, R), K = O(n) (the 1 H case corresponding to θ = 2 ), or G = GL(n, ), K = Sp(2n) (the case corresponding to θ = 2), then the zonal spherical functions for the Gelfand pairs (GL(n, R),O(n)), (GL(n, H),Sp(2n)) are given by the same equation (3.14) with the Schur polynomials replaced by the Jack polynomials with θ =1/2 or θ = 2 respectively, see [32, Chapter VII (3.24)] and [32, Chapter VII (6.20)]. As a result, one gets θ = 1/2 and θ = 2 analogues of equation (3.16) which can be interpreted as in the statement of Lemma 3.2.  The second key ingredient is a set of integral formulas known as the Selberg integral and its generalizations. These identities can be found in [16, (12.3),(12.46),(12.143)]. Lemma 3.3. For any a, b, θ > 0, we have 1 1 n S (a, b, θ) := u u 2θ (u )a(1 u )bdx n ··· | j − k| j − j j 0 0 1 j k n j=1 Z Z ≤Y≤ ≤ Y n 1 − Γ(a +1+ jλ)Γ(b +1+ jλ)Γ(1 + (j + 1)θ) = . Γ(a + b +2+(N + j 1)θ)Γ(1 + θ) j=0 Y − 20 ANDREW AHN AND EUGENE STRAHOV

Furthermore if κ Y, then ∈ 1 n 1 a b 2θ du1 duN (uj) (1 uj) Jκ(u1,...,un; θ) uj uk Sn(a, b, θ) ··· − | − | 0 j=1 1 j

Lemma 3.4. Let n, n , n , m Z such that n n m, i = 1, 2. Suppose T and T are 1 2 ∈ + ≤ i ≤ respectively n2 n1 and n2 n truncations of a random Haar O(m), U(m), or Sp(2m) matrix if × × 1/2 N θ = 1/2, 1, 2 respectively. Let X be a fixed n n matrix and X = (X∗X) . If σ(A) e R 1 × ∈ denotes the singular values of a matrix A, then σ(TX) =d σ(T X). e Remark 3.5. The result of Lemma 3.4 was proved in a more general setting in Ipsen and Kieburg e e [21]. We give an independent proof below for the reader’s convenience.

Proof. Let Pa b denote the a b matrix with the min(a, b) min(a, b) in the upper left corner and× 0 elsewhere. The× singular value decomposition× of X gives

X = UPn1 nΣV ∗ × where Σ = diag(σ(X)), U, V are Haar distributed orthogonal, unitary, or symplectic matrices depending on whether θ =1/2, 1, or 2 with the appropriate dimensions. Then

(TX)∗(TX)= V ΣPn n1 U ∗T ∗TUPn1 nΣV ∗. × × Observe that d d Pn n1 U ∗T ∗TUPn1 n = T ∗T = V ∗T ∗TV × × using the fact that the distributions of T and T are invariant under right translation by orthogonal, unitary, symplectic matrices for θ =1/2, 1, 2 respectively.e e Thuse e

d e (TX)∗(TX) = V ΣV ∗T ∗TV ΣV ∗ =(T X)∗T X which proves the lemma.  e e e e e e We are now ready to prove Proposition 3.1. Proof of Proposition 3.1. We first recast the statement of Proposition 3.1 in terms of square ma- trices. By Lemma 3.4, the distribution for the squared singular values of TX is the same as that of T X where 1/2 X =(X∗X) e e and T is a (n + ν) n truncation of S. The squared singular values of X are (x ,...,x ) and the × e 1 n squared singular values (u1,...,un) of T are distributed as the Jacobi ensemble (2.1). By right invariance,e e e T =d TV,

e e PRODUCT MATRIX PROCESSES VIA SYMMETRIC FUNCTIONS 21 where V is a Haar distributed orthogonal, unitary, or symplectic matrix depending on θ = 1/2,

θ = 1, or θ = 2 respectively. By Lemma 3.2, the squared singular values (y1,...,yn) of T X satisfy J (x ,...,x ; θ) E κ 1 n E e e Jκ(y1,...,yn; θ)= n Jκ(u1,...,un; θ), Jκ(1 ; θ) where the expectation in the right-hand side is over the Jacobi ensemble. By Lemma 3.3, the latter expectation is

1 1 n θ(ν+1) 1 θ(m 2n ν+1) 1 2θ Jκ(u1,...,un; θ) j=1(uj) − (1 uj) − − − dui uj uk − ··· | − | Sn(θ(ν + 1) 1, θ(m 2n ν + 1) 1, θ) 0 0 1 j

lim E Jκ(y1(ǫ),...,yn(ǫ); θ)= E Jκ(y1,...,yn; θ), ǫ 0 → where y1(ǫ),...,yn(ǫ) in the left-hand side is the random configuration introduced in the statement of Proposition 3.1. Using the fact that

n 1 n 2 lim Pκ(z1t − (ǫ), z2t − (ǫ),...,zn; q(ǫ), t(ǫ)) = Jκ(z1,...,zn; θ) ǫ 0 → uniformly over compact sets, we see that it is enough to show

n 1 (3.18) lim E Pκ(y1(ǫ)t − ,...,yn(ǫ); q, t)= E Jκ(y1,...,yn; θ). ǫ 0 → κ1 n 1 κn where E Jκ(y1,...,yn; θ) is given explicitly by equation (3.17). Let Cκ =(q t − ,...,q ) for any Young diagram κ with n rows, and compute

n 1 1 Pλ(B) E Pκ(y1(ǫ)t − ,...,yn(ǫ)) = Pκ(Cλ) Qλ/µ(A) Π(A; B) Pµ(B) λ Y X∈ 1 Pκ(B) = Pλ(Cκ)Qλ/µ(A) Π(A; B) Pµ(B) λ Y X∈ Π(A; Cκ) Pκ(B) Π(A; Cκ) = Pµ(Cκ)= Pκ(Cµ) Π(A; B) Pµ(B) Π(A; B) where A, B are defined by equation (3.10), the second equality uses (3.9), the third uses the specialization of (3.8) to the case ν = , and the final uses (3.9) again. We now observe that the right hand side converges to the right hand∅ side of (3.17). Thus we have shown (3.18), completing our proof.  22 ANDREW AHN AND EUGENE STRAHOV

3.3. Proof of Theorem 2.7. If m n = ν + 1 (as we have assumed in the statement of Theorem 2.7), then A turns into the list containing− tν+1 only. By Proposition 3.1, we establish the Markov kernel formula by showing that ν+1 n 1 (t ; q) Qλ(1,t,...,t − ) ν+1 ∞ (3.19) ν+n+1 n 1 Pλ/µ(t ) (t ; q) · Qµ(1,t,...,t − ) ∞ 1 1 ε θ converges to (2.14) as ε 0, where λi = ε− log yi, µi = ε− log yi′, q = e− , and t = q . A similar computation can→ be found in [10] for− the β-Jacobi corners− process. We use the following facts repeatedly: as q 1 we have → a a (q u; q) b a (q ; q) Γ(b) b a (3.20) (i) ∞ (1 u) − , (ii) ∞ ε − (qbu; q) → − (qb; q) → Γ(a) ∞ ∞ where the former holds uniformly over compact subsets of 0

