arXiv:1811.06592v1 [math.CA] 15 Nov 2018 e fprmtr novd ti hna motn question important an then is It involved. parameters theor of representation in set interpretation [39 an [38], have [28], [31], [27], [30], [29], [23], [1], connection to in refer motivat we been orthogonal particular have valued papers th of These to tion related operators. wa polynomials differential step orthogonal using first valued next A matrix and time. studied polynomials, later orthogonal much [ vector-valued in a references realizing of see 1950s, is the polynomials in Godement orthogonal gene of the work though refer the Even to and least functions. theory spherical of valued overview matrix an via that for as [7] developed al. fully et as Damanik not th to is for refer theory true the particularly but is s polynomials, as This such etc. r topics, physics, analyis, other numerical mathematical of functions, study hypergeometric the of to terms in related speci or as The sca in such arising etc. methods as i.e. properties, be approximation classical, can general methods of problem, general divide study More roughly the be classes. can in special this As and d approaches, and possible [18]. several theory [17], o scattering [16], study recurrences, relations [14], the order as in higher such of as polynomials, well theory sever orthogonal as introduction valued its problem matrix Since moment to [36]. matrix [35], the indices of deficiency study a in Date arxvle oyoil aealn itr,itoue i introduced history, long a have polynomials valued Matrix ti elkonta ntesaa-aemn oyoil in polynomials many scalar-case the in that well-known is It ari that ones the are cases special the of examples Important oebr1,2018. 19, November : oml,tececet ntetretr eurne diffe pol recurrence, Laguerre three-term valued the matrix formulas. in these coefficients of the shift structures formula, The the polynomials. for these sions for P operators matrix satisfies natur shift function matrix introduce weight the matrix the using parameters freedom the of degrees several involving Abstract. ARXVLE AUREPOLYNOMIALS LAGUERRE VALUED MATRIX namrto fhsmteais esnllaesi n c and leadership personal , his of admiration In arxvle aureplnmasaeitoue i mat a via introduced are polynomials Laguerre valued Matrix eiae oBnd atro h caino i retirement. his of occasion the on Pagter de Ben to Dedicated RKKEIKADPBOROM PABLO AND KOELINK ERIK 1. Introduction 1 ]. ,bttpclyol o eylimited very a for only typically but y, linked been have applications other al peetto hoy plctosin applications theory, epresentation prtr edt xlctexpres- explicit to lead operators eta prtr,adexpansion and operators, rential arxvle aeo orthogonal of case valued matrix e asneutos hc lo to allow which equations, earson .Udrsial odtoson conditions suitable Under e. ngnrlmtosadstudying and methods general in d ru bu,Pcaoi ia [20] Tirao Gr¨unbaum, Pacharoni, lcassaetpclyintroduced typically are classes al a-aud oyoil hr are there polynomials lar-valued, nmas uha Rodrigues a as such ynomials, ffrniloeaos e ..[3], e.g. see operators, ifferential a peia ucin obc at back go functions spherical ral ltost ieeta operators, differential to olutions n plctost ..spectral e.g. to applications and 5,terlto omti valued matrix to relation the 15], orpeetto hoy n in and theory, representation to ymti ar(SU(3) pair symmetric e AN ´ fotooa oyoil.We polynomials. orthogonal of oa oudrtn h situa- the understand to ional ences. tdigtegnrlmoment general the studying aeb orwne [34] Koornwinder by made s o oetn oageneral a to extend to how h 90 yMG Krein M.G. by 1940s the n ei ersnaintheory representation in se h se cee e [4], see scheme, Askey the almness. prtr ihhigher with operators f i egtfunction weight rix , U(2)) 2 ERIK KOELINK AND PABLO ROMAN´ set of parameters, and this is a non-trivial question since usually the limited set is finite. An important example is the extension of multivariable spherical functions on Riemannian symmetric spaces to the general solutions of integrable systems in terms of Heckman-Opdam functions, see Heckman’s lectures in [22]. For the case of the extension of [29], [30] from matrix valued Chebyshev polynomials to matrix valued Gegenbauer polynomials, this is done in [32]. In [32] it turns out that one of the essential features that make the analytic extension work is the appropriate LDU-decomposition of the matrix weight. Here L is an lower consisting of Gegenbauer polynomials, and the D is intimately related to the weight of the classical Gegenbauer polynomials. These features make it possible to introduce matrix valued analogues of shift operators, which make it possible to derive properties in an explicit fashion, cf. [22] for the importance of shift operators in the multivariable setting. Moreover, it turns out that the inverse matrix L has a simple expression in terms of Gegenbauer polynomials, due to Cagliero and Koornwinder [5]. It is then a natural question to find the appropriate analogues for the simplest classical polynomials in the Askey scheme, i.e. the Hermite and Laguerre polynomials. For the Hermite polynomials it is possible to use a limit from Gegenbauer polynomials, and this suggests how to deal with the appropriate matrix valued analogue of the Hermite polynomials, see [25]. It tuns out that then the matrix polynomial L is, up to a constant matrix, a , and it is shown that, for certain values of the parameters, it is related to examples by Gr¨unbaum and Dur´an in [12], [11] and [10]. Again, as shown in [25], the shift operators allow us to be explicit in describing the properties of these matrix polynomials. The purpose of this paper is to describe explicitly the matrix valued analogue of the La- guerre polynomials following the approach sketched above. So we introduce in Section 3 the matrix weight function in terms of an LDU-decomposition where the matrix polynomial L is an unipotent lower triangular matrix with entries given in terms of Laguerre polynomials in analogy with the LDU-decomposition of the matrix weight in [32]. The corresponding monic matrix valued Laguerre polynomials are eigenfunctions of a second order matrix valued dif- ferential operator which is symmetric with respect to the weight, see Section 4. In Section 5 we use the degrees of freedom introduced in the matrix weight to impose two non-linear conditions, which allow us to derive Pearson equations for the weight. For specific values of the parameters, the matrix-valued Laguerre polynomials are closely related to the family studied in [13]. In turn, in Section 6 this allows us to calculate the squared norm explicitly, and we give an Rodrigues formula to generate the polynomials. We naturally obtain from the Pearson equations a matrix valued differential operator which is factored as lowering and raising operator. We calculate its Darboux transform in terms of the matrix differential op- erator obtained before. As an application of the shift operators, we give expressions for the matrix coefficients in the three-term recurrence relation and we apply it to find an expansion formula for the matrix valued Laguerre polynomials. We briefly describe three families of so- lutions to the conditions. In particular, this family of matrix valued orthogonal polynomials fits into the setting of Cantero, Moral and Vel´azquez [6], since the derivative of the matrix valued Laguerre polynomials form a family of matrix valued orthogonal polynomials as well. However, [6] does not contain any explicit family of arbitrary size of comparable complexity as in this paper. For convenience, Section 2 describes some basic results on matrix valued orthogonal polynomials. As mentioned above, the matrix valued analogue of the Laguerre polynomials in this paper is not motivated by a limit transition of the matrix valued Gegenbauer polynomials [32], MATRIX VALUED LAGUERRE POLYNOMIALS 3 which find their origin in group theory and the use of shift operators. This is the case for the Hermite polynomials of [25], and it turns out that the matrix valued Gegenbauer polynomials and matrix valued Hermite polynomials share some features, such as the simple structure of the matrices in the three-term recurrence relation. Indeed, the Bn and Cn of (2.2) are tridiagonal and diagonal, respectively. This allows, with the results of [33, Lemma 3.1], to give a simple proof of the triviality of W and AW , see Section 2. In the case of the A matrix valued Laguerre polynomials the expressions for Bn and Cn are explicit, but more complex, see Proposition 6.6 and a proof of the irreducibility along these lines does not seem to work. We conjecture that W and AW are trivial as well, but we have no full proof. Another important difference isA that in the case of the matrix valued Hermite, respectively Gegenbauer, polynomials, the entries of these polynomials can be explicitly calculated in terms of the scalar Hermite polynomials and Hahn polynomials, respectively Gegenbauer and Racah polynomials, see [25, Thm. 3.13], [32, Thm. 3.4]. In the Laguerre setting of this paper, this is not possible because the eigenvalue matrix of Proposition 4.3 is not diagonal. Acknowledgements. Erik Koelink gratefully acknowledges the support and hospitality of FaMAF at Universidad Nacional de C´ordoba and the support of an Erasmus+ travel grant. The work of Pablo Rom´an was supported by Radboud Excellence Fellowship, CONICET grant PIP 112-200801-01533, FONCyT grant PICT 2014-3452 and by SeCyT-UNC.

