arXiv:1506.00087v1 [math.CA] 30 May 2015 ilctv nes,dntdby denoted inverse, tiplicative mletinteger smallest variable where eae oGgnae oyoil,Ceyhvpolynomials Chebyshev a , sequences Gegenbauer to special related more that so sequences Sheffer eune ytmtclyb h hoyo oenuba cal umbral modern of theory the by systematically sequences Abstract: eune n pelsqecsa w ucass Roman subclasses. two as sequences Appell and sequences preliminaries and Introduction 1. considered. Keywords: also are family this to belonging members quasi- new definition, propert determinantal other function, and generating determina quasi-monomial The a The arrays, established. Riordan is and mials t sequences the Sheffer Using the representation. between operational and function ating l,Hrieplnmas aurepolynomials, Eu polynomials, Laguerre , by Hermite formed sequen als, those Sheffer as of class such the sequences is sequences polynomial of classes nivril eis h series The series. invertible an f ( f * ¯ eodato;Emi:mmartgalcm(utzRiyas (Mumtaz [email protected] E-mail: author; Second orsodn uhr -al uui06galcm(Sub [email protected] E-mail: author; Corresponding ( t Let oa 2]frhrdvlpdteter fuba aclsa calculus umbral of theory the developed further [26] Roman eune fplnmaspa udmna oei mathema in role fundamental a play polynomials of Sequences )= )) a k t K ∈ over f ¯ ( nti ril,te2ieae hffrplnmasaeintro are polynomials Sheffer 2-iterated the article, this In eafil fcaatrsi eo Let zero. characteristic of field a be f K -trtdSee oyoil;Dffrnileuto;Deter equation; Differential polynomials; Sheffer 2-iterated ( t K o all for )= )) k hsa lmn of element an Thus . o hc h offiin of coefficient the which for t fadol fO( if only and if , k ∈ -trtdSee polynomials Sheffer 2-iterated uuiKhan Subuhi N := f f ( ( { t eateto Mathematics of Department t lgr ulmUniversity Muslim Aligarh 0 a opstoa nes,dntdby denoted inverse, compositional a has ) ) , − 1 1 , F 2 or f . . . , f ( lgr,India Aligarh, t ( a h form the has )=1 Then 1. = )) f t * = ) ( 1 } t n utzRiyasat Mumtaz and ) h re O( order The . fadol fO( if only and if , X k t ∞ =0 k F osntvns.Teseries The vanish. not does a etc. k etesto l omlpwrsre nthe in series power formal all of set the be t k h Khan) uhi , f ooiladohrpoete o some for properties other and monomial at) n otistecasso associated of classes the contains and ( t er fRodnary n relations and arrays Riordan of heory scle et series. delta a called is ) t al. et. e fteeplnmasaederived. are polynomials these of ies e,wihcnan ayimportant many contains which ces, f n aoipolynomials. Jacobi and culus. eicue,sc stesequences the as such included, re ( e oyoil,Ae polynomi- Abel polynomials, ler t tldfiiinfrteepolyno- these for definition ntal )o oe series power a of )) f ( dgnrlzdtecnetof concept the generalized nd n[72]suidteSheffer the studied [27,28] in t is n ftems famous most the of One tics. )=0 Then 0. = )) ⋆ ue ymaso gener- of means by duced iatldefinition. minantal f ¯ ( t f n satisfying and ) ( t f a mul- a has ) ( t f scalled is ) ( t sthe is ) (1 . 1) Let f(t) be a delta series and g(t) be an invertible series of the following forms:

∞ tn f(t)= f , f = 0, f 6= 0 (1.2a) n n! 0 1 nX=0 and ∞ tn g(t)= g , g 6= 0. (1.2b) n n! 0 n=0 X According to Roman [27, p.18 (Theorem 2.3.4)], the sn(x) is uniquely determined by two (formal) power series given by equations (1.2a) and (1.2b). The exponential of sn(x) is then given by

n 1 ¯ ∞ t ex(f(t)) = s (x) , (1.3) g(f¯(t)) n n! Xn=0 for all x in C, where f¯(t) is the compositional inverse of f(t). The sequence sn(x) in equation (1.3) is the Sheffer sequence for the pair (g(t),f(t)).

It has been given in [18] that the Sheffer polynomials sn(x) are ‘quasi-monomial’, for this see [11,29]. The associated raising and lowering operators are given by g (∂ ) 1 Mˆ = x − ′ x (1.4a) s g(∂ ) f (∂ )  x  ′ x and Pˆs = f(∂x), (1.4b) ∂ respectively, where ∂x := ∂x .

Also, for an invertible series g(t) and a delta series f(t), the sequence (sn(x))n N is Sheffer ∈ for the pair (g(t),f(t)) if and only if

k hg(t)f(t) |sn(x)i = cnδn,k, (1.5) for all n, k ≥ 0, where δn,k is the kronecker delta. Particularly, the Sheffer sequence for (1,f(t)) is called the associated sequence for f(t) defined by the generating function of the form:

n ¯ ∞ t exf(t) = s˜ (x) (1.6) n n! n=0 X and the Sheffer sequence for (g(t),t) is called the [3] for g(t) defined by the generating function of the form [27]:

1 ∞ tn ext = A (x) . (1.7) g(t) n n! n=0 X Now, we recall some preliminaries from [32].

2 Let P be the algebra of polynomials in the variable x over K and P ⋆ be the dual vector space of all functionals on P . Let (cn)n N be a fixed sequence of nonzero constants. Then for ∈ each f(t) in F, we define a linear functional f(t) in P ⋆ is defined by

n hf(t)|x i = cnan (1.8) and define a linear operator f(t) on P is defined by

n n cn n k f(t)x = akx − . (1.9) cn k Xk=0 − Now, we give the concept of Riordan arrays, which was introduced by Shapiro et. al. [23] and further studied by many authors, for example see [4, 8,15–17,24, 25].

For an invertible series g(t) and a delta series f(t), a generalized Riordan array with re- spect to the sequence (cn)n N is a pair (g(t),f(t)), which is an infinite, lower triangular array ∈ (an,k)0 k n< according to the following rule: ≤ ≤ ∞ tn (f(t))k a = g(t) , (1.10) n,k c c  n  k k where the functions g(t)(f(t)) /ck are called the column generating functions of the Riordan array. Particularly, the classical Riordan arrays corresponds to the case cn = 1 and the exponen- tial Riordan arrays corresponds to the case of cn = n!. One of the most important applications n of the theory of the Riordan arrays is to deal with summations of the form an,khk [24, 25]. k=0 P Also, for any fixed sequence (cn)n N, the set of all Riordan arrays (g(t),f(t)) is a ∈ under matrix multiplication and is called a Riordan group with respect to (cn)n N. The identity ∈ of this group is (1,t) and the inverse of the array (g(t),f(t)) is (1/g(f¯(t)), f¯(t)).

Costabile et. al. in [9] proposed a determinantal definition for the classical Bernoulli poly- nomials. Further, Costabile and Longo in [10] introduced the determinantal definition for the Appell sequences. Later on, by using the theory of Riordan arrays and relation between the Shef- fer sequences and Riordan arrays, the determinantal definition of Appell sequences is extended to Sheffer sequences [32]. The relations between the Sheffer sequences and Riordan arrays for the case of classical Sheffer sequences and classical Riordan arrays are given in [16] and for the case of generalized Sheffer sequences and generalized Riordan arrays are given in [15, 31].

