검색
Article Search

JMB Journal of Microbiolog and Biotechnology

OPEN ACCESS eISSN 0454-8124
pISSN 1225-6951
QR Code

Article

Kyungpook Mathematical Journal 2020; 60(4): 731-752

Published online December 31, 2020

Copyright © Kyungpook Mathematical Journal.

The p-deformed Generalized Humbert Polynomials and Their Properties

Rajesh V. Savalia*, B. I. Dave

Department of Mathematical Sciences, P. D. Patel Institute of Applied Sciences, Faculty of Applied Sciences, Charotar University of Science and Technology, Changa-388 421, DistAnand, Gujarat, India
Department of Mathematics, Faculty of Science, The Maharaja Sayajirao University of Baroda, Vadodara-390 002, Gujarat, India
e-mail: rajeshsavalia.maths@charusat.ac.in
Department of Mathematics, Faculty of Science, The Maharaja Sayajirao University of Baroda, Vadodara-390 002, Gujarat, India
e-mail: bidavemsu@yahoo.co.in

Received: September 10, 2019; Revised: July 29, 2020; Accepted: August 4, 2020

We introduce the p-deformation of generalized Humbert polynomials. For these polynomials, we derive the differential equation, generating function relations, Fibonacci-type representations, and recurrence relations and state the companion matrix. These properties are illustrated for certain polynomials belonging to p-deformed generalized Humbert polynomials.

Keywords: p-Gamma function, p-Pochhammer symbol, differential equation, generating function relations, mixed relations

In 2007, Díaz and Pariguan introduced the one-parameter deformation of the classical gamma function in the form [7]:

Γp(z)=0 t z1e tp p dt,

where z,(z)>0 and p>0. In fact, the occurrence of the product of the form x(x+p)(x+2p)(x+(n1)p) in the combinatorics of creation and annihilation operators [6, 8] and the perturbative computation of Feynman integrals [5] led them to generalize the gamma function in the above stated form and to generalize the Pochhammer p-symbol in the following form:

(z)n,p=z(z+p)(z+2p)(z+(n1)p),

in which z,p and n. These generalizations lead to the following elementary properties.

Γp(z+p)=zΓp(z),Γp(p)=1,(z)k,p=Γp(z+kp)Γp(z),(z)nk,p=(1)k(z)n,p(pznp)k,p,(z)mn,p=mmnj=1mz+jppmn,p.

For p>0,a and |x|<1p, Diaz et al. [7] demonstrated that

n=0 (a) n,pn!xn=(1px)ap.

This may be regarded as the p-deformed binomial series. The radius of convergence of this series can be enlarged or diminished by choosing a smaller or larger p; unlike in the classical theory of the radius of convergence of the binomial series which is fixed. This motivated us to study the p-deformation of certain Special functions, particularly, the polynomial system formed by generalized Humbert polynomials.

Our objective is to extend generalized Humbert polynomials according to Gould [9] by involving a new parameter: p(>0). We call this extension a p-deformation of the polynomial. We study its properties, namely, the differential equation, generating function relations (GFRs), differential recurrence relations, and mixed relations, and we illustrate the companion matrix.

The companion matrix of the monic polynomial is defined as follows.

Definition 1.1.

Let f(x)[X] be a monic polynomial given by f(x)=δ0+δ1x+δ2x2++δk1xk1+xk. Then the k×k matrix, called the companion matrix of f(x), is denoted and defined as follows [14, p. 39]:

C(f(x))=010000100001δ0δ1δ2δk1.

We have the following lemma [14, Proposition 1.5.14, p.39].

Lemma 1.1.

If f ∈ K[x] is nonconstant and A=C(f(x)), then f(A)=O, the null matrix.

The class {Pn(m,x,γ,s,c);n=0,1,2,} of generalized Humbert polynomials is defined below [9.Eq.5.11, p.707]

Definition 1.2.

For m,n{0}, and x,

Pn(m,x,γ,s,c)= k=0 [n/m]sn+mkksnmkcsn+mkk×γk(mx)nmk, 

where γ, c, and s are generally unrestricted.

This class of polynomials is generated by the following relation:

(cmxt+γtm)s= n=0Pn(m,x,γ,s,c)tn.

The substitution s=-ν, γ=1, and c=1 in (1.2) result in Humbert polynomials, according to Humbert [11]:

Πn,mν(x)= k=0 n/m(mx)nmkΓ(1νn+(m1)k)(nmk)!k!.