λi λj j i Qλ f(q − t − ) (tu; q) = bλ = , f(u)= ∞ . λi λj+1 j i Pλ f(q − t − ) (qu; q) 1 i j ℓ(λ) ∞ ≤ ≤Y≤ We obtain n 1 θ 1 θ n n(1 θ) n f(1) (1 yi/yi+1) − (1 yi/yn) − f(1) ε − θ 1 b − − = (1 y ) − . λ ∼ (1 y /y )1 θ (1 y /y )1 θ ··· (1 y )1 θ (1 y )1 θ ∼ Γ(θ)n − i i=1 i i+1 − i i+2 − i − i=1 i − i=1 Y − − − Y − Y These asymptotics imply θn(n 1)/2+n(1 θ) n n 1 ε− − − Γ(θ(j i)) θ θ 1 (3.23) Q (1,...,t − ) − (y y ) (1 y ) − . λ ∼ Γ(θ)n Γ(θ(j i + 1)) j − i − i 1 i

ν+1 In order to find the asymptotics of Pλ/µ (t ) we use the combinatorial formula for the skew Macdonald symmetric functions Pλ/µ (see [32, Chapter VI, (6.19)]) representing these functions as sums over all column-strict (skew) tableaux of shape λ µ. Restricting Pλ/µ to the Macdonald polynomial in a single variable x, we obtain −

Pλ/µ(x; q, t)= ψλ/µ(x)δµ λ ≺ where δµ λ is equal to 1 in case condition (3.24) is satisfied, and is equal to zero otherwise. Here ≺

µi µj j i λi λj+1 j i λ µ f(q − t − )f(q − t − ) ψλ/µ(x)= x| |−| | , λi µj j i µi λj+1 j i f(q − t − )f(q − t − ) 1 i j ℓ(µ) ≤ ≤Y≤ where λ denotes the number of boxes in the Young diagram λ. Taking ℓ(λ)= ℓ(µ)= n, we may write | |

n λi n i µi µj j i λi λj j i 1 λ µ n f(q t − ) 1 f(q − t − )f(q − t − − ) ψλ/µ(x)= x| |−| |f(1) . f(qµi tn i) f(qλi µi ) f(qλi µj tj i)f(qµi λj tj i 1) i=1 − − 1 i

3.4. Proof of Theorem 2.11. Set X = T1 in the statement of Theorem 2.7, and apply Theorem 2.7 to find the joint density of squared singular values of T1, T2T1, T3T2T1, ..., Tp ...T1. Taking into account that the distribution of the squared singular values of T1 is given by equation (2.1) (θ)  (with the normalization constant Zn,Jacobi given by equation (2.2)), we obtain formula (2.19).

4. Proof of Theorem 2.12 We use an integration identity equivalent to one derived by Dixon [12]. A proof is provided in [16, Exercise 4.2.2] based on unpublished work by Eric Rains. 24 ANDREW AHN AND EUGENE STRAHOV

n Proposition 4.1. Let α0,...,αn, β0,...,βm have positive real parts, and suppose i=0 αi = m βj. Then j=0 P P n n m αj 1 βj dx dx x x a x − b x − 1 ··· n | j − i| | j − i| | j − i| Rn 1 i

a0 x1 a1 an 1 xn an, b0 x1 b1 bm 1 xm bm, ≤ ≤ ≤···≤ − ≤ ≤ ≤ ≤ ≤···≤ − ≤ ≤ respectively. By sending b , we obtain 0 → −∞ n m Corollary 4.2. Let α0,...,αn, β0,...,βm have positive real parts, and suppose i=0 αi = j=0 βj. Then n n m P P αj 1 βj dx dx x x a x − b x − 1 ··· n | j − i| | j − i| | j − i| Rn 1 i

a0 x1 a1 an 1 xn an, < x1 b1 bm 1 xm bm, ≤ ≤ ≤···≤ − ≤ ≤ −∞ ≤ ≤···≤ − ≤ ≤ respectively. To prove Theorem 2.12, observe that the p = 1 case is clear. For our induction step, we apply the Markov kernel θ(νp+1+1) 1 n n (p+1) − 1 Γ(θ(ν + n + 1)) ∆ y(p+1) θ 1 yi p+1 y(p) y(p+1) − n 2θ 1 i j θ(ν +1+2) 1 Γ(θ) Γ(θ(ν + 1)) (p) · −  (p)  p p+1 ∆(y ) − i,j=1 i=1 − yi Y Y to the density given by equation (2.22), integrating out y(p). We then apply Corollary 4.2 with (p+1) (p+1) a0 = y1 an 1 = yn an =1 ··· − , α0 = θ αn 1 = θ αn = θ(m1 2n ν1 + p) ··· − − − m = p, and

b0 = b1 = vp 1,1 bp 1 = vp 1,p 1 bp =0 −∞ − ··· − − − β0 = θ(m1 n νp+1) β1 = θ βp 1 = θ βp = θ(νp+1 ν1 + 1) − − ··· − − PRODUCT MATRIX PROCESSES VIA SYMMETRIC FUNCTIONS 25 where we require the condition m1 n > νp+1 so that β0 > 0. By our assumption, βp > 0. Thus θ,n,p+1 −  we obtain Im1;ν1,...,νp+1 as desired.

5. Proof of Theorem 2.14 Assume ν+1 m1 n νp+1 mp n A =(t ,...,t − ,...,t ,...,t − ), and suppose T1, T2,... be independent matrices where Ti is a (n + νi) (n + νi 1) truncation of × − a Haar distributed matricese taken from O(mi), U(mi), or Sp(2mi) for θ = 1/2, 1, 2 respectively. Further assume that n, mi, νi satisfy (2.4) and (2.5).

Proposition 5.1. Let λ be a random Young diagram whose distribution is defined by MMacdonald(λ; A, B). Assume that the Macdonald parameters q and t depend on ǫ> 0, and are given by ǫ θ θǫ e (5.1) q = q(ǫ)= e− , t = t(ǫ)=(q(ǫ)) = e− .

Let x(ǫ)=(x1(ǫ),...,xn(ǫ)) be a random configuration associated with the Young diagram λ as

ǫλ1 ǫλn x1(ǫ)= e− ,...,xn(ǫ)= e− .