Let (sn(x))n N be Sheffer for (g(t),f(t)) and suppose ∈ n n x = an,ksk(x), (1.11) Xk=0 then an,k is the (n, k) entry of the Riordan array (g(t),f(t)) [31].

3 In view of above fact, the following determinantal definition for the Sheffer sequences holds true [32]:

Let (sn(x))n N be Sheffer for (g(t),f(t)), then we have ∈ s (x)= 1 , (1.12) 0 a0,0 2 n 1 n 1 x x · · · x − x

a0,0 a1,0 a2,0 · · · an 1,0 an,0 −

( 1)n 0 a1,1 a2,1 · · · an 1,1 an,1 sn(x)= − − , a0,0a1,1...an,n

0 0 a2,2 · · · an 1,2 an,2 − . . . · · · . .

. . . · · · . .

0 0 0 · · · an 1,n 1 an,n 1 − − −

( 1)n Xn+1 = a a− ...a det , 0,0 1,1 n,n Sn (n+1)  ×  (1.13) 2 n where Xn+1 = (1,x,x ,...,x ), Sn (n+1) = (aj 1,i 1)1 i n, 1 j n+1 and an,k is the (n, k) entry − − ≤ ≤ ≤ ≤ of the Riordan array (g(t),f(t)). ×

Also, let (sn(x))n N be the sequence of polynomials defined by equations (1.12) and (1.13), ∈ where an,k is the (n, k) entry of the Riordan array (g(t),f(t)), then

n k sn(x)= bn,kx , (1.14) Xk=0 where bn,k is the (n, k) entry of the Riordan array (1/g(f¯(t)), f¯(t)) and (sn(x))n N be Sheffer ∈ for (g(t),f(t)) [31].

Motivated by the theory of Riordan arrays and relation between the Sheffer sequences and Riordan array, in this paper, a new family of the 2-iterated Sheffer polynomials is introduced by means of generating function. The determinantal definition of the 2-iterated Sheffer poly- nomials is established by using the relation between Sheffer sequences and Riordan array. The quasi-monomial properties of these polynomials are derived. The 2-iterated associated Sheffer polynomials are deduced and their properties are considered. Examples of some members be- longing to these families are given.

2. 2-iterated Sheffer polynomials

The 2-iterated Sheffer polynomials (2ISP in the following) are introduced by means of gen- erating function. Further, a determinantal definition of the 2ISP is given.

4 In order to obtain the generating function of the 2ISP, we prove the following result:

Theorem 2.1. The 2ISP are defined by the following generating function:

n 1 1 ¯ ¯ ∞ [2] t exp(xf2(f1(t))) = sn (x) . (2.1) g1(f¯1(t)) g2(f¯2(f¯1(t))) n! Xn=0 (1) (2) Proof. Let sn (x) and sn (x) be Sheffer for (g1(t),f1(t)) and (g2(t),f2(t)), respectively be two different polynomials defined by the generating functions of the forms:

n 1 ¯ ∞ (1) t exp(x(f1(t))) = sn (x) (2.2a) g1(f¯1(t)) n! Xn=0 and n 1 ¯ ∞ (2) t exp(x(f2(t))) = sn (x) , (2.2b) g2(f¯2(t)) n! nX=0 respectively.

Expanding the exponential function and then replacing the powers x0, x1,...,xn by the (2) (2) (2) polynomials s0 (x), s1 (x),...,sn (x), respectively in both sides of equation (2.2a), so that we have 1 f¯ (t) (f¯ (t))n ∞ tn 1+ s(2)(x) 1 + . . . + s(2)(x) 1 = s(1)(s(2)(x)) . (2.3) g (f¯ (t)) 1 1! n n! n 1 n! 1 1 n=0   X Using equation (2.2b) with t replaced by f¯1(t) in the l.h.s. and denoting the resultant 2ISP in the r.h.s. of equation (2.3) by

[2] (1) (2) sn (x)= sn (s1 (x)), (2.4) we get assertion (2.1).

Remark 2.1. We remark that equation (2.4) is the operational correspondence between the [2] (1) 2ISP sn (x) and Sheffer polynomials sn (x).

Remark 2.2. We know that the Sheffer sequence for (g(t),t) becomes the Appell sequence 1 1 1 1 An(x). Therefore, taking f1(t) = f2(t) = t which gives = and = g1(f¯1(t)) g1(t) g2(f¯2(f¯1(t))) g2(t) in equation (2.1) yields the generating function for the 2-iterated Appell polynomials (2IAP) [2] An (x) [19].

Remark 2.3. We know that the Sheffer sequence for (1,f(t)) becomes the associated Sheffer 1 1 sequence sn(x). Therefore, taking g1(t) = g2(t) = 1 which gives = = 1 in g1(f¯1(t)) g2(f¯2(f¯1(t))) equation (2.1) yields the following consequence of Theorem 2.1: e

5 [2] Corollary 2.1. The 2-iterated associated Sheffer polynomials (2IASP) sn (x) are defined by the following generating function: e ∞ tn exp(xf¯ (f¯ (t))) = s[2](x) . (2.5) 2 1 n n! nX=0 e [2] Remark 2.4. The following operational correspondence between the 2IASP sn (x) and associ- ated Sheffer sequences sn(x) holds:

[2] (1) (2) e e sn (x)= sn (s1 (x)). (2.6) (1) (2) It is shown in [32] that for sn (xe) and sn e(x)e be Sheffer for (g1(t),f1(t)) and (g2(t),f2(t)), respectively be two different Sheffer sequences defined by n (1) k sn (x)= bn,kx (2.7a) Xk=0 and n (2) k sn (x)= dn,kx , (2.7b) Xk=0 where bn,k is the (n, k) entry of the Riordan array (1/g1(f¯1(t)), f¯1(t)) and dn,k is the (n, k) entry (1) (2) of the Riordan array (1/g2(f¯2(t)), f¯2(t)), then the umbral composition of sn (x) and sn (x) is (2) (1) the sequence (sn (s (x)))n N defined by ∈ n n k n n n s(2)(s(1)(x)) = d s (x)= d b xj = d b xj = p xj, (2.8) n n,k k n,k k,j  n,k k,j n,j Xk=0 Xk=0 Xj=0 Xj=0 Xk=j Xj=0   1 (2) (1) where pn,j is the (n, j) entry of the Riordan array , f¯1(f¯2(t)) and sn (s (x)) g2(f¯2(t))g1(f¯1(f¯2(t))) is Sheffer for (g1(t)g2(f1(t)),f2(f1(t))).  

[2] (1) Remark 2.5. We remark that, the 2ISP sn (x) are actually the composition of sn (x) and (2) sn (x) and are defined by n [2] sn (x)= dn,ksk(x), (2.9) Xk=0 1 where dn,k is the is the (n, k) entry of the Riordan array , f¯1(f¯2(t)) . g2(f¯2(t))g1(f¯1(f¯2(t)))   The series definition (2.9) can also be obtained by replacing the powers x1 and xk by the (2) (2) polynomials s1 (x) and sk (x), respectively, in equation (2.7b) and then using equation (2.4) in the l.h.s. of resultant equation.

