This content was uploaded by our users and we assume good faith they have the permission to share this book. If you own the copyright to this book and it is wrongfully on our website, we offer a simple DMCA procedure to remove your content from our site. Start by pressing the button below!
1.
(44)
(45)
(46)
Rl,λ is the convergence radius of the series Nl,λ (x). Remark that the q-deformed coherent states introduced in [94] are recovered as a particular case corresponding to l = 1 and λ = 0. We now aim at showing that the coherent states (44) satisfy the Klauder’s criteria [72, 71]. To this end let us first prove the following lemma: Lemma 2.7 If q > 1, then 1 Nl,λ (x) = , −1 Nl,λ(q x) 1 − (q − 1)x/(l2 q λ ) 1 Nl,λ(x) = , ((q − 1)x/(l2 q λ ); q −1 )∞ Z Rl,λ −1 l,λ xn Nl,λ(q −1 x) dq x = (l2 q λ )n q −n(n+1)/2 [n]q !.
(47) (48) (49)
0
Proof: We use the (q; l, λ)-derivative defined by ∂ql,λ f (x) = l2 q λ
f (x) − f (q −1 x) (q − 1)x
(50)
to obtain Nl,λ(x) = ∂ql,λ Nl,λ (x) = l2 q λ
Nl,λ (x) − Nl,λ(q −1 x) (q − 1)x
which leads to (47) and Nl,λ (q −n x) , n = 1, 2, ... −k x/(l 2 q λ )) k=0 (1 − (q − 1)q
Nl,λ (x) = Qn−1
(51)
Letting n to +∞ and taking into account the fact that Nl,λ (0) = 1 lead to (48). Next, we use the (q; l, λ)-integration given by Z
0
∞
a
f (x)dl,λ q x =
q − 1 X −k q f (aq −k ) a l2 q λ k=0
to get Z
Rl,λ
n
x 0
−1
Nl,λ (q x)
−1
dl,λ q x
∞ X
(l2 q λ )n −(k+1) −1 (q ; q )∞ (q − 1)n k=0 ∞ (l2 q λ )n −1 −1 X q −(n+1)k (q ; q )∞ = (q − 1)n (q −1 ; q −1)k k=0 =
q −(n+1)k
14
(52)
(l2 q λ )n (q −1 ; q −1 )∞ (q − 1)n (q −(n+1) ; q −1 )∞ (l2 q λ )n −1 −1 (q ; q )n = (l2 q λ )n q −n(n+1)/2 [n]q !. = (q − 1)n
=
✷
Proposition 2.8 The coherent states defined in (44) (i) are normalized eigenvectors of the operator a with eigenvalue z, i.e. a|zil,λ = z|zil,λ ,
l,λ hz|zil,λ
= 1;
(53)
(ii) are not orthogonal to each other, i.e. l,λ hz1 |z2 il,λ
6= 0, when z1 6= z2 ;
(54)
(iii) are continuous in their labels z; (iv) resolve the identity, i.e. 1=
Z
Dl,λ
dµl,λ(¯ z , z)|zil,λl,λ hz|,
(55)
where dµl,λ(¯ z , z) =
1−q Nl,λ(¯ z z) d2 z , if 0 < q < 1, l2 q λ ln q −1 Nl,λ(¯ z z/q) π
(56)
and dl,λ 1 q x dθ dµ(¯ z , z) = , 2π 1 − (q − 1)x/(l2 q λ ) with 0 < x <
l2 q λ q−1
x = |z|2 , θ = arg(z),
(57)
and 0 ≤ θ ≤ 2π for q > 1.
Proof: •Non orthogonality and normalizability l,λ hz1 |z2 il,λ
=
Nl,λ (¯ z1 z2 )
(Nl,λ (|z1 |2 )Nl,λ (|z2 |2 ))1/2
imply that the coherent states are not orthogonal. •Normalizability From the above relation taking z1 = z2 = z we obtain a|zil,λ =
−1/2 Nl,λ (|z|2 )
6= 0
l,λ hz|zil,λ
∞ X q n(n+1)/4 z n p a|ni (l2 q λ )n [n]q ! n=0
15
(58)
= 1. Also,
∞ X
q n(n−1)/4 z n |n − 1i (l2 q λ )n−1 [n − 1]q ! n=1 ∞ X q n(n+1)/4 z n −1/2 2 p |ni. = zNl,λ (|z| ) (l2 q λ )n [n]q ! n=0 =
−1/2 Nl,λ (|z|2 )
p
•Continuity in the labels z
|||z1 il,λ − |z2 il,λ||2 = 2 (1 − Rel,λ hz1 |z2 il,λ ) . So, |||z1il,λ − |z2 il,λ||2 → 0 as |z1 − z2 | → 0, since l,λ hz1 |z2 il,λ → 1 as |z1 − z2 | → 0. •Resolution of the identity The computation of the RHS of (55) gives Z Z X q [n(n+1)+m(m+1)]/4 dµl,λ(¯ z , z) |nihm| p dµl,λ(¯ z , z)|zil,λl,λ hz| = . z¯n z m Nl,λ (|z|2 ) (l2 q λ )n+m [n]q ![m]q ! Dl,λ Dl,λ n,m
So, in order to satisfy (55) it is required Z dµl,λ(¯ z , z) = δmn (l2 q λ )n q −n(n+1)/2 [n]q !, n, m = 0, 1, 2, ... (59) z¯n z m 2) N (|z| l,λ Dl,λ √ Upon passing to polar coordinates, z = x eiθ , dµl,λ (¯ z , z) = dωl,λ (x)dθ where 0 ≤ θ ≤ 2π, 0 < x < Rl,λ and ωl,λ is a positive valued function, this is equivalent to the classical Stieltjes power moment problem when 0 < q < 1 or the Hausdorff power moment problem when q > 1 [2, 98]: Z Rl,λ 2π dωl,λ(x) = (l2 q λ )n q −n(n+1)/2 [n]q !, n = 0, 1, 2, ... (60) xn N (x) l,λ 0 If 0 < q < 1, then we have the following Stieltjes power moment problem: Z +∞ 2π dωl,λ (x) = (l2 q λ )n q −n(n+1)/2 [n]q !, (61) xn Nl,λ (x) 0 or, equivalently, Z
0
+∞
yn
2π dωl,λ(l2 q λ y) = q −n(n+1)/2 [n]q !, Eq ((1 − q)qy)
(62)
where the change of variable y = l2xqλ has been made. Atakishiyev and Atakishiyeva [8] have proved that Z +∞ ln q −1 −n(n−1)/2 y n−1dy = q [n − 1]!q . (63) gq (n) = Eq ((1 − q)y) 1−q 0 16
Therefore we deduce dωl,λ(l2 q λ y) = or
1 1 − q Eq ((1 − q)qy)dy 2π ln q −1 Eq ((1 − q)y)
1 − q Eq ((1 − q)qx/(l2 q λ ))dx 1 dωl,λ(x) = 2π l2 q λ ln q −1 Eq ((1 − q)x/(l2 q λ )) 1 1 − q Nl,λ (x)dx = . 2 2π l q λ ln q −1 Nl,λ (x/q)
(64)
Hence 1−q Nl,λ (¯ z z) d2 z dµl,λ(¯ z , z) = 2 λ . l q ln q −1 Nl,λ(¯ z z/q) π
(65)
On the other hand, if q > 1, then combining (60), (47) and (48) of the Lemma 2.7 we get dl,λ 1 q x dθ , dµ(¯ z , z) = 2π 1 − (q − 1)x/(l2 q λ ) where 0 < x <
2.3
l2 q λ q−1
x = |z|2 , θ = arg(z),
and 0 ≤ θ ≤ 2π.
(66)
Statistics and geometry of coherent states |zil,λ
The conventional boson operators b and b† may be expressed in terms of the deformed operators a and a† as s s N N and b† = a† , ϕ(N) 6= ϕ(0) (67) b=a ϕ(N) ϕ(N) and their actions on the states |ni are given by √ √ b|ni = n|n − 1i, and b† |ni = n + 1|n + 1i.
(68)
Besides, br |ni =
s
n! |n − ri, (n − r)!
0≤r≤n
(69)
and (b† )s |ni =
r
(n + s)! |n + si. n! 17
(70)
2.3.1
Quantum statistics of the coherent states |zil,λ
Proposition 2.9 The expectation value of monomials of boson creation and annihilation operators b† , b in the coherent states |zil,λ are given by s ∞ s r X q [(n+s)(n+s+1)+(n+r)(n+r+1)]/2 (n + r)!(n + s)! |z|2n z¯ z , (71) h(b† )s br i = Nl,λ(|z|2 ) n=0 (l2 q λ )(n+s)+(n+r) [n + s]q ![n + r]q ! n! where s = 0, 1, 2, · · · and r = 0, 1, 2, · · ·. In particular, r xr d † r r h(b ) b i = Nl,λ (x), x = |z|2 , Nl,λ(x) dx
r = 0, 1, 2, · · · ,
(72)
and ′ Nl,λ (x) hNi = x , Nl,λ (x)
(73)
′ where Nl,λ (x) denotes the derivative with respect to x.
Proof: Indeed, for s = 0, 1, 2, · · · and r = 0, 1, 2, · · ·, we have h(b† )s br i := l,λ hz|(b† )s brs |zil,λ ∞ ∞ XX q [m(m+1)+n(n+1)]/2 n!(n − r + s)! m n 1 z¯ z hm|n + s − ri = Nl,λ (|z|2 ) m=0 n=r (l2 q λ )m+n [m]q ![n]q !(n − r)!(n − r)! s ∞ X 1 q [(n+s−r)(n+s−r+1)+n(n+1)]/2 n!(n − r + s)! = z¯n+s−r z n 2 2 λ n+s−r+n Nl,λ (|z| ) n=r (l q ) [n + s − r]q ![n]q !(n − r)!(n − r)! s ∞ z¯s z r X q [(n+s)(n+s+1)+(n+r)(n+r+1)]/2 (n + r)!(n + s)! |z|2n = , Nl,λ (|z|2 ) n=0 (l2 q λ )(n+s)+(n+r) [n + s]q ![n + r]q ! n! In the special case s = r, we have ∞
xr X q (n+r)(n+r+1)/2 (n + r)! xn h(b ) b i = Nl,λ (x) n=0 (l2 q λ )(n+r) [n + r]q ! n! ∞ xr X q n(n+1)/2 (n)! xn−r = Nl,λ (x) n=r (l2 q λ )(n) [n]q ! (n − r)! r xr d = Nl,λ(x), x = |z|2 . Nl,λ (x) dx † r r
18
In particular hNi ≡ hb† bi = x
′ Nl,λ (x) . Nl,λ (x)
The probability of finding n quanta in the deformed state |zil,λ is given by q n(n+1)/2 xn Pl,λ (n) := |hn|zil,λ| = 2 λ n . (l q ) [n]q !Nl,λ(x) 2
(74)
The Mendel parameter measuring the deviation from the Poisson statistics is defined by the quantity Ql,λ :=
hN 2 i − hNi2 − hNi . hNi
(75)
Let us evaluate it explicitly. From the expectation value of the operator N 2 = (b† )2 b2 + N provided by hN 2 i = x2
′ ′′ Nl,λ (x) Nl,λ (x) +x , Nl,λ(x) Nl,λ (x)
(76)
we readily deduce Ql,λ = x
′ ′′ (x) Nl,λ (x) Nl,λ − ′ Nl,λ (x) Nl,λ(x)
!
.
(77)
It is then worth noticing that for x << 1, Ql,λ = −
q(1 − q) x + o(x2 ) + q)
l2 q λ (1
(78)
meaning that the Pl,λ (n) is a sub-Poissonian distribution [71]. 2.3.2
Geometry of the states |zil,λ
The geometry of a quantum state space can be described by the corresponding metric tensor. This real and positive definite metric is defined on the underlying manifold that the quantum states form, or belong to, by calculating the distance function (line element) between two quantum states. So, it is also known as a Fubini-Study metric of the ray space. The knowledge of the quantum metric enables one to calculate quantum mechanical transition probability and uncertainties 19
In the case q < 1, the map from z to |zil,λ defines a map from the space C of complex numbers onto a continuous subset of unit vectors in Hilbert space and generates in the latter a two-dimensional surface with the following Fubini-Study metric: dσ 2 := ||d|zil,λ||2 − |l,λ hz|d|zil,λ|2
(79)
Proposition 2.10 The above Fubini-Study metric is reduced to dσ 2 = Wl,λ (x)d¯ z dz,
(80)
where x = |z|2 and Wl,λ (x) =
′ Nl,λ (x) x Nl,λ(x)
′
=
d hNi. dx
(81)
In polar coordinates, z = reiθ , dσ 2 = Wl,λ (r 2 )(dr 2 + r 2 dθ2 ).
(82)
Proof: Computing d|zil,λ by taking into account the fact that any change of the form d|zil,λ = α|zil,λ , α ∈ C, has zero distance, we get d|zil,λ = Nl,λ(|z|2 )−1/2 Then, ||d|zil,λ||
2
∞ X q n(n+1)/4 nz n−1 p |ni dz. 2 q λ )n [n] ! (l q n=0
2 −1
= Nl,λ(|z| )
= Nl,λ(|z|2 )−1 +|z|2
∞ X q n(n+1)/2 n2 |z|2(n−1) n=0 ∞ X n=0
(l2 q λ )n [n]q !
q n(n+1)/2 n|z|2(n−1) (l2 q λ )n [n]q !