Apart from the polynomials (1.4), according to Humbert [12, p. 75], Humbert functions also exits. They have explicit representations in a double infinite series, given by the following:

Ψ1(a;b;c,d;x,y)=r=0s=0 (a) r+s (b)r (c)r (d)s r!s!xrys,Ψ2(a;b,c;x,y)=r=0s=0 (a) r+s (b)r (c)s r!s!xrys,Ξ1(a,b;c;d;x,y)=r=0s=0 (a)r (b)s (c)r (d) r+s r!s!xry

and

Ξ2(a,b;c;x,y)= r=0s=0 (a)r (b)r (c) r+sr!s!xrys,

where |x|<1,|y|< and c,d0,1,2,. In recent years, these functions have been increasingly used in various fields, for example, in theoretical physics [3, 13] and communication theory [1, 2, 18]. Moreover, for specific values of parameters and variables, their reduced forms have also been found useful, especially in connection with simplification algorithms in computer algebra systems (see [3]). Therefore, we propose the following extension of the polynomials Pn(m,x,γ,s,c).

Definition 1.3.

For γ,s,c,m,x,n{0}, and p>0,

Pn,p(m,x,γ,s,c)= k=0 n/mγkcsn+mkk(s+p)mkkn,p(nmk)!k!(mx)nmk,

in which the floor function r=floorr, represents the greatest integer ≤ r.

We call these polynomials p-deformed generalized Humbert polynomials or pGHPs. When p=1, it coincides with the polynomial (1.2). The particular polynomials belonging to these general p-polynomials provide an extension to the Humbert polynomials (1.4), Kinney polynomials, Pincherle polynomials, Gegenbauer polynomials, and Legendre polynomials (see [9]). For instance, the substitutions γ=1, c=1, and s=-ν in (1.5) yield p-deformed Humbert polynomials:

Πn,m,pν(x)= k=0 n/m(mx)nmkΓp(pνnp+(m1)kp)(nmk)!k!.

For p=1, this coincides with (1.4). If we substitute ν=1/m,m, in (1.6), then we obtain the p-deformed Kinney polynomial:

Pn,p(m,x)= k=0 n/m(mx)nmkΓp(p1/mnp+(m1)kp)(nmk)!k!.

For m=3 and ν=1/2, (1.6) reduces to the p-deformed Pincherle polynomial:

Pn,p(x)= k=0 n/3(3x)n3kΓp(p1/2np+2kp)(n3k)!k!.

The p-deformed Gegenbauer polynomial is the special cases where m=2 of (1.6) which occurs in the following form:

Cn,pν(x)= k=0 n/2(2x)n2kΓp(pνnp+kp)(n2k)!k!.

Further, if ν=1/2 then (1.7) is reduced to the p-deformed Legendre polynomial given by the following:

Pn,p(x)= k=0 n/2(2x)n2kΓp(p1/2np+kp)(n2k)!k!.

All these polynomials reduce to their classical forms when p=1 [9, p.697].

In this section, we derive the differential equation of the polynomial (1.5). Costa et al. [4] demonstrated that the homogeneous differential equation:

(1xN)y(N)+ k=1 NANkxNky(Nk)=0,

has a polynomial solution if and only if 0r<N,r, and n0 such that n mod N =r, where n is a root of the recurrence relation, and y(j) is the jth derivative of y with respect to x for j=1,2,,N.

Let the sequence {fr}r=0n be given by fr=f(r), where

f(r)=(nr)s+rp+nrmpm1,p.

We use the forward difference operator 'Δ' and the shift operator 'E' which are defined as follows [10, Eq. (5.2.13), p. 178]:

Δft=ft+1ft,Ekft=ft+k.

The relation between Δ and E is given by [10, Eq. (5.2.14), p. 178] Δ=E1, where 1 is the identity operator defined by 1f=f. In (1.5), we use the following formula:

(p+s)n+mkk,p=(p+s)(nmk+k),p=(1)nmk+k(pps)nmk+k,p,

to obtain the alternative form:

Pn,p(m,x,γ,s,c)= k=0 n/m(1)kγkcsn+mkk(s)nmk+k,p(nmk)!k!(mx)nmk.

The differential equation for this explicit form is derived in Theorem 2.1.

Theorem 2.1.Let s,p>0, and m. Then, the polynomial y=Pn,p(m,x,γ,s,c) satisfies the following equation:

γcm1y(m)+ r=0marxry(r)=0,

where ar=mm1r!Δrf0.

Proof. Let n=ml+q, where n/m=l and 0qm1. The rth derivative of (2.1) is given by the following:

DrPn,p(m,x,γ,s,c)= k=0 (nr)/m(1)kγkcsn+mkk(s)nmk+k,p(nmkr)!k!mnmk×xnmkr. 