2. Preliminaries In this section we give some background on matrix valued orthogonal polynomials and we fix the notation, for more information we refer to e.g. [7], [12], [21]. We consider a complex N N matrix valued integrable function W : (a, b) MN (C), and we assume that W (x) > 0 almost× everywhere and such that the weight has finite→ moments of all orders. Here MN (C) is the algebra of complex N N-matrices. Integration of a matrix valued function is separate integration of each matrix× entry, so that the integral is a matrix. Here we allow the endpoints a and b to be infinite. By MN (C)[x] we denote the algebra over C of all polynomials in x with coefficients in MN (C). The weight matrix W induces a Hermitian sesquilinear form b ∗ (2.1) P,Q = P (x)W (x)Q(x) dx MN (C), h i ∈ Za such that for all P, Q, R MN (C)[x], T MN (C) and a, b C the following properties are satisfied ∈ ∈ ∈ aP + bQ, R = a P,R + b Q, R , TP,Q = T P,Q , P,Q ∗ = Q, P . h i h i h i h i h i h i h i Moreover P, P = 0 P = 0, h i ⇐⇒ see for instance [21]. Given a weight matrix W there exists a unique sequence of monic orthogonal polynomials Pn n≥ in MN (C)[x]. Another sequence of Rn n≥ of orthogonal { } 0 { } 0 polynomials in MN (C)[x] is of the form Rn(x)= AnPn(x) for some An GLN (C). Orthogonal polynomials satisfy a three-term recurrence relation and∈ for monic orthogonal polynomials Pn n≥ we have { } 0 (2.2) xPn(x)= Pn (x)+ Bn(x)Pn(x)+ CnPn− (x), n 0, +1 1 ≥ where P−1 = 0 and Bn, Cn are matrices depending on n and not on x. 4 ERIK KOELINK AND PABLO ROMAN´

We say that a matrix weight W is reducible to weights of smaller size if there exists a ∗ constant matrix M such that MW (x)M = diag(W (x),...,Wk(x)) for all x [a, b], where 1 ∈ W1,...,Wk, k > 1, are weights of size less than N. In such a case, the real vector space ∗ W = Y MatN (C) YW (x)= W (x)Y , for all x [a, b] , A { ∈ | ∈ } is non-trivial. On the other hand, if the commutant algebra

AW = Y MatN (C) YW (x)= W (x)Y, for all x [a, b] , { ∈ | ∈ } is nontrivial, then the weight W is reducible via a M. The relation between W and AW is the following; if W is -invariant, the weight W reduces to weights of smaller A A ∗ size if and only if the commutant algebra AW is not trivial, see [33].

3. Matrix valued Laguerre-type polynomials We introduce a family of matrix valued weight functions by giving an explicit LDU- decomposition, and in Section 5 we give conditions on the parameters involved. The matrix polynomial L in this decomposition is a unipotent lower triangular matrix polynomial, whose entries are multiples of Laguerre polynomials independent of the dimension of the matrix. This structure is inspired by a family of matrix valued orthogonal polynomials related to the the symmetric pair (G, K) = (SU(2) SU(2), SU(2)) (K diagonally embedded), which involves a lower triangular matrix whose entries× are Gegenbauer polynomials, see [29], [30], [32], as well as by the extension to matrix valued q-orthogonal polynomials related to the quantum analogue of (G, K), see [2]. In the paper the integer N 1 is fixed. Let µ = (µ1,...,µN ) be a sequence of non-zero ≥(α) coefficients and α> 0. Then Lµ is the N N unipotent lower triangular matrix defined by × µm L(α+n)(x), m n, (α) µn m−n (3.1) Lµ (x)m,n = ≥ (0 n

(α) N dL (α) (α) (α) xA (3.3) (x)= L (x) A, A = Ek,k− = L (x)= L (0)e . dx − 1 ⇒ Xk=2 (α) −1 dLµ (α) (α) (α) xAµ By conjugation we have Aµ = SµASµ , dx (x) = Lµ (x) Aµ and Lµ (x) = Lµ (0)e . Because of the dependence of the parameter of the Laguerre polynomial on n in (3.1), L(α)(x) does not commute with A, see Lemma 4.2. Remark 3.1. Comparing the expression L(α)(x) of (3.1), (3.3) with the corresponding L op- erator in the matrix valued Hermite case, see [25, (3.1), Prop. 3.1], we see that the exponential part of L(α)(x) and of the L in [25] is the same up to a factor 2. So, denoting L of [25] for − MATRIX VALUED LAGUERRE POLYNOMIALS 5 the moment by H, we see that L(α)( 2x)= L(α)(0)H(0)−1H(x). Taking the (m,n)-entry we get connection coefficients relating Laguerre− polynomials and Hermite polynomials, explicitly m (α+n) (α) −1 1 L ( 2x)= L (0)H(0) Hp−n(x) m−n − m,p (p n)! p=n − X  −1 where Hn are the standard Hermite polynomials [28, 1.13]. Since H(0) is known by [25, (α) −1 § Prop. 3.1], we can rewrite the L (0)H(0) m,p explicitly as a sum, and we obtain m  1 1 (α+n) (α + p + 1)m−p 2 (p m), 2 (p m + 1) (3.4) L ( 2x)= F − − ; 1 Hp−n(x). m−n − (m p)! (p n)! 2 2 1 1 p=n 2 (α + p + 1), 2 (α + p + 2) ! X − − (α) Observe that the inverse of Lµ is a unipotent lower triangular polynomial matrix function. Hence, its inverse is a unipotent lower triangular polynomial matrix function as well. In order (α) −1 to obtain an explicit expression for (Lµ (x)) , we need the following result of [26, 1], which can be obtained from the generating function (3.2) for the Laguerre polynomials. § Lemma 3.2. For i, j N, i j, the following inversion formula for the Laguerre polynomials holds true. ∈ ≥ i (−α−i−1) (α+j) L ( x)L (x)= δi,j. i−k − k−j Xk=j (α) Corollary 3.3. The inverse of Lµ is given explicitly by − − − µm L( α m 1)( x), m n, (α) −1 µn m−n (Lµ (x) )m,n = − ≥ (0 n

(α) Lemma 3.4. Let Lµ (x) be the matrix polynomial as in (3.1) and let λ> 0. Then (α) (α,λ) (λ) (α,λ) (α) (λ) −1 (α) (λ) −1 Lµ (x)= Mµ Lµ (x), Mµ = Lµ (x)Lµ (x) = Lµ (0)Lµ (0) . (α,λ) Moreover, Mµ is explicitly given by N−1 (α λ)k M (α,λ) = − ( 1)kAk . µ k! − µ Xk=0 6 ERIK KOELINK AND PABLO ROMAN´

(α,λ) Note that the series for Mµ terminates, since Aµ is nilpotent. Using the binomial theorem it can also be expressed as (α,λ) λ−α Mµ = (1+ Aµ) .