As ǫ 0+, the distribution of x1(ǫ), ..., xn(ǫ) will coincide with that of squared singular values of T → T . p ··· 1 Proof. By iterative applications of Lemma 3.2 and Lemma 3.4, we have for any Young diagram κ with length n ≤ p Jκ(T ∗ T ∗Tp T1; θ) J (T T ; θ) E 1 ··· p ··· = E κ i∗ i J (1n; θ) J (1n; θ) κ i=1 κ Y where we used the fact that the distributions of Ti are invariant under right translation by Haar unitary matrices. Since the squared singular values of Ti are Jacobi distributed, Lemma 3.3 implies p n Jκ(T ∗ T ∗Tp T1; θ) Γ(θ(ν + n j +1)+ κ ) Γ(θ(m j + 1)) (5.2) E 1 ··· p ··· = i − j i − . J (1n; θ) Γ(θ(m j +1)+ κ ) Γ(θ(ν + n j + 1)) κ i=1 j=1 i j i Y Y − − Since P (u ,...,u ) J (u ,...,u ) κ 1 n → κ 1 n uniformly over u ,...,u [0, 1] as q, t 1 with t = qθ, it suffices to show that 1 n ∈ → λ1 n 1 λ2 n 2 λ P (q t , q t ,...,q n ) Jκ(T ∗ T ∗Tp T1; θ) E κ − − E 1 ··· p ··· n Pκ(B) → Jκ(1 ; θ) for every κ with length n where the expectation is with respect to MMacdonald(λ; A, B) — as in the proof of Proposition≤ 3.1, this follows from the Stone-Weierstrass theorem. Indeed, we have P (qλ1 tn 1, qλ2 tn 2,...,qλn ) P (qκ1 tn 1, qκ2 tn 2,...,qκn ) e E κ − − = E λ − − Pκ(B) Pλ(B)

1 κ1 n 1 κ2 n 2 κn = Qλ(A)Pλ(q t − , q t − ,...,q ) Π(A; B) λ Y X∈ Π(A; qκ1 tn 1, qκ2 ten 2,...,qκn ) = e − − Π(A; B) e e 26 ANDREW AHN AND EUGENE STRAHOV where the first equality follows from (3.9) and the last equality from the Cauchy identity for Macdonald symmetric functions. The right hand side of the above converges as ǫ 0 to the right hand side of (5.2). →  To complete the proof of Theorem 2.14 let us define λ and A as in Proposition 5.1 which states that ǫλ1 ǫλn x1(ǫ)= e− ,...,xn(ǫ)= e− e converges in distribution to the squared singular values of Tp T1. Thus we establish the theorem by showing that ··· 1 Pλ(A)Qλ(B) Π(A; B) 1 ǫ θ converges to (2.29) as ǫ 0, where λi = ǫ− log xi(eǫ), q = e− , and t = q . By our assumption → e µ1 M 1 µ2 M 2 µM A =(q t − , q t − ,...,q ) up to reordering. By (3.9), we have e λ1 M 1 λ2 n 2 λn M n M n 1 Pλ(A)Qλ(B) 1 Pµ(q t − , q t − ...,q t − , t − − ,..., 1) M 1 M 2 − − = M 1 M 2 Pλ(t , t ,..., 1)Qλ(B). Π(A; B) Π(A; B) Pµ(t − , t − ,..., 1) e Let λ(ǫ) be a family of Young diagrams with n rows. Assume that λ(ǫ) depends on a positive e e parameter ǫ in such a way that ǫλj(ǫ) log xj as ǫ 0+, for some values 0 < x1 ... xn < 1. Then → − → ≤ ≤ P (qλ1 tM 1, qλ2 tn 2 ...,qλn tM n, tM n 1,..., 1) J (x ,...,x , 1M n; θ) lim µ − − − − − = µ 1 n − ǫ 0+ P (tM 1, tM 2,..., 1) J (1M ; θ) →  µ − −  µ In addition, we use several asymptotics from the proof of Theorem 2.7. From (3.21) and (3.22), under the same assumptions on the family λ(ǫ) of Young diagrams as above we have n M 1 M 2 θ((M n)n+n(n 1)/2) Γ(θ(j i)) θ θ(M n) P (t − , t − ,..., 1) ǫ− − − − (x x ) (1 x ) − . λ ∼ Γ(θ(j i + 1)) j− i − i i

n p mr n n p mr n − j+i 1 − 1 (t − ; q) θMn Γ(θ(j + i)) = ∞ ǫ , (tj+i; q) ∼ Γ(θ(j + i 1)) Π(A; B) i=1 r=1 j=ν +1 i=1 r=1 j=ν +1 Y Y Yr ∞ Y Y Yr − n 1 and (3.23) gives the relevant asymptotics of Qλ(1,...,t − ). Combining the asymptotics above e 1 1 and the fact that dλi ǫ− xi− dxi, we obtain the desired result after simplifying the Gamma factors. ∼ 

6. The derivation of formulas for the singular values of the product of two truncated symplectic matrices In this Section we provide the derivations of different formulas stated in Section 2.4. These formulae describe the distribution of singular values of the product of two truncated symplectic matrices. We start from the proof of Proposition 2.18 which gives the density of squared singular values as a determinant. PRODUCT MATRIX PROCESSES VIA SYMMETRIC FUNCTIONS 27

6.1. Proof of Proposition 2.18. Let Y = T T , where T is the (n + ν ) n truncation of a Haar 2 1 1 1 × distributed symplectic matrix S1 of size m1 m1, and T2 is the (n + ν2) (n + ν1) truncation of a Haar distributed symplectic matrix S of size× m m . The distribution× of the squared singular 2 2 × 2 values (x1,...,xn) of T1 is the Jacobi ensemble with the parameters θ = 2, ν = ν1, and m = m1 given by equation (2.1). Clearly, the density of the squared singular values (y1,...,yn) of the product matrix T2T1 can be obtained using Theorem 2.7 (with θ = 2). Namely, the probability measure defined by equation (2.14) leads to a Markov kernel for the product matrix process formed by truncated symplectic matrices. It is not hard to see that this Markov kernel (for θ = 2) can be written as n Γ(2(ν + n + 1)) n (y) y2ν2+1 (6.1) 2 (x y ) △ det j 1[x y ] , Γ(2(ν + 1)) i − j (x)3 2ν2+3 i ≥ j 2 i,j=1 "xi # Y △ i,j=1 where 0 < x1 <...