Hence,

xrDrPn,p(m,x,γ,s,c)= k=0 (nr)/m(1)kγkcsn+mkk(s)nmk+k,p(nmkr)!k!×(mx)nmk. 

Next, we have

DmPn,p(m,x,γ,s,c)= k=0 (nm)/m(1)kγkcsn+mkk(s)nmk+k,p(nmkm)!k!×mnmkxnmkm = k=0 l1(1)kγkcsn+mkk(s)nmk+k,p(nmkm)!k!mm ×(mx)nmkm, 

where

nrm=l,ifrql1,ifr>q.

Substituting the equation (2.3) and (2.4) on the left-hand side of the differential equation (2.2) and comparing the corresponding coefficients of x, we find that

r=0mnmkrr!ar = m mk(s)nmk+k+m1,p(s)nmk+k,p  =mmk(s+npmkp+kp)m1,p, 

where k=0,1,2,,l1, and

r=0qnmlrr!ar =mml(s+npmlp+lp)m1,p. 

Because n=ml+qnml=q, we have the following:

r=0qqrr!ar =mm1(nq)s+qp+ nqmpm1,p.

By substituting ar=mm1Δrf0r! in (2.6), we obtain

r=0qqrΔrf0 =(nq)s+qp+ nqmpm1,p,

that is,

(1+Δ)qf0=(nq)s+qp+nqmpm1,p.

Hence, 1+Δ=E, the shift operator, in (2.7) becomes the following:

Eqf0=f(q)=(nq)s+qp+nqmpm1,p.

For k=0,1,2,,l1, (2.5) can be written in the following form:

r=0mnmkrΔrf0=fnmk=mk(s+npmkp+kp)m1,p.

As tf(t) is a polynomial of degree m, the equation (2.8) is a forward difference formula for f at the point t=n-mk. Thus, the proof is completed for the choice:

ar=mm1Δrf0r!=mm1r!Δrnnpmsm m1,p.

2.1. Particular Cases

We illustrate special instances of the differential equation (2.2). In particular, the equations of the Pincherle, Gegenbauer, and Legendre polynomials. We choose r=0, 1, 2 and 3 to obtain the following:

a0=mm1Δ0f0=mm1nnpmsmm1,p, a1=mm1Δf0=mm1(E1)f(0)=mm1(f(1)f(0))=mm1(n1)(n1)p+m(s+p)mm1,pnnpmsmm1,p, a2=mm1Δ2f02!=mm12!(E1)2f0=mm12!(E22E+1)f(0)=mm12!(f(2)2f(1)+f(0))=mm12!(n2)(n2)p+m(s+2p)mm1,p2(n1)×(n1)p+m(s+p)mm1,p+nnpmsmm1,p,

and

a3=mm1Δ3f03!=mm13!f(3)3f(2)+3f(1)f(0)=mm13!(n3)(n3)p+m(s+3p)m m1,p3(n2)× (n2)p+m(s+2p) m m1,p+3(n1)×(n1)p+m(s+p)m m1,pnnpmsm m1,p.

By choosing m=3 in (2.9), (2.10), (2.11) and (2.12), we obtain the following:

a0=32nnp3s32,p=n(np3s)(np3(sp)), a1=32(n1)(n1)p+3(s+p)32,p(n)np3s32,p=3np(np2s+p)(3s2p)(3s5p), a2=12ps18p2, a3=4p2.

Further, after entering m=3, γ=1, c=1, and s=-λ into the differential equation (2.2), from the particular values (2.13), (2.14), (2.15) and (2.16), we arrive at the differential equation of the p-deformed Pincherle polynomial. From the following general form:

y(3)+ r=03arxry(r)=0,

that is,

(1+a3x3)y(3)+a0y+a1xy(1)+a2x2y(2)=0,

we obtain the following equation:

14p2x3y(3)62pλ+3p2x2y(2)+3np(np+2λ+p)(3λ+2p)(3λ+5p)xy(1)+n(np+3λ)(np+3(λ+p))y=0.

The choice p=1 yields the differential equation of the Pincherle polynomial according to Humbert [11,p. 23].

Next, to obtain the equation for the p-deformed Gegenbauer polynomial, we enter m=2, γ=1, c=1, and s=-ν into (2.2) to obtain

y(2)+ r=02arxry(r)=0,

or equivalently,

(1+a2x2)y+a1xy+a0y=0.

From (2.9), (2.10) and (2.11), we have the following:

a0=nnp2s,a1=2(n1)(n1)p+2(s+p)21,p(n)np2s21,p=2(n1)(n1)p+2(s+p)2(n)np2s2=2sp,a2=p.