Proof. Note that it suffices to deal with the case µi = 1 for all i, and next conjugate by Sµ. Since L(α)(x) and (L(λ)(x))−1 are unipotent lower triangular matrices, M (α,λ) is unipotent lower triangular. Moreover, M (α,λ) is constant, since the exponentials cancel. Now the (r, s)- entry of M (α,λ)L(λ)(x)= L(α)(x), for r>s, is r r α,λ λ s α s (α λ)r−k λ s M ( )L( + )(x)= L( + ) = − L( + ), r,k k−s r−s (r k)! k−s k s k s X= X= − where we use [8, 18.18.18] in order to write the Laguerre polynomials with parameter λ + s in terms of Laguerre polynomials with paramater α + s. Note that the required identity [8, 18.18.18] is a direct consequence of the same generating function (3.2) and the binomial theorem. We obtain α,λ (α λ)r−k M ( ) = − . r,k (r k)! − Conjugating by Sµ, the lemma follows.  We are now ready to introduce the weight matrix. For ν > 0 we define the matrix N (α,ν) (α) (ν) (α) ∗ (ν) −x ν+k (ν) (3.5) Wµ (x)= Lµ (x) T (x) Lµ (x) , T (x)= e x δk Ek,k. k X=1 Here the parameters δ(ν), 1 k N, are to be determined later, see conditions (5.2) and k ≤ ≤ (5.4) in Section 5. For now we assume the condition δ(ν) > 0, 1 k N, so that the weight k ≤ ≤ is positive definite. Moreover, since ν > 0 the factor e−xxν+k in the entries of the diagonal matrix T (ν) guarantees that all the moments exist. Note that even ν > 2 suffices for the existence of moments as well, but we require ν > 0 in all cases in Examples− 6.9, 6.10, 6.11. By (α,ν) (α,ν) Pn n we denote the sequence of monic orthogonal polynomials with respect to Wµ . We{ suppress} the µ-dependence in the notation for the polynomials and the related quantities, such as the squared norms, coefficients in the three-term recurrence relation, etc. Note that the structure in T (ν) is motivated by the results of [30], [32]. Observe that for real µi’s we have ∗ (α,ν) (α) (ν) (α) ∗ (α) xAµ (ν) xAµ (α) ∗ (3.6) Wµ (x)= Lµ (x)T (x)Lµ (x) = Lµ (0)e T (x)e Lµ (0) .

From now on we assume that the coefficients µi are real and non-zero for all i. Using Lemma 3.4 we obtain (α,ν) (α,λ) (λ,ν) (α,λ) ∗ (3.7) Wµ = Mµ Wµ (Mµ ) , (α,ν) for α,λ > 0. Denote by Hn the n-th squared norms for the monic orthogonal polynomials: ∞ (α,ν) (α,ν) (α,ν) (α,ν) ∗ Hn = Pn (x)Wµ (x)Pn (x) dx, Z0 so that by (3.7) we have the following relation for the squared norms with different parameters (α,ν) (α,λ) (λ,ν) (α,λ) ∗ (3.8) Hn = Mµ Hn (Mµ ) . MATRIX VALUED LAGUERRE POLYNOMIALS 7

(α,ν) Recall that the squared norms satisfy Hn > 0. (ν,ν) Proposition 3.5. If we let α = ν, then the 0-th moment H0 is the diagonal matrix ∞ 2 j (ν) (ν,ν) (ν) µj Γ(ν + j + 1) δk k+1 (H )j,j = (W (x))j,j dx = ( 1) ( j + 1)k− . 0 µ (j 1)! µ2 − − 1 0 k k Z − X=1 Proof. It follows from the definition of the weight (3.5) and the orthogonality relations for the Laguerre polynomials that

∞ min(i,j) ∞ (ν,ν) (ν,ν) µiµj (ν) (ν+k) (ν+k) ν+k −x (H )i,j = (W (x))i,j dx = δ L (x)L (x)x e dx 0 µ µ2 k i−k j−k 0 k k 0 Z X=1 Z N (ν) 2 δk = δi,jµj Γ(ν + j + 1) 2 . µk(j k)! Xk=1 − Now the proposition follows by rewriting the factor (j k)!.  − 4. A symmetric second order differential operator A standard technique in order to deal with matrix valued polynomials is to obtain a matrix valued differential operator having the matrix valued orthogonal polynomials as eigenfunc- tions, see e.g. [9], [12], [19], [25], [32]. We obtain a second-order matrix valued differential (ν) operator which is symmetric with respect to Wµ and which preserves polynomials and its degree, by establishing a conjugation to a diagonal matrix differential operator using the approach of [32]. Let F2, F1, F0 be matrix valued polynomials of degrees two, one and zero respectively. Assume that we have a matrix valued second-order differential operator D which acts on a matrix valued C2([0, ))-function Q by ∞ d2Q dQ (4.1) QD = (x) F (x)+ (x) F (x)+ Q(x)F (x). dx2 2 dx 1 0     For a positive definite matrix valued weight W with finite moments of all orders, we say that D is symmetric with respect to W if for all matrix valued C2([0, ))-functions G, H we have ∞ ∞ ∞ (GD)(x)W (x)(H(x))∗ dx = G(x)W (x)((HD)(x))∗dx. Z0 Z0 By [12, Thm 3.1], D is symmetric with respect to W if and only if the boundary conditions

(4.2) lim F2(x)W (x)=0= lim F2(x)W (x), x→a x→b d(F2W ) d(F2W ) (4.3) lim F1(x)W (x) (x)=0= lim F1(x)W (x) (x) x→a − dx x→b − dx and the symmetry conditions

∗ d(F W ) ∗ (4.4) F (x)W (x)= W (x) F (x) , 2 2 (x) F (x)W (x)= W (x) F (x) , 2 2 dx − 1 1 2 d (F W ) d(F W ) ∗  (4.5) 2 (x) 1 (x)+ F (x)W (x)= W (x) F (x) dx2 − dx 0 0 for almost all x (a, b) hold.  ∈ 8 ERIK KOELINK AND PABLO ROMAN´

Remark 4.1. Suppose that the differential operator D is symmetric with respect to a weight (ν) ∗ d2 d matrix of the form Wµ (x)= L(x) T (x) L(x) , and let D = dx2 F2 + dx F1 + F0 be the second- order differential operator obtained by conjugation of D by L. Then for all C2-matrix valued functions Q we have e e e e d2(QL) d(QL) (4.6) (x)F (x)+ (x)F (x) + (QL)(x)F (x)= QD (x)L(x), dx2 2 dx 1 0  where the coefficients Fei and Fi are relatede by e dL d2L dL (4.7) F2L = LF2, F1Le= 2 F2 + LF1, F0L = F2 + F1 + LF0, dx dx2 dx see the discussion in [32, 4]. Moreover, it follows from [32, Proposition 4.2] that D is e § e e e e e symmetric with respect to W if and only if D is symmetric with respect to T . In order to obtain the explicit expression for a symmetric differential operator having the (ν) e matrix valued orthogonal polynomials Pn as eigenfunctions, we need to control commutation relations with some explicit matrices. In particular, J is the diagonal matrix Jk,k = k. Note that JSµ = SµJ. (α) Lemma 4.2. The following commutation relations for Lµ , Aµ and J hold. (α) (α) −1 −1 L (x)J(L (x)) = x(Aµ + 1) x + (α + J)Aµ + J, µ µ − (α) −1 (α) 2 (L (x)) JL (x)= J (α + J x)Aµ xA , µ µ − − − µ (α) (α) −1 −1 L (x)(1 Aµ)(L (x)) = (1+ Aµ) . µ − µ (α) (α) Note that in particular, the last equality gives [Aµ,Lµ (x)] = AµLµ (x)Aµ.

Proof. It suffices to prove the lemma for µi = 1 for all i. We first prove the last identity. If we multiply the third equation by L(α)(x) on the right and by A + 1 on the left, we see that it suffices to show (A + 1)L(α)(x)A = AL(α)(x). The (r, s)-entry of this equation is given by L(α+s+1)(x) L(α+s+1)(x)= L(α+s) (x), r−s−2 − r−s−1 − r−s−1 which is e.g. [8, 18.9.14]. Similarly, the first equation is equivalent to (A + 1)L(α)(x)J = ((α + J x)A + AJ + J + αA2 + AJA)L(α)(x). − The (r, s)-entry of this equation, after simplifying and regrouping terms, gives sL(α+s) (x)+ L(α+s)(x)= rL(α+s)(x) + (x α 2r + 1)L(α+s) (x) + (α + r 1)L(α+s) (x). − r−s−1 r−s r−s − − r−s−1 − r−s−2 Observe that all the Laguerre polynomials have the same parameter and this follows from the three-term recurrence relation of the Laguerre polynomials, see e.g. [28, (1.11.3)]. The second equation follows from the other two. Use the third equation in the first one twice to rewrite the first equation as L(α)(x)J(L(α)(x))−1 = xL(α)(x)A(L(α)(x))−1 + αA + JL(α)(x)(1 A)−1(L(α)(x))−1 − − and isolating J from the last term gives J = L(α)(x) J(1 A)+ xA(1 A) α(L(α)(x))−1AL(α)(x)(1 A) (L(α)(x))−1 − − − −  MATRIX VALUED LAGUERRE POLYNOMIALS 9 and use the last equation once more to see (L(α)(x))−1AL(α)(x)(1 A)= A. Rewriting gives the second equation. −  Now we are ready to write explicitly a symmetric second order differential operator. Proposition 4.3. Let D(α,ν) be given by d2 d D(α,ν) = F (α,ν)(x)+ F (α,ν)(x)+ F (α,ν)(x), dx2 2 dx 1 0     (α,ν) with F2 (x)= x and (α,ν) −1 (α,ν) −1 F (x)= x(Aµ + 1) + ν + J +1+(α + J)Aµ, F (x) = (α ν)(Aµ + 1) J. 1 − 0 − − (α,ν) (α,ν) Then D is symmetric with respect to Wµ . Moreover, (α,ν) (α,ν) (α,ν) (α,ν) (α,ν) −1 P D =Γ P , Γ = ( n + α ν)(Aµ + 1) J, n N. n n n n − − − ∈ (α,ν) Note that the eigenvalue matrix Γn is a lower triangular matrix, which also depends on the choice of the sequence µ.