∞ X q n(n+1)/2 n(n − 1)|z|2(n−2)
(l2 q λ )n [n]q ! ′ ′′ Nl,λ (x) + xNl,λ (x) d¯ z dz ′ ′ xNl,λ (x) d¯ z dz
n=0 −1
= Nl,λ(x) = Nl,λ(x)−1 and |l,λ hz|d|zil,λ|2
d¯ z dz
!
d¯ z dz
2 ∞ n(n+1)/2 2(n−1) X q n|z| 2 −1 z¯dz = Nl,λ(|z| ) 2 λ n (l q ) [n]q ! n=0
20
′ = xNl,λ (x)−2 Nl,λ (x)
Therefore, dσ
2
2
d¯ z dz.
2 −1 ′ ′′ −2 ′ d¯ z dz = Nl,λ (x) Nl,λ(x) + xNl,λ (x) − xNl,λ (x) Nl,λ (x) ′ ′ Nl,λ (x) d = x d¯ z dz = hNi d¯ z dz. Nl,λ (x) dx
For x << 1, we have q 2q(1 − q) 2 Wl,λ (x) = 2 λ 1 − 2 λ x + o(x ) . l q l q (1 + q)
3
(83)
On generalized oscillator algebras and their associated coherent states
A unified method of calculating structure functions from commutation relations of deformed single-mode oscillator algebras is presented. A natural approach to building coherent states associated to deformed algebras is then deduced [16]. Known deformed algebras are given as illustration and such mathematical properties as the continuity in the label, normalizability and resolution of the identity of the corresponding coherent states are discussed.
3.1
Unified deformed single-mode oscillator algebras
Definition 3.1 We call deformed Heisenberg algebra, an associative algebra generated by the set of operators {1, a, a† , N} satisfying the relations [N, a† ] = a† ,
[N, a] = −a,
(84)
such that there exists a non-negative analytic function f , called the structure function, defining the operator products a† a and aa† in the following way: a† a := f (N),
aa† := f (N + 1),
(85)
where N is a self-adjoint operator, a and its Hermitian conjugate a† denote the deformed annihilation and creation operators, respectively.
21
Afore-mentioned deformed Heisenberg algebras have a common property characterized by the existence of a self-adjoint number operator N, a lowering operator a and its formal adjoint, called raising operator, a† and differ by the expression of the structure function f . The associated Fock space F is now spanned by the orthonormalized eigenstates of the number operator N given by:
where
|ni = p
1 f (n)!
(a† )n |0i,
n ∈ N ∪ {0},
f (n)! = f (n)f (n − 1)...f (1)
(86)
with f (0) = 0.
(87)
p f (n + 1)|n + 1i.
(88)
Moreover, a|ni =
p
f (n)|n − 1i, a† |ni =
We emphasize that the structure function f is a key unifying methods of coherent state construction corresponding to deformed algebras. To this end let us first recall the definition of the canonical coherent states. Definition 3.2 The canonical coherent states (CS) are normalized states |zi ∈ H satisfying one of the following three equivalent conditions [48, 49, 69, 70, 97, 102]: (i) they saturate the Heisenberg inequality: (∆ˆ q )(∆ˆ p) =
~ , 2
(89)
where (∆A)2 := hz|A2 − hAi2 |zi with hAi := hz|A|zi; (ii) they are eigenvectors of the annihilation operator, with eigenvalue z ∈ C: b|zi = z|zi;
(90)
(iii) they are obtained from the ground state |0i of the harmonic oscillator by a unitary action of the Weyl-Heisenberg group: † −¯ zb
|zi = ezb
|0i.
(91)
From (91) and using the famous Baker-Campbell-Hausdorff formula 1
eA+B = e− 2 [A,B] eA eB 22
(92)
whenever [A, [A, B]] = [B, [A, B]] = 0, one obtains |zi = e−|z|
2 /2
∞ X zn 2 † √ |ni = e−|z| /2 ezb |0i, z ∈ C. n! n=0
(93)
The important feature of these coherent states resides in the partition (resolution) of identity: Z [d2 z] |zihz| = 1, (94) C π where we have put [d2 z] = d(Rez)d(Imz) for simplicity. Suppose that f (0) = 0, f (n) > 0 for all n ∈ N and denote Df = {z ∈ C : |z|2 < Rf }, where Rf is the radius of convergence of the series ( called deformed exponential function): Nf (x) :=
∞ X n=0
xn . [f (n)]!
(95)
Then, the following holds: Proposition 3.3 The states 2
|z, f i := (Nf (|z| ))
−1/2
∞ X n=0
= (Nf (|z|2 ))−1/2
∞ X n=0
zn (a† )n |0i [f (n)]! zn p
[f (n)]!
|ni, z ∈ Df ,
(96)
are normalized eigenstates of the raising operator a with eigenvalue z. They are not orthogonal to each other. Moreover, the map z 7→ |z, f i from Df ⊂ C to the Fock space F is continuous. Proof: The first assertion is true by definition of states |z, f i and the action of the raising operator a. To prove the non orthogonality, let z1 , z2 ∈ Df . Then, hz1 , f |z2, f i =
Nf (¯ z1 z2 )
(Nf (|z1 |2 )Nf (|z2 |2 ))1/2
6= 0,
when z1 6= z2 .
(97)
Furthermore, |||z1 , f i − |z2 , f i||2 = 2 (1 − Rehz1 , f |z2 , f i) → 0 as |z1 − z2 | → 0 that means the map Df ∋ z 7→ |z, f i ∈ F is continuous. 23
(98)
The family {|z, f i : z ∈ Df } will be called coherent states whether there exists a positive measure µf such that [72]: Z
Df
dµf (¯ z , z)|z, f ihz, f | =
∞ X n=0
|nihn| = 1,
thus forming an overcomplete set of states, or equivalently Z dµf (¯ z , z) = δnm [f (n)]!, n, m = 0, 1, 2, ... z¯n z m Nf (|z|2 ) Df
(99)
(100)
√ Passing to polar coordinates, z = xeiθ , where 0 ≤ θ ≤ 2π, 0 < x < Rf , and dµ(¯ z , z) = dωf (x)dθ, the latter equation leads to the following classical Stieltjes (for Rf = ∞) or Hausdorff (Rf < ∞) power-moment problem [2, 98]: Z Rf 2π dωf (x) = [f (n)]!, n = 0, 1, 2, ... (101) xn Nf (x) 0
Note immediately that not all deformed algebras lead to coherent states because the moment problem (101) does not always have solution [2, 98]. Nevertheless, it is remarkable that the structure function f plays an important role in the construction of coherent states associated to an algebra. So, the question arises is then how to determine the structure function corresponding to a given algebra. Many techniques have been proposed in literature [9, 13, 19, 76, 83]. A more general answer to this question can be given starting from the Meljanac et al [83] point of view. Indeed, these authors introduced the generalized q-deformed singlemode oscillator algebra through the identity operator 1, a self-adjoint number operator N, a lowering operator a and an operator a¯ which is not necessarily conjugate to a satisfying [N, a] = −a, [N, a¯] = a ¯, a¯a − F (N)¯aa = G(N)
(102) (103)
where F and G are arbitrary complex analytic functions. Such an algebra furnishes an appropriate approach for the unification of classes of deformed algebras known in the literature. For the purpose, let us start from the relations (102) to get [N, a¯a] = 0 = [N, a¯a]
(104)
implying the existence of a complex analytic function ϕ such that a ¯a = ϕ(N) and
a¯a = ϕ(N + 1). 24
(105)
Therefore, Eq.(103) can be rewritten as follows ϕ(N + 1) − F (N)ϕ(N) = G(N).
(106)
Denote now a† the Hermitian conjugate of the operator a. Then, [N, a† ] = a† ,
a ¯ = c(N)a† ,
and
(107)
where c(N) is a complex function. For convenience take c(N) = ei arg ϕ(N ) . Therefore, from (105) and the fact that a† a and aa† are Hermitian operators we necessarily have a† a = |ϕ(N)|
aa† = |ϕ(N + 1)|.
and
(108)
We now assume the existence of a ”vacuum state” |0i such that N|0i = 0,
a|0i = 0 and h0|0i = 1,
(109)
and construct the non normalized eigenvectors (a† )n |0i of the operator N. It follows that m
† n
h0|a (a ) |0i = δmn
n Y
k=1
|ϕ(k)| =: (|ϕ(n)|!)δmn , m, n = 0, 1, 2, · · ·
(110)
and we have the following proposition Proposition 3.4 Suppose that the initial condition ϕ(0) = 0 is satisfied. Then
ϕ(n) = [F (n − 1)]!
n−1 X G(k) , [F (k)]! k=0
n ≥ 1,
(111)
where [F (k)]! =
F (k)F (k − 1) · · · F (1) 1
if if
k≥1 . k=0
(112)
Proof: Applying (106) to the vector (a† )n |0i, we obtain ϕ(n + 1) − F (n)ϕ(n) = G(n),
n = 0, 1, 2 · · ·
(113)
Then the result follows. Notice that for all n = 1, 2, ... each ”excited” state (a† )n |0i is a eigenstate of the operator N corresponding to the eigenvalue n with norm p ||(a† )n |0i|| = |ϕ(n)|! . (114) 25
Of course, these states are orthogonal, i.e. h0|am (a† )n |0i = 0 for m 6= n.
(115)
Now, if ϕ(n) 6= 0 for all n ≥ 1, one normalizes the eigenstates (a† )n |0i of N and gets the vectors |ni ∈ F as |ni = p
(a† )n [|ϕ(n)|]!
|0i.
(116)
In the opposite, if ϕ(n0 ) = 0 for some n0 , then the state (a† )n0 |0i has zero norm and, consistently, we can put |n0 , ϕi ≡ 0. The corresponding Hilbert space is the finite-dimensional space Cn0 . Besides, the following relations hold: p p a|n, ϕi = |ϕ(n)||n − 1, ϕi, a† |n, ϕi = |ϕ(n + 1)||n + 1, ϕi, (117) a† a|n, ϕi = |ϕ(n)||n, ϕi and aa† |n, ϕi = |ϕ(n + 1)||n, ϕi (118) showing that the structure function characterizing a given deformation is defined as follows: f (n) = ϕ(n) if ϕ(n) ≥ 0, and f (n) = |ϕ(n)|, in general. Moreover, the conventional boson operators b and b† may be expressed in terms of the deformed operators a and a† as s s N N and b† = a† . (119) b=a f (N) f (N) Thus the actions of b and b† on the states are as usual √ √ b|n, ϕi = n|n − 1, ϕi, and b† |n, ϕi = n + 1|n + 1, ϕi.
(120)
So, for simplicity of notations we set |n, ϕi = |ni. Note also that the WeylHeisenberg algebra oscillator corresponds to F (N) = G(N) = 1. In the next section, we analyse known algebras in the light of the above developed formalism.
3.2
Application to known deformed algebras
3.3
The Tamm-Dancoff deformed algebra
This algebra appeared in the frame of Tamm-Dancoff method [27, 104], in quantum field theory, and was defined by the commutation relations [N, a† ] = a† , [N, a] = −a, † † N aa − qa a = q , 26
(121) (122)
where q is an arbitrary complex non-zero number. This corresponds to the case F (N) = q and G(N) = q N , and yields ϕ(n) = nq n−1
and
f (n) = n|q|n−1 .
(123)
In this case, the exponential function (95) written as Nq (x) =
∞ X n=0
xn n!|q|n(n−1)/2
(124)
converges on the whole complex plane C for |q| ≥ 1. Proposition 3.5 The moment problem (101) with Rf = +∞ has the following solution o n p p ln2 (x/t) Z ∞ |q| exp − |q|t + 2 ln |q| Nq (x) dx p dωq (x) = dt. (125) 2π x 2π ln |q| 0
˜ q (x)dx = 2π dωq (x) , the moment problem (101) is then written Proof: Setting W Nq (x) as follows: Z ∞ ˜ q (x)dx = n!|q|n(n−1)/2 , n = 0, 1, 2, ... xn W (126) 0
2 The inverse Mellin transforms of f˜1 (s) = |q|s /2 and f˜2 (s) = |q|−s/2Γ(s) gives [93, 100] ln2 x
√ e− 2 ln |q| −1 ˜ − |q|x ˜ p =: f1 (x) and M {f2 (s)} = e M {f1 (s)} = =: f2 (x) (127) 2π ln |q| −1
respectively. Thus, the solution of the integral equation Z ∞ ˜ q (x)dx = |q|s(s−1)/2 Γ(s + 1), xs W
(128)
0
can be obtained using the Mellin integral formula [91] Z ∞ M f1 (x/t)f2 (t)dt = f˜1 (s)f˜2 (s + 1) 0
which, in particular, gives Z ∞ − ln22ln(x/t) Z ∞ √ |q| e p e− |q|t dt dx = |q|−1/2 |q|s(s−1)/2 Γ(s + 1) xs−1 2π ln |q| 0 0 27
(129)
or equivalently n p p Z ∞ Z ∞ |q| exp − |q|t + p xs x 2π ln |q| 0 0
ln2 (x/t) 2 ln |q|
o
dt dx = |q|s(s−1)/2 Γ(s + 1).(130)
Therefore,
˜ q (x) = W
Z
0
∞
n p p |q| exp − |q|t + p x 2π ln |q|
ln2 (x/t) 2 ln |q|
o
dt.
Then follows the result. Hence, the states defined in C by
|z, qi = (Nq (|z|2 )−1/2 = (Nq (|z|2 )−1/2
∞ X n=0
∞ X n=0
constitute a family of coherent states. 3.3.1
zn n!|q|n(n−1)/2
(131)
(a† )n |0i
zn p |ni n!|q|n(n−1)/2
(132)
The Arick-Coon-Kuryskin deformed algebra (1976)
Arick and Coon first introduced this algebra [7] whose generators satisfy the following relations: [N, a† ] = a† , [N, a] = −a, † † aa − qa a = 1, .