With a0, a1, and a2, the above equation takes the precise form:

(1px2)y+n(np+2ν)y(2ν+p)xy=0,

where y=Cn,pν(x) is given by (1.7). When p=1, this reduces to the differential equation of the Gegenbauer polynomial [17, Eq.(1.4), p.279]. The well-known special case ν=1/2 of this equation is the following differential equation:

(1px2)Pn,p(x)(1+p)xPn,p(x)+n(np+1)Pn,p(x)=0

of the p-deformed Legendre polynomial (1.8). In addition, for p=1, this reduces to the differential equation of the Legendre polynomial Pn(x) (cf. [17, Eq.(5), p.161]).

We derive GFR of the pGHP (1.5). This is accomplished with the help of the p-deformed version of the identity [16, Ex. 212 and 216, p. 146]:

(1+z)a+11zb= n=0a+bn+nnwn,

where a,b and w=z(1+z)b1. It is given as explained in Theorem 3.1.

Theorem 3.1. For p>0, a,b, and w=z(1+z)b+1.

(1+z)a/p+11zb= n=0Γp(a+bnp+np+p)Γp(a+bnp+p)n!wnpn.

Proof. We use the Lagrange's series [19, Eq. (3), p. 354]:

f(z)1wg(z)= n=0wnn!Dznf(z)(g(z))nz=z0,D=ddz,

where w=zz0g(z).

To derive (3.1), we take z0=0,f(z)=(1+z)a/p,g(z)=(1+z)b+1, and w=z(1+z)b1 in the left-hand side of Lagrange's series gives the following:

f(z)1wg(z)=(1+z)a/p1w(b+1)(1+z)b=(1+z)a/p1z(1+z)(b+1)=(1+z)a/p+11zb.

The same substitution on the right-hand side of the Lagrange's series gives the following:

n=0wnn!Dznf(z)(g(z))nz=z0=n=0wnn!Dzn(1+z)ap+bn+nz=0=n=0wnn!ap+bn+nap+bn+n1×ap+bn+nn+1=n=0wn(1)npnn!(abnpnp)n,p=n=0wnpnn!(a+bnp+p)n,p=n=0Γp(a+bnp+np+p)Γp(a+bnp+p)n!wnpn.

This completes the proof.

We define the function as follows:

R(An,α,γ,r,m,p)= k=0 n/mΓp (α+mrk+p)Γp (α+mrkkp+p)k!γkpkAnmk,

which is required in deriving the following GFR.

For m,w=t(1+γwm)β/p,p>0,G(z)= n=0Anzn where A00,

n=0R(An,α+βn,γ,r,m,p)tn=(1+γwm)(pα)/p1+ βmp+1γwmGw (1+γwm ) r/p,

where {An} is an arbitrary sequence such that i=0|Ai|< and other parameters are generally unrestricted.

Proof. We begin with the following:

n=0R(An,α+βn,γ,r,m,p)tn=n=0k=0n/m Γp (pαβnnr+mrk) Γp (pαβnnr+mrkkp)k!γkpkAnmktn=n=0k=0 Γp (pαβnβmknr) Γp (pαβnβmknrkp)k!γkpkAntn+mk=n=0Antnk=0 Γp(pαβnβmknr) Γp(pαβnβmknrkp)k!γkpktmk.

In view of the sum in (3.1), the inner series simplifies to the following:

n=0R(An,α+βn,γ,r,m,p)tn=n=0Antn (1+v) (αβnnr)/p+1 1+ βmp+1v= (1+v) (pα)/p 1+ βmp+1vn=0An t (1+v) β/p n (1+v) nr/p,

where v=γtm(1+v)βm/p. If we replace v with γ wm, then w=t(1+γwm)β/p and, consequently, we obtain

n=0R(An,α+βn,γ,r,m,p)tn= (1+γwm) (pα)/p1+ βmp+1γwmn=0An t (1+γwm ) β/p n (1+γwm) nr/p= (1+γwm) (pα)/p1+ βmp+1γwmn=0An w (1+γ w m )r/p n.

This completes the proof of GFR (3.3).

The replacement of γ with γ p/c, and the substitutions α=s,r=p, and

An=(m)ncsnΓp(p+s)Γp(p+snp)n!xn

in (3.3), yield the GFR of the pGHP in the following form:

n=0Pn,p(m,x,γ,s+βn,c)tn=(1+γpwm/c)(p+s)/p1+γpwmc βmp+1Gw 1+ γpwm c ,

where w=t1+γpwmcβ/p and

G(u)= n=0(c)snΓp(p+s)Γp(p+snp)n!(mxu)n.