[2] Now, we derive the determinantal definition for the 2ISP sn (x). For this we prove the following result:

6 [2] Theorem 2.2. The 2ISP sn (x) of degree n are defined by

s[2](x)= 1 , (2.10) 0 a0,0 (2) (2) (2) (2) 1 s1 (x) s2 (x) · · · sn 1(x) sn (x) −

a0,0 a1,0 a2,0 · · · an 1,0 an,0 −

n [2] ( 1) 0 a1,1 a2,1 · · · an 1,1 an,1 sn (x)= − , a0,0a1,1...an,n −

0 0 a2,2 · · · an 1,2 an,2 − . . . · · · . .

. . . · · · . .

0 0 0 · · · an 1,n 1 an,n 1 − − −

( 1)n Sn+1(x) = a a− ...a det , 0,0 1,1 n,n Mn (n+1)  ×  (2.11) (2) (2) where Sn+1(x) = (1,s1 (x),...,sn (x)), Mn (n+1) = (aj 1,i 1)1 i n, 1 j n+1 and an,k is the × − − ≤ ≤ ≤ ≤ (n, k) entry of the Riordan array (g1(t)g2(f1(t)),f2(f1(t))).

n (2) (2) Proof. Replacing the power x by the polynomial sn (x) in the l.h.s. and x by s1 (x) in r.h.s. of equation (1.11) and then using operational correspondence (2.4) in the r.h.s. of resultant equation, we find n (2) [2] sn (x)= an,ksk (x), (2.12) Xk=0 where an,k is the (n, k) entry of the Riordan array (g1(t)g2(f1(t)),f2(f1(t))).

[2] The above identity leads to the following system of infinite equations in the unknown sn (x) for n = 0, 1,... :

[2] a0,0s0 (x) = 1,   [2] [2] (2) a1,0s (x)+ a1,1s (x)= s (x),  0 1 1   (2.13)  [2] [2] [2] (2) a2,0s0 (x)+ a2,1s1 (x)+ a2,2s2 (x)= s2 (x), . .   [2] [2] [2] [2] (2) an,0s0 (x)+ an,1s1 (x)+ an,2s2 (x)+ . . . + an,nsn (x)= sn (x).   From first equation of system (2.13), we get assertion (2.10). Also, the special form of system [2] (2.13) (lower triangular) allows us to work out the unknown sn (x). Operating with the first n + 1 equations simply by applying the Cramer’s rule, we have

7 a0,0 0 0 · · · 0 1

(2) a1,0 a1,1 0 · · · 0 s1 (x)

(2) 1 a a a · · · 0 s (x) s[2](x)= 2,0 2,1 2,2 2 . (2.14) n a a . . . a 0,0 1,1 n,n . . . · · · . .

(2) an 1,0 an 1,1 an 1,2 · · · an 1,n 1 sn 1(x) − − − − − −

(2) an,0 an,1 an,2 · · · an,n 1 sn (x) −

Then, bringing (n + 1)-th column to the first place by n transpositions of adjacent column and noting that the determinant of a square matrix is the same as that of its transpose, we obtain assertion (2.11).

Remark 2.6. We know that for (1,f(t)), the Sheffer sequences sn(x) become the associated sequences sn(x), i.e., for cn = n! in equation (1.9), (sn(x))n N is associated to f(t) and an,k ∈ is then the (n, k) entry of the Riordan array (1,f(t)) and coefficients of the associated Sheffer polynomialse obtained are of the form: tn (f(t))0 c a = = n [tn]1 = δ , (2.15) n,0 c c c n,0  n  0 0 In view of equation (2.15), the following determinantal definition of the associated Sheffer sequences sn(x) holds true [32]:

e s (x)= 1 , (2.16) 0 a0,0 2 n 1 n x x · · · x − x e

a1,1 a2,1 · · · an 1,1 an,1 − ( 1)n sn(x)= a a− ...a , 0,0 1,1 n,n 0 a2,2 · · · an 1,2 an,2 − . . · · · . .

e . . · · · . .

0 0 · · · a a n 1,n 1 n,n 1 − − −

( 1)n Sn(x) = − det , a0,0a1,1...an,n M(n 1) n ! − × e (2.17) n f where Sn(x) = (x,...,x ), M(n 1) n = (aj,i)1 i n 1, 1 j n and an,k is the (n, k) entry of the ≤ ≤ − ≤ ≤ Riordan array (1,f(t)). − × e f

8 (2) (2) Using equation (2.15) and replacing sn (x) by sn (x) in the r.h.s. of the determinantal def- [2] inition (2.10) and (2.11) of the 2ISP sn (x), we obtain the following consequence of Theorem 2.2: e [2] Corollary 2.2. The 2IASP sn (x) of degree n are defined by

e s[2](x)= 1 , (2.18) 0 a0,0 (2) (2) (2) (2) s1 (x) s2 (x) · · · sn 1(x) sn (x) e −

ea1,1 ea2,1 · · · ean 1,1 ean,1 − [2] ( 1)n sn (x)= − , a0,0a1,1...an,n 0 a2,2 · · · an 1,2 an,2 − . . · · · . . e . . · · · . .

0 0 · · · an 1,n 1 an,n 1 − − −

( 1)n Sn(x) = − det , a0,0a1,1...an,n M(n 1) n ! − × e (2.19) (2) (2)f where Sn(x) = (s1 (x),..., sn (x)), M(n 1) n = (aj,i)1 i n 1, 1 j n and an,k is the (n, k) entry − × ≤ ≤ − ≤ ≤ of the Riordan array (1,f2(f1(t))). e e e f Remark 2.7. We know that for (g(t),t), the Sheffer sequences sn(x) become Appell sequences An(x). The coefficients an,k in equation (1.13) are then the (n, k) entry of the exponential Riordan array (g(t),t) as follows:

n k t t n! n k n an,k = g(t) = [t − ]g(t)= gn k. (2.20) n! k! k! k −     Therefore, using equation (2.20) in the determinantal definition (2.10) and (2.11) of the [2] [2] 2ISP sn (x) yields the determinantal definition of the 2IAP An (x), which is given in [20].

3. Quasi-monomial and other properties

[2] In this section, we frame the 2ISP sn (x) within the context of monomiality principle and [2] derive certain other properties of the 2ISP sn (x).

[2] In order to derive the multiplicative and derivative operators for the 2ISP sn (x), we prove the following result:

[2] Theorem 3.1. The 2ISP sn (x) are quasi-monomial with respect to the following multiplicative and derivative operators:

g1′ (f2(∂x)) g2′ (∂x) 1 Mˆ [2] = x − f ′ (f (∂ )) − (3.1a) s g (f (∂ )) 1 2 x g (∂ ) f (f (∂ )) f (∂ )  1 2 x 2 x  1′ 2 x 2′ x 9 and ˆ Ps[2] = f1(f2(∂x)), (3.1b) ∂ respectively, where ∂x := ∂x .