(133) (134)
The same algebra was examined independently by other authors like Kuryshkin[78], Jannussis [59], etc., and has gained popularity because of its connection to the developed mathematical q-analysis theory. One can check that F (N) = q and G(N) = 1 leading to ϕ(n) =
qn − 1 =: [n]q and f (n) = ϕ(n) for q ∈ [−1, 1[∪]1, ∞). q−1
(135)
wihch is one of the forms of the so called q-numbers. This result is also obtained replacing (134) by [a, a† ] = q N .
28
(136)
Then follows the series ∞ X xn . Nq (x) = [n]q ! n=0
(137)
Arick and Coon have considered the case 0 < q < 1 for which Nq (x) = eq ((1 − q)x) where eq (x) is one of the famous Jackson q-exponential functions, which converges for |x| < 1 [42]. In this case the radius of convergence ofNq (x) is 1 Rq = 1−q . Note also that, ∂q Nq (x) = Nq (x),
(138)
where the q-derivative ∂q is defined as follows ∂q f (x) =
f (x) − f (qx) . (1 − q)x
(139)
The following statements hold: Lemma 3.6 1 1 Nq (x) , = , Nq (x) = Nq (qx) 1 − (1 − q)x ((1 − q)x; q)∞ ∞ X xn 1 eq (x) = = Nq (x/(1 − q)) = , (q; q) (x; q) n ∞ k=0 where (a; q)n =
(141)
n−1 Y k=0
Z
(140)
(1 − aq k ). Moreover,
(1−q)−1
xn (Nq (qx))−1 dq x = [n]q !
n = 0, 1, 2, ...,
(142)
0
where Z
0
a
f (x)dq x = a(1 − q)
∞ X
f (aq n )q n
(143)
k=0
defines the Jackson integral [42] of a function f .
Proof: Equations (138) and (139) lead to (140)-(141). Using the Jackson integral we obtain Z (1−q)−1 ∞ X −1 q (n+1)l −1 l+1 n N (q /(1 − q)) x (Nq (qx)) dq x = q (1 − q)n 0 l=0 29
∞ ∞ X (q; q)∞ X q (n+1)l q (n+1)l l+1 (q ; q)∞ = = (1 − q)n (1 − q)n (q; q)l l=0 l=0 (q; q)∞ (q; q)n 1 = = . n n+1 (1 − q) (q ; q)∞ (1 − q)n
Then (142).
✷
1 Proposition 3.7 The solution of the moment problem (101) with the Rq = 1−q is given by dq x 1 , x = |z|2 (144) dωq (x) = 2π 1 − (1 − q)x
Proof: In this case the moment problem (101) becomes Z
1 1−q
xn
0
2π dωq (x) = [n]q !, Nq (x)
n = 0, 1, 2, ...
(145)
The comparison of moment problem (145) and (142) with the use of the first equality of previous relations (140) leads to (144). Thus, the states |ziq
∞ X zn † n = (Nq (|z| )) (a ) |0i [n]q ! n=0 ∞ X zn 2 −1/2 p |ni = (Nq (|z| )) [n]q ! n=0 2
−1/2
(146)
define a family of coherent states on Dq = z ∈ C : |z| < (1 − q)−1/2 . 3.3.2
The Feinsilver deformed algebra (1987)
The generators of the algebra by Feinsilver algebra [36, 37] verify [a, a† ] = q −2N ,
[N, a] = −a,
[N, a† ] = a† ,
(147)
where q is non-zero real number. It follows that F (N) = 1 and G(N) = q −2N implying f (n) = ϕ(n) =
1 − q −2n = [n]q−2 . 1 − q −2
(148)
The change of parameters q˜ = q −2 , with q > 1, leads to the previous Arick-CoonKuryshkin algebra. 30
3.3.3
The Biedenharn-Macfarlane oscillator algebra (1989)
The generators of the deformed algebra introduced by Biedenharn [12] and independently by Macfarlane [79], in the context of oscillator realization of the quantum algebra suq (2), satisfy [N, a† ] = a† , [N, a] = −a, † † −N aa − qa a = q or aa† − q −1 a† a = q N ,
2
q 6= 1.
(149) (150)
Here F (N) = q and G(N) = q −N . Therefore, ϕ(n) = q
n−1
n−1 −j X q j=0
qj
=
q n − q −n 2 , q 6= 1 q − q −1
(151)
Then, the structure function is given by q n − q −n B =: [n]q , f (n) = −1 q−q
q ∈ R∗+ \ {1}.
(152)
Hence, considering the series ∞ X xn NB (x) = , [n]Bq ! n=0
(153)
one notices that its radius of convergence RB = +∞. Using the deformed derivative and integration defined by ∂qB f (x) :=
f (qx) − f (q −1 x) (q − q −1 )
(154)
and Z
xj
xi
−1 f (x)dB q x := (q − q )x
f X
q 2l f (q 2l x), xi = xq i , xf = xq f
(155)
l=i
respectively, You-quan and Zheng-mao [110] showed that Z ∞ B xn NB (−x)dB n = 0, 1, 2, · · · q x = [n]q !,
(156)
0
Therefore, the power-moment problem (101) has a solution given by dωB (x) =
1 NB (x)NB (−x)dB q x 2π 31
(157)
and the states |ziB = (NB (|z|2 ))−1/2
∞ X n=0
define a family of coherent states with z ∈ C.
zn q |ni B [n]q !
(158)
Remark 3.8 • El Baz and Hassouni [32] demonstrated, for |q| = 1 and using the Fourrier transforms, that the power-moment problem (101) has the solution dωB (x) =
1 ˜ ˜ B (x)dx N (x)W 2π B
(159)
˜ (x) is the Fourier transform of the series where W B W B (y) =
∞ X [n]B q ! n=0
(iy)n
(160)
e−ixy W B (y)dy.
(161)
n!
i.e. ˜ (x) = 1 W B 2π
Z
∞ −∞
Notice that in this case the function NB (x) is replaced by N˜B (x) =
∞ X n=0
xn . |[n]Bq !|
• Yan [109] has later examined this algebra with relations a† a = [N], 3.3.4
aa† = [N + 1],
[a, a† ] = [N + 1] − [N].
(162)
The Calogero-Vasiliev oscillator algebra (1991)
In 1991 Vasiliev [105] introduced a deformed algebra whose generators satisfy [N, a] = −a, [N, a† ] = a† , [N, K] = 0, aK = −Ka, a† K = −Ka† , K 2 = 1, [a, a† ] = 1 + νK,
(163) (164) (165)
where ν is a real such that ν > − 21 and K = (−)N is the Klein operator interpreted as the generator of the symmetric group S2 . From (165), we have F (N) = 1 and G(N) = 1 + ν(−)N . Therefore ϕ(2n) = 2n and ϕ(2n + 1) = 2(n + ν) + 1, n = 0, 1, 2, ... 32
(166)
The exponential function (95) written as ∞ X xn Nν (x) = ϕ(n)! n=0
(167)
converges everywhere x. However, the corresponding moment problem (101) Z ∞ 2π dων (x) xn = ϕ(n)!, n = 0, 1, 2, · · · (168) Nν (x) 0 remains to solve. 3.3.5
The (p, q)-Chakrabarti-Jagannathan oscillator algebra (1991)
The two-parameter quantum algebra was first introduced by Chakrabarti and Jagannathan [24] in order to generalize or/and unify the Arick-Coon-Kurskin oscillator algebra (p = 1) and Biedenharn-Macfarlane oscillator algebra (p = q). The generators satisfy [N, a] = −a, [N, a† ] = a† , aa† − qa† a = p−N , or aa† − p−1 a† a = q N ,
(169) (170)
where p, q ∈ R∗+ . From the first relation of (170) we deduce F (N) = q and G(N) = p−N . Therefore, ϕ(n) = q n−1
n−1 −j X p j=0
qj
= q n−1
1 − ((pq)−1)n p−n − q n = =: [n]p−1 ,q . 1 − (pq)−1 p−1 − q
(171)
Notice that the second relation of (170) gives the same result. Hence, it suffices to consider only one of the two relations (170). The Fock space of the Bose oscillator FCJ associated to this deformation is generated by the orthonormalized states
where
|ni = p [n]p−1 ,q ! =
(a† )n |0i, [n]p−1 ,q !
n = 0, 1, 2, ...
1 [n]p−1 ,q [n − 1]p−1 ,q ...[1]p−1 ,q 33
if n = 0 . if n ≥ 1
(172)
(173)
3.3.6
The Kalnins-Mukherjee-Miller oscillator algebra (1993)
This q-oscillator algebra generated by four elements H, E+ , E− and E satisfying [H, E+ ] = E+ [H, E− ] = −E− −H [E+ , E− ] = −q E [E, E± ] = 0 = [E, H],
(174) (175)
where 0 < q < 1, was introduced by Kalnins et al [66]. The elements C = qq −H E + (q − 1)E+ E− and E lie in the center of this algebra. It admits a class of irreducible representations for C = l2 1 and E = l2 q λ−1 1, where l and λ are real numbers with l 6= 0. Setting N = H, a = E− and a† = E+ , we get [N, a] = −a, [N, a† ] = a† , aa† − a† a = l2 q −N +λ−1 .
(176) (177)
Hence, F (N) = 1 and G(N) = l2 q −N +λ−1 leading to ϕ(n) =
n−1 X
l2 q −k+λ−1 = l2 q λ
k=0
1 − q −n = l2 q λ−n [n]q q−1
(178)
Suppose q < 1. Then, the series Nl,λ(x) =
∞ ∞ X X q n(n−1)/2 (1 − q)qx n q n(n+1)/2 xn = (l2 q λ )n [n]q ! n=0 (q; q)n l2 q λ n=0
= Eq ((1 − q)qx/(l2 q λ )).
(179)
has a radius of convergence Rl,λ = ∞ [42]. While in the case q > 1, the factors ϕ(n) remain positive for every n ≥ 0, but the series Nl,λ(x) converges only for 2 qλ |x| < lq−1 := Rl,λ . We have the following result: Proposition 3.9 The power-moment problem (101) has a solution given by dωl,λ (x) = and dωl,λ (x) =
1−q Nl,λ(x) 1 dx, for 0 < q < 1, 2 λ −1 2π l q ln q Nl,λ(x/q)
dl,λ 1 q x , 2π 1 − (q − 1)x/(l2 q λ )
Proof: See proof of proposition 2.8 Hence, the states |zil,λ =
−1/2 Nl,λ (|z|2 )
0<x<
l2 q λ for q > 1. q−1
∞ X q n(n+1)/4 z n p |ni (l2 q λ )n [n]q ! n=0
form a family of coherent states for z ∈ Dl,λ = {z ∈ C; |z|2 < Rl,λ }. 34
(180)
(181)
(182)
3.3.7
The Chung-Chung-Nam-Um oscillator algebra (1993)
The generalized (q, α, β)-algebra was introduced by Chung et al [26] with the generators satisfying [N, a] = −a, [N, a† ] = a† , aa† − qa† a = q αN +β
(183) (184)
where q ∈ R∗+ and α, β are real parameters. One notices that F (N) = q and G(N) = q αN +β . Therefore, n−1+β nq if α = 1 (185) ϕ(n) = f (n) =: Fα,β (n; q) = β q n −q αn q q−qα if α 6= 1. This algebra generalizes the algebras introduced by: • Tamm-Dancoff, with α = 1, β = 0; • Arick-Coon-Kuryshkin, with α = β = 0; and • Biedenharn-Macfarlane, with α = −1, β = 0. 3.3.8
The Borzov-Damasky-Yegorov oscillator algebra (1993)
γ The generalized Wα,β (q)-algebra was introduced by Borzov et al [13] in order to unify a large class of konwn q-deformed oscillator algebras. The generators satisfy
[N, a] = −a, [N, a† ] = a† , aa† − q γ a† a = q αN +β
(186) (187)
where q ∈ R∗+ and α, β, γ are real parameters. Here F (N) = q γ and G(N) = q αN +β leading to γ(n−1)+β nq if α = γ γ ϕ(n) = f (n) = Fα,β (n; q) = (188) β q γn −q αn if α 6= γ. q qγ −qα 3.3.9
The Brzezi´ nski - Egusquinza - Macfarlane oscillator algebra (1993)
Brzezi´ nski et al [14] introduced this algebra as q-deformation `a la Biedenharn - Macfarlane of the Calogero-Vasiliev oscillator algebra. It is governed by the relations: [N, a] = −a, [N, a† ] = a† , [N, K] = 0, aK = −Ka, a† K = −Ka† , K 2 = 1, aa† − qa† a = q −N (1 + 2αK), 35
(189) (190) (191)
where α ∈ R∗ , q ∈ R+ and K = (−)N is the Klein operator. Here F (N) = q and G(N) = q −N (1 + 2αK) leading to ϕ(n) =: fα (n) = 3.3.10
q n − (−1)n q −n q n − q −n + 2α . q − q −1 q + q −1
(192)
The Quesne oscillator algebra (2002)
The coherent states introduced by Quesne [94] may be associated with the q-deformed algebra satisfying the relations [52]: [N, a] = −a, [N, a† ] = a† , aa† − a† a = q −N −1 or qaa† − a† a = 1
(193) (194)
where 0 < q < 1. The first relation of (194) suggests F (N) = 1 and G(N) = q −N −1 leading to ϕ(n) =
1 − q −n = q −n [n]q =: [n]Q q . q−1
(195)
It is a particular case of the Kalnins-Miller-Mukherjee algebra developed above with l = 1, λ = 0. 3.3.11
The (q; α, β, γ; ν)-Burban oscillator algebra (2007)
In 2007, Burban [19] introduced the (q; α, β, γ; ν)-oscillator algebra whose generators satisfy the relations [N, a] = −a, [N, a† ] = a† , [N, K] = 0, aK = −Ka, a† K = −Ka† , K 2 = 1, aa† − q γ a† a = q αN +β (1 + 2νK),
(196) (197) (198)
where ν ∈ R∗ , α, β ∈ R, q ∈ R+ and K = (−)N is the Klein operator. Here F (N) = q γ and G(N) = q αN +β (1 + 2νK) leading to ϕ(n) = q =
γ(n−1)
(
For α = γ and 1 + 2ν > 0
n−1 αj+β X q (1 + 2ν(−1)j )
q γj
j=0 γ(n−1)+β
q (n + ν(1 − (−1)n )) if α = γ γn q γn −(−1)n q αn −q αn if α 6= γ. q β q qγ −q α + 2ν q γ +q α
γ f (n) = Fα,β;ν (n; q) = q γ(n−1)+β (n + ν(1 − (−1)n )).