We note that β=0w=t; and hence, using the p-binomial series (1.1), we find the following:

n=0Pn,p(m,x,γ,s,c)tn=(1+γptm/c)s/pGt1+γptmc=c(11/p)s(c+γptm)s/pn=0(m)nΓp(p+s)Γp(p+snp)n!×xtc+γptmn=c(11/p)s(c+γptm)s/pn=0(m)n(p+s)n,pn!×xtc+γptmn=c(11/p)s(c+γptm)s/pn=0(s)n,pn!×mxtc+γptmn=c(11/p)s(c+γptm)s/p1pmxtc+γptms/p=c(11/p)sc+γptmpmxts/p.

This generalizes the GFR (1.3).

The GFR of the p-deformed Humbert polynomials occurs as a special case of (3.4) with the substitutions γ=1,c=1, and s=μ given by

n=0Πn,m,pμ+βn(x)tn=(1+pwm)(pμ)/p1+βm+pwmGw 1+pwm ,

where w=t1+pwmβ/p and

G(z)= n=0Γp(pμ)Γp(pμnp)n!(mxz)n.

The case β =0 yields the following GFR:

n=0Πn,m,pμ(x)tn=1+ptm pmxtμ/p.

This extends the GFR according to Humbert [11, p.24] (also see [15, Eq.(1.15), p.5] with p=1).

The GFR of the p-deformed Kinney polynomial occurs with γ=1,c=1, and s=-1/m from (3.4) given by

n=0Pn,p(m,βn,x)tn=(1+pwm)(p1/m)/p1+βm+pwmGw 1+pwm ,

where w=t1+pwmβ/p and

G(z)= n=0Γp(p1m)Γp(p1mnp)n!(mxz)n.

The GFR of the p-deformed Pincherle polynomial is obtained by substituting m=3, γ=1, c=1, and s=-λ in (3.4), and it is given by the following:

n=0Pn,pλ+βn(x)tn=(1+pw3)(pλ)/p1+ β3p+1pw3Gw 1+pw3 ,

where w=t1+pw3β/p and

G(z)= n=0Γp(pλ)Γp(pλnp)n!(3xz)n.

Similarly, taking m=2, γ=1, c=1, and s= -ν in (3.4), we obtain

n=0Cn,pν+βn(x)tn=(1+pw2)(pν)/p1+ β2p+1pw2Gw 1+pw2 ,

where w=t1+pw2β/p and

G(z)= n=0Γp(pν)Γp(pνnp)n!(2xz)n,

which is the GFR of the p-deformed Gegenbauer polynomial. Furthermore, entering ν=1/2 into (3.7), we obtain the GFR of the p-deformed Legendre polynomial or briefly pLP given by

n=0Pn,p(x)tnn=(1+pw2)(2p1)/2p1+ β2p+1pw2Gw 1+pw2 ,

with w=t1+pw2β/p and

G(z)= n=0Γp(p12)Γp(p12np)n!(2xz)n.

3.1. Fibonacci-type Polynomials

We provide a computation formula of Fibonacci-type polynomials of order n (cf. [15, Theorem 2.2, p.6]) in Theorem using (3.5).

Theorem 3.3. For the pGHP defined by (1.5),

Pn,p(m,x,γ,s1+s2,c)= k=0nPnk,p(m,x,γ,s1,c)Pk,p(m,x,γ,s2,c).

Proof. Replacing s with s1+s2 in GFR (3.5) provides the following result:

n=0Pn,p(m,x,γ,s1+s2,c)tn=(c)(11/p)(s1+s2)c+γptmpmxts1+s2/p=(c)(11/p)s1c+γptmpmxts1/p×(c)(11/p)s2c+γptmpmxts2/p=n=0Pn,p(m,x,γ,s1,c)tnk=0Pk,p(m,x,γ,s2,c)tk=n=0k=0Pn,p(m,x,γ,s1,c)Pk,p(m,x,γ,s2,c)tn+k=n=0k=0nPnk,p(m,x,γ,s1,c)Pk,p(m,x,γ,s2,c)tn.

Comparing the coefficients of tn, this yields (3.8).

The GFR (3.6) leads to the computation formula of Fibonacci-type polynomials of order n, stated as Corollary 3.4.

Corollary 3.4. In the usual notations and meaning,

Πn,m,pμ1+μ2(x)= k=0nΠnk,m,pμ1(x)Πk,m,pμ2(x).

In this section, the differential recurrence relations and mixed relations of the pGHP are derived.

First, we denote c+γptmpmxts/p as A(t;m,x,γ,s,c,p) and rewrite (3.5) in the following form:

A(t;m,x,γ,s,c,p)=c(1/p1)s n=0Pn,p(m,x,γ,s,c)tn.