We apply this kernel to the density of squared singular values of T1 given by equation (2.1) with the parameters θ = 2, ν = ν1, m = m1, and obtain

n ∞ ∞ 2ν2+1 n const yj ... dx1 ...dxn det [ϕj (y1) ,...,ϕj (yn) ,ϕj (x1) ,...,ϕj (xn)]1 j 2n det [ψi (xj)]i,j=1 , ≤ ≤ j=1 Y Z0 Z0 where

j 1 2(ν1 ν2 1) 2(m1 2n ν1)+1 ϕ (x)= x − , 1 j 2n; ψ (x)= x − − (1 x) − − 1 [y x 1] , 1 i n. j ≤ ≤ i − i ≤ ≤ ≤ ≤ Representing the determinants as sums over permutations we can rewrite the expression above as a single determinant, namely as

ϕ1(y) ... ϕ2n(y1) . . . . n   ϕ (y ) ... ϕ (y ) (6.3) const y2ν2+1 det 1 n 2n n . j  ∞ ∞   0 ψ1(t)ϕ1(t)dt . . . 0 ψ1(t)ϕ2n(t)dt  j=1  . .  Y  . .   R R   ∞ ψ (t)ϕ (t)dt . . . ∞ ψ (t)ϕ (t)dt   0 n 1 0 n 2n    Changing variables t = yi/τ, we have R R

1 2(ν1 ν2 1)+j 2(m1 2n ν1)+1 ∞ yi − − yi − − dτ ψi(t)ϕj(t)dt = 1 . 0 τ − τ τ Z yZi     Recall that b a b 1 x (1 x) − − a − 1 (x)= G1,0 x , Γ(a b) [0,1] 1,1 b   −

28 ANDREW AHN AND EUGENE STRAHOV see, for example, Luke [31], Section 6.6, equation (3). Therefore,

2ν1+j 1 2(m1 2n ν1)+1 yi − yi − − yi 1,0 2(m1 2n)+ j +1 yi 1 1[0,1] = Γ(2(m1 2n ν1 + 1)) G1,1 − , τ − τ τ − − j + ν1 1 τ  −        which gives

1

2ν2+1 ∞ dτ 2ν2+1 1,0 2(m1 2n)+ j +1 yi yi ψi(t)ϕj(t)dt = Γ(2(m1 2n ν1 + 1)) τ G1,1 − . − − τ j + ν1 1 τ Z0 Z  −  0

The following formula holds true 1 β α β 1 1,0 a y dx 2,0 α a x (1 x) − − G = Γ(α β)G y , − 1,1 b x x − 2,2 β b Z     0 see, for example, Luke [31], Section 5.6, equation (6). Taking this into account we see that the integrals in the determinant in expression (6.3) can be rewritten in terms of the corresponding Meijer G-functions. This gives the formula in the statement of Proposition 2.18. 

6.2. A matrix representation for the correlation kernel. Proof of Proposition 2.19.

6.2.1. The formula for the correlation kernel for a general symplectic-type ensemble. Proposition 2.18 implies that squared singular values (x1,...,xn) of X = T2T1 form a symplectic-type ensemble (2) in the sense of Section 2.4. Indeed, the density Pn,Product (x1,...,xn) of squared singular values (x1,...,xn) of X = T2T1 can be written as (2) Pn,Product (x1,...,xn) = const det (φj(xk) ψj(xk))1 k n, 0 j 2n 1 , ≤ ≤ ≤ ≤ − where 2ν +2 2(m 2n +1)+ j φ (x)= xj, ψ (x)= G2,0 2 1 x , j j 2,2 2ν +1− 2ν + j  2 1  and 0 j 2n 1. The next Proposition gives a 2 2 matrix representation for the correlation kernel≤ of a≤ general− symplectic-type ensemble. × Proposition 6.1. Consider a symplectic-type ensemble defined by probability measure (2.39), where φ (x), ψ (x) are certain functions, 0 j 2n 1, and Z is the normalizing constant. The j j ≤ ≤ − n correlation kernel Kn(x, y) of this ensemble defined by equation (2.40) can be written as

2n 1 2n 1 − − ψk(x)qk,lφl(y) ψk(x)qk,lψl(y) k,l=0 − k,l=0 (6.4) Kn(x, y)=   , 2Pn 1 2Pn 1 − φ (x)q φ (y) − φ (x)q ψ (y)  k k,l l − k k,l l   k,l=0 k,l=0  2n 1  P 2n 1 P  where Q =(qi,j)i,j=0− is the inverse of C =(ci,j)i,j=0− defined by

c = (φ (x)ψ (x) φ (x)ψ (x)) dx. i,j i j − j i Z Proof. See Tracy and Widom [37].  PRODUCT MATRIX PROCESSES VIA SYMMETRIC FUNCTIONS 29

6.2.2. Proof of Proposition 2.19. The distribution of squared singular values for the product of two truncated symplectic matrices is the symplectic-type ensemble (2.39) with 2ν +2 2(m 2n +1)+ j (6.5) φ (x)= xj, ψ (x)= G2,0 2 1 x , j j 2,2 2ν +1− 2ν + j  2 1  2n 1 where 0 j 2n 1. The matrix C = (c ) − in Proposition 6.1 can be computed explicitly, ≤ ≤ − i,j i,j=0 the computation gives equation (2.46). 

6.3. A general formula for the inverse of a skew-symmetric Hankel-type matrix. Proof of Proposition 2.20. In order to obtain explicit formulae for the matrix entries of the kernel 2n 1 K Product Product − n,Product(x, y) of Proposition 2.19 we need to find explicitly the inverse of C = ci,j i,j=0 Product defined by equation (2.46). The matrix C can be understood as a skew-symmetric Hankel type matrix, see Definition 6.2 below. In this Section we first derive a general formula for the inverse of a skew-symmetric Hankel type matrix, see Proposition 6.3 below. Then we apply Proposition 6.3 to derive the formulae stated in Proposition 2.20. Definition 6.2. Let µ be a positive measure on R with finite moments,

k (6.6) hk = x µ(dx), k =0, 1, 2,.... ZR We will refer to the skew- H of size 2n 2n defined by × 2n 1 (6.7) H = ((j i)hi+j 1) − − − i,j=0 as to a skew-symmetric Hankel type matrix. Here we find a general formula for the inverse of H. Proposition 6.3. Let ., . denote the skew inner product defined in terms of µ, h iµ 1 (6.8) f,g = (f(x)g′(x) g(x)f ′(x)) µ(dx). h iµ 2 − ZR n 1 Let q2k(x), q2k+1(x) k=0− be a system of skew orthogonal polynomials satisfying the following con- dition{ }

rk, if i =2k, j =2k +1 for some k 0, 1,...,n 1 , (6.9) q , q = r , if i =2k +1, j =2k for some k ∈{0, 1,...,n − 1}, h i jiµ  − k ∈{ − }  0, otherwise. Then the 2n 2n skew-symmetric matrix Q be defined by ×  2n 1 1 − (6.10) Q = (q q ) , 2 j,i − i,j  i,j=0 where 2n 1 − 1 1 di 1 dj (6.11) qi,j = i q2k(x) j q2k+1(y) , rk i! dx j! dy k=0   x=0  y=0 X is the inverse of the skew-symmetric Hankel-type matrix H defined by equations (6.6) and (6.7). 30 ANDREW AHN AND EUGENE STRAHOV