For each of the following cases 36
(199)
(200)
γ
α
γ
α
+q 1. (q < 1, α < γ and −1 < 2ν < − qqγ −q α ); +q 2. (q > 1, α > γ and −1 < 2ν < − qqγ −q α );
3. (q < 1, α > γ and 1 + 2ν > 0); 4. q > 1, α < γ and 1 + 2ν > 0) we have f (n) = 3.3.12
γ Fα,β;ν (n; q)
=q
β
q γn − q αn q γn − (−1)n q αn + 2ν qγ − qα qγ + qα
.
(201)
The (p, q) and (p, q; µ, ν, f )-oscillator algebras (2007)
In 2007, our group [52] introduced an algebra generalizing the Quesne oscillator algebra: [N, a] = −a, −1 p aa† − a† a = q −N −1 ,
[N, a† ] = a† , qaa† − a† a = pN +1
(202) (203)
From the first relation (203), we get F (N) = p and G(N) = pq −N −1 and then ϕ(n) = pn−1
n−1 X pq −j−1 j=0
pj
=
pn − q −n =: [n]Q p,q . q − p−1
(204)
The (p, q; µ, ν, f )-oscillator algebra is defined through the following commutation relations [N, a† ] = a† , ν N q pµ−1 † −1 † aa − p a a = f (p, q) ν−1 q pµ−1
[N, a] = −a,
(205) (206)
where p, q, µ, ν are real numbers such that 0 < pq < 1, pµ < q ν−1 , p > 1 and f a well behaved real and non-negative function of deformation parameters p and q, satisfying lim f (p, q) = 1 as (p, p) → (1, 1). ν N q ν−1 q q ν−1 Here, F (N) = µ and G(N) = f (p, q) µ−1 , so that p p pµ−1 ν n n p − q −n q =: [n]µ,ν (207) ϕ(n) = f (p, q) µ p,q,f . −1 p q−p
37
The series µ,ν Np,q,f (x)
=
∞ X n=0
xn [n]µ,ν p,q,f !
(208)
has a radius of convergence R = +∞. It had also shown in [52] that the moment problem (101) has, for µ = 1 and ν = 0, the following solution 1,0 dωp,q,f (x)
Finally, the states
1,0 Np,q,f (x) 1 p−1 − q = dx. 1,0 2π f (p, q) ln (pq)−1 Np,q,f (x/(pq))
1,0 2 |zi1,0 p,q,f = Np,q,f (|z| )
∞ −1/2 X n=0
form a family of coherent states. 3.3.13
(209)
zn q |ni, z ∈ C, 1,0 [n]p,q,f !
(210)
Unified (p, q; α, β, ν; γ)-deformed oscillator algebra (2012)
More recently, Balo¨ıtcha et al [9] introduced the unified (p, q; α, β, ν; γ)-deformed oscillator algebra whose generators satisfy: [N, a] = −a, [N, a† ] = a† , [N, K] = 0, aK = −Ka, a† K = −Ka† , K 2 = 1, aa† − pν a† a = (1 + 2γK)q αN +β ,
(211) (212) (213)
where, α, β, γ, ν ∈ R, p, q ∈ R+ and K = (−)N is the Klein operator. Here, F (N) = pν and G(N) = (1 + 2γ(−)N )q αN +β and pνn −(−1)n q αn q β pνnν −qαn + 2γ if pν 6= q α α ν α p −q p +q . ϕ(n) = q β+α(n−1) n + 2γ 1−(−)n if pν = q α . 2
(214)
For pν = q α and 1 + 2γ > 0 f (n) =:
ν Fα,β;γ (n; p, q)
=q
β+α(n−1)
For each of the following cases:
1 − (−)n . n + 2γ 2
(215)
1. pν > q α and 1 + 2γ > 0, and ν
α
+q 2. pν < q α and −1 < 2γ < − ppν −q α
we have
f (n) =:
ν Fα,β;γ (n; p, q)
=q
β
pνn − q αn pνn − (−1)n q αn + 2γ pν − q α pν + q α 38
.
(216)
4
R(p, q)-deformed quantum algebras: coherent states and special functions
We provide with a generalization of well known (p, q)-deformed Heisenberg algebras, called R(p, q)-deformed quantum algebras, and study the corresponding R(p, q)-series. A general formulation of the binomial theorem is given. Special functions are obtained as limit cases. This work well prolongs a previous work by Odzijewicz [87]. Known results in the literature are recovered.
4.1
Theoretical framework
The development displayed in this section is essentially based on the formalism elaborated by Odzijewicz [87] in a nice, mathematically based work published in 1998, but unfortunately hushed up in the recent literature on the topic. In the mentioned work, this author investigated the quantum algebras generated by the coherent state maps of the disc, leading to a generalized analysis which includes standard analysis as well as q-analysis. He provided with the meromorphic continuation of the generalized basic hypergeometric series and constructed a reproducing measure, when the series is treated as a reproducing kernel. Indeed, much to our very great surprise, most all the remarkable coherent state generalizations, performed from the generalization of exponential function by different authors, can be generated from this more general theory. Let H be an infinite dimensional separable Hilbert space and {|ni}∞ n=0 its canon∞ ical basis. Assume that there exists a sequence {fn }n=0 in H such that fn = cn C|ni
(217)
where C and its inverse C −1 are bounded operators on H, and cn (n = 0, 1, 2, · · ·) are real positive numbers satisfying the conditions sup n∈N
Definition 4.1
√ cn−1 < +∞ and R−1 = lim sup n cn . n→∞ cn
(218)
[87] A coherent states map is a complex analytic map K : DR −→ H \ {0} z
֒→
K(z) =
∞ X n=0
39
fn z n
(219)
where DR ={z ∈ C : |z| < R}. The states K(z) are called coherent states and the operator A admitting these states as eigenstates with eigenvalues z ∈ DR , i.e. AK(z) = zK(z),
(220)
is said to be the annihilation operator. The relations (217)-(220) lead to C −1 AC|0i = 0 and
C −1 AC|ni =
cn−1 |n − 1i ∀n ≥ 1. cn
(221)
Therefore, kC −1 ACk < +∞ meaning that A is a bounded operator. Its Hermitian conjugate, called the creation operator and denoted by A† , is also bounded. The algebra closure, spanned by the operators {A, A† }, in their norm topology, gives the so-called C ∗ −algebra AK . Proposition 4.2 Let O(DR ) be a set of holomorphic functions defined on the disc DR . The map IK : H −→ O(DR ) such as IK (v) :=< v | K(·) >,
v∈H
(222)
is an antilinear monomorphism of complex vector spaces. Moreover, considering the topology of simple convergence, IK (H) is dense on O(DR ). Proof 4.1 From (219) fN +1 = lim
z→0
1
z
(K(z) − N +1
N X
fk z k )
and
f0 = K(0),
(223)
k=0
it results, by induction, that fN belongs to the closure of the linear space spanned by the elements of the subset {K(z), |z| < ǫ < R} ⊂ H. Otherwise, the coherent states K(z), z ∈ DR , form a linearly subset dense in H. One can easily check that IK is antilinear, i.e. , ¯ K (v), IK (αu + βv) = αI ¯ K (u) + βI
∀α, β ∈ C u, v ∈ H,
(224)
and implying z n ∈ IK (H) ,
† −1 z n = IK c−1 n (C ) |ni
∀n ≥ 0.
40
(225)
Therefore, IK (H) inherits the Hilbert space structure endowed with the scalar product hIK (v1 )|IK (v2 )i := hv1 |v2 i ∀v1 , v2 ∈ H. (226) Thus, the Hilbert space H can be identified with the dense subspace IK (H) of −1 O(DR ). Hence, IK ◦ AK ◦ IK is the analytic representation of the algebra AK as shown in the following diagram:
H
X
/
H IK
IK
IK (H)
IK (H)
Id
−1 βX = IK ◦ X ◦ IK
Id
O(DR )
βX
/
O(DR )
So, given an operator X ∈ AK , we find a unique linear operator βX into O(DR ) −1 such that βX = IK ◦X ◦IK . Consequently, the action of the analytic representation −1 † (IK ◦ A ◦ IK ) of the creation operator A† on ϕ ∈ O(DR ) yields: −1 IK ◦ A† ◦ IK ϕ(z) = zϕ(z), (227)
−1 while that of the analytic representation ∂ := (IK ◦A◦IK ) of the annihilation operator, the so-called K−derivative [87], depends on the operator C and parameters cn . On the basis elements z n , it gives n
∂z =
∞ X k=0
−1 † ck c−1 A C|kiz k . n hn|C
(228)
Without loss of generality and as a matter of convenience, we restrict, in the sequel, the analysis to a unity operator C, (i.e. C = I) leading to ∂z n = [n]z n−1
(229)
where [n] = Therefore,
(
0 if n = 0 2 cn−1 if n ≥ 1. cn
c0 cn = p [n]!
(230)
(231) 41
with [0]! = 1 and [n]! = [n]([n − 1]!), and
AA† |ni = [n + 1]|ni, cn ≥
A† A|ni = [n]|ni,
c0 for kAk ≥ R. kAkn
(232) (233)
The sequence {[n]}n≥0 converges and R = lim
n→∞
p n
[n].
(234)
Similarly, the K−exponential function is defined by Exp(¯ v , z) := hK(v)|K(z)i
(235)
∂Exp(¯ v, z) = v¯Exp(¯ v, z).
(236)
and satisfies the equation
Setting v = 1, one has Exp(z) := Exp(1, z)
(237)
∂Exp(z) = Exp(z).
(238)
which satisfies the equation
Hence, provided the K−derivative (∂) realization, we can determine the analytic −1 representation, IK ◦ AK ◦ IK , of the C ⋆ − algebra AK . Moreover through the equalities (229) and (231), the coherent states map K : DR −→ H \ {0} generate the coherent states ∞ ∞ X X c p 0 z n |ni, cn z n |ni = K(z) = (239) [n]! n=0 n=0 and the exponential functions (235) and (237) are reduced to Exp(¯ v , z) = and, after normalization (i.e. c0 = 1), to
∞ X c20 (¯ v z)n [n]! n=0
Exp(z) =
∞ X z2 , [n]! n=0
(240)
(241)
respectively. In the next section, we aim at extending the above formalism to a general (p, q)analysis, constructing a (R, p, q)− derivative, where R is a meromorphic function defined on C × C and can be, in particular cases, a rational function. 42
4.2
R(p, q)−basic hypergeometric series related to meromorphic functions
The problem we set here consists in defining the derivative which leads to a generalization of (p, q)−algebras and (p, q)−basic hypergeometric series. Let p and q be two positive real numbers such that 0 < q < p. Consider a meromorphic function R, defined on C × C by R(x, y) =
∞ X
rkl xk y l
(242)
k,l=−L
with an eventual isolated singularity at the zero, where rkl are complex numbers, L ∈ N ∪ {0}, R(pn , q n ) > 0 ∀n ∈ N, and R(1, 1) = 0, and the following linear operators defined on O(DR ) [24, 57, 87] by: Q : ϕ 7−→ Qϕ(z) = ϕ(qz) P : ϕ 7−→ P ϕ(z) = ϕ(pz) ϕ(pz) − ϕ(qz) ∂p,q : ϕ 7−→ ∂p,q ϕ(z) = . z(p − q)
(243)
Then, we define the analytic representation of the annihilation operator A, called the R(p, q)− derivative, by ∂R(p,q) := ∂p,q
p−q p−q R(P, Q) = R(pP, qQ)∂p,q . P −Q pP − qQ
(244)
Using relations (229), (231), and (239), we define the R(p, q)-factors (also called R(p, q)-numbers) by R(pn , q n ), n = 0, 1, 2, · · · from which we deduce the R(p, q)factorials 1 for n = 0 n n (245) R!(p , q ) := R(p, q) · · · R(pn , q n ) for n ≥ 1, and inducing the coefficients cn and the coherent states map K in the form c20 R!(pn , q n ) ∞ X p KR(p,q) (z) =
c2n =
n=0
(246) c0 z n |ni. R!(pn , q n )
(247)
Besides, the relations (240) and (241) can be readily generalized to take the form ExpR(p,q) (¯ v, z) =
∞ X n=0
43
c20 (¯ v z)n R!(pn , q n )
(248)
and ExpR(p,q) (z) =
∞ X n=0
1 zn, n n R!(p , q )
(249)
respectively, with the virtue that c20 = ExpR(p,q) (0) = 1. Unless we say otherwise R will represent the radius of convergence of the series. The equations (236) and (238) remain valid and (249) is a solution of the R(p, q)−difference equation R(P, Q)ExpR(p,q) (z) = zExpR(p,q) (z).