Then, with Dx=d/dx, we have

Dx(A(t;m,x,γ,s,c,p))=Dxc+γptmpmxts/p=mts c+γptm pmxt s/p1.

Setting sq=p,q, this results in the following:

DxA(t;m,x,γ,p/q,c,p)=mtpqc+γptmpmxts/p+sq/p=mtpqA(t;m,x,γ,p/q,c,p)1+q.

The successive differentiation yields the following:

Dx2(A(t;m,x,γ,p/q,c,p))=DxmtpqA(t;m,x,γ,p/q,c,p)1+q=mtpqDxA(t;m,x,γ,p/q,c,p)1+q=(mtp)(1+q)q A(t;m,x,γ,p/q,c,p) q×DxA(t;m,x,γ,p/q,c,p)=(mtp)2(1+q)q2× A(t;m,x,γ,p/q,c,p) 1+2q, Dx3(A(t;m,x,γ,p/q,c,p))=mtpq3(1+q)(1+2q)A(t;m,x,γ,p/q,c,p)1+3q, 

and in general,

Dxj(A(t;m,x,γ,p/q,c,p))=mtpqj{ i=0 j1(1+iq)}A(t;m,x,γ,p/q,c,p)1+jq. 

Next, taking the jth derivative with respect to x in (4.1) yields tche following:

DxjA(t;m,x,γ,s,c,p)=c(1/p1)sn=0tnDxjPn,p(m,x,γ,s,c)=c(1/p1)sn=0tnk=0n/m(γ)kcsn+mkk(s)nmk+k,pk!(nmk)!mnmkDxjxnmk=c(1/p1)sn=jtnk=0(nj)/m(γ)kcsn+mkk(s)nmk+k,pk!(nmkj)!mnmkxnmkj=c(1/p1)sn=0tn+jk=0n/m(γ)kcsnj+mkk(s)n+jmk+k,pk!(nmk)!mj(mx)nmk.

However, because,

DxjPn+j,p(m,x,γ,s,c)=k=0(n+j)/m(1)kγkcsnj+mkk(s)n+jmk+k,pk!(n+jmk)!mn+jmkDxj(x)n+jmk=k=0n/m(1)kγkcsnj+mkk(s)n+jmk+k,pk!(nmk)!mn+jmkxnmk,

from (4.3), we have the following:

DxjA(t;m,x,γ,s,c,p)=c(1/p1)s n=0tn+jDxjPn+j,p(m,x,γ,s,c).

Inserting s=-p/q into (4.4) and employing (4.3), we obtain

c(1/p1)s n=0tnDxjPn+j,p(m,x,γ,p/q,c)= mpq j{ i=0 j1(1+iq)} A(t;m,x,γ,p/q,c,p)1+jq.

Replacing A(t;m,x,γ,p/q,c,p) with its series expansion from (4.1), it becomes

(1/p1)s n=0tnDxjPn+j,p(m,x,γ,p/q,c)= mpq j{ i=0 j1(1+iq)} c (1/p1)s n=0 P n,p (m,x,γ,p/q,c) t n 1+jq,

that is,

n=0tnDxjPn+j,p(m,x,γ,p/q,c)=mpq jc(1/p1)sjq{i=0j1(1+iq)} n=0 Pn,p (m,x,γ,p/q,c) t n 1+jq=mpq jc(1/p1)sjq{i=0j1(1+iq)}n=0 i1 +i2 ++i 1+jq =nP i 1,pPi2,pPi 1+jq,ptn,

where q and Pir,p=Pir,p(m,x,γ,p/q,c). Comparing the coefficients of tn. this yields the following:

DxjPn+j,p(m,x,γ,p/q,c)= mpqjc(1/p1)sjq{ i=0 j1(1+iq)} i1 +i2 ++i 1+jq =nP i 1,pPi2,pPi1+jq,p.

This provides the p-deformed version of the result according to Gould [9, Eq.(3.4), p.702](cf. with p=1). Further, multiplying (4.3) by tn and then taking the summation from n=0 to ∞, produces the following:

n=0DxjPn+j,p(m,x,γ,s,c)tn=n=0k=0n/m(γ)kcsnj+mkk(s)n+jmk+k,pk!(nmk)!×mn+jmkxnmktn=n=0cjpj(s)j,pmjk=0n/m(γ)kcsjpn+mkk×(s+jp)nmk+k,pk!(nmk)!(mx)nmktn=n=0c(p1)j(s)j,pmjPn,p(m,x,γ,sjp,c)tn=n=0c(p1)j(s)j,pmjPn,p(m,x,γ,sjp,c)tn.