Proof. Differentiating equation (2.1) partially with respect to t, we find ¯ ¯ ¯ g2′ (f2(f1(t))) g1′ (f1(t)) xf¯′ (f¯2(f¯1(t))) f¯′ (f¯1(t)) − f¯′ (f¯2(f¯1(t))) f¯′ (f¯1(t)) − f¯′ (f¯1(t)) 2 1 g (f¯ (f¯ (t))) 2 1 g (f¯ (t)) 1  2 2 1 1 1  n 1 1 ∞ [2] t exp(xf¯2(f¯1(t))) = s (x) , (3.2) g (f¯ (t)) g (f¯ (f¯ (t))) n+1 n! 1 1 2 2 1 n=0 X which can also be simplified as ¯ ¯ ¯ g2′ (f2(f1(t))) g1′ (f1(t)) ¯ ¯ 1 1 1 x − ¯ ¯ − ¯ f2′ (f2(f1(t))) ¯ ¯ ¯ ¯ ¯ ¯ g2(f2(f1(t))) g1(f1(t)) f ′2(f2(f1(t))) f1′ (f1(t)) g1(f1(t)) g2(f2(f1(t)))   n ∞ tn exp(xf¯ (f¯ (t))) = s[2] (x) . (3.3) 2 1 n+1 n! o nX=0 g′ (t) g′ (t) Since g (t) and g (t) are invertible series of t, therefore 1 and 2 possess power series 1 2 g1(t) g2(t) [2] expansions of t. Thus, in view of the following identity for the 2ISP sn (x):

1 1 1 1 ∂ exp(xf¯ (f¯ (t))) = f¯ (f¯ (t)) exp(xf¯ (f¯ (t))) , x g (f¯ (t)) g (f¯ (f¯ (t))) 2 1 2 1 g (f¯ (t)) g (f¯ (f¯ (t))) 2 1  1 1 2 2 1   1 1 2 2 1  (3.4a) or, equivalently

1 1 1 1 f (f (∂ )) exp(xf¯ (f¯ (t))) = t exp(xf¯ (f¯ (t))) , 1 2 x g (f¯ (t)) g (f¯ (f¯ (t))) 2 1 g (f¯ (t)) g (f¯ (f¯ (t))) 2 1  1 1 2 2 1   1 1 2 2 1  (3.4b) equation (3.3) becomes

g1′ (f2(∂x)) g2′ (∂x) 1 1 1 x − f1′ (f2(∂x)) − ¯ ¯ ¯ g1(f2(∂x)) g2(∂x) f1′ (f2(∂x)) f2′ (∂x) g1(f1(t)) g2(f2(f1(t)))   n ∞ tn exp(xf¯ (f¯ (t))) = s[2] (x) . (3.5) 2 1 n+1 n! n=0 o X Again, using generating function (2.1) in the l.h.s. of equation (3.5) and then rearranging the summation yields

n n ∞ g1′ (f2(∂x)) g2′ (∂x) 1 [2] t ∞ [2] t x − f ′ (f (∂ )) − s (x) = s (x) . g (f (∂ )) 1 2 x g (∂ ) f (f (∂ )) f (∂ ) n n! n+1 n! n=0  1 2 x 2 x  1′ 2 x 2′ x    n=0 X X (3.6)

10 Equating the coefficients of the same powers of t in both sides of the above equation, we find

g1′ (f2(∂x)) g2′ (∂x) 1 [2] [2] x − f1′ (f2(∂x)) − sn (x) = sn+1(x), (3.7) g1(f2(∂x)) g2(∂x) f1′ (f2(∂x)) f2′ (∂x)    n o [2] which in view of monomiality principle equation Mˆ {sn(x)} = sn+1(x) [18] for sn (x) yields as- sertion (3.1a).

In order to prove assertion (3.1b), we use generating function (2.1) in both sides of the identity (3.4b), so that we have

n n ∞ [2] t ∞ [2] t f1(f2(∂x)) sn (x) = sn 1(x) . (3.8) ( n!) − (n − 1)! nX=0 nX=1 Rearranging the summation in the l.h.s. of equation (3.8) and then equating the coefficients of the same powers of t in both sides of the resultant equation, we find

[2] [2] f1(f2(∂x)) sn (x) = n sn 1(x), n ≥ 1, (3.9) − n o [2] which in view of monomiality principle equation Pˆ{sn(x)} = n sn 1(x) [18] for sn (x) yields − assertion (3.1b).

[2] Theorem 3.2. The 2ISP sn (x) satisfy the following differential equation: x g (f (∂ )) f (f (∂ )) g (∂ ) f (f (∂ )) − 1′ 2 x 1 2 x − 2′ x 1 2 x − n s[2](x) = 0. (3.10) f (f (∂ )) g (f (∂ )) f (∂ ) g (∂ ) f (f (∂ )) f (∂ ) n  1′ 2 x 1 2 x  2′ x 2 x 1′ 2 x 2′ x 

Proof. Using equations (3.1a) and (3.1b) in monomiality principle equation Mˆ Pˆ{sn(x)} = [2] nsn(x) [18] for sn (x), we get assertion (3.10).

Remark 3.1. We know that the Sheffer sequence for (1,f(t)) becomes the associated Sheffer sequence sn(x). Therefore, taking g1(t)= g2(t) = 1 which implies g1′ (t)= g2′ (t) = 0 in equations (3.1a), (3.1b) and (3.10), we get the following consequence of Theorem 3.1 and Theorem 3.2: e [2] Corollary 3.1. The 2IASP sn (x) are quasi-monomial with respect to the following multiplicative and derivative operators: ˆ x e Mse[2] = (3.11a) f1′ (f2(∂x)) f2′ (∂x) and ˆ Pse[2] = f1(f2(∂x)), (3.11b) respectively.

[2] Corollary 3.2. The 2IASP sn (x) satisfy the following differential equation: x f (f (∂ )) e 1 2 x − n s[2](x) = 0. (3.12) f (f (∂ )) f (∂ ) n  1′ 2 x 2′ x  e 11 Remark 3.2. We know that for (g(t),t), the Sheffer sequences sn(x) reduce to the Appell se- quences An(x). Therefore, taking f1(t)= f2(t)= t which implies f1′ (t)= f2′ (t) = 1 in equations (3.1a), (3.1b) and (3.10) yields the multiplicative and derivative operators and differential equa- [2] tion for the 2IAP An (x), which are given in [19].

[2] Now, we prove the conjugate representation of the 2ISP sn (x). For this, we prove the following result:

[2] Theorem 3.2. Let sn (x) be Sheffer for (g1(t)g2(f1(t)),f2(f1(t))), then the following conjugate [2] representation for the 2ISP sn (x) holds true:

n ¯ ¯ ¯ 1 ¯ ¯ k [2] h(g2(f2(t))g1(f1(f2(t))))− (f1(f2(t))) | sn(x)i sn (x)= sk(x). (3.13) ck Xk=0 [2] Proof. Since sn (x) are defined by equation (2.9), where dn,k is the (n, k) entry of the Riordan 1 array , f¯1(f¯2(t)) . Then, according to the definition of Riordan arrays, we g2(f¯2(t))g1(f¯1(f¯2(t))) have   tn 1 (f¯ (f¯ (t)))k d = 1 2 . (3.14) n,k c g (f¯ (t))g (f¯ (f¯ (t))) c  n  2 2 1 1 2 k Simplifying the above equation yields

n 1 t 1 k d = (g (f¯ (t))g (f¯ (f¯ (t))))− ((f¯ (f¯ (t)))) n,k c c 2 2 1 1 2 1 2 k  n  1 1 k = h(g2(f¯2(t))g1(f¯1(f¯2(t))))− (f¯1(f¯2(t))) | sn(x)i, (3.15) ck Now, using equation (3.15) in equation (2.9), we are led to assertion (3.13).