(α,ν) d2 (α,ν) d (α,ν) (α,ν) Proof. Let us consider the differential operator D = dx2 F2 + dx F1 +F0 obtained (α,ν) (α) by conjugation of D by the matrix Lµ . Then it follows from (4.7) that the coefficients (α,ν) (α,ν) e e e e Fi are given by F2 (x)= x and (α) e (α,νe ) (α) −1 (ν) (α) dLµ F1 = (Lµ ) F1 Lµ 2x , − dx ! (4.8) e 2 (α) (α) (α,ν) (α) −1 (α,ν) (α) d Lµ dLµ (α,ν) F0 = (Lµ ) F0 Lµ x 2 F1 . − dx − dx ! e e (α,ν) The derivatives in (4.8) can be evaluated by (3.3). It follows from the definition of F1 that

(α,ν) (α) −1 −1 (α) (α) −1 (α) F (x)= ν + 1 xL (x) (Aµ + 1) L (x)+ L (x) JL (x) 1 − µ µ µ µ (α) −1 (α) + L (x) (α + J)AµL (x) 2xAµ. e µ µ − (α) −1 (α) We use the first equation of Lemma 4.2 to rewrite the term Lµ (x) (α + J)AµLµ (x). (α,ν) Similarly the last equation of Lemma 4.2 is used, and we obtain F1 (x)= ν + J + 1 x. (α,ν) − Similarly, using (3.3) and Lemma 4.2 and the expression for F1 we obtain e (α,ν) (α) −1 (α) 2 F (x) = (α ν)(1 Aµ) Lµ (x) JLµ (x) xAµ Aµ(νe+ J + 1 x) = (α ν) J. 0 − − − − − − − − (α,ν) (α,ν) Bye Remark 4.1, in order to prove that D is symmetric with respect to Wµ it is enough to prove that D(α,ν) is symmetric with respect to T (ν), i.e. we need to show that the boundary conditions (4.2), (4.3) and the symmetry conditions (4.4), (4.5) hold true with W replaced by T (ν) ande the F ’s replaced by the corresponding F (α,ν)’s. Since F (α,ν)’s are polynomials, the weight involves the exponential e−x we see that for ν > 1, so in particular − for ν > 0, the boundary conditions (4.2), (4.3) are satisfied. e e The symmetry equation (4.4) and (4.5) are diagonal conditions, and can be verified by a simple calculation. 10 ERIK KOELINK AND PABLO ROMAN´

(α,ν) (α,ν) (α,ν) Hence D is symmetric with respect to the weigh matrix Wµ . Since D preserves (α,ν) (α,ν) polynomials and the degree of the polynomials, Pn D are also orthogonal with respect (α,ν) (α,ν) (α,ν) (α,ν) (α,ν) (α,ν) to Wµ , so that Pn D = Γn Pn for some matrix Γn , which is obtained by considering the leading coefficient.  Remark 4.4. It follows from the proof Proposition (4.3), that matrix valued polynomials (α,ν) (α) Pn Lµ are polynomial eigenfunctions of the diagonal second-order differential operator (α,ν) (α,ν) −1 −1 D with eigenvalue Γn = n(Aµ + 1) + (α ν)(Aµ + 1) J. More precisely, we have − − − e 2 (α,ν) (α) (α,ν) (α) d (Pn Lµ ) d(Pn Lµ ) x (x)+ (x)(ν+J+1 x) (P (α,ν)L(α))(x)J =Γ(α,ν)(P (α,ν)L(α))(x). dx2 dx − − n µ n n µ (α,ν) (α,ν) Observe that although the differential operator D is diagonal, the eigenvalue Γn is a full lower triangular matrix so that the previous equation gives a coupled system of differential (α,ν) (α) equations for the entries of Pn Lµ . This ise in contrast with the case of matrix valued Gegenbauer polynomials studied in [32]. An analogous result for a differential operator and (α,ν) (α) involving a diagonal eigenvalue allows to determine the entries of the analogue of Pn Lµ as a single Gegenbauer polynomial. This allowed to find explicit expressions for the polynomials, see [32, 5.2]. The situation of the Gegenbauer setting of [32] is repeated for the matrix valued Hermite§ polynomials [25, 3]. §

5. The matrix valued Pearson equation In order to establish the existence of shift operators, the Pearson equations are essential. (α,ν) We derive the matrix valued Pearson equations for the family of weights Wµ under explicit non-linear conditions relating the coefficients of the sequence µ and the coefficients in T (ν). First we need certain relations involving the function exAµ . Recall the diagonal matrix J; ∗ ∗ Jk,l = δk,lk. Then [J, Aµ]= Aµ and [J, A ]= A , so that µ − µ ∗ ∗ −xAµ xAµ −xA xA ∗ (5.1) e Je = xAµ + J, e µ Je µ = xA + J, − µ For this we use that the left hand side of the first expression is a matrix valued polynomial −xAµ xAµ in x, since Aµ is nilpotent. Its derivative e [J, Aµ]e = Aµ is constant, and the first formula follows. For the second equation of (5.1) we take adjoints and replace x by x in the first formula. − (ν) Now we need to impose conditions on the sequence µi i and the coefficients δk . We (ν) (ν) (ν) { } (ν) −x ν+k (ν) consider the diagonal matrix ∆ = diag(δ1 ,...,δN ), so that (T )k,k = e x (∆ )k,k. (ν) (ν) (ν+1) (ν) (ν) (ν) We assume that there exist coefficients c and d such that δk = (k d + c ) δk for all k = 1,...,N. In other words, we assume that (5.2) ∆(ν+1) = (d(ν)J + c(ν))∆(ν).

Note that d(ν), c(ν) 0, since δ(ν) > 0. In view of (5.1), the condition (5.2) implies that ≥ k ∗ ∗ ∗ ∗ (5.3) e−xAµ (∆(ν))−1∆(ν+1)exAµ = e−xAµ (d(ν)J + c(ν))exAµ = d(ν)( xA∗ + J)+ c(ν). − µ This is the main ingredient in Proposition 5.1. MATRIX VALUED LAGUERRE POLYNOMIALS 11

(α,ν) Proposition 5.1. Let Wµ be the weight matrix given in (3.5) and assume that (5.2) holds true. Then (α,ν) (α,ν) −1 (α,ν+1) Φ (x) = (Wµ (x)) Wµ (x), is a matrix valued polynomial of degree two.

Proof. Using (3.6), the fact that T (ν)(x)−1T (ν+1)(x)= x(∆(ν))−1∆(ν+1) and (5.3), we obtain

(α,ν) −1 (α,ν+1) (α) ∗ −1 (ν) −1 (ν+1) (α) ∗ (Wµ (x)) Wµ (x)= x(Lµ (x) ) (∆ ) ∆ Lµ (x) ∗ ∗ (α) ∗ −1 −xAµ (ν) (ν) xAµ (α) ∗ = x(Lµ (0) ) e (d J + c )e Lµ (0) = d(ν)x2(L(α)(0)∗)−1A∗ L(α)(0)∗+ − µ µ µ (ν) (α) ∗ −1 (α) ∗ (ν)  + x d (Lµ (0) ) JLµ (0) + c .   (ν) Now we assume that the coefficients µk and δk satisfy the relation ν µ2 δ( ) (5.4) k+1 = d(ν)k(N k) k+1 , k = 1,...,N 1. µ2 − (ν+1) − k δk

Note that, since the coefficients µk are independent of ν, we require the right hand side of (5.4) to be independent of ν.