From this, it follows that

DxjPn+j,p(m,x,γ,s,c)=c(p1)j(s)j,pmjPn,p(m,x,γ,sjp,c).

This generalizes the formula given by Gould [9, Eq.(3.5), p.702](cf. with p=1). In addition, setting j=1, m=2, γ=1, c=1, and s=-ν and replacing n with n-1 in (4.5) yields

DxCn,pν(x)=2νCn1,pν(x),

involving the p-Gegenbauer polynomial. This generalizes the familiar formula of Whittaker et al. [20, (III), p. 330](cf. with p=1). For ν=1/2, this further reduces to the following:

DxPn,p(x)=Pn1,p(x),

involving the p-Legendre polynomial.

For the recurrence relations, we first obtain the following:

c+γptmpmxttDtA(t;m,x,γ,s,c,p)=c+γptmpmxttDt c+γp t m pmxts/p=c+γptmpmxttsp c+γp t m pmxts/p1(γpmtm1pmx)=(ms)(xtγtm)A(t;m,x,γ,s,c,p).

From (4.1), we have the following:

c+γptmpmxttDtc(1/p1)s n=0Pn,p(m,x,γ,s,c)tn=(ms)(xtγtm)c(1/p1)s n=0Pn,p(m,x,γ,s,c)tn.

Simplifying this and abbreviating Pn,p(m,x,γ,s,c) by Pn,p(x), we obtain

c+γptmpmxt n=0nPPn,p(x)tn=ms(xtγtm) n=0Pn,p(x)tn n=0cnPn,p(x)tn+γptm n=0nPn,p(x)tnpmxt n=0nPn,p(x)tn=msxt n=0Pn,p(x)tn+msγtm n=0Pn,p(x)tn n=0cnPn,p(x)tn+γp n=0nPn,p(x)tn+mpmx n=0nPn,p(x)tn+1=msx n=0Pn,p(x)tn+1+msγ n=0Pn,p(x)tn+m n=0cnPn,p(x)tn+γp n=m(nm)Pnm,p(x)tnpmx n=1(n1)Pn1,p(x)tn=msx n=0Pn1,p(x)tn+msγ n=0Pnm,p(x)tn n=m(cnPn,p(x)+mx(snp+p)Pn1,p(x)+γ(npmpms)Pnm,p(x))tn=0,

because nm1. The recurrence relation is given as follows:

cnPn,p(x)+mx(snp+p)Pn1,p(x)+γ(npmpms)Pnm,p(x)=0.

This identity provides the p-deformed version of a recurrence relation derived by Gould [9, Eq.(2.3), p.700](cf. with p=1).

Taking derivative of (3.5) with respect to x, we obtain

n=0DxPn,p(m,x,γ,s,c)tn=smt(c)(11/p)sc+γptm pmxts/p1,

and differentiating (3.5) with respect to t and using (4.7), we find

n=0P n,p(m,x,γ,s,c)ntn1=spc(11/p)s(γpmtm1pmx)c+γp tm pmxts/p1=sc(11/p)s(γmt m1mx)smt (c) (11/p)sn=0DxPn,p(m,x,γ,s,c)tn=(xγt m1)tn=0DxPn,p(m,x,γ,s,c)tn.

Thus, we obtain

n=0nPn,p(m,x,γ,s,c)tn=xn=0DxPn,p(m,x,γ,s,c)tnγtm1n=0DxPn,p(m,x,γ,s,c)tn=xn=0DxPn,p(m,x,γ,s,c)tnγn=0DxPn,p(m,x,γ,s,c)tn+m1=xn=0DxPn,p(m,x,γ,s,c)tnγn=m1DxPnm+1,p(m,x,γ,s,c)tn.

After equating the coefficients of tn on both sides for nm1, we obtain

nPn,p(m,x,γ,s,c)=xDxPn,p(m,x,γ,s,c)γDxPnm+1,p(m,x,γ,s,c). 

This provides a p-deformation of the differential recurrence relation according to Gould [9, Eq.(2.5), p.700]. The other recurrence relations involving the pGHP may be obtained similarly.

Let P˜n,p(m,x,γ,s,c) be the monic polynomial obtained from (2.1) and it is defined by the following:

P˜n,p(m,x,γ,s,c)= k=0Nδkxnmk,

where

δk=(1)k+mk(n)mkγkcmkk(s+np)mk+k,pmmkk!.

Then, with this δk,p,CP˜n,p(m,x,γ,s,c) assumes the form stated in Definition 1.1. The eigen values of this matrix are thus precisely the zeros of P˜n,p(m,x,γ,s,c) (see [14, p. 39]).