Proof. It can be checked that the kernel n 1 − 1 (6.12) Sn(x, y)= (q2k(x)q2k+1(y) q2k+1(x)q2k(y)) rk − Xk=0 has the reproducing property (6.13) yk = S (x, y), xk h n iµ Write n 1 2n 1 − − 1 i j j i (6.14) (q2k(x)q2k+1(y) q2k+1(x)q2k(y)) = qi,j x y x y . rk − − k=0 i,j=0 X X  The equation above defines the coefficients qi,j. Taking into account this equation we see that the inner product S (x, y), xk can be rewritten as h n iµ 2n 1 − k 1 i (6.15) Sn(x, y), x µ = (qi,j qj,i)(k i)hk+i 1y h i 2 − − − i,j=0 X which implies 2n 1 1 − (6.16) (i k)hk+i 1 (qj,i qi,j)= δk,j. 2 − − − i=0 X Therefore, the matrix Q defined by equations (6.10) and (6.14) is the inverse of the skew-symmetric Hankel-type matrix H defined by equations (6.6) and (6.7). Moreover, it is not hard to see that qi,j (defined by (6.14)) can be determined from equation (6.11)  If dµ(x)=(1 x)a+1(1 + x)b+1dx (this is a measure on [ 1, 1]), then we have − − 1 2a+b+k+3 Γ(a + k + 2)Γ(b + 2) (6.17) h = (x 1)kdµ(x)= , k − ( 1)k Γ(a + b + k + 4) Z1 − − where k =0, 1,....

Jacobi Proposition 6.4. Let qk be the family of the skew-orthogonal polynomials with respect to ., . , where µ is defined{ by equation} (6.17), and let h iµ n 1 − 1 (6.18) SJacobi(x, y)= qJacobi(x)qJacobi(y) qJacobi(x)qJacobi(y) n rJacobi 2k 2k+1 − 2k+1 2k k=0 k X  Jacobi be the corresponding reproducing kernel. Define the coefficients qi,j from the expansion 2n 1 − (6.19) SJacobi(x, y)= qJacobi (x 1)i(y 1)j (x 1)j(y 1)i . n i,j − − − − − i,j=0 X  2n 1 Then the matrix qJacobi = 1 qJacobi qJacobi − is the inverse of 2 j,i − i,j i,j=0 2n 1 2a+b+i+j+2 Γ(a + i + j + 1)Γ(b + 2) − CJacobi = (j i) . − ( 1)i+j 1 Γ(a + b + i + j + 3)  − − i,j=0 PRODUCT MATRIX PROCESSES VIA SYMMETRIC FUNCTIONS 31

Jacobi The coefficients qj,i can be written as

n 1 − 1 1 di 1 dj (6.20) qJacobi = qJacobi(x) qJacobi(y) . i,j rJacobi i! dxi 2k j! dyj 2k+1 k=0 k   x=1  y=1 X

Proof. The proof of Proposition 6.4 is very similar to that of Proposition 6.3. The only difference is that we use (y 1)k = SJacobi(x, y), (x 1)k − h n − iµ instead of (6.13).  Proposition 6.5. We have (6.21) n 1 k − 4k Jacobi 1 2 (a + b +4l + 1)(a + b +4k + 3)Γ(a +1+2l)Γ(a +2+2k) qi,j = a+b+i+j+1 4l a b a b 2 2 Γ(l + 1)Γ 2 + 2 + l +1 Γ 2 + l +1 Γ 2 + l +1 Xk=0 Xl=0 a b a b Θ(k +1)Θ + k +1 Θ + k +1 Θ + + k +1   × 2 2 2 2       Γ(a + b +2l + i + 1)Γ(a + b +2k + j + 2) 2l 2k +1 , × Γ(a + i + 1)Γ(a + j + 1) i j     where Θ(x)=Γ(x)/Γ(2x).

Jacobi ∞ Proof. The skew-orthogonal polynomials qk k=0 are given by (6.22)  k Γ a + b + k +1 Γ a + k +1 Γ b + k +1 Γ(a + b +4l + 2) pJacobi(x) qJacobi(x)=26kk! 2 2 2 2 2l , 2k Γ a + b + l +1 Γ a + l +1 Γ b + l +1 Γ(a + b +4k + 2) l!26l l=0 2 2  2  2  X and    Jacobi Jacobi (6.23) q2k+1 (x)= p2k+1 (x),

Jacobi ∞ see Adler, Forrester, Nagao, and van Moerbeke [1]. Here pk (x) k=0 are polynomials defined by  Γ(a + b + k + 1) (6.24) pJacobi(x)=2kk! P (a,b)(x), k Γ(a + b +2k + 1) k

(a,b) where Pk (x) are the Jacobi polynomials. Since k d (a,b) (n + a + b + 1)k (a+k,b+k) (6.25) k Pn (x)= k Pn k (x), dx 2 − and (a + 1) (6.26) P (a,b)(1) = n , n n! we obtain 1 di Γ(a +1+ k)Γ(a + b +1+ k + i) (6.27) P (a,b)(x) = , i! dxi k 2ii!(k i)!Γ(a + i + k)Γ(a + b +1+ k) x=1 −

32 ANDREW AHN AND EUGENE STRAHOV which gives 1 di k 26k 4l ik! Γ a + b + k +1 Γ a + k +1 Γ b + k +1 qJacobi(x) = − − 2 2 2 2 i! dxi 2k l! Γ a + b + l +1 Γ a + l +1 Γ b + l +1 x=1 l=0 2 2  2  2  (6.28) X Γ(a +1+2 l)Γ(a + b +4l + 2)Γ(a + b +2l + i + 1) 2l  ,   × Γ(a + b +4l + 1)Γ(a + b +4k + 2)Γ(a + i + 1) i   and j 1 d Jacobi 2k+1 j 2k +1 Γ(a +2k + 2)Γ(a + b +2k + j + 2) (6.29) q (y) =2 − j! dyj 2k+1 j Γ(a + j + 1)Γ(a + b +4k + 3) y=1  

Taking into account that

1 Γ(a + b +4k + 2)Γ(a + b +4k + 4) (6.30) Jacobi = a+b+4k+2 , rk 2 (2k + 1)!Γ(a +2k + 2)Γ(b +2k + 2)Γ(a + b +2k + 2) we obtain the formula in the statement of the Proposition.  Proof of Proposition 2.20. Proposition 2.20 is an immediate Corollary of Proposition 6.4 and Proposition 6.5. 

6.4. Proof of Theorem 2.21. Now we are ready to derive explicit formulae for the matrix entries of the correlation kernel Kn,Product(x, y) stated in Theorem 2.21. We use equations (2.42), (2.43), Product (2.44), and (2.45). All these formulae involve the coefficients qk,l which can be understood as 2n 1 Product Product − matrix entries of the inverse of C = ci,j i,j=0 defined by equation (2.46). Proposition Product 2.20 can be used to find the inverse of C explicitly. As a result, we obtain the desired formulae for the matrix entries of the correlation kernel Kn,Product(x, y). 