(α,ν) Proposition 5.2. Let Wµ be the weight matrix given in (3.5) and assume that the condi- tions (5.2) and (5.4) hold true. Then

dW (α,ν+1) Ψ(α,ν)(x) = (W (α,ν)(x))−1 µ (x), µ dx is a matrix valued polynomial of degree one.

Proof. Using (3.6) we obtain

∗ ∗ (α) ∗ (α,ν) (α) ∗ −1 −xAµ (ν) −1 (ν+1) xAµ Lµ (0) Ψ (x)(Lµ (0) ) =e (T (x)) AµT (x)e (ν+1) ∗ dT ∗ ∗ ∗ +e−xAµ (T (ν))−1 (x)exAµ + e−xAµ (T (ν))−1T (ν+1)exAµ A∗ . dx µ It follows from (5.2) and (5.1) that

∗ ∗ ∗ ∗ −xAµ (ν) −1 (ν+1) xAµ ∗ −xAµ (ν) −1 (ν+1) xAµ ∗ e T (x) T (x)e Aµ = xe (∆ ) ∆ e Aµ (5.5) = x2d(ν)(A∗ )2 + xd(ν)JA∗ + xc(ν)A∗ , − µ µ µ and (ν+1) ∗ dT ∗ e−xAµ (T (ν))−1 (x)exAµ = x2d(ν)(A∗ + (A∗ )2) (5.6) dx µ µ x((A∗ + 1)(d(ν)J + c(ν))+ d(ν)(ν + J + 1)A∗ ) + (ν + J + 1)(d(ν)J + c(ν)). − µ µ 2 (ν) ∗ 2 We observe that the term x d (Aµ) of the right hand side of (5.6) cancels with the term of degree two in (5.5). 12 ERIK KOELINK AND PABLO ROMAN´

(ν) −1 (ν+1) (ν) −1 (ν+1) Now we note that (T (x)) Aµ T (x) = x(∆ ) Aµ ∆ . On the other hand, (ν) −1 (ν+1) ∗ the matrix [(∆ ) Aµ ∆ , Aµ] is a diagonal matrix whose k-th diagonal entry is given by 2 (ν+1) 2 (ν+1) (ν) −1 (ν+1) ∗ µk+1δk µkδk−1 [(∆ ) Aµ ∆ , A ]k,k = . 2 (ν) − 2 (ν) µkδk+1 µk−1δk (ν) −1 (ν+1) ∗ (ν) (ν) By (5.4) we verify that [(∆ ) Aµ ∆ , A ] = 2d J d (N + 1). This leads to µ − d ∗ ∗ ∗ ∗ e−xAµ (∆(ν))−1 A ∆(ν+1)exAµ = e−xAµ (∆(ν))−1 A ∆(ν+1), A∗ exAµ dx µ µ µ ∗ ∗  = e−xAµ (2d(ν)J d(ν)(N +h 1))exAµ i − = 2xd(ν)A∗ + 2d(ν)J d(ν)(N + 1). − µ − Therefore we have that −xA∗ (ν) −1 (ν+1) xA∗ (ν) 2 ∗ (ν) (ν) −1 (ν+1) (5.7) e µ (∆ ) Aµ ∆ e µ = d x A +xd (2J N 1)+(∆ ) Aµ ∆ − µ − − Adding (5.5), (5.6) and (5.7) shows that Ψ(α,ν) is a polynomial of degree one.  For future reference we state Corollary 5.3 as an immediate consequence of the proofs of Propositions 5.1 and 5.2. Corollary 5.3. Assuming the conditions (5.2) and (5.4), the matrix valued polynomials L(α)(0)∗Φ(α,ν)(x)(L(α)(0)∗)−1 = d(ν)x2A∗ + x d(ν)J + c(ν) , µ µ − µ   L(α)(0)∗Ψ(α,ν)(x)(L(α)(0)∗)−1 = x d(ν)(J A∗ (J + ν + 1) N 1) c(ν) µ µ − µ − − −  (ν) (ν) (ν) −1 (ν+1) + (ν + J + 1)(d J + c )+(∆ ) Aµ ∆ . satisfy the Pearson equations   (α,ν+1) dWµ Φ(α,ν)(x) = (W (α,ν)(x))−1W (α,ν+1)(x), Ψ(α,ν)(x) = (W (α,ν)(x))−1 (x). µ µ µ dx (ν+1) (ν) (ν) (ν) Remark 5.4. Upon replacing δk = (kd + c )δk in (5.4), we can iterate the resulting identity to obtain (ν) c(ν) (ν) δk (1 + d(ν) )k−1 δ1 (5.8) 2 = 2 . µ (k 1)!(N k + 1)k− µ k − − 1 1 (ν,ν) This relation can now be used to evaluate explicitly the the 0-th moment H0 given in Proposition (3.5). Indeed, 2 (ν) N c(ν) µ δ Γ(ν + j + 1) (1 + )k− ( j + 1)k− (ν,ν) j 1 d(ν) 1 − 1 (H0 )j,j = 2 µ1 (j 1)! (N k + 1)k−1(k 1)! − Xk=1 − − 2 (ν) c(ν) µj δ Γ(ν + j + 1) 1+ , (j 1) = 1 F d(ν) − − ; 1 µ2 (j 1)! 2 1 N + 1 1 − − ! 2 (ν) c(ν) µ δ Γ(ν + j + 1)( N )j− j 1 − − d(ν) 1 = 2 , µ (j 1)!( N 1)j− 1 − − − 1 MATRIX VALUED LAGUERRE POLYNOMIALS 13 where the 2F1 is summed by the Chu-Vandermonde Identity.

6. Shift Operators In this section we use the Pearson equations to give explicit lowering and rising operators (α,ν) for the polynomials Pn . Next we exploit the existence of the shift operators to give an explicit Rodrigues formula, to calculate the squared norms as well as the coefficients in the three-term recurrence relation. Moreover, we find another matrix valued differential operator to which the matrix polynomials are eigenfunctions. For this explicit matrix valued differential operator it is possible to perform a Darboux transform, and we give an explicit expression for the Darboux transformation. We end by obtaining a Burchnall type identity, see [25], for the matrix valued Laguerre polynomials, and by showing that there are at least three families of solutions to the non-linear conditions (5.2) and (5.4). In this section we assume that these conditions are satisfied, and hence that the Pearson equations of Corollary 5.3 hold. For matrix valued functions P and Q, we denote by ∞ P,Q (α,ν) = P (x)W (α,ν)(x)Q(x)∗ dx, h i µ Z0 whenever the integral converges. Moreover, we have ∞ dP dP ∗ ,Q (α,ν+1) = (x)W (α,ν+1)(x) Q(x) dx h dx i dx µ Z0 ∞ ∞ dQ∗ = P (x)W (α,ν)(x)Ψ(α,ν)(x)Q(x)∗ dx P (x)W (ν)(x)Φ(α,ν)(x) (x) dx − µ − µ dx Z0 Z0 = P,QS(α,ν) (α,ν), − h i where S(α,ν) is the first order matrix valued differential operator dQ (6.1) (QS(α,ν))(x)= (x)(Φ(α,ν)(x))∗ + Q(x)(Ψ(α,ν)(x))∗. dx Note that we have to assume that the decay at 0 and at is sufficiently large, which is the case for e.g. polynomials P and Q. ∞ (α,ν) k (α,ν) In particular, if we set P (x)= Pn (x) and Q(x) = x , considering the degrees of Φ (α,ν) (α,ν) dPn (α,ν+1) N and Ψ we obtain dx ,Q = 0 for all k , k

(α,ν) where the matrices Kn are invertible and are explicitly given by (α) −1 (α,ν) (α) (ν) (ν) L (0) K L (0) = d (J (J + ν + n)Aµ N 1) c . µ n µ − − − − (α,ν) dPn Proof. Taking into account the preceding discussion, we have that dx is a multiple (α,ν+1) (α,ν+1) (α,ν) (α,ν) of Pn−1 and Pn−1 S ) is a multiple of Pn (x), these multiples follow from the (α,ν) leading coefficients. Now we only need to show that Kn is invertible. Observe that (α) −1 (α,ν) (α) Lµ (0) Kn Lµ (0) is a lower triangular matrix whose j-th diagonal entry is given by (α) −1 (α,ν) (α) (ν) (ν) (L (0) K L (0))j,j = d (j (N + 1)) c . µ n µ − − These entries are strictly negative, since c(ν) and d(ν) are positive. So invertibility follows. 