(250)
The following two statements are essential to perform a generalization of the (p, q)-binomial theorem. Lemma 4.3 Let z , z − R(1, 0) p(Q − P )R(pP, qQ) + (pP − qQ)R(1, 0) G(P, Q) = pQR(pP, qQ)
F (z) =
(251)
if R(1, 0) 6= 0, and F (z) = z, G(P, Q) =
qQ − pP pQR(pP, qQ)
if R(1, 0) = 0. Then the exponential function in (249) satisfies q z . ExpR(p,q) (z) = [1 − F (z)G(P, Q)] ExpR(p,q) p
(252)
(253)
Proof 4.2 From definitions (243) and (244), we deduce 1=
Q pP − qQ Q +z ∂R(p,q) . P pQR(pP, qQ) P
(254)
The action of the operator (254) on the exponential function (249) gives q qQ − pP ExpR(p,q) z ExpR(p,q) (z) = 1 − z pQR(pP, qQ) p which corresponds to the case R(1, 0) = 0. Now, if R(1, 0) 6= 0, the identity in (254) can be rewritten 1 Q 1 Q 1−z ∂R(p,q) = −z ∂R(p,q) R(1, 0) P R(1, 0) P 44
p(Q − P )R(pP, qQ) + (pP − qQ)R(1, 0) Q ∂R(p,q) +z pQR(pP, qQ)R(1, 0) P (255)
which, acting on ExpR(p,q) (z), leads, after a short computation, to the result: p(Q − P )R(pP, qQ) + (pP − qQ)R(1, 0) z ExpR(p,q) (z) = 1 − z − R(1, 0) pQR(pP, qQ)R(1, 0) q ×ExpR(p,q) z . p Lemma 4.4 Under assumptions of the Lemma 4.3, k n n−1 Y q q 1−F ExpR(p,q) (z) = z G(P, Q) ExpR(p,q) z . k n p p k=0
(256)
Proof 4.3 : It is immediate by induction using Lemma 4.3.
Finally, the R(p, q)-binomial formula is given through the following statement. Theorem 4.5 Let Lemmas 4.3 and 4.4 be satisfied. Then, a generalization of the (p, q)−binomial theorem can be expressed as: n ∞ ∞ Y X q 1 n (z) = 1−F z G(P, Q) · 1. (257) R!(pn , q n ) pn n=0 n=0 Proof of theorem 4.1 : The result follows from ( 256), tending n to +∞. by
Finally, let us consider the particular case when R is a rational function defined r Rs (x, y)
=
x1+s−r (x − y)(c1p−1 x − d1 q −1 y) . . . (cs p−1 x − ds q −1 y) (−y)1+s−r (a1 p−1 x − b1 q −1 y) . . . (ar p−1 x − br q −1 y)
(258)
where a1 , . . . ar ; b1 , · · · br ; c1 , · · · , cs ; d1 , · · · , ds are complex numbers, r and s being non negative integers. Then, the r Rs (p, q)−numbers and factorials, as well as the coefficients cn of the corresponding exponential function are readily found to be (pn − q n )(c1 pn−1 − d1 q n−1 ) . . . (cs pn−1 − ds q n−1 ) , (259) ( − (q/p)n )1+s−r (a1 pn−1 − b1 q n−1 ) . . . (ar pn−1 − br q n−1 ) ((p, q), (c1, d1 ), · · · , (cs , ds ); (p, q))n [n]!r Rs (p,q) = 1+s−r , n ≥ 1, (260) n ((a1 , b1 ), · · · , (ar , br ); (p, q))n (−1)n (q/p)(2 ) ((a1 , b1 ), · · · , (ar , br ); (p, q))n n 1+s−r (−1)n (q/p)(2 ) (261) c2n = ((p, q), (c1, d1 ), · · · , (cs , ds ); (p, q))n [n]r Rs (p,q) =
45
where ((a1 , b1 ), · · · , (ar , br ); (p, q))n = ((a1 , b1 ); (p, q))n · · · , ((ar , br ); (p, q))n , ((ai , bi ); (p, q))0 = 1 and n−1 Y (ai pk − bi q k ) for n ≥ 1. (262) ((ai , bi ); (p, q))n = k=0
Therefore,
∞ X ((a1 , b1 ), · · · , (ar , br ); (p, q))n z n n 1+s−r (−1)n (q/p)(2 ) Expr Rs (p,q) (z) = ((p, q), (c1, d1 ), · · · , (cs , ds ); (p, q))n n=0
proving that the r Rs (p, q)−exponential function corresponds to the twin-basic hypergeometric series r Φs [57, 24]: Expr Rs (p,q) (z) = r Φs ((a1 , b1 ), · · · , (ar , br ); (c1 , d1 ), · · · , (cs , ds ); (p, q); z).
(263)
For r = s + 1, the coherent states map Kr Rs (p,q) and the exponential function Expr Rs (p,q) are defined on the unit disc D1 while, for r < s + 1, they are defined on the whole complex plane. The exponential function ExpR(p,q) thus appears as a natural generalization of twin-basic (or (p, q)−) hypergeometric series.
4.3
R(p, q)− deformed quantum algebras
In this section, we deal with the study of the R(p, q)− deformed quantum algebra. Relations between the annihilation and creation operators, and the operators Q and P as well as the algebra generated by the meromorphic function R are obtained. Some relevant particular representations recovered in this framework are also investigated. The use of the relation (232) and the (R(p, q)-factors engenders †
AA K(z) =
∞ X n=0
giving
R(pn+1 , q n+1 )cn z n |ni = R(pP, qQ)K(z)
By analogy, one can show that
AA† = R(pP, qQ).
(264)
A† A = R(P, Q).
(265)
If one passes to analytic representation in which Q, P and ∂p,q are given by (243), and A† acts as mutliplication by z, one obtains the relations: AQ = qQA,
AP = pP A, 46
(266)
QA† = qA† Q,
P A† = pA† P.
(267)
It is clear that QP = P Q.
(268)
So, the R(p, q)-deformed quantum algebra is generated by the operators {1, A, A† , Q, P } that verify the relations (264)-(268). Let us determine the analytic representation of the number operator N. Taking into account N|ni = n|ni and (222), one obtains n dz n 1 N|ni (z) = IK |ni (z) = nz n = z (269) IK cn cn dz implying
n dz n −1 . IK ◦ N ◦ IK (z ) = z dz Therefore, ∀f ∈ O(DR ) we infer d −1 IK ◦ N ◦ IK f (z) = z f (z) dz
and
pN f (z) = P f (z) = f (pz) and q N f (z) = Qf (z) = f (qz)
(270)
(271)
(272)
for f ∈ O(DR ). To sum up, the R(p, q)−deformed quantum algebra is then generated by the set of operators {1, A, A† , N} and the commutation relations [N, A] = −A and N, A† = A† (273)
and
Remarks
A, A† = R(pN +1 , q N +1 ) − R(pN , q N ).
(274)
Particulars cases are readily recovered. (i). Odziejewicz generalization [87] is recovered for 0 < q < p = 1. Consider the meromorphic function R defined on C by R(z) =
∞ X
rk z k
(275)
k=−L
which may have an isolated singularity at the zero and such that L ∈ N∪{0}, R(q n ) > 0 for n > 0, R(0) > 0, and R(1) = 0. The following results hold: 47
• The R(q)-derivative is given by ∂R(q) = ∂q
1−q 1−q R(Q) = R(qQ)∂q 1−Q 1 − qQ
where Qϕ(z) = ϕ(qz) and ∂q =
(276)
1−Q . (1−q)z
• The R(q)-factors and R(q)-factorials are given by R(q n ) i.e. ∂R z n = R(q n )z n−1 for n ≥ 0, 1 for n = 0 n R!(q ) = R(q) · · · R(q n ) for n ≥ 1,
(277) (278)
respectively. • The coefficients cn , the coherent states map KR and the exponential functions ExpR are given by c20 , R!(q n ) ∞ X c0 p z n ni, KR (z) = n R!(q ) n=0 ∞ X 1 ExpR (¯ v , z) = (¯ v z)n n R!(q ) n=0
c2n =
ExpR (z) =
∞ X n=0
1 zn, R!(q n )
(279) (280) (281) (282)
respectively. • The latter exponential function is solution of the q−difference equation [R(Q)ExpR ](z) = zExpR (z),
(283)
and satisfies the following generalization of the q−binomial theorem ExpR (z) =
∞ X n=0
∞ Y 1 n z = [1 − F (zq n )G(Q)].1 R!(q n ) n=0
(284)
where z , z − R(0) (Q − 1)R(qQ) + (1 − qQ)R(0) G(Q) = QR(qQ)
F (z) =
48
(285)
for L = 0, and F (z) = z, qQ − 1 G(Q) = QR(qQ)
(286)
for L > 0. • The quantum algebra AKR(q) generated by the set of operators {1, A, A†, Q} satisfies the relations:
AA† = R(qQ), A† A = R(Q),
(287) (288)
from which one obtains r kAk = kA k = sup R(q n ), †
(289)
n∈N
for L = 0, i.e. A and A† are bounded. Due to (287) and (288), A and A† are unbounded if L > 0. Indeed, setting R(z) = R(1, z), (275) readily follows from (242). The derivative (276) is then deduced from (244). Using (245), one gets (277) and (278). Coefficients (279) and coherent states maps (280) are obtained from (246) and (247), while the exponential functions in (248) and (249) are reduced to (281) and (282), respectively. In the other hand, the q−difference equation (283) is directly obtained from (250). In the same way and using Theorem 4.5, we get the generalization of the q−binomial theorem (284). Finally, the relations (287) and (288) between A, A† , and Q simply follow from (264) and (265). (ii). Jagannathan’s generalization [57] The main results summarized as follows are particular cases of Theorem 4.5: • The (p, q)−binomial theorem is: 1 Φ0 ((a, b); −; (p, q); z; )
=
((p, bz); (p, q))∞ . ((p, az); (p, q))∞
(290)
• The exponential functions, denoted by ep,q and Ep,q , are ep,q (z) := 1 Φ0 ((1, 0); −; (p, q); z) = 49
∞ X n=0
pn(n−1)/2 zn ((p, q); (p, q))n
(291)
Ep,q (z) := 1 Φ0 ((0, 1); −; (p, q); −z) = and
∞ X n=0
q n(n−1)/2 zn ((p, q); (p, q))n
ep,q (z)Ep,q (−z) = 1. Indeed, if we consider 1 R0 (x, y)
(292)
(293)
x−y − qb y
(294)
pn − q n , apn−1 − bq n−1
(295)
=
a x p
where 0 < q < p and a, b are complex numbers, then we get n n 1 R0 (p , q ) =
((p, q); (p, q))n , ((a, b); (p, q))n ((a, b); (p, q))n c2n = ((p, q); (p, q))n 1 R0 !(p
n
, qn) =
(296) (297)
and Exp1 R0 (p,q) = =
∞ X (a − b)(ap − bq)...(apn−1 − bq n−1 ) n=0 ∞ X n=0
(p − q)...(pn − q n )
zn
((a, b); (p, q))n n z , ((p, q); (p, q))n
where ((a, b); (p, q))n = (a − b)(ap − bq)...(apn−1 − bq n−1 ), meaning that Exp1 R0 (p,q) = 1 Φ0 ((a, b); −; (p, q); z).
(298)
In the other hand, since 1 R0 (1, 0) = ap , we get F (z) =
a−b az and G(P, Q) = . az − p a
Then, the application of Theorem 4.5 yields ∞ Y z(q/p)n 1− Exp1 R0 (p,q) = (a − b) p − az(q/p)n n=0 ∞ Y p − bz(q/p)n = p − az(q/p)n n=0 ∞ Y ppn − bzq n = . ppn − azq n n=0 50
(299)
Thus, Exp1 R0 (p,q) =
((p, bz); (p, q))∞ . ((p, az); (p, q))∞
(300)
So, (298) and (300) lead to (290). Finally, (291), (292) and (293) are straightforwardly obtained. It is worth noticing that one can also consider the meromorphic functions 1 R0 (x, y)
=
x−y
(301)
1 R0 (x, y)
=
x−y , − yq
(302)
x p
and use Theorem 4.5 to immediately obtain (291), (292) and (293).
4.4
R(p, q)−coherent states
This section aims at proving that the coherent states derived from the coherent states map (247) satisfy the following conditions [4, 69, 70]: (i). normalizability (as any vector of Hilbert space) (ii). continuity in the label z, and (iii). existence of a resolution of identity with a positive definite weight function, implying that the states form an overcomplete set. 4.4.1
Normalizability
The coherent states defined as 2
|ziR(p,q) := ExpR(p,q) (|z| )
∞ −1/2 X n=0
are normalized so that R(p,q) hz|ziR(p,q)
4.4.2
1 p
R!(pn , q n )
z n |ni.
= 1.
(303)
(304)
Continuity in z
The coherent states |ziR(p,q) are continuous in z. Indeed,
|ziR(p,q) − |z ′ iR(p,q) 2 = 1 − 2Re 51
R(p,q) hz|z
′
iR(p,q) + 1,
(305)
where R(p,q) hz|z
′
iR(p,q) = ExpR(p,q) (|z|2 )ExpR(p,q) (|z ′ |2 )
So, |z − z ′ | → 0 4.4.3
−1/2
ExpR(p,q) (¯ z , z ′ ).