7. Extension to Arbitrary β > 0 In this section, we describe a generalization of our model to arbitrary θ > 0. We write β =2θ, to indicate the connection with β-ensembles from random matrix theory. Further details about these β-deformed models can be found in [14] and [6]. We begin by recalling the β-Jacobi ensemble with parameters m, ν. Suppose T is an (n + ν) n-truncation of a Haar distributed matrix S taken from U(m), O(m), Sp(2m) for β = 1, 2,×4 respectively. Recall that (see Proposition 2.1) the distribution on [0, 1]n defined by the density (β/2) Pn,Jacobi (x1,...,xn) dx1 ...dxn n (7.1) 1 β (β/2)(ν+1) 1 (β/2)(m 2n ν+1) 1 = x x (x ) − (1 x ) − − − dx ...dx , (β/2) | j − k| j − j 1 n Zn,Jacobi 1 j 0. The resulting distribution is referred to as the β-Jacobi ensemble and is among the three classical β-ensembles alongside the β-Laguerre and β-Hermite ensembles. For an introduction to β-ensembles and further literature, we refer the reader to [3, Chapter 20] and references therein. PRODUCT MATRIX PROCESSES VIA SYMMETRIC FUNCTIONS 33

We can go further and consider a β > 0 extension of the Markov transition kernel from Theorem 2.7. Recall that (see Theorem 2.7) if we assume m = n + ν + 1, then for a deterministic n n n × matrix X with squared singular values x =(x1 < < xn) [0, 1] , we have that the joint density of the ordered squared singular values y =(y < ··· 0, just as with the β-Jacobi ensemble. From another perspective, the density (7.3) can also be obtained by scaling limit of the transition KMarkov(λ,µ; A, B) as in Proposition 3.1 by taking θ = β/2. Indeed, the proof of Proposition 3.1 follows verbatim. By viewing (7.3) itself as a transition kernel, we can iterate the kernel and define a β > 0 generalization of the matrix product process:

1 1 1 n Definition 7.1. Fix β > 0, νk > 1 for k =2, 3,..., and x =(x1 < < xn) [0, 1] . Define the 1 ··· ⊂ β-Jacobi product process with initial state x and parameters (νk)k∞=2 to be the Markov process in k k 1 k (β/2) k k 1 discrete time (x )k∞=1 where the Markov transition kernel from x − to x is given by pn,νk (x x − ) as in (7.3). | The limit shape and fluctuations of the β-Jacobi product process, where the initial state was distributed as a β-Jacobi ensemble, were studied in [6]. We can further interpret the density (7.3) as coming from an operation on vectors which is some generalization of products of O(m), U(m), and Sp(2m) invariant matrices at the level of singular values. In particular, many properties from the β = 1, 2, 4 cases generalize readily to arbitrary β > 0. For details, we refer the reader to [6]. We indicate one of these properties here. By Lemma 3.4, the squared singular values of TX are distributed as the squared singular values 1/2 of (T ∗T ) X. Since m = n + ν + 1, the condition m 2n + ν is not satisfied. In particular, n 1 ≥ − of the eigenvalues of T ∗T must be 1, and only one eigenvalue is deterministic. Nonetheless, it is known (see e.g. Forrester [16], Section 3.8.3.) that if (u1, 1,..., 1) are the eigenvalues of T ∗T , then (β/2) the density of u1 is P1,Jacobi with parameters m = n + ν + 1 and ν. By Lemma 3.2, we have J (y; β/2) J (x; β/2) J (u , 1n 1; β/2) E κ = κ E κ 1 − , κ Y, ℓ(κ) n J (1n; β/2) J (1n; β/2) J (1n; β/2) ∈ ≤  κ  κ  κ  (β/2) where the left hand side expectation is over y ,...,y distributed as pn (y x) and the right hand 1 n | side expectation is over u1. It turns out that we can extend this identity to arbitrary β > 0. 34 ANDREW AHN AND EUGENE STRAHOV

n Proposition 7.2. Fix β > 0. If x = (x1 < < xn) [0, 1] and y = (y1 < < yn) is distributed as p (y x), then ··· ⊂ ··· n,ν | J (y; β/2) J (x; β/2) J (u , 1n 1; β/2) E κ = κ E κ 1 − , κ Y, ℓ(κ) n J (1n; β/2) J (1n; β/2) J (1n; β/2) ∈ ≤  κ  κ  κ  (β/2) where u1 is distributed as P1,Jacobi with parameters m = n + ν +1 and ν.

(β/2) Proof. From the discussion above, pn,ν (y x) is the limit of KMarkov(λ,µ; A, B) under the scaling limit in Proposition 3.1 with θ = β/2. Then| this proposition is a special case of [6, Proposition 3.9] (see the discussion following the proof in the reference). 

8. Crystallization and Gaussianity at β = ∞ In this section, we consider the β limit of the β-Jacobi product process introduced in Section 7. We find that the particles freeze→ ∞ at deterministic positions, but fluctuate as correlated Gaussians. We precisely describe this limiting object, the -Jacobi product process, in Section 8.3. This crystallization phenomenon and associated Gaussianity∞ also arise in the closely related -Hermite corners process introduced in [14]. We note that the n limit of the -Hermite ∞corners process was studied in [13]. For us, we consider β and keep→ ∞n (the number∞ of particles at each step in the Markov chain) fixed. However, we expect→∞ that the methods of [13] may be generalized to study the n limit of the -Jacobi product process. →∞ ∞

n n 8.1. Warmup: β Crystallization of the β-Jacobi Ensemble. Let := [0, 1] x1 x . We can→∞ express the density for the β-Jacobi ensemble (7.1) U ∩{ ≥ ···≥ n} 1 n dx (8.1) P (β/2) (x ,...,x )dx dx = h(x ,...,x ; m, ν)β/2 i n,Jacobi 1 n 1 ··· n (β/2) 1 n x (1 x ) Zn,m,ν i=1 i i Y − n (β/2) supported on where Zn,m,ν is a normalization constant and U n 2 ν+1 m 2n ν+1 h(x ,...,x ; m, ν)=( (x)) x (1 x ) − − 1 n △ i − i i=1 Y for some m, n, ν 0 with m 2n ν 0. ≥ − − ≥ n By [33], we know that (˜x1,..., x˜n) is the unique maximizer of h(x1,...,xn, m, ν) in . It follows that as β , the random particle system (x ,...,x ) converges to the deterministicU → ∞ 1 n configuration (˜x1,..., x˜n). As an example, consider the simple case where n = 1, then

ν+1 m ν 1 h(x ; m, ν)= x (1 x ) − − 1 1 − 1 where we need ν +1 < m. The log-derivative gives ν +1 m ν 1 − − x − 1 x 1 − 1 which has a unique rootx ˜ = ν+1 [0, 1]. 1 m ∈ PRODUCT MATRIX PROCESSES VIA SYMMETRIC FUNCTIONS 35