In the next section, examples of some members belonging to the 2ISP and 2IASP are con- sidered.

4. Examples

(α) We recall that the Ln (x) of order α form the Sheffer sequence for 1 t the pair (1 t)α+1 , t 1 [2, 22,27] are defined by the generating function: − −   1 −xt ∞ exp = L(α)(x)tn, (4.1) (1 − t)α+1 1 − t n n=0   X which for α = 0 gives the generating function of the Laguerre polynomials Ln(x) as [2]:

1 −xt ∞ exp = L(α)(x)tn. (4.2) (1 − t) 1 − t n n=0   X

12 λ Also, the polynomials of the Gegenbauer case denoted by −n sn(x) considered in [32] form λ0 the Sheffer sequence for the pair 2 , t and are defined by the generating 1+√1 t2 1+√−1 t2 − − function of the form:   

∞ 2 λ λ0 2 λ −λ n (1 + t ) − (1 − 2xt + t )− = s (x)t , (4.3) n n n=0 X   (λ) which for λ0 = λ gives the generating function of the Gegenbauer polynomials Cn (x) as [1,22, 26]:

2 λ ∞ (λ) n (1 − 2xt + t )− = Cn (x)t . (4.4) nX=0 We note that corresponding to each member belonging to the Sheffer family, there exists a new special polynomial belonging to the 2ISP family. Thus, by making suitable choice for the functions g(t) and f(t) in equation (2.1), we get the generating function for the corresponding member belonging to the 2ISP family. The other properties of these special polynomials can be obtained from the results derived in previous sections.

We consider the following examples:

1 t Example 4.1. Taking g1(t)= g2(t)= (1 t)α+1 and f1(t)= f2(t)= t 1 in the l.h.s. of generating function (2.1), we find that the resultant− 2-iterated Laguerre polynomials− (2ILP) of order α, (α)[2] which may be denoted by Ln (x) in the r.h.s. are defined by the following generating function:

1 ∞ (α)[2] n α+1 α 1 exp(xt)= Ln (x)t . (4.5) (1 − t) (1 − t)− − nX=0 (α)[2] The operational correspondence between the 2ILP of order αLn (x) and Laguerre poly- (α) nomials of order, α Ln (x) is given by:

(α)[2] (α,1) (α,2) L (x)= Ln (L1 (x)). (4.6)

(α)[2] The 2ILP of order α, Ln (x) are quasi-monomial with respect to the following multi- plicative and derivative operators:

3 (α + 1) Mˆ (α)[2] = x − (α + 1)(∂ − 1) − , (4.7a) L x (∂ − 1)  x  ˆ PL(α)[2] = ∂x. (4.7b) (α)[2] The 2ILP of order α, Ln (x) satisfy the following differential equation:

(α + 1)∂ x∂ − (α + 1)(∂ − 1)3∂ − x − n L(α)[2](x) = 0. (4.8) x x x (∂ − 1) n  x 

13 (α) Consider the following series definition for the Laguerre polynomials of order α, Ln (x) [27]:

n (−1)k n + α L(α)(x)= xk. (4.9) n k! n − k Xk=0   k (α) (α) Replacing the power x by the polynomial Lk (x) in the r.h.s. and x by L1 (x) in the l.h.s. of equation (4.9) and then using equation (4.6) in the l.h.s. of resultant equation yields (α)[2] the following series definition for the 2ILP of order α, Ln (x):

n (−1)k n + α L(α)[2](x)= L(α)(x). (4.10) n k! n − k k Xk=0   k n! n+α It has been shown in [32] that for an,k = (−1) k! n k the determinantal definition of the Sheffer polynomials given by equations (1.12) and (1.13)− reduces to the determinantal definition (α)  of the Laguerre polynomials of order α, Ln (x).

(α) k n! n+α Therefore, taking sn(x)= Ln (x) and an,k = (−1) k! n k in equations (2.10) and (2.11), − (α)[2] we find that the following determinantal definition of the 2IL P of order α, Ln (x) holds true:

(α)[2] L0 (x) = 1, (4.11)

(α) (α) (α) (α) 1 L1 (x) L2 (x) . Ln 1(x) Ln (x) −

1 α + 1 (α + 2)2 . (α + n − 1)n 1 (α + n)n −

(α)[2] n(n+3) 0 −1 −2(α + 2) . −(n − 1)(α + n − 1)n 2 −n(α + n)n 1 L (x) = (−1) 2 − − , n (n 1)(n 2) n(n 1) 0 0 1 . − − (α + n − 1)n 3 − (α + n)n 2 2 − 2 − ......

0 0 0 . (−1)n 1 (−1)n 1n(n + α) − −

n(n+3) Sn+1(x) = (−1) 2 det , Mn (n+1)  ×  (4.12) (α) (α) (α) where Sn+1(x) = (1,L1 (x),...,Ln (x)), Mn (n+1) = (aj 1,i 1)1 i n, 1 j n+1 and Ln (x) (n = − − ≤ ≤ ≤ ≤ 0, 1,...) are the Laguerre polynomials of order× α.

Remark 4.1. We remark that by taking α = 0 in the results derived above for the 2-iterated Laguerre polynomials (2ILP) of order α, we get the corresponding results for the 2-iterated [2] Laguerre polynomials (2ILP), which may be denoted by Ln (x).

14 λ0 Example 4.2. Taking g (t)= g (t)= 2 and f (t)= f (t)= t in the l.h.s. of 1 2 1+√1 t2 1 2 1+√−1 t2 − − generating function (2.1), we find that the resultant 2-iterated polynomials of the Gegenbauer λ s[2] case (2IPoGc), which may be denoted by −n n (x) in the r.h.s. are defined by the following generating function:  λ 1+ t2 0 4xt(1 + t2) ∞ −λ exp = s[2](x)tn. (4.13) 1 + 6t2 + t4 1 + 6t2 + t4 n n n=0     X   [2] The operational correspondence between the 2IPoGc sn (x) and polynomials of the Gegen- bauer case sn(x) is given by: s[2] s(1) s(2) n (x)= n ( 1 (x)). (4.14) λ s[2] The 2IPoGc −n n (x) are quasi-monomial with respect to the following multiplicative and derivative operators: 

λ0 1 2 λ0+2 λ02 − ∂x(1 + 1 − ∂x) Mˆ s(α)[2] = x − 2 2 2 2 2 2 λ0+1  2(1 − ∂x + 1 − ∂x) 1+ 1 −p∂x + 2(1 − ∂x + 1 − ∂x) λ ∂q p  p 1+ q1 − ∂2 p  − 0 x x . 2 2 1 − ∂x(1 + 1 − ∂x) 2 2 2 2 2  2(1 − ∂x + 1 − ∂x) (1+ p1 − ∂x + 2(1 − ∂x + 1 − ∂x)) p 1 p q q × p p p (4.15) 2 (1 − ∂x)

p −∂x Pˆs(α)[2] = . (4.16) 2 2 2 1+ 1 − ∂x + 2(1 − ∂x + 1 − ∂x) q λ s[2] p p The 2IPoGc −n n (x) satisfy the following differential equation:

 λ 2λ0 1∂ (1 + 1 − ∂2)λ0+2 x + 0 − x x 2 2 2 2 2 2 λ0+1  2(1 − ∂x + 1 − ∂x) 1+ 1 −p∂x + 2(1 − ∂x + 1 − ∂x) q λ ∂ p  p q∂ (1 + 1 −p∂2)  + 0 x x x 2 2 1 − ∂x(1 + 1 − ∂x) 2 2 2 2 2 2  2(1 − ∂x + 1 − ∂x) (1+ 1p− ∂x + 2(1 − ∂x + 1 − ∂x)) p p q q 1 s[2] p p p × − n n (x) = 0. (4.17) 1 − ∂2 x  p It has been shown in [32] that by taking

0 if n − k is odd, k an,k = cn( 1) λ0+k λ0+n (4.18) ( c −2nλ +n n−k if n − k is even, k 0 2

λ  where cn = 1/ n , the determinantal definition of the Sheffer polynomials given by equations (1.12) and (1.13) reduces to the determinantal definition of the polynomials of Gegenbauer case 

15 λ s −n n(x).