2 =⇒ |ziR(p,q) − |z ′ iR(p,q) → 0.
(306)
Resolution of identity
Assume that there exists a positive measure µ on the disc DR for which the resolution of the identity Z |K(z)ihK(z)| dµ(¯ z , z) (307) I= z , z) DR ExpR(p,q) (¯ holds. Proposition 4.6 IK (H) is a subspace of the Hilbert space L2 (DR , Exp 1 dµ). R(p,q) Moreover, for ϕ, ψ ∈ IK (H) Z 1 hϕ|ψi = ϕ(z)ψ(z) dµ(¯ z , z), (308) ExpR(p,q) (¯ z , z) DR and ϕ(z) =
Z
ϕ(v)ExpR,p,q (¯ v , z) DR
1 ExpR(p,q) (¯ v, v)
dµ(¯ v, v).
(309)
Proof 4.4 : Let ϕ ∈ IK (H). There exists ζ ∈ H such that IK (ζ) = ϕ. So, Z 1 hϕ|ϕi = dµ(¯ z , z) ϕ(z)ϕ(z) ExpR(p,q) (¯ z , z) DR Z 1 IK (ζ)(z)IK (ζ)(z) = dµ(¯ z , z) ExpR(p,q) (¯ z , z) DR Z 1 dµ(¯ z , z) = hζ|K(z)ihK(z)|ζi ExpR(p,q) (¯ z , z) DR Z |K(z)ihK(z)| = hζ| dµ(¯ z , z)|ζi z , z) DR ExpR(p,q) (¯ = hζ|ζi = kζk2 < ∞.
Therefore, ϕ ∈ L2 (DR , Exp 1 dµ). R(p,q) Thus, since IK (H) is a vector space as subspace of O(DR ), we conclude that IK (H) is a subspace of the Hilbert space L2 (DR , Exp 1 dµ). R(p,q)
52
Hence, the scalar product (308) is well defined. Moreover, for ϕ ∈ IK (H) and z ∈ DR , we get −1 ϕ(z) = IK ◦ IK ◦ ϕ (z) = hIK −1 ◦ ϕ|K(z)i Z |K(v)ihK(v)| −1 dµ(¯ v, v)|K(z)i = hIK ◦ ϕ| v , v) DR ExpR(p,q) (¯ Z hIK −1 ◦ ϕ|K(v)ihK(v)|K(z)i dµ(¯ v, v) = ExpR(p,q) (¯ v , v) DR Z ExpR(p,q) (¯ v , z) = ϕ(v) dµ(¯ v, v). ExpR(p,q) (¯ v, v) DR Proposition 4.7 For n, m ∈ N ∪ {0} Z z¯n z m dµ(¯ z , z) = R!(pn , q n ) δm,n . (310) Exp (¯ z , z) DR R(p,q) √ Passing to polar coordinate z = xeiϕ , we get Z R2 dν(x) R!(pn , q n ) xn = , for n = 0, 1, 2 · · · . (311) ExpR(p,q) (x) 2π 0 p Proof 4.5 : For n ∈ N ∪ {0}, z n ∈ IK (H), since z n = IK ( R!(pn , q n )|ni)(z). So, Z z¯n z m n m dµ(¯ z , z) hz |z i = z , z) DR ExpR(p,q) (¯ p p Z IK ( R!(pn , q n )|ni)(z)IK ( [m]!R(p,q) |mi)(z) = dµ(¯ z , z) ExpR(p,q) (¯ z , z) DR Z q p hK(z)|nihm|K(z)i n n dµ(¯ z , z) R!(p , q ) [m]!R,p,q = ExpR(p,q) (¯ z , z) DR Z q p |K(z)ihK(z)| n n dµ(¯ z , z)|ni R!(p , q ) [m]!R(p,q) hm| = z , z) DR ExpR(p,q) (¯ q p = R!(pn , q n ) [m]!R(p,q) hm|ni √ iϕ xe , 0 ≤ x ≤ R2 and 0 ≤ ϕ ≤ 2π, we from which (310) follows. Setting z = √ −iϕ √ iϕ get dµ(¯ z , z) = dµ( xe , xe ) = dν(x)dϕ and ExpR(p,q) (¯ z , z) = ExpR,p,q (x). Then, from (310) and taking m = n we have Z 2π Z R2 Z R2 dν(x) dν(x) n n n R!(p , q ) = dϕ x = 2π . xn ExpR(p,q) (x) ExpR(p,q) (x) 0 0 0 53
Dividing the left and right sides of the above equalities by 2π we obtain (311). As in [87] the quantities (311) may also be treated as defining properties of the measure ν. This is the famous moment problem of finding ν from the knowledge of the fixed moments R!(pn , q n ), n = 0, 1, · · ·. It may be interesting to formulate the moment problem in a way more adequate is replaced for K−analysis. So, the integration with respect to the measure Expdν R(p,q) by the K− integration Ixn :=
1
xn+1 ,
R(pn+1 , q n+1)
(312)
with some unknown analytic weight function σ such that its Taylor expansion σ(x) =
∞ X
ak xk
(313)
k=0
has R as its convergence radius. The K− integration is just the right inverse of the K− differentiation, i.e. I.∂R(p,q) = I. Then, instead of looking for a measure dν which satisfies (311), the problem is to find an analytic function σ which satisfies the moment conditions Ixn σ(x)|x=R2 = i.e.
∞ X k=0
R!(pn , q n ) 2π
for n = 0, 1, 2, · · · ,
R2(n+k+1) R!(pn , q n ) a = k R(pn+k+1 , q n+k+1) 2π
for n = 0, 1, 2, · · · .
(314)
(315)
Remarks Special cases can be recovered: (i). The (p, q)− algebra of Chakrabarty and Jagannathan [24] AA† − qA† A = q −N [N, A] = −A affords the associated coherent states
† † N AA −† qA A† = p N, A = A
∞ −1/2 X zn p |ni |zi = Np,q (|z|2 ) [n] ! p,q n=0
54
(316)
(317)
where
∞ X zn , Np,q (z) = [n]p,q ! n=0
(318)
as well as the (p, q)−number and the (p, q)−factorial given by [n]p,q =
p−n − q n p−1 − q
(319)
and [n]p,q ! = [1]p,q [2]p,q ...[n]p,q , respectively, taking R(x, y) =
(320)
1−xy . (p−1 −q)x
Indeed, one obtains relations in (316) through relations (273) and (274, whereas coherent states (317) become particular cases of coherent states (303) with the function N in (318) as the analog of the exponential function (249); the (p, q)−number and (p, q)−factorial are deduced from the R(p, q)factors and (245), respectively. (ii). (p, q)−generalization of q−Quesne algebra [52, 95]: p−1 AA† − A† A = q −N −1 [N, A] = −A generates the associated coherent states
2
|zi = Np,q (|z| ) where
† † N +1 qAA †− A A† = p N, A = A
∞ −1/2 X n=0
zn q |ni [n]Q ! p,q
∞ X zn Np,q (z) = Q n=0 [n]p,q !
(321)
(322)
(323)
while the (p, q)−number is given by
[n]Q p,q = setting R(x, y) =
p−n − q n , p−1 − q
(324)
xy−1 . (q−p−1 )y
(iii). (p, q; µ, ν, f )−deformed states of Hounkonnou and Ngompe [52]: ∞ X zn 2 −1/2 q |ziµ,ν = N (|z| ) |ni p,q p,q,f µ,ν [n]p,q,f ! n=0
55
(325)
where µ,ν Np,q,f (z)
=
with the (p, q)−number given by
∞ X
zn , [n]µ,ν p,q,f !
(326)
q νn pn − q −n , pµn q − p−1
(327)
n=0
[n]µ,ν p,q,f = f (p, q)
such that 0 < pq < 1, pµ < q ν−1 , p > 1, and f a well behaved real and non-negative function of deformation parameters p and q, satisfying lim
(p,q)→(1,1)
f (p, q) = 1,
(328)
becomes a particular case in the generalization provided in this work, setting νn n −y −n R(x, y) = f (p, q) xyµn xq−p −1 , f being meromorphic. From the above mentioned (p, q; µ, ν, f )−deformed states, other deformed states known in the litterature can be easily recovered. See [52] for more details.
4.5
R(p, q)-trigonometric and hyperbolic functions
From the expression (249) of the exponential function, we obtain ExpR(p,q) (iz) =
∞ X n=0
(iz)n R!(pn , q n )
∞ ∞ X X (−1)n z 2n (−1)n z 2n+1 = + i R!(p2n , q 2n ) R!(p2n+1 , q 2n+1 ) n=0 n=0
(329)
and ExpR(p,q) (−iz) =
∞ X (−iz)n R!(pn , q n ) n=0
∞ ∞ X X (−1)n z 2n (−1)n z 2n+1 = − i . 2n , q 2n ) 2n+1 , q 2n+1 ) R!(p R!(p n=0 n=0
(330)
We then define the R(p, q)−cosine, sinus, hyperbolic cosine and sinus functions by ∞ X (−1)n z 2n , cosR(p,q) (z) = 2n , q 2n ) R!(p n=0
56
(331)
sinR(p,q) (z) =
∞ X n=0
coshR(p,q) (z) =
(−1)n z 2n+1 , R!(p2n+1 , q 2n+1 )
∞ X
z 2n R!(p2n , q 2n )
(333)
z 2n+1 , R!(p2n+1 , q 2n+1 )
(334)
n=0
and sinhR(p,q) (z) =
∞ X n=0
(332)
respectively. It is readily checked that 1 ExpR(p,q) (iz) + ExpR(p,q) (−iz) , cosR(p,q) (z) = 2 1 i sinR(p,q) (z) = ExpR(p,q) (iz) − ExpR(p,q) (−iz) , 2 ExpR(p,q) (iz) = cosR(p,q) (z) + i sinR(p,q) (z).
(335) (336) (337)
In particular, the Euler formula is expressed as follows: ExpR(p,q) (iθ) = cosR(p,q) (θ) + i sinR(p,q) (θ). Besides, there follow the relations 1 ExpR(p,q) (z) + ExpR(p,q) (−z) , coshR(p,q) (z) = 2 1 sinhR(p,q) (iz) = ExpR(p,q) (z) − ExpR(p,q) (−z) , 2 ExpR(p,q) (z) = coshR(p,q) (z) + sinhR(p,q) (z).
(338)
(339) (340) (341)
The derivatives are immediately expressible from their definition: ∂R(p,q) sinhR(p,q) (az) = a coshR(p,q) (az),
(342)
∂R(p,q) coshR(p,q) (az) = −a sinhR(p,q) (az),
(343)
∂R(p,q) sinR(p,q) (az) = a cosR(p,q) (az) and ∂R(p,q) cosR(p,q) (az) = −a sinR(p,q) (az),
Therefore, the R(p, q)−oscillator equation
a ∈ C.
2 ∂R(p,q) f (z) + ω 2 f (z) = 0
(344) (345)
(346)
can be solved to give the solution f (z) = C1 cosR(p,q) (ωz) + C2 sinR(p,q) (ωz). 57
(347)
4.6
Modified (p, q)-Bessel functions
The (p, q)−analogues of the q−Bessel functions [55, 56] Js(1) (z; q) Js(2) (z; q)
∞ (q s+1 ; q)∞ X (−1)n ( z2 )2n+s = (q; q)∞ n=0 (q, q s+1 ; q)n
(348)
∞ (q s+1 ; q)∞ X q n(n−1) (−1)n ( 2z )2n+s q n(s+1) = (q; q)∞ n=0 (q, q s+1; q)n
(349)
can be defined by Js(1) (z|p, q) Js(2) (z|p, q)
= B(s + 1|p, q) = B(s + 1|p, q)
∞ X
n=0 ∞ X n=0
pn(n−1) (−1)n ( 2z )2n+s ((p, q), (ps+1, q s+1 ); (p, q))n q n(n−1) (−1)n ( z2 )2n+s ( pq )n(s+1) ((p, q), (ps+1, q s+1 ); (p, q))n
(350) ,
(351)
where 0 < q < p < 1, z, s ∈ C with 0 < |z| < 1, and B(s|p, q) =
((ps , q s ); (p, q))∞ . ((p, q); (p, q))∞
(352)
Remark 4.8 (k)
(k)
• Js (z; pq ) 6= Js (z|p, q) for k = 1, 2 since: pns (( pq )s ; q/p)n ((ps , q s ); (p, q))n = ((p, q); (p, q))n ( pq ; pq )n ∞ ∞ X (−1)n (z/2)2n+s X pn(n−1) (−1)n (z/2)2n+s = 6 ( pq , ( pq )s+1 ; pq )n ((p, q), (ps+1, q s+1); (p, q))n n=0 n=0 ∞ X ( pq )n(n−1)+s+1 (−1)n ( z2 )2n+s n=0
• For p = 1,
( pq , ( pq )s+1 ; pq )n
6=
∞ X n=0
Js(k) (z|p, q) = Js(k) (z; q)
q n(n−1) (−1)n ( 2z )2n+s . ((p, q), (q s+1, q s+1); (p, q))n
for k = 1, 2.