(α,ν) (α) −1 (α,ν) (α) We note that the matrix Kn = Lµ (0) Kn Lµ (0) is actually a function of ν + n so (α,ν+j) (α,ν) (α) −1 that Kn−j = Kn for all j n. Now conjugating with Lµ (0) we obtain that e ≤ (6.2) K(α,ν+j) = K(α,ν), j n. e e n−j n ≤ (α,ν) Theorem 6.2. Assume that the conditions of Proposition 6.1. The polynomials Pn satisfy the following Rodrigues formula

n (α,ν+n) (α,ν) (α,ν) d Wµ (α,ν) −1 Pn (x)= Gn n (x) Wµ (x) , dx !

(α,ν) (α,ν) −1 (α,ν+n−1) −1 (α,ν) −n where Gn = (Kn ) (K1 ) = (Kn ) . Moreover, the squared norm (α,ν) · · · Hn is given by H(α,ν) = ( 1)n n! (K(α,ν))−nH(α,ν+n) n − n 0 = ( 1)n n! M (α,ν+n)(K(ν+n,ν))−nH(ν+n,ν+n)(M (α,ν+n))∗, − µ n 0 µ (ν+n,ν) where Mµ is the matrix given in (3.7).

(α,ν+1) (α,ν) d(QWµ ) (α,ν) −1 Proof. Observe that (QS )(x)= dx (x) Wµ (x) by Corollary 5.3. Iterating gives  n (α,ν+n) d (QWµ ) − QS(α,ν+n−1) S(α,ν) (x)= (x) W (α,ν)(x) 1. · · · dxn µ (α,ν+n)   Now taking Q(x) = P0 (x) = 1 and using Proposition 6.1 repeatedly we obtain the Rodrigues formula. Finally, for the squared norm we observe that

(α,ν) dPn nH(α,ν+1) = n P (α,ν+1), P (α,ν+1) (α,ν+1) = , P (α,ν+1) (α,ν+1) n−1 h n−1 n−1 i h dx n−1 i = P (α,ν), P (α,ν+1)S(α,ν) (α,ν) = P (α,ν), P (α,ν) (α,ν)(K(α,ν))∗ = H(α,ν)(K(α,ν))∗, h n n−1 i h n n i n n n (α,ν) (α,ν) −1 (α,ν+1) (α,ν) so that Hn = n (Kn ) Hn−1 , where we have used that Hn is self-adjoint for all n N. Iterating and using (6.2) and (3.8) gives the expressions for the squared norm.  ∈ MATRIX VALUED LAGUERRE POLYNOMIALS 15

Corollary 6.3. Assume the conditions of Proposition 6.1. The second-order differential operator d d2 d (α,ν) = S(α,ν) = Φ(α,ν)(x)∗ + Ψ(α,ν)(x)∗, D ◦ dx dx2 dx     (α,ν) is symmetric with respect to the weight Wµ . Moreover, for all n N we have ∈ P (α,ν) (α,ν) = Λ(α,ν)P (α,ν), Λ(α,ν) = nK(α,ν). n D n n n n Moreover, the operators (α,ν) and D(α,ν) commute. D (α,ν) (α,ν) Proof. The fact that is symmetric with respect to Wµ follows directly from the D factorization (α,ν) = S(α,ν) d and Proposition 6.1. Then the orthogonal polynomials D ◦ dx (α,ν) (α,ν) Pn are eigenfunctions of and the eigenvalue is obtained by looking at the leading coefficients. D In order to prove that (α,ν) and D(α,ν) commute, we will show that the corresponding (α,ν) (α,νD) eigenvalues Γn and Λn commute, see [37, Cor. 4.4]. It is then enough to show that the following matrices commute; (α,ν) (α) −1 (α,ν) (α) (α,ν) (α) −1 (α,ν) (α) Λn = (Lµ (0)) Λn Lµ (0), Γn = (Lµ (0)) Γn Lµ (0).

(α,ν) Using the explicite expressions of Γn , Propositione 4.2 and Proposition 6.1 we obtain (α,ν) (ν) (α,ν) Λn c Γ (α ν n) = (J + ν + n)Aµ J = − N 1. n − − − − − d(ν) − − (α,ν) (α,ν) e So Γn ande Λn commute for all n.  Remark 6.4. Corollary 6.3 states that the differential operator (α,ν) has a factorization, d D (α,ν) where first the lowering operator dx is applied and next the raising operator S . The Darboux transform of such a differential operator is obtained by interchanging the order of (α,ν) d (ν) the lowering and raising operator. This gives a differential operator = dx S which (α,ν+1) D ◦ has the orthogonal polynomials Pn as eigenfunctions. Explicitly we have e 2 (α,ν+1) (α,ν+1) (α,ν) (α,ν+1) ˜(α,ν) d Pn (α,ν) ∗ dPn dΦ ∗ (α,ν) ∗ Pn (x)= 2 (x)Φ (x) + (x) (x) +Ψ (x) D dx dx dx !   dΨ(α,ν) + P (α,ν+1)(x) (x)∗ =Ξ(α,ν+1)P (α,ν+1)(x), n dx n n (α,ν+1) where the eigenvalue Ξn is given by (α,ν+1) 2 (α,ν) ∗ (α,ν) ∗ Ξn = n lc(Φ ) + (n + 1) lc(Ψ ) . Proposition 6.5. With the notations as in Proposition 4.3, Corollary 6.3 and Remark 6.4 ν and assuming c( ) for some constant we have d(ν) = ν + ρ ρ 1 1 P ˜(α,ν) = P (α,ν+1) P D(α,ν+1) + α N 2ν 2 ρ. d(ν) D d(ν+1) D − − − − − Note that the assumption on the quotient is satisfied in Examples 6.9, 6.10 and 6.11. 16 ERIK KOELINK AND PABLO ROMAN´

Proof. The proof is a bit involved, but essentially straightforward. First, to use the explicit expression of Remark 6.4 we need the explicit expressions forΦ(α,ν)(x)∗,Ψ(α,ν)(x)∗ and com- pare the difference for ν and ν + 1. From Corollary 5.3 we have

(ν) (ν+1) 1 (α,ν) ∗ 1 (α,ν+1) ∗ c c ν Φ (x) ν Φ (x) = x ν ν I = xI, d( ) − d( +1) d( ) − d( +1) ! − ∗ (α,ν) 1 dΦ (α) (α) −1 ν (x) = Lµ (0) 2xAµ + J + ν + ρ Lµ (0) d( ) dx ! −   −1 = 2x 1 (1 + Aµ) + (α + J)Aµ + J + ν + ρ − − (α,ν) ∗ 1 (ν) −1 (ν+1) using Lemma 4.2. To do the same for Ψ (x) we first observe that d(ν) (∆ ) Aµ∆ is actually independent of ν, because of (5.4). So we find from Corollary 5.3 and Lemma 4.2

1 (α,ν) ∗ 1 (α,ν+1) ∗ (α) (α) −1 Ψ (x) Ψ (x) = L (0) x(1 + Aµ) 2J 2 2ν ρ L (0) d(ν) − d(ν+1) µ − − − − µ −1   = x + x(1 (1 + Aµ) ) 2J 2(α + J)Aµ 2 2ν ρ. − − − − − − Finally, by Corollary 5.3 and Lemma 4.2 ∗ (α,ν) (ν) 1 dΨ (α) c (α) −1 ν (x) = Lµ (0) J (J + ν + 1)Aµ N 1 ν Lµ (0) d( ) dx ! − − − − d( )   −1 = (α + J)Aµ + J ((α + J)Aµ + J + ν + 1)(1 (1 + Aµ) ) N 1 ν ρ − − − − − − −1 = ((α + J)Aµ + J + ν + 1)(1 + Aµ) N 2ν 2 ρ − − − − −1 = J + (αAµ + ν + 1)(1 + Aµ) N 2ν 2 ρ. − − − − Collecting these expressions in the differential operator in Remark 6.4 gives 1 1 P ˜(α,ν)(x) P (α,ν+1)(x)= d(ν) D − d(ν+1) D 2 d P dP −1 (x)( xI)+ (x) x(1 + Aµ) J (α + J)A 2 ν + dx2 − dx − − − − −1 P (x) J + (αAµ + ν+ 1)(1 + Aµ) N 2ν 2 ρ =  − − − − α,ν (P D( +1))(x)+ α N 2ν 2 ρ.   − − − − − As a next application of the shift operators, we calculate the matrix coefficients in the three-term recurrence for the monic polynomials explicitly. As in (2.2), the monic matrix (α,ν) valued Laguerre-type polynomials Pn satisfy a three-term recurrence relation of the form

(α,ν) (α,ν) (α,ν) (α,ν) (α,ν) (α,ν) xPn (x)= P (x)+ Bn Pn (x)+ Cn Pn−1 (x).