 (2) Therefore, taking sn (x) = snx and using equation (4.18) in equations (2.10) and (2.11), λ s[2] we find that the following determinantal definition of the 2IPoGc −n n (x) holds true: 

s[2] 0 (x) = 1, (4.19)

1 s1(x) s2(x) . sn 1(x) sn(x) −

λ n!λ0(λ0+n 1)! 1 0 0 . 0 − 2λ(λ+1) n n n 2 !λ(λ+1)...(λ+n 1) λ0+ ! 2 − 2

  0 − 1 0 . 0 [2] n(n+3) n(n+1) 2 sn (x) = (−1) 2 2 2 ,

1 n(n 1)...3(λ0+2)(λ0+n 1)! 0 0 . 0 − − 4 n n−2 n+2 2 !(λ+2)...(λ+n 1) λ0+ ! 2 − 2 ......  . 

n 1 0 0 0 . 1 − 0 −2

 n(n+3) n(n+1) Sn+1(x) = (−1) 2 2 2 det , Mn (n+1)  ×  (4.20) where Sn+1(x) = (1, s1(x),..., sn(x)), Mn (n+1) = (aj 1,i 1)1 i n, 1 j n+1 and sn(x) (n = − − ≤ ≤ ≤ ≤ 0, 1,...) are the polynomials of Gegenbauer× case.

Remark 4.2. We remark that, by taking λ0 = λ in the results derived above for the 2IPoGc λ s[2] −n n (x) we get the corresponding results for the 2-iterated Gegenbauer polynomials (2IGP) (λ)[2] which  may be denoted by Cn (x).

Remark 4.3. It is given in [1] that, for λ = 1, the Gegenbauer polynomials becomes the Cheby- (1) shev polynomials of the second kind Un(x), i.e., Un(x) = Cn (x) and for λ = 1/2, the Gegen- (1/2) bauer polynomials becomes the Legendre polynomials Pn(x), i.e., Pn(x)= Cn (x). Therefore, taking λ = 1 and λ = 1/2 in the results of the 2-iterated Gegenbauer polynomials (2IGP) (λ)[2] Cn (x), we obtained the corresponding results of the 2-iterated Chebyshev polynomials of the [2] [2] second kind Un (x) and 2-iterated Legendre polynomials Pn (x).

Now, we proceed to introduce certain members belonging to the 2IASP. We recall that the x associated Sheffer family contains the falling factorials and exponential polynomials φn(x) a n as the important members.  

The falling factorial x associated to f(t) = eat − 1 [27] is defined by the generating a n   16 function: n 1 ∞ x t exp(xa− log(1 + t)) = (4.21) a n n! n=0 X   and the exponential polynomials φn(x) associated to f(t) = log(1 + t) [5, 27] defined by the generating function: ∞ tn exp(x(et − 1)) = φ (x) . (4.22) n n! n=0 X Also, for each member belonging to the associated Sheffer family, there exists a new special polynomial belonging to the 2IASP family. Thus, by making suitable choice for the functions f1(t) and f2(t) in equation (2.5), we get the generating function for the corresponding member belonging to the 2IASP family. The other properties of these special polynomials can be ob- tained from the results derived in previous sections.

We consider the following examples:

at Example 4.3. Taking f1(t)= f2(t)= e − 1 in the l.h.s. of generating function (2.5), we find x [2] that the resultant 2-iterated falling factorial (2IFF), denoted by a n in the r.h.s. are defined by the following generating function:  [2] n 1 1 ∞ x t exp(xa− log(1 + a− log(1 + t))) = . (4.23) a n n! Xn=0   [2] The operational correspondence between the 2IFF x and falling factorial x is given a n a n by:     (2) x [2] x (1) = a 1 . (4.24) a n a n [2]     The 2IFF x are quasi-monomial with respect to the following multiplicative and deriva- a n tive operators:   ˆ x M [2] = a∂ , (4.25a) x ae x a∂x a n ae ae a(ea∂x 1) ˆ [2] P x = e − − 1. (4.25b) a n x [2]  The 2IFF a n satisfy the following differential equation:

 a(ea∂x 1) x e − − 1 x [2] a∂x − n = 0. (4.26) aeae aea∂x ! a n   Consider the following series definition for the falling factorial x [27]: a n   n x = ans(n, k)xk, (4.27) a n   Xk=0 17 n where s(n, k) are the Stirling numbers of the first kind defined by s(n, k) = , k, n ∈ k N, 1 ≤ k

Replacing the power xk by the polynomial x in the r.h.s. and x by x in the l.h.s. a k a 1 of equation (4.27) and then using equation (4.24) in the l.h.s. of resultant equation  yields the x [2] following series definition for the 2IFF a n :  n x [2] x = ans(n, k) . (4.28) a n a k   Xk=0   n It has been shown in [32], that for an,k = a S(n, k), where S(n, k) are the Stirling numbers of the second kind in equations (2.16) and (2.17), the determinantal definition of the associated Sheffer polynomials reduces to the determinantal definition of falling factorials x . a n   x n Therefore, taking sn(x) = a n and an,k = a S(n, k) in the r.h.s. of equations (2.18) and [2] (2.19), we find that the 2IFF x  are defined by the following determinantal definition: e a n   [2] x = 1, (4.29) a 0   x x · · · x x a 1 a 2 a n 1 a n       −   2 n 1 n aS(1, 1) a S(2, 1) · · · a − S(n − 1, 1) a S(n, 1) [2] n+1 x ( 1) a = − n+1 , n a( 2 ) 0 a2S(2, 2) · · · an 1S(n − 1, 2) anS(n, 2)   − . . · · · . .