(353)
So, the (p, q)−Bessel functions generalize the q−Bessel functions [74]. (1)
Proposition 4.9 : The following relation between Js
(2)
and Js
holds:
2
Js(2) (z|p, q)
((ps+2, z4 ); (p, q))∞ (1) = J (z|p, q). ((ps+2 , 0); (p, q))∞ s 58
(354)
Proof 4.6 ∞ X n=0
2
∞ X (−1)n ( 4pzs+2 )n pn(n−1) (−1)n ( 2z )2n = ((p, q), (ps+1, q s+1); (p, q))n ( q , ( pq )s+1 ; qp )n n=0 p ! s+1 q z2 q = 2 φ1 0, 0; ; ; − s+2 p p 4p ! s+1 1 q z 2 q s+1 q = z 2 q 0 φ1 −; ; , − 2s+3 p p 4p ( 4ps+2 ; p )∞ ! ∞ 2 s+1 X (−1)n ( z4pq2s+3 )n [(−1)n ( pq )n(n−1)/2 ]2 1 = lim 2 n→∞ ( pq , ( pq )s+1 ; pq )n ( 4pzs+2 ; pq )n n=0 s+1
∞ n n(n−1) z 2n q ( 2 ) ( ps+1 )n (ps+2 , 0); (p, q)∞ X (−1) q = 2 (ps+2 , z4 ); (p, q)∞ n=0 ((p, q), (ps+1, q s+1); (p, q))n
using the following Heine’s transformation [74]: 1 0 φ1 (−; c, z; q, cz). 2 φ1 (a, b; c; q, z) = (z; q)∞ Thus s+1 ∞ X (−1)n q n(n−1) ( z2 )2n ( pq s+1 )n = ((p, q), (ps+1, q s+1); (p, q))n n=0
(355)
∞
2
(ps+2 , z4 ); (p, q)∞ X pn(n−1) (−1)n ( 2z )2n . (356) (ps+2 , 0); (p, q)∞ n=0 ((p, q), (ps+1, q s+1); (p, q))n
Multiplying both sides of this equality by B(s + 1|p, q)( 2z )2 leads to (354).
Proposition 4.10 The following three-term recursion relation: s 1 s 1 s s z 2 (1) +1 12 +1 12 2 2 2 2 2 2 (p P − q Q )(p P − q Q ) + Js (z|p, q) = 4 h i s+3 s 1 s 1 z s+3 (1) (1) (p 2 − q 2 )(p 2 +1 P 2 − q 2 +1 Q 2 ) Js+1 (z|p, q) + Js−1 (z|p, q) (357) 2 (1)
is satisfied for the (p, q)−Bessel function Js (z|p, q).
Proof 4.7 : We have s 1 s 1 2 2 2 2 p P − q Q Js(1) (z|p, q)
∞ X pn(n−1) (−1)n ( z2 )2n+s (ps+n − q s+n ) z = B(s + 1|p, q) 2 ((p, q), (ps+1, q s+1); (p, q))n n=0 z z (1) = Js−1 (z|p, q) (358) 2 2
59
and s 1 s 1 −1 1 Js(1) (z|p, q) = B(s + 1|p, q) p 2 +1 P 2 − q 2 +1 Q 2 2 ∞ n(n−1) X p (−1)n ( 2z )2n+s × ((p, q), (ps+1, q s+1 ); (p, q))n(ps+n+1 − q s+n+1)−1 n=0 z (1) = Js+1 (z|p, q). 2 Adding (358) and (359), we obtain (357).
5
(359)
R(p, q)-calculus: differentiation, integration and Hopf algebras
In this section we build a framework for R(p, q)-deformed calculus, which provides a method of computation for a deformed R(p, q)-derivative, generalizing known deformed derivatives of analytic function defined on a complex disc as particular cases corresponding to conveniently chosen meromorphic functions. Under prescribed conditions, we define the R(p, q)-derivative. The main result resides in the proof that R(p, q)-algebra is a Hopf algebra. Relevant examples are also given.
5.1
R(p, q)−factors and their associated quantum algebras
In the previous chapter (see also [15, 18, 51]) we have built the R(p, q)−factors which are a generalization of Heine q−factors (also called Heine q−number in physics literature) 1 − qn , [n]q = 1−q
n = 0, 1, 2, · · ·
(360)
and Jagannathan-Srinivasa (p, q)−factors [57] [n]p,q =
pn − q n , p−q
n = 0, 1, 2, · · ·
(361)
as follows. Let p and q be two positive real numbers such that 0 < q < p ≤ 1. Consider, as in the previous chapter, a meromorphic function R, defined on C × C by ∞ X rkl xk y l (362) R(x, y) = k,l=−L
60
with an eventual isolated singularity at the zero, where rkl are complex numbers, L ∈ N ∪ {0}, R(pn , q n ) > 0 ∀n ∈ N, and R(1, 1) = 0 by definition. Then, the R(p, q)-factors denoted by R(pn , q n ), n = 0, 1, 2, · · · are used to deduce the R(p, q)−factorial 1 for n = 0 n n R!(p , q ) = (363) R(p, q) · · · R(pn , q n ) for n ≥ 1, the R(p, q)−binomial coefficient R!(pm , q m ) m = , n R(p,q) R!(pn , q n )R!(pm−n , q m−n )
m, n = 0, 1, 2, · · · ,
m ≥ n, (364)
and the R(p, q)-exponential function ExpR(p,q) (z) =
∞ X n=0
1 zn. R!(pn , q n )
(365)
Denote by DR ={z ∈ C : |z| < R} a complex disc and by O(DR ) the set of holomorphic functions defined on DR , where R is the radius of convergence of the series (365). We then define the following linear operators on O(DR ) by (see [15, 51] and references therein): Q : ϕ 7−→ Qϕ(z) = ϕ(qz), P : ϕ 7−→ P ϕ(z) = ϕ(pz), ϕ(pz) − ϕ(qz) , ∂p,q : ϕ 7−→ ∂p,q ϕ(z) = z(p − q)
(366)
ϕ ∈ O(DR ), 0 < q < p ≤ 1, and the R(p, q)-derivative by ∂R(p,q) := ∂p,q
p−q p−q R(P, Q) = R(pP, qQ)∂p,q . P −Q pP − qQ
(367)
Note that the R(p, q)-exponential function is invariant under the action of the R(p, q)-derivative since 0 for n = 0 n ∂R(p,q) z = (368) R(pn , q n )z n−1 for n ≥ 1. In [51], we also studied the R(p, q)−deformed quantum algebra generated by the set of operators {1, A, A† , N} and the commutation relations [N, A] = −A and N, A† = A† (369) 61
with AA† = R(pN +1 , q N +1 ),
This algebra is defined on O(DR ) as: A† := z, where ∂z :=
∂ ∂z
and A† A = R(pN , q N ).
A := ∂R(p,q) ,
N := z∂z ,
(370)
(371)
is the usual derivative on C. Therefore, the following holds:
Proposition 5.1 P = pz∂z ,
Q = q z∂z
(372)
and the algebra AR(p,q) generated by {1, z, z∂z , ∂R(p,q) } satisfies the relations: z ∂R(p,q) = R(P, Q), ∂R(p,q) z = R(pP, qQ), [z∂z , z] = z, [z∂z , ∂R(p,q) ] = −∂R(p,q) .
(373)
Proposition 5.2 If there exist two functions Ψ1 and Ψ2 : C × C −→ C such that Ψ (p, q) > 0 for i = 1, 2 (374) i n+1 n n = Ψk1 (p, q) + Ψ2n+1−k (p, q) ,(375) k k k − 1 R(p,q) R(p,q) R(p,q) ba = Ψ1 (p, q)ab, xy = Ψ2 (p, q)yx, and [i, j] = 0 for i∈{a, b}, j∈{x, y} (376) for quantities a, b, x, y, then n X n an−k bk y k xn−k . (ax + by) = k R(p,q) n
(377)
k=0
Proof: By induction over n. Indeed, the equality (377) holds for n = 1 since 1 1 1 1 0 0 2 (ax + by) = ax + by = aby x + a0 b1 y 1 x0 0 R(p,q) 1 R(p,q) 1 X 1 a1−k bk y k x1−k . = k R(p,q) k=0
Suppose that the equality (377) holds for n ≤ m, this means in particular for n = m, m X m m am−k bk y k xm−k , (378) (ax + by) = k R(p,q) k=0
62
and let us prove that it remains valid for n = m + 1. Indeed, (ax + by)m+1 = (ax + by)m (ax + by) m X m am−k bk y k xm−k (ax + by) = k R(p,q) k=0 m m X X m m m−k k k m−k am−k bk y k xm−k by a b y x ax + = k R(p,q) k R(p,q) k=0 k=0 m X m Ψk (p, q)am+1−k bk y k xm+1−k = k R(p,q) 1 k=0 m X m Ψm−k (p, q)am−k bk+1 y k+1xm−k + k R(p,q) 2 k=0 m X m m+1 m+1 Ψk (p, q)am+1−k bk y k xm+1−k =a x + k R(p,q) 1 k=1 m−1 X m Ψm−k (p, q)am−k bk+1 y k+1xm−k + bm+1 y m+1 + k R(p,q) 2 k=0 m X m m+1 m+1 k =a x + Ψ1 (p, q) am+1−k bk y k xm+1−k k R(p,q) k=1 m X m m+1−k am+1−k bk y k xm+1−k + bm+1 y m+1 Ψ2 (p, q) + k − 1 R(p,q) k=1 m X m k m+1 m+1 m+1 m+1 Ψ1 (p, q) =a x +b y + k R(p,q) k=1 ! m m+1−k + Ψ2 (p, q) am+1−k bk y k xm+1−k k − 1 R(p,q) m X m m+1 m+1 m+1 m+1 am+1−k bk y k xm+1−k . =a x +b y + k R(p,q) k=1
5.2 5.2.1
✷
R(p, q)−differential and integration calculi Differential calculus
We define a linear operator dR(p,q) on AR(p,q) by dR(p,q) = (dz)∂R(p,q) . It follows that dR(p,q) 1 = 0,
dR(p,q) z = (dz)R(p, q), 63
2 dR(p,q) ∂R(p,q) = (dz)∂R(p,q)
(379)
dR(p,q) (z∂z ) = (dz)(z∂z + 1)∂R(p,q)
and d2R(p,q) = 0.
(380)
Hence, the set of zero-forms Ω0 (AR(p,q) ) is naturally AR(p,q), while a one form ω, element of Ω1 (AR(p,q) ), is given by ω = (dz)ω0 (z, z∂z , ∂R(p,q) ), (381) P i j k where ω0 (z, z∂z , ∂R(p,q) ) = ∞ i,j,k=0 αijk (z) (z∂z ) (∂R(p,q) ) with αijk belonging to C. Therefore, dω = 0 for ω ∈ Ω1 (AR(p,q) ). Proposition 5.3 For a nonnegative integer n, the following equalities hold: dR(p,q) (z n ) = (dz)R(pn , q n )z n−1 , dR(p,q) (z∂z )n = (dz)(z∂z + 1)n ∂R(p,q) , n n+1 dR(p,q) (∂R(p,q) ) = (dz)∂R(p,q) .
(382)
Moreover if f ∈ O(DR ) then dR(p,q) f (z) = (dz)∂R(p,q) f (z).
(383)
Proof: The equalities in (382) follow from the definition of the R(p, q)− derivative (367), the commutation relations (373) and the definition of the differential (379). Then, (383) follows by definition (367). ✷ Proposition 5.4 The differential dR(p,q) obeys the two following equivalent Leibniz rules dR(p,q) (f g) = (dz)
p−q R(pP, qQ) {∂p,q (f ))(P g) + (Qf )(∂p,q (g))} , pP − qQ
(384)
dR(p,q) (f g) = (dz)
p−q R(pP, qQ) {(∂p,q (f ))(Qg) + (P f )(∂p,q (g))} . pP − qQ
(385)
for f, g ∈ O(DR ).