(α,ν) (α,ν) Proceeding as in [32, 5.3], the coefficients Bn and Cn , are given by § (6.3) B(α,ν) = X(α,ν) X(α,ν), C(α,ν) = H(α,ν)(H(α,ν))−1, n n − n+1 n n n−1 (α,ν) (α,ν) where Xn is the one-but-leading coefficient of Pn . MATRIX VALUED LAGUERRE POLYNOMIALS 17

Proposition 6.6. Assume the conditions of Proposition 6.1. The coefficients of the three- term recurrence relation for the monic Laguerre-type orthogonal polynomials are given by − B(α,ν) = n (K(α,ν+n 1))−1(Ψ(α,ν+n−1)(0))∗ (n + 1) (K(α,ν+n))−1(Ψ(α,ν+n)(0))∗, n 1 − 1 − − C(α,ν) = nM (α,ν+n)(K(ν+n,ν))−nH(ν+n,ν+n)(1 A∗ )(H(ν+n 1,ν+n 1))−1 n − µ n 0 − µ 0 − (K(ν+n 1,ν))n−1(M (α,ν+n−1))−1 × n−1 µ (α,ν) Proof. By taking the derivative of Pn with respect to x and using Proposition 6.1, we find (α,ν) (α,ν) (α,ν) (α,ν+n−1) that (n 1)Xn = nX which gives Xn = nX . Using the Rodrigues formula − n+1 1 (α,ν) (α,ν) (α,ν) ∗ we obtain P1 (x)= G1 (Ψ (x)) . Evaluating at x = 0 gives (α,ν) (α,ν+n−1) −1 (α,ν+n−1) ∗ Xn = n (K1 ) (Ψ (0)) . (α,ν) (α,ν) Replacing Xn in (6.3) we obtain the expression for Bn . On the other hand, using the expression for the norm in Theorem 6.2, we find that

H(α,ν)(H(α,ν))−1 = nM (α,ν+n)(K(ν+n,ν))−nH(ν+n,ν+n)(M (α,ν+n))∗((M (α,ν+n−1))∗)−1 n n−1 − µ n 0 µ µ − − − (H(ν+n 1,ν+n 1))−1(K(ν+n 1,ν))n−1(M (α,ν+n−1))−1. × 0 n−1 µ (ν+n) ∗ (ν+n−1) ∗ −1 ∗ Now we use Lemma 3.4 to write Lµ (0) (Lµ (0)) ) = (1 Aµ) and this completes the proof of the proposition. −  As a final application of the shift operators we discuss briefly an expansion formula for the matrix valued Laguerre polynomials arising from the Burchnall formula for matrix valued polynomials satisfying a Rodrigues formula. Note that the matrix valued Laguerre polynomi- als form an example of the general conditions in [25, 4]. In particular, Burchnall’s formula c(ν) § [25, Thm. 4.1] applies. Assuming d(ν) = ν + ρ as in Proposition 6.5, we see from Corollary 5.3 that − ∗ k 1 (α,ν) (α,ν+k−1) k k (ν+p) Φ (x) Φ (x) = ( 1) x d (J xAµ + ν + ρ)k. · · · −   − p=0   Y (α,ν)   Moreover, since the matrices Gn in Theorem 6.2 are powers, the result of [25, Cor 4.2] simplifies and we obtain Corollary 6.7. Corollary 6.7. Under the conditions of Proposition 6.5 we have the following expansion formula for the matrix valued Laguerre polynomials. m∧n k−1 n m (K(α,ν))−nP (α,ν)(x)= ( 1)k k! d(ν+p) xk n+m n+m − k k k p=0 X=0    Y  (α,ν+n−k) (α,ν) n−k (α,ν+k) (α) (α) −1 P (x)(K ) P (x)L (0)(J xAµ + ν + ρ)kL (0) × m−k n n−k µ − µ Remark 6.8. In the scalar case the Burchnall identities for some subclasses of the Askey scheme, notably Hermite, Laguerre, Meixner-Pollaczek, Krawtchouk, Meixner and Charlier polynomials, can be used to find expressions for the orthogonal polynomials for the corre- sponding Toda modification of the weight, i.e. multiplication by e−xt, see [24, Prop. 7.1]. These are precisely the cases where it is easy to “glue” on the exponent e−xt to the classical weight function. In the matrix case, the Toda modification of the matrix weight leads to 18 ERIK KOELINK AND PABLO ROMAN´ solutions of the non-abelian Toda lattice, see [25, 2.1] and references given there, and this is worked out in detail for the matrix valued Hermite§ polynomials in [25, 5]. In the case of the matrix valued Laguerre polynomials, however, we are not lead to a corresponding§ solution of the non-abelian Toda lattice. Essentially, the Burchnall approach fails due to the fact that Φ(α,ν) Φ(α,ν+k−1) is a polynomial of degree 2k (instead of k), see the discussion in [24]. Even though· · · in the matrix valued case it is straightforward to glue on the exponential e−xt (α,ν) (ν) to the matrix weight Wµ (x), it leads to a linear change in the parameters µ and δk , and since, the conditions (5.2) and (5.4) form non-linear conditions, the Pearson equations do not hold for the Toda modification. So in particular, Proposition 6.6 is no longer valid for the Toda modified matrix weight. 6.1. Examples. We cannot give all the solutions to the (5.2) and (5.4) in general, due to the non-linearity of these relations. However, if we choose coefficients µk such that the quotient µ2 /µ2 coincides with some of the factors in the product k(N k) of the right hand side of k+1 k − (5.4), we can give explicit examples of Laguerre-type matrix valued orthogonal polynomials. In the case of Hermite-type matrix valued orthogonal polynomials given in [25] there are two nonlinear conditions. The first one is the same as (5.2) and the second one differs from (5.4) 1 by a factor of 2 on the right-hand side. The examples in this paper are obtained in the same way as in [25].

Example 6.9. If we assume that µk = √N kµk, then the left hand side of (5.4) coincides +1 − with the factor N k on the right hand side. Thus µk = (N k + 1)k for k = 1,...,N. Then (5.2) and (5.4)− give the following recurrence relations − p (ν) (ν+1) (ν) (ν) (ν) (ν) c (ν) (6.4) δk = (d k + c )δk , δk+1 = k + ν δk . d( ) ! One solution to (6.5) is given by

(ν) (ν) (ν) δk = Γ(ν)(ν)k, c = ν, d = 1.

Example 6.10. If we assume that µk = k(N k) µk, then the left hand side of (5.4) co- +1 − incides with the factor k(N k) on the right hand side. Then µ = (k 1)!(N k + 1) , p k k−1 k = 1,...,N. The relations−(5.2) and (5.4) give the following recurrence− relations− p (ν) (ν+1) (ν) (ν) (ν) (ν) c (ν) (6.5) δk = (d k + c )δk , δk+1 = k + ν δk . d( ) ! A solution to (6.5) is given by

(ν) ν ν (ν) (ν) δk = λ Γ(ν + k) = (ν)kλ Γ(ν), c = νλ, d = λ. for some fixed λ> 0.