Iillustation 5.1. To illustrate the companion matrix of the above monic polynomial, we take n=3, p=2, and m=2 into (5.1) to obtain

P˜3,2(2,x,γ,s,c)= k=01δkx32k=δ0x3+δ1x=δ0x3+0x2+δ1x+0,

where δ0=1 and δ1=32γcs4.

Thus, the companion matrix is as follows:

CP˜3,2(2,x,γ,s,c)=0100010δ10

where the eigen values are determined from the following determinant equation:

λ100λ10δ1λ=0.

From this, we obtain the eigen values λ=0,δ1, and δ1, that satisfy the equation P˜3,2(2,x,γ,s,c)=0.

Figure 1. .

The authors express their sincere thanks to the reviewers and the editor-in-chief for their valuable suggestions for the improvement of the manuscript.

  1. Yu. A. Brychkov, On some properties of the Nuttall function Qμ, ν(a,b), Integral Transforms Spec. Funct., 25(1)(2014), 33-43.
    CrossRef
  2. Yu. A. Brychkov and N. V. Savischenko, A special function of communication theory, Integral Transforms Spec. Funct., 26(6)(2015), 470-484.
    CrossRef
  3. V. V. Bytev, M. Yu. Kalmykov and S.-O. Moch, HYPERDIRE: HYPERgeometric functions DIfferential REduction: MATHEMATICA based packages for differential reduction of generalized hypergeometric functions: FD and FS Horn-type hypergeo-metric functions of three variables, Comput Phys. Commun., 185(2014), 3041-3058.
    CrossRef
  4. G. B. Costa and L. E. Levine, Nth-order differential equations with finite polynomial solutions, Int. J. Math. Educ. Sci. Tech., 29(6)(1998), 911-914.
  5. P. Deligne, P. Etingof, D. S. Freed, L. C. Jeffrey, D. Kazhdan, J. W. Morgan, D. R. Morrison and E. Witten, Quantum fields and strings: a course for mathematicians, American Mathematical Society, 1999.
  6. R. Diaz and E. Pariguan, Quantum symmetric functions, Comm. Algebra, 33(6)(2005), 1947-1978.
    CrossRef
  7. R. Diaz and E. Pariguan, On hypergeometric functions and Pochhammer k-symbol, Divulg. Mat., 15(2)(2007), 179-192.
  8. R. Diaz and C. Teruel, q; k-generalized gamma and beta function, J. Nonlinear Math. Phys., 12(1)(2005), 118-134.
    CrossRef
  9. H. W. Gould, Inverse series relations and other expansions involving Humbert polynomials, Duke Math. J., 32(4)(1965), 697-711.
    CrossRef
  10. F. B. Hildebrand, Introduction to numerical analysis, Second edition, Dover Publications INC., New York, 1987.
  11. P. Humbert, Some extensions of Pincherle's polynomials, Proc. Edinburgh Math. Soc., 39(1)(1920), 21-24.
    CrossRef
  12. P. Humbert, The conuent hypergeometric functions of two variables, Proc. R. Soc. Edinb., 41(1922), 73-96.
    CrossRef
  13. B. A. Kniehl and O. V. Tarasov, Finding new relationships between hypergeometric functions by evaluating Feynman integrals, Nuclear Phys. B, 854(3)(2012), 841-852.
    CrossRef
  14. M. Mignotte and D. Stefanescu, Polynomials: an algorithmic approach, Springer- Verlag, New York, 1999.
  15. G. Ozdemir, Y. Simsek, and G. V. Milovanović, Generating functions for special poly- nomials and numbers including Apostol-type and Humbert-type polynomials, Mediterr. J. Math., 14(3)(2017), Paper No. 117, 17 pp.
    CrossRef
  16. G. PÓlya and G. Szegő, Problems and theorems in analysis I, Springer-Verlag, New York-Heidelberg and Berlin, 1972.
  17. E. D. Rainville, Special functions, The Macmillan Company, New York, 1960.
  18. P. C. Sofotasios , T. A. Tsiftsis, Yu. A. Brychkov, et al., Analytic expressions and bounds for special functions and applications in communication theory, IEEE Trans. Inform. Theory, 60(12)(2014), 7798-7823.
    CrossRef
  19. H. M. Srivastava and H. L. Manocha, A treatise on generating functions, Ellis Hor- wood Limited, John Willey and Sons, 1984.
  20. E. T. Whittaker and G. N. Watson, A course of modern analysis, 4th ed., Cambridge University Press, Cambridge, 1927.