. . · · · . .

n 1 n 0 0 · · · a − S(n − 1,n − 1) a S(n,n − 1)

( 1)n+1 Sn(x) = − det , (n+1) a 2 M(n 1) n ! − × e (4.30) x f x x where Sn(x)= a ,..., a , M(n 1) n = (aj,i)1 i n 1, 1 j n and a (n = 0, 1,...) are 1 n − × ≤ ≤ − ≤ ≤ n the falling factorials.       e f

Example 4.4. Taking f1(t) = f2(t) = log(1 + t) in the l.h.s. of generating function (2.5), we [2] find that the resultant 2-iterated exponential polynomials (2IEP) denoted by φn (x) in the r.h.s. are defined by the following generating function:

n (et 1) ∞ [2] t e − − 1= φ (x) . (4.31) n n! nX=0

18 [2] The operational correspondence between the 2IEP φn (x) and exponential polynomials φn(x) is given by: [2] (1) (2) φn (x)= φn (φ1 (x)). (4.32) [2] The 2IEP φn (x) are quasi-monomial with respect to the following multiplicative and deriva- tive operators: x Mˆ [2] = , (4.33a) φ (1 + log(1 + t))(1 + t) ˆ Pφ[2] = log(1 + log(1 + t)). (4.33b) [2] The 2IEP φn (x) satisfy the following differential equation:

x log(1 + log(1 + t)) − n φ[2](x) = 0. (4.34) (1 + log(1 + t))(1 + t) n  

Consider the following series definition for the exponential polynomials φn(x) [27]:

n k φn(x)= S(n, k)x , (4.35) Xk=0 where S(n, k) are the Stirling numbers of the second kind defined by

k n 1 k j k n S(n, k)= = (−1) − j . (4.36) k k! j   Xj=0   k Replacing the power x by the polynomial φk(x) in the r.h.s. and x by φ1(x) in the l.h.s. of equation (4.30) and then using equation (4.35) in the l.h.s. of resultant equation yields the [2] following series definition for the 2IEP φn (x):

n [2] φn (x)= S(n, k)φk(x), (4.37) Xk=0

It has been shown in [32], that for an,k = s(n, k), where s(n, k) are the Stirling numbers of the first kind in equations (2.16) and (2.17), the determinantal definition of the associated Sheffer polynomials reduces to the determinantal definition of exponential polynomials φn(x).

Therefore, taking sn(x) = φn(x) and an,k = s(n, k) in equations (2.18) and (2.19), we [2] find that the 2-iterated exponential polynomials (2IEP) φn (x) are defined by the following determinantal definition:e

19 [2] φ0 (x) = 1, (4.37) φ1(x) φ2(x) · · · φn 1(x) φn(x) −

s(1, 1) s(2, 1) · · · s(n − 1, 1) s(n, 1)

[2] φ (x) = (−1)n+1 , n 0 s(2, 2) · · · s(n − 1, 2) s(n, 2)

. . · · · . .

. . · · · . .

0 0 · · · s(n − 1,n − 1) s(n,n − 1)

S (x) = (−1)n+1det n , M(n 1) n ! − × e (4.38) f where Sn(x) = φ1(x),...,φn(x) , M(n 1) n = (aj,i)1 i n 1, 1 j n and φn(x) (n = 0, 1,...) − × ≤ ≤ − ≤ ≤ are the exponential polynomials.  e f Similar results can also be proved for certain other known members belonging to the Sheffer family. These members are listed in Table 1:

Table 1. Some known Sheffer polynomials.

S.No. A(t); H(t) g(t); f(t) Generating Functions Polynomials 2 2 t 2 n t 4 t 2xt t t I. e− ; 2t e ; 2 e − = n∞=0 Hn(x) n! Hn(x) [2] m t m P n t ( ν ) t − m t II. e− ; νt e ; ν exp(νxt t ) = n∞=0 Hn,m,ν (x) n! Generalized Hermite polynomials Hn,m,ν (x) [21] P − t 1 − t 1 1 xt L x tn III. (1 )− ; (1 )− (1 t) exp t 1 = n∞=0 n( ) Laguerre polynomials − − ! t t P t 1 t 1 n!Ln(x) [2] − − t t 1+t 2 e 1 t 1+t x tn IV. ; ln ; − = ∞ Pn(x) Pidduck polynomials 1 t 1 t et 1 et+1 1 t 1 t n=0 n! − − − − −     Pn(x) [7, 13] P n βt − t − β − t (β) t V. e ; 1 e (1 t)− ; exp βt + x(1 e ) = n∞=0 an (x) n! Actuarial polynomials (β) ln(1 − t)   P an (x) [7] x n t t − t t t VI. e− ; exp a(e 1) ; e− 1+ a = n∞=0 cn(x ; a) n! Poisson-Charlier polynomials t t−    ln 1+ a a(e 1) P cn(x ; a) [12, 14, 30] µ µ µ VII. 1+(1+ t)λ − ; 1+ eλt ; 1+(1+ t)λ − (1 + t)x Peters polynomials n  t  et −   s x λ, µ t s x λ, µ ln(1 + ) 1 = n∞=0 n( ; ) n! n( ; ) [7] t t t t x tn VIII. ;ln(1 + t) ; e − 1 (1 + t) = ∞ bn(x) Bernoulli polynomials ln(1+t) et 1 ln(1+P t) n=0 n! − of the second kind bn(x) [14] P n 2 1 t t − 2 x t IX. 2+t ; ln(1 + t) 2 (1 + e ); e 1 2+t (1 + t) = n∞=0 rn(x) n! Related polynomials r (x) [14] P n 1 1 tn X. ; arctan(t) sect; tan t exp(x arctan(t)) = n∞=0 Rn(x) n! Hahn polynomials 1+t2 1+t2 q q P Rn(x) [6]

1 1 1+t 2 a 1 XI. (1 − 4t)− 2 ; (1 − 4t)− 2 − Shively’s pseudo-Laguerre (1 t)a 1+√1 4t − 2  −  2 a 1 1 1 1+t 4xt n × − ; − × exp − = ∞ Rn(a,x)t polynomials Rn(a,x) [22] 1+√1 4t 4 4 1 t (1+√1 4t ) 2 n=0 − − ! − !  4t  − P (1+√1 4t ) 2 −

20 1+α+β 2+α+β 1+α+β , ; 2xt 1 2 − 1 α β 2 2 (1 t)2 XII. 1+α+β ; ; (1 t)− − − 2F1 The polynomials of (1 t) 1+√1+2t " 1+ α − # −   2t t ∞ tn = Jn(x) the Jacobi case [26] (1 t)2 1+t+√1+2t cn − n=0 2 P 2 (1 t ) 1 1 t ∞ − nt x tn XIII. (1+− t) ; ; − 2 = ( 1) n( ) The polynomials of 1 t2 1 2xt+t n=0 − − 2t q t P − 2 − the Chebyshev case [26] (1+t) 1+ 1 t2 − q (α)[2] [2] Now, we proceed to plot the graphs related to the 2ILP of order, αLn (x), 2ILP Ln (x), [2] [2] (α) 2IFF (x)n and 2IEP φn (x) for n = 4. For this, we need the first few values of Ln (x), Ln(x), (α) (x)n and φn(x). We give the list of first five Ln (x), Ln(x), (x)n and φn(x) in Table 2.

(α) Table 2. First five Ln (x), Ln(x), (x)n and φn(x).

n 0 1 2 3 4 L(α) x α − x 1 α α 1 α α α 1 α α α α − x − 1 α α α − x x n ( ) 1 1+ 2 (1 + )(2 + ) 6 (1 + )(2 + )(3 + ) 24 (1 + )(2 + )(3 + )(7 + ) 8 (2 + )(3 + )(7 + ) − α x 1 x2 − 1 α α x 1 α α − x x2 − 1 α − x x3 − 1 α α α (2 + ) + 2 2 (2 + )(3 + ) 8 (3 + )(7 + ) 24 (7 + ) 8 (1 + )(2 + )(3 + ) 1 α x2 − 1 x3 1 α α x − 1 α x + 2 (3 + ) 6 + 24 (2 + )(3 + ) 8 (3 + ) − 1 2 − 1 − 3 2 − 1 4 − 3 2 − Ln(x) 1 1 x 2 (x 4x + 2) 6 ( x + 9x 18x + 6) 24 (x 16x + 72x 96x + 24) 2 3 2 4 3 2 (x)n 0 x x − x x − 3x + 2x x − 6x + 11x − 6x 2 3 2 4 3 2 φn(x) 1 x x + x x + 3x + x x + 6x + 7x + x

In view of equations (4.10), (4.28), (4.37) and Table 2, we find the first few values of (3)[2] [2] [2] [2] (3)[2] [2] [2] Ln (x), Ln (x), (x)n and φn (x). We present the values of first five Ln (x), Ln (x), (x)n [2] and φn (x) in Table 3.