8.2. β Crystallization of the β-Jacobi Product Process. We now turn to the transition kernel (7.3).→∞ Fixing 0 < x < < x < 1, we can see that the transition density has the form 1 ··· n (β/2) 1 β/2 (8.2) pn,ν (y x)= (β/2) fn(y; x)gn,ν(y; x) | Zx,n,ν (β/2) where Zx,n,ν is a normalization constant and (y) n 1 f (y; x)= △ n n x y y i,j=1 | i − j| i=1 i n n Y Q g (y; x)= x y yν+1 n,ν | i − j| i i,j=1 i=1 Y Y supported in n := [0, 1]n x y x y . Ux ∩{ 1 ≤ 1 ≤···≤ n ≤ n} As β , we see that y converges in distribution to a deterministic configurationy ˜ = (˜y1 ... y˜ →) which ∞ satisfies ≤ ≤ n y˜ = argmax gn,ν(y; x) y where the argmax is over y n. The maximizer solves ∈ Ux n 1 ν +1 (8.3) + =0, i =1, . . . , n. y˜ x y˜ i=1 i j i X − n In words, it is a root of the log-gradient of gn,ν; clearly log gn,ν is strictly concave on x so the maximizer is unique. U Lemma 8.1. The maximizer y˜ above satisfies 1 n n (z ν+1 x ) (z x )= (z y˜ ). n − n+ν+1 i − j − i i=1 j=i i=1 X Y6 Y Proof. Following the proof of Theorem 1.1 in [14] by taking κ in Proposition 7.2 to be a partition of the form (1,..., 1, 0,..., 0), we can show that n 1 E (z y ) = E (z u x ) − i n! − σ(i) i "i=1 # σ Sn Y X∈   is independent of β > 0, where u2 = = un = 1 and u1 is distributed as the β-Jacobi ensemble with parameters m = n + ν + 1 and ···ν. We refer the reader to [14, Section 2] for details on this argument. After evaluating the right hand side expectation, we obtain 1 n (z E[u ]x ) (z x ). n − 1 i − j i=1 j=i X Y6 Sending β and recalling the discussion from (8.1), we have →∞ ν +1 lim E[u1] β n + ν +1 →∞ → which establishes the lemma. We note that in fact E[u1] is independent of β > 0, see e.g. [7].  36 ANDREW AHN AND EUGENE STRAHOV

8.3. Gaussianity and the -Jacobi Product Process. We are now ready to introduce the β limit of the β-Jacobi∞ product process. →∞ 1 1 1 n Definition 8.2. Fix parameters νk > 0 for k = 2, 3,... andx ˜ = (˜x1 < < x˜n) [0, 1] . The 1 ··· ∈ -Jacobi product process with initial state x˜ with parameters (νk)k∞=2 is a deterministic sequence ∞k k k k (˜x )∞ , wherex ˜ = (˜x < < x˜ ) recursively satisfies k=1 1 ··· n 1 n n (z νk+1+1 x˜k) (z x˜k)= (z x˜k+1), k =2, 3,..., n − n+νk+1+1 i − j − i i=1 j=i i=1 X Y6 Y k k k 1 1 equipped with a Gaussian field ξ = (ξ1 ,...,ξn) k∞=1 such that ξ1 = ξn = 0 and the joint density of ξ2,...,ξp is proportional{ to } ··· (8.4) p k 1 k 1 2 n k 1 k 2 n k 1 2 k 2 1 (ξ − ξ − ) 1 (ξ − ξ ) ν +2 (ξ − ) ν +1 (ξ ) exp i − j i − j + k i k i . k 1 k 1 2 k 1 k 2 k 1 2 k 2 2 (˜x − x˜ − ) − 4 (˜x − x˜ ) 4 (˜x − ) − 4 (˜x ) "k=2 1 i 0 for k =2, 3,... and x =(x1 < < xn) [0, 1] . Suppose k ···1 ∈ (x )k∞=1 is distributed as the β-Jacobi product process with initial state x and parameters (νk)k∞=2. k 1 1 Let (˜x )k∞=1 be the deterministic part of the -Jacobi product process with initial condition x˜ = x k k ∆xk ∞ k k and parameters (νk)k∞=2. If x =x ˜ + √β , then (˜x )k∞=2, (∆x )k∞=1 converges to the Gaussian field k 1 (ξ )∞ associated to the -Jacobi product process with initial state x˜ as β . k=1 ∞ →∞ Proof. We proceed as in [14, Section 3.4]. The density of (x1,...,xp) is proportional to p n n k k β/2(νk+1) 1 (x ) k 1 k β/2 1 (xi ) − k △ xi − xj − k 1 dxi . ( (xk 1))β 1 | − | β/2(νk+2) 1 − − i,j=1 i=1 (xi − ) − kY=2 △ Y Y 1 1 k k k 1 Sequentially definex ˜ = x andx ˜ to be the maximizer of gn,νk (x ;˜x − ) for k =2, 3,.... Observe that n 1 n (8.5) (z x˜k+1)= (z νk+1+1 x˜k) (z x˜k), k =2, 3,... − i n − n+νk+1+1 i − j i=1 i=1 j=i Y X Y6 k by Lemma 8.1. In other words (˜x )k∞=1 is the deterministic part of a -Jacobi product process. Set ∞ ∆xk xk =x ˜k + i . i i √β The density (ignoring the differentials) then becomes

p n n k k k β/2(νk+1) 1 1 x˜i x˜j k 1 k β/2 1 (˜xi ) − k |1 − k 1| x˜i − x˜j − k 1 Z (β) β 1 | − | β/2(νk+2) 1 n i

k k k β/2(νk+1) 1 ∆xi ∆xj β/2 1 ∆x − p 1+ − n k 1 k − n i √β(˜xk x˜k) ∆x − ∆x 1+ √βx˜k  i − j 1+ i − j i  β 1 k 1 k k−1 β/2(ν +2) 1 × k−1 k−1 √ ∆x k ∆xi ∆xj − β(˜xi − x˜j ) i − k=2 i

PRODUCT MATRIX PROCESSES VIA SYMMETRIC FUNCTIONS 37

The second line can be rewritten as

p k−1 k−1 n k−1 k n k k−1 ∆xi ∆xj √β ∆xi ∆xj √β ∆xi ∆xi exp β − + − + (νk + 1) (νk + 2) − k−1 k−1 2 k−1 k 2 x˜k − k−1  k=2 i 0 independent of ζ1,...,ζn. Thus we can integrate the density of the (ξ2,...,ξp) in the -Jacobi corners process sequentially in k starting from p and descending using the identity above.∞  38 ANDREW AHN AND EUGENE STRAHOV