Proof: This follows from ∂p,q (f g) = (∂p,q (f ))(Qg) + (P f )(∂p,q (f ) = (∂p,q (f )(P g) + (Qf )(∂p,q (g)). ✷
64
5.2.2
R(p, q)-integration
We define the operator IR(p,q) over O(DR ) as the inverse image of the R(p, q)derivative. For elements z n of the basis of O(DR ), IR(p,q) acts as follows: IR(p,q) z n := ∂R(p,q)
−1
zn =
1 R(pn+1 , q n+1 )
z n+1 + c,
(386)
where n ≥ 0 and c is an integration constant. Hence, if f ∈ O(DR ) then
IR(p,q) ∂R(p,q) f (z) = f (z) + c and ∂R(p,q) IR(p,q) f (z) = f (z) + c′ , (387) where c and c′ are integration constants. Provided that R(P, Q) is invertible, one can define the R(p, q)-integration by the following formula IR(p,q) = R−1 (P, Q) z, with c = c′ = 0. One can also derive the definite integrals: Z β f (z)dR(p,q) z = IR(p,q) (β) − IR(p,q) f (α), α, β ∈ DR ; α Z +∞ Z pn /qn f (z)dR(p,q) z = lim f (z)dR(p,q) z; n→∞ α α Z +∞ Z pn /qn f (z)dR(p,q) z = lim f (z)dR(p,q) z. −∞
n→∞
(388)
(389) (390) (391)
−pn /q n
Moreover, the Eqs. (384) and (385) lead to the following formulae:
and
IR(p,q) ∂R(p,q) (f (z)g(z)) = f (z)g(z) + c (392) p−q R(pP, qQ)(∂p,q (f ))(P g) = IR(p,q) pP − qQ p−q +IR(p,q) R(pP, qQ) {(Qf )(∂p,q (g))} pP − qQ IR(p,q) ∂R(p,q) (f (z)g(z)) = f (z)g(z) + c (393) p−q R(pP, qQ)(∂p,q (f ))(Qg) = IR(p,q) pP − qQ p−q +IR(p,q) R(pP, qQ) {(P f )(∂p,q (g))} , pP − qQ
respectively. These relation can be viewed as formulae of integration by parts. 65
5.3
R(p, q)-Hopf algebra
The aim of this section is to establish conditions for which the R(p, q)-algebra carries a Hopf algebra structure. This is summarized in a theorem given below. Remind first that the algebra AR(p,q) will be a Hopf algebra if it admits operations of homomorphisms of a co-product ∆, a counit ǫ and an anti-homomorphism of an antipode S [1]: ∆ : AR(p,q)
ǫ : AR(p,q)
S : AR(p,q)
/
AR(p,q) ⊗ AR(p,q) ,
C, / /
∆(Ω1 Ω2 ) = ∆(Ω1 )∆(Ω2 );
ǫ(Ω1 Ω2 ) = ǫ(Ω1 )ǫ(Ω2 );
AR(p,q) ,
(394)
S(Ω1 Ω2 ) = S(Ω2 )S(Ω1 )
satisfying (id ⊗ ∆)∆(Ω) = (∆ ⊗ id)∆(Ω), (id ⊗ ǫ)∆(Ω) = Ω = (ǫ ⊗ id)∆(Ω), m((id ⊗ S)∆(Ω)) = m((S ⊗ id)∆(Ω)) = ǫ(Ω)1,
(395) (396) (397)
for all Ω, Ω1 , Ω2 ∈ AR(p,q) . To prove this it is sufficient to show that these relations are satisfied by the generators governing the considered algebra. See [25] and references therein. Let the Leibniz rule be written as (∂R(p,q) f g)(z) = (∂R,p, f (z))(Ψ(pz∂z , q z∂z )g(z) ˜ z∂z , q z∂z )f (z))∂R(p,q) g(z), +(Ψ(p
(398)
˜ .) are meromorphic functions. Let the for f, g ∈ O(DR ), where Ψ(., .) and Ψ(., coproduct ∆, the counit ǫ, and the antipode S be defined as follows: ˜ α˜1 N , q α˜2 N ) ⊗ A, ∆(A) = αA ⊗ Ψ(pα1 N , q α2 N ) + α ˜ Ψ(p ˜ β˜1 N , q β˜2 N ) ⊗ A† , ∆(A† ) = βA† ⊗ Ψ(pβ1 N , q β2N ) + β˜Ψ(p
∆(N) = N ⊗ 1 + 1 ⊗ N + τ 1 ⊗ 1, ∆(1) = 1 ⊗ 1, ǫ(A) = 0, ǫ(A† ) = 0, ǫ(N) = −τ, ǫ(1) = 1 S(A) = −s1 A, S(A† ) = −˜ s1 A† , S(N) = −N − 2τ 1,
(399) (400) (401) (402) (403) S(1) = 1,(404)
where αi , α ˜ i , βi , β˜i (i = 1, 2) and s1 , s˜1 α, α, ˜ β, β˜ and τ are real constants such that the following equations hold: ˜ −τ α˜1 , q −τ α˜ 2 ) = 1, α ˜ Ψ(p
αΨ(p−τ α1 , q −τ α2 ) = 1, 66
(405)
˜ −τ β˜1 , q −τ β˜2 ) = 1, bΨ(p−τ β1 , q −τ β2 ) = 1, β˜Ψ(p αβ = 1, α ˜ β˜ = 1, ∆(Ψ(pα1 N , q α2 N )) = αΨ(pα1 N , q α2 N ) ⊗ Ψ(pα1 N , q α2 N ), ˜ α˜1 N , q α˜2 N )) = α ˜ α˜1 N , q α˜2 N ) ⊗ Ψ(p ˜ α˜1 N , q α˜2 N ) ∆(Ψ(p ˜ Ψ(p ∆(Ψ(pβ1 N , q β2N )) = βΨ(pβ1N , q β2 N ) ⊗ Ψ(pβ1 N , q β2 N ), ˜ β˜1 N , q β˜2 N ), ˜ β˜1 N , q β˜2 N ) ⊗ Ψ(p ˜ β˜1 N , q β˜2N )) = β˜Ψ(p ∆(Ψ(p ˜ −α˜1 (N +2τ ) , q −α˜2 (N +2τ ) ), s1 αΨ(pα1 (N +1) , q α2 (N +1) ) = α ˜ Ψ(p ˜ α˜1 (N −1) , q α˜2 (N −1) ), αΨ(p−α1 (N +2τ ) , q −α2 (N +2τ ) ) = s1 α ˜ Ψ(p
(406) (407) (408) (409) (410)
˜ β˜1 (N +1) , q β˜2 (N +1) ), βΨ(p−β1(N +2τ ) , q −β2(N +2τ ) ) = s˜1 β˜Ψ(p ˜ β˜1 (N +2τ ) , q β˜2 (N +2τ ) ).. s˜1 βΨ(p−β1 (N −1) , q −β2 (N −1) ) = β˜Ψ(p
(414)
(411) (412) (413) (415)
Then the following main statement is true: Theorem 5.5 The more general deformed algebra generated by {1, N, A, A† } satisfying: [N, A] = −A,
[N, A† ] = A†
[A, A† ]γ = AA† + γA† A,
(416)
where γ is a real constant such that ˜ β˜1 N , q β˜2N ) ⊗ Ψ(pα1 N , q α2 N ) = (417) Ψ(p ˜ ˜ β (N −1) α (N −1) α (N −1) β (N −1) ˜ 1 ) ⊗ Ψ(p 1 ,q 2 ), ,q 2 γ Ψ(p ˜ α˜1 (N +1) , q α˜2 (N +1) ) ⊗ Ψ(pβ1 (N +1) , q β2 (N +1) ) = Ψ(p ˜ α˜1 N , q α˜2 N ) ⊗ Ψ(pβ1 N , q β2 N ), γ Ψ(p
(418)
is a Hopf algebra. Proof: Notice first that AθλN = θλ(N +1) A, θλ(N ) A = Aθλ(N −1) , A† θλ(N ) = A† θλ(N −1) , θλ(N ) A† = A† θλ(N +1) ,
(419)
˜ = α, ˜ where θ = p, q, so that, for λ = α,β and λ ˜ β. AΨ(pλ1 N , q λ2 N ) = Ψ(pλ1 (N +1) , q λ2 (N +1) )A, ˜ λ˜1 (N +1) , q λ˜2 (N +1) )A, ˜ λ˜1 N , q λ˜2 N ) = Ψ(p AΨ(p A† Ψ(pλ1 N , q λ2 N ) = Ψ(pλ1 (N −1) , q λ2 (N −1) )A† , ˜ λ˜1 N , q λ˜2 N ) = Ψ(p ˜ λ˜1 (N −1) , q λ˜2 (N −1) )A† . A† Ψ(p 67
(420)
Let us now prove that the above definitions of coproduct, counit and antipode satisfy the properties (395)-(397) for Ω ∈ {A, A† , N, 1}. Indeed, z For Ω = N and using (401) and (402), we have (id ⊗ ∆)∆(N) = = = = = =
(id ⊗ ∆)(N ⊗ 1 + 1 ⊗ N + τ 1 ⊗ 1) N ⊗ ∆(1) + 1 ⊗ ∆(N) + τ 1 ⊗ ∆(1) N ⊗ 1 ⊗ 1 + 1 ⊗ N ⊗ 1 + 1 ⊗ 1 ⊗ N + 2τ 1 ⊗ 1 ⊗ 1 ∆(N) ⊗ 1 + ∆(1) ⊗ N + τ ∆(1) ⊗ 1 (∆ ⊗ id)(N ⊗ 1 + 1 ⊗ N + τ 1 ⊗ 1) (∆ ⊗ id)∆(N).
So, (395) is satisfied. Also, (id ⊗ ǫ)∆(N) = (id ⊗ ǫ)(N ⊗ 1 + 1 ⊗ N + τ 1 ⊗ 1) = N ⊗ ǫ(1) + 1 ⊗ ǫ(N) + τ 1 ⊗ ǫ(1) = N ⊗ 1 − τ1 ⊗ 1 + τ1 ⊗ 1 = N and (ǫ ⊗ id)∆(N) = (ǫ ⊗ id)(N ⊗ 1 + 1 ⊗ N + τ 1 ⊗ 1) = ǫ(N) ⊗ 1 + ǫ(1) ⊗ N + τ ǫ(1) ⊗ 1 = −τ ⊗ 1 + 1 ⊗ N ⊗ τ ⊗ 1 = N, where we use (401) and the fact that ǫ(N) = −τ . Hence, N satisfies (396). Next, m((id ⊗ S)∆(N)) = = = =
m((id ⊗ S)(N ⊗ 1 + 1 ⊗ N + τ 1 ⊗ 1)) m(N ⊗ S(1) + 1 ⊗ S(N) + τ 1 ⊗ S(1)) m(N ⊗ 1 − 1 ⊗ N − 2τ 1 ⊗ 1 + τ 1 ⊗ 1) −τ m(1 ⊗ 1) = −τ 1 = ǫ(N)1,
and similarly m((S ⊗ id)∆(N)) = −τ m(1 ⊗ 1) = −τ 1 = ǫ(N)1, where we use (401) and the fact that S(N) = −N − 2τ 1. Therefore N satisfies (397). z For Ω = A, we have ˜ α˜1 N , q α˜2 N ) ⊗ A) (id ⊗ ∆)∆(A) = (id ⊗ ∆)(αA ⊗ Ψ(pα1 N , q α2 N ) + α ˜ Ψ(p ˜ α˜1 N , q α˜2 N ) ⊗ ∆(A) = αA ⊗ ∆(Ψ(pα1 N , q α2 N )) + α ˜ Ψ(p = αA ⊗ ∆(Ψ(pα1 N , q α2 N )) 68
=
= = =
˜ α˜1 N , q α˜2 N ) ⊗ A ⊗ Ψ(pα1 N , q α2 N ) +αα ˜ Ψ(p ˜ α˜1 N , q α˜2 N ) ⊗ Ψ(p ˜ α˜1 N , q α˜2 N ) ⊗ A +α ˜ 2 Ψ(p α2 A ⊗ Ψ(pα1 N , q α2 N ) ⊗ Ψ(pα1 N , q α2 N ) ˜ α˜1 N , q α˜2 N ) ⊗ A ⊗ Ψ(pα1 N , q α2 N ) +αα ˜ Ψ(p ˜ α˜1 N , q α˜2 N )) ⊗ A +α∆( ˜ Ψ(p ˜ α˜1 N , q α˜2 N )) ⊗ A α∆(A) ⊗ Ψ(pα1 N , q α2 N ) + α∆( ˜ Ψ(p ˜ α˜1 N , q α˜2 N ) ⊗ A) (∆ ⊗ id)(αA ⊗ Ψ(pα1 N , q α2 N ) + α ˜ Ψ(p (∆ ⊗ id)∆(A),
where we use (399) and the fact that α, α ˜ , αi and α ˜ i (i = 1, 2) satisfy equations (408) and (409). Hence, (395) holds. The property (396) also holds since (ǫ ⊗ id)∆(A) = = = = = =
˜ α˜1 N , q α˜2 N ) ⊗ A) (αǫ ⊗ id)(A ⊗ Ψ(pα1 N , q α2 N ) + α ˜ Ψ(p ˜ α˜1 N , q α˜2 N )) ⊗ A αǫ(A) ⊗ Ψ(pα1 N , q α2 N ) + αǫ( ˜ Ψ(p ˜ −α˜1 τ , q −α˜2 τ )A = A = αΨ(p−α1 τ , q −α2 τ )A α ˜ Ψ(p ˜ α˜1 N , q α˜2 N ) ⊗ ǫ(A) αA ⊗ ǫ(Ψ(pα1 N , q α2 N )) + α ˜ Ψ(p ˜ α˜1 N , q α˜2 N ) ⊗ A) (id ⊗ ǫ)(αA ⊗ Ψ(pα1 N , q α2 N ) + α ˜ Ψ(p (id ⊗ ǫ)∆(A),
where the use of (399), (403) and (405) has been made. A satisfies also property (397) since m((S ⊗ id)∆(A)) = m((S ⊗ id)(αA ⊗ Ψ(pα1 N , q α2 N ) ˜ α˜1 N , q α˜2 N ) ⊗ A) +α ˜ Ψ(p ˜ α˜1 N , q α˜2 N )) ⊗ A) = m(αS(A) ⊗ Ψ(pα1 N , q α2 N ) + αS( ˜ Ψ(p ˜ α˜1 N , q α˜2 N ))A = −s1 αAΨ(pα1 N , q α2 N ) + α ˜ S(Ψ(p h i ˜ −α˜1 (N +2τ ) , q −α˜2 (N +2τ ) ) A = −s1 αΨ(pα1 (N +1) , q α2 (N +1) ) + α ˜ Ψ(p = 0.A h = ǫ(A)1 = A.0 i ˜ α˜1 (N −1) , q α˜2 (N −1) ) = A αΨ(p−α1 (N +2τ ) , q −α2 (N +2τ ) ) − s1 α ˜ Ψ(p ˜ α˜1 N , q α˜2 N )A = αAS(Ψ(pα1 N , q α2 N )) − s1 α ˜ Ψ(p ˜ α˜1 N , q α˜2 N ) ⊗ S(A)) = m(αA ⊗ S(Ψ(pα1 N , q α2 N )) + a ˜Ψ(p ˜ α˜1 N , q α˜2 N ) ⊗ A)) = m((id ⊗ S)(αA ⊗ Ψ(pα1 N , q α2 N ) + α ˜ Ψ(p = m((id ⊗ S)∆(A)),
where we use (399), (404) and the fact that s1 , α, α, ˜ αi and α ˜ i (i = 1, 2) satisfy equations (412) and (413). 69
z For Ω = A