(α)[2] [2] [2] [2] Table 3. First five Ln (x), Ln (x), (x)n and φn (x).

n 0 1 2 3 4 2 3 4 3 2 (3)[2] x 5x − 11x − 2 − 5x − x 7x 35x − 35x − 175 L n (x) 1 x 4 + 2 5 36 x 2 30 576 + 48 + 16 24 8 2 3 2 4 2 L[2] x x x x − 1 x x x − 13 x 5x − x − 15 n ( ) 1 4 + 2 36 + 2 + 2 6 576 + 8 6 24 [2] 2 3 2 4 3 2 (x)n 0 x x − 2x x − 6x + 7x x − 12x + 40x − 35x [2] 2 3 2 4 3 2 φn (x) 1 x x + 2x x + 6x + 5x x + 12x + 32x + 15x

From Table 2, we get the following graphs:

Graph of L(3)[2](x) Graph of L[2](x) 4 4 20 12

15 10 10 8 5

0 6

−5 4

−10 2 −15 0 −20

−25 −2 −4 −3 −2 −1 0 1 2 3 4 −4 −3 −2 −1 0 1 2 3 4

21 Graph of (x)[2] Graph of φ[2](x) 4 4 2000 1600

1400

1500 1200

1000 1000 800

600 500 400

200 0

0

−500 −200 −4 −3 −2 −1 0 1 2 3 4 −4 −3 −2 −1 0 1 2 3 4

References

[1] M. Abramowitz, I.A. Stegun, Handbook of Mathematical Functions with Formulas, Graphs and Mathematical Tables, reprint of the 1972 edition, Dover Publications, Inc., New York, 1992.

[2] L.C. Andrews, Special Functions for Engineers and Applied Mathematicians, Macmillan Publishing Company, New York, 1985.

[3] P. Appell, Sur une classe de polynˆomes, Ann. Sci. E´cole. Norm. Sup. 9(2) (1880) 119-144.

[4] P. Barry, A. Hennessy, Riordan arrays and the LDU decomposition of symmetric Toeplitz plus Hankel matrices, Linear Algebra Appl. 437(6) (2012) 13801393.

[5] E.T. Bell, Exponential polynomials, Ann. Math. 35 (1934) 258-277.

[6] C.M. Bender, Solution of operator equations of motion (Rigorous Results in Quantum Dynamics) Eds. J. Dittrich and P. Exner (World Scientific, Singapore, 1991) 99-112.

[7] R.P. Boas, R.C. Buck, Polynomial Expansions of Analytic Functions, Springer-Verlag, Berlin, Gottingen, Heidelberg, 1958.

[8] G.-S. Cheon, H. Kim, L.W. Shapiro, Combinatorics of Riordan arrays with identical A and Z sequences, Discrete Math. 312(12-13) (2012) 2040-2049.

[9] F.A. Costabile, F. Dell’Accio, M.I. Gualtieri, A new approach to Bernoulli polynomials, Rendiconti di matematica e delle sue applicazioni 26 (1) (2006) 1-12.

[10] F.A. Costabile, E. Longo, A determinantal approach to Appell polynomials, J. Comput. Appl. Math. 234(5) (2010) 1528-1542.

[11] G. Dattoli, Hermite-Bessel and Laguerre-Bessel functions: a by-product of the monomiality principle, Advanced Special functions and applications, (Melfi, 1999), 147-164, Proc. Melfi Sch. Adv. Top. Math. Phys., 1, Aracne, Rome, 2000.

22 [12] A. Erd´elyi, W. Magnus, F. Oberhettinger, F.G. Tricomi, Higher Transcendental Functions, Vol. II, McGraw-Hill Book Company, New York, Toronto and London, 1953.

[13] A. Erd´elyi, W. Magnus, F. Oberhettinger, F.G. Tricomi, Higher Transcendental Functions, Vol. III, McGraw-Hill Book Company, New York, Toronto and London, 1955.

[14] C. Jordan, Calculus of Finite Differences. Third Edition, Chelsea Publishing Company, Bronx, New York, 1965.

[15] H.W. Gould, T.-X. He, Characterization of (c)-Riordan arrays, Gegenbaur-Humbert-type polynomial sequences and (c)-Bell polynomials, J. Math. Res. Appl. 33(5) (2013) 505-527.

[16] T.-X. He, L.C. Hsu, P.J.-S. Shine, The Sheffer group and the Riordan group, Discrete Appl. Math. 155(15) (2007) 1895-1909.

[17] T.-X. He, A symbolic operator approach to power series transformation-expansion formulas, J. In-teger Seq. 11(2) (2008), Article 08.2.7, 19 pp.

[18] Subuhi Khan, M.W.M. Al-Saad, G. Yasmin, Some properties of Hermite based Sheffer polynomials, Appl. Math. Comput. 217(5) (2010) 2169-2183.

[19] Subuhi Khan, N. Raza, 2-Iterated Appell polynomials and related numbers, Appl. Math. Comput. 219(17) (2013), 9469-9483.

[20] Subuhi Khan, M. Riyasat, Certain results for the 2-iterated q-Appell polynomials. (Sub- mitted for publication in Journal of the Franklin Institute)

[21] M. Lahiri, On a generalisation of Hermite polynomials, Proc. Amer. Math. Soc. 27 (1971) 117-121.

[22] E.D. Rainville, Special functions, Reprint of 1960 First Edition. Chelsea Publishing Com- pany, Bronx, New York, 1971.

[23] L.W. Shapiro, S. Getu, W.J. Woan, L.C. Woodson, The Riordan group, Discrete Appl. Math. 34 (1-3) (1991) 229-239.

[24] R. Sprugnoli, Riordan arrays and combinatorial sums, Discrete Math. 132(1-3) (1994) 267- 290.

[25] R. Sprugnoli, Riordan arrays and Abel-Gould identity, Discrete Math. 142(1-3) (1995) 213-233.

[26] S. Roman, The theory of the , I. J. Math. Anal. 87(1) (1982) 58-115.

[27] S. Roman, The Umbral Calculus, Academic Press, New York, 1984.

[28] S. Roman, G. Rota, The Umbral Calculus, Advances Math. 27 (1978) 95-188.

[29] J.F. Steffensen, The poweriod, an extension of the mathematical notion of power, Acta. Math. 73 (1941) 333-366.

23 [30] G. Szeg¨o, , Amer. Math. Soc., Providence, Rhode Island, 1978.

[31] W. Wang, T. Wang, Generalized Riordan arrays, Discrete Math. 308(24) (2008) 6466-6500.

[32] W. Wang, A determinantal approach to Sheffer sequences, Linear Algebra Appl. 463 (2014) 228-254.

24