References [1] Adler, M.; Forrester, P. J.; Nagao, T.; van Moerbeke, P. Classical skew orthogonal polynomials and random matrices, J. Stat. Phys. 99 (2000), 141–170. [2] Akemann G.; Burda Z.; Kieburg, M. From integrable to chaotic systems: Universal local statistics of Lyapunov exponents EPL (Europhysics Letters), Volume 126, Number 4 (2019) 40001. [3] Akemann, G.; Baik, J.; Di Francesco P. The Oxford handbook of random matrix theory, Oxford University Press, Oxford, 2011. [4] Akemann, G.; Kieburg M.; Wei, L. Singular value correlation functions for products of Wishart random matrices. J. Phys. A. 46 (2013) 275205. [5] Akemann, G.; Ipsen, J.; Kieburg M. Products of rectangular random matrices: singular values and progressive scattering. Phys. Rev. E 88 (2013) 052118. [6] Ahn, A. Fluctuations of β-Jacobi Product Processes, 2019, Preprint, arXiv:1910.00743. [7] Aomoto, K. Jacobi polynomials associated with Selberg integrals. SIAM J. Math. Anal. 18 (1987), no. 2, 545–549. [8] Borodin, A.; Corwin, I. Macdonald processes. Probab. Theory Related Fields 158 (2014), no. 1-2, 225–400. [9] Borodin, A.; Gorin, V.; Strahov, E. Product Matrix Processes as Limits of Random Plane Partitions, Int. Math. Res. Not. IMRN 2019, doi:10.1093/imrn/rny297. [10] Borodin, A.; Gorin, V. General β-Jacobi corners process and the Gaussian free field, Comm. Pure Appl. Math. 68 2015, no. 10, 1774–1844. [11] Cuenca, C. Pieri integral formula and asymptotics of Jack unitary characters. Selecta Math. (2018), Volume 24, 2737–2789. ′ ′ 1 [12] Dixon, A. L. Generalizations of Legendre’s formula ke (k e)k = 2 π. Proc. London Math. Soc. 3 1905, 206–224. − − [13] Gorin, Vadim; Kleptsyn, Victor. Universal objects of the infinite beta random matrix theory, 2020, Preprint, arXiv: 2009.02006. [14] Gorin, Vadim; Marcus, Adam. Crystallization of random matrix orbits. Internat. Math. Res. Not. 2018, doi:10.1093/imrn/rny052. [15] Gorin, V.; Panova, G. Asymptotics of symmetric polynomials with applications to statistical mechanics and representation theory. Ann. Probab. (2015), Volume 43, 3052–3131. [16] Forrester, P.J. Log-Gases and Random Matrices, Princeton University Press, 2010. [17] Forrester, P. J.; Zhang, J. Co-rank 1 projections and randomised Horn problem. arXiv: 1905.05314v3. [18] Furstenberg, H.; Kesten, H., Ann. Math. Statist. 31 (1960), 457–469. [19] Hanin, B.; Nica, M. Products of Many Large Random Matrices and Gradients in Deep Neutral Networks. Comm. Math. Phys. 376 (2020), no. 1, 287 – 322. [20] Horn, A. Eigenvalues of sums of Hermitian matrices. Pacific J. Math. 12 (1962), 225–241. [21] Ipsen, J. R.; Kieburg, M. Weak commutation relations and eigenvalue statistics for products of rectangular random matrices. Phys. Rev. E 89 (2014) 3, 032106. [22] Kieburg, M. Additive Matrix Convolutions of P´olya Ensembles and Polynomial Ensembles. arXiv:1710.09481 [23] Kieburg, M.; Forrester, P. J.; Ipsen, J. R. Multiplicative convolution of real asymmetric and real anti-symmetric matrices. Adv. Pure Appl. Math. 10 (2019), no. 4, 467–492. [24] Kieburg, M.; Kuijlaars, A. B. J.; Stivigny, D. Singular value statistics of matrix products with truncated unitary matrices. Int. Math. Res. Not. IMRN 2016, no. 11, 3392–3424. [25] Kuijlaars, A. B. J. Transformations of polynomial ensembles. Modern trends in constructive function theory, 253-268, Contemp. Math., 661, Amer. Math. Soc., Providence, RI, 2016. [26] Kuijlaars, A. B. J.; Rom´an, P. Spherical functions approach to sums of random Hermitian matrices. Int. Math. Res. Not. IMRN 2019, no. 4, 1005–1029. [27] Kuijlaars, A. B. J.; Stivigny, D. Singular values of products of random matrices and polynomial ensembles. Random Matrices Theory Appl. 3 (2014), no. 3, 1450011. [28] Kuijlaars, A. B. J.; Zhang, L. Singular values of products of Ginibre random matrices, multiple orthogonal polynomials and hard edge scaling limits. Comm. Math. Phys. 332 (2014), no. 2, 759–781. [29] Liu, D.-Z.; Wang, D.; Zhang, L. Bulk and soft-edge universality for singular values of products of Ginibre random matrices. Ann. Inst. Henri Poincar´eProbab. Stat. 52 (2016), no. 4, 1734–1762. PRODUCT MATRIX PROCESSES VIA SYMMETRIC FUNCTIONS 39

[30] Liu, D.-Z.; Wang, D.; Wang, Y. Lyapunov exponent, universality and phase transition for products of random matrices. arXiv181000433 [31] Luke, Y.L. The special functions and their approximations. Academic Press, New York 1969. [32] Macdonald, I. Symmetric Functions and Hall Polynomials. Oxford Mathematical Monographs. Oxford Univer- sity Press, USA, 2 edition, 1995. [33] Marcell´an, F.; Mart´ınez-Finkelshtein, A.; Mart´ınez-Gonz´alez, P. Electrostatic models for zeros of polynomials: old, new, and some open problems. J. Comput. Appl. Math. 207 (2007), no. 2, 258-272. [34] Muirhead, R. J. Aspects of Multivariate Statistical Theory. Wiley Series in Probability and Statistics, 2005. [35] Pollicott, M. Maximal Lyapunov exponents for random matrix products. Invent. Math. 181 (2010), 209–226. [36] Strahov, E. Dynamical correlation functions for products of random matrices. Random Matrices Theory Appl. 4 (2015), no. 4, 1550020. [37] Tracy, C. A.; Widom, H. Correlation functions, cluster functions and spacing distributions for random matrices, J. Stat. Phys. 92 (1998), 809–835. [38] Van Peski, R. Limits and fluctuations of p-adic random matrix products, 2020, Preprint, arXiv:2011.09356. [39] Zhang, J.; Kieburg, M.; Forrester, P. J. Harmonic analysis for rank-1 Randomised Horn Problems. arXiv: 1911.11316v1.

Department of Mathematics, Massachusetts Institute of Technology, 77 Massachusetts Av- enue Cambridge, MA 02139-4307, USA Email address: [email protected] Department of Mathematics, The Hebrew University of Jerusalem, Givat Ram, Jerusalem 91904, Israel Email address: [email protected]