Example 6.11. Now we take µk = 1 for all k = 1,...,N. Therefore the relations (5.2) and (5.4) are given by

(ν) (ν) d k(N k) δ ν ν 1= − k+1 , δ( +1) = (d(ν)k + c(ν))δ( ) (d(ν)k + c(ν)) (ν) k k+1 δk MATRIX VALUED LAGUERRE POLYNOMIALS 19 for which

(ν) (ν) (ν) (1 + ν + C/ρ)k−1 ν d = ρ, c = C + νρ, δk = ρ Γ(1 + ν + C/ρ) (k 1)! (N k + 1)k− − − 1 with ρ> 0, ν > 0 and C 0 gives a solution meeting all the conditions. ≥ References [1] N. Aldenhoven, Explicit matrix inverses for lower triangular matrices with entries involving continuous q-ultraspherical polynomials, J. Approx. Theory 199 (2015), 1–12. [2] N. Aldenhoven, E. Koelink, P. Rom´an, Matrix valued orthogonal polynomials related to the quantum analogue of (SU(2) × SU(2),diag), Ramanujan J. 43 (2017), 243–311. [3] A.I. Aptekarev, E.M. Nikishin, The scattering problem for a discrete Sturm-Liouville operator. Mat. USSR Sbornik 49 (1984), 325–355. [4] R. Askey, J. Wilson, Some basic hypergeometric orthogonal polynomials that generalize Jacobi polynomials, Mem. Amer. Math. Soc. 54 (1985), no. 319. [5] L. Cagliero, T.H. Koornwinder, Explicit matrix inverses for lower triangular matrices with entries involv- ing Jacobi polynomials, J. Approx. Theory 193 (2015), 20–38. [6] M.J. Cantero, L. Moral, L. Vel´azquez, Matrix orthogonal polynomials whose derivatives are also orthog- onal, J. Approx. Theory 146 (2007), 174–211. [7] D. Damanik, A. Pushnitski, B. Simon, The analytic theory of matrix orthogonal polynomials, Surveys in Approx. Th. 4 (2008), 1–85. [8] Digital Library of Mathematical Functions, (eds. F.W.J. Olver, A.B. Olde Daalhuis, D.W. Lozier, B.I. Schneider, R.F. Boisvert, C.W. Clark, B.R. Miller, and B.V. Saunders), http://dlmf.nist.gov. [9] A.J. Dur´an, A method to find weight matrices having symmetric second-order differential operators with matrix leading coefficient, Constr. Approx. 29(2009), 181–205. [10] A. J. Dur´an, A method to find weight matrices having symmetric second-order differential operators with matrix leading coefficient, Constr. Approx. 29 (2009), 2, 181–205. [11] A. J. Dur´an, F. A. Gr¨unbaum, Structural formulas for polynomials satisfying second- order differential equations. I, Constr. Approx. 22 (2005), 2, 255–271. [12] A.J. Dur´an, F.A. Gr¨unbaum, Orthogonal matrix polynomials satisfying second-order differential equations, Int. Math. Res. Not. 2004 (2004), 461–484. [13] A.J. Dur´an, M. de la Iglesia, Some examples of orthogonal matrix polynomials satisfying odd order differ- ential equations, J. Approx. Theory 150 (2008), 2, 153–174. [14] A.J. Dur´an, W. Van Assche, Orthogonal matrix polynomials and higher-order recurrence relations, Appl. 219 (1995), 261–280. [15] R. Gangolli, V.S. Varadarajan. Harmonic Analysis of Spherical Functions on Real Reductive Groups, Ergeb. Math. 101, Springer, 1988. [16] J.S. Geronimo, Scattering theory and matrix orthogonal polynomials on the real line, Circuits Systems Signal Process. 1 (1982), 471–495. [17] W. Groenevelt, M.E.H. Ismail, E. Koelink, Spectral theory and matrix valued orthogonal polynomials, Adv. Math. 244 (2013), 91–105. [18] W. Groenevelt, E. Koelink, A hypergeometric function transform and matrix valued orthogonal polynomi- als, Constr. Approx. 38 (2013), 277–309. [19] F.A. Gr¨unbaum, M.D. de la Iglesia, A. Mart´ınez-Finkelshtein, Properties of matrix orthogonal polynomi- als via their Riemann-Hilbert characterization, SIGMA Symmetry Integrability Geom. Methods Appl. 7 (2011), paper 098, 31 p. [20] F.A. Gr¨unbaum, I. Pacharoni, J. Tirao, Matrix valued spherical functions associated to the complex pro- jective plane, J. Funct. Anal. 188 (2002), 350–441. [21] F.A. Gr¨unbaum, J. Tirao, The algebra of differential operators associated to a weight matrix, Integral Eq. Operator Theory 58 (2007), 449–475. [22] G. Heckman, H. Schlichtkrull, Harmonic Analysis and Special Functions on Symmetric Spaces, Perspec- tives Math. 16, Academic Press, 1994. [23] G. Heckman, M. van Pruijssen, Matrix valued orthogonal polynomials for Gelfand pairs of one, Tohoku Math. J. (2) 68 (2016), 407–436. 20 ERIK KOELINK AND PABLO ROMAN´

[24] M.E.H. Ismail, E. Koelink, P. Rom´an, Generalized Burchnall-type identities for orthogonal polynomials and expansions, SIGMA Symmetry Integrability Geom. Methods Appl. 14 (2018), 072, 24 p. [25] M.E.H. Ismail, E. Koelink, P. Rom´an, Matrix valued Hermite polynomials, Burchnall formulas and non- abelian Toda lattice, to appear. [26] R. Koekoek, Inversion formulas involving orthogonal polynomials and some of their applications, pp. 166–180 in “Special Functions” (eds. M.E.H. Ismail, C. Dunkl, R. Wong), World Sci., 2000. [27] R. Koekoek, P.A. Lesky, R.F. Swarttouw, Hypergeometric Orthogonal Polynomials and their q-Analogues, Springer, 2010. [28] R. Koekoek, R.F. Swarttouw, The Askey-scheme of hypergeometric orthogonal polynomials and its q- analogue, online at http://aw.twi.tudelft.nl/~koekoek/askey.html, Report 98-17, Technical Univer- sity Delft, 1998. [29] E. Koelink, M. van Pruijssen, P. Rom´an, Matrix-valued orthogonal polynomials related to (SU(2) × SU(2), diag), Int. Math. Res. Not. 2012 (2012), 5673–5730. [30] E. Koelink, M. van Pruijssen, P. Rom´an, Matrix-valued orthogonal polynomials related to (SU(2) × SU(2), diag),II, Publ. Res. Inst. Math. Sci. 49 (2013), 271–312. [31] E. Koelink, M. van Pruijssen, P. Rom´an, Matrix elements of irreducible representations of SU(n + 1) × SU(n + 1) and multivariable matrix valued orthogonal polynomials, arXiv:1706.01927. [32] E. Koelink, A.M. de los R´ıos, P. Rom´an, Matrix-valued Gegenbauer-type type polynomials, Constr. Approx. 46 (2017), 459–487. [33] E. Koelink, P. Rom´an, Orthogonal vs. non-orthogonal reducibility of matrix valued measures, SIGMA Symmetry Integrability Geom. Methods Appl. 12 (2016), 008, 9 p. [34] T.H. Koornwinder, Matrix elements of irreducible representations of SU(2) × SU(2) and vector-valued orthogonal polynomials, SIAM J. Math. Anal. 16 (1985), 602–613. [35] M.G. Krein, Infinite J-matrices and a matrix moment problem, Dokl. Akad. Nauk SSSR 69 (1949), 125– 128. [36] M.G. Krein, Fundamental aspects of the representation theory of hermitian operators with deficiency index (m,m), AMS Translations ser. 2 97 (1971), 75–143. [37] I. Pacharoni, P. Rom´an, A sequence of matrix valued orthogonal polynomials associated to spherical func- tions, Constr. Approx. 28 (2008), 127–147. [38] M. van Pruijssen, Multiplicity free induced representations and orthogonal polynomials, Int. Math. Res. Not. 2018 (2018), 2208–2239. [39] M. van Pruijssen and P. Rom´an, Matrix-valued classical pairs related to compact Gelfand pairs of rank one, SIGMA Symmetry Integrability Geom. Methods Appl. 10 (2014), 113, 28 p.

IMAPP, Radboud Universiteit, Heyendaalseweg 135, 6525 GL Nijmegen, the Netherlands E-mail address: [email protected]

CIEM, FaMAF, Universidad Nacional de Cordoba,´ Medina Allende s/n Ciudad Universitaria, Cordoba,´ Argentina E-mail address: [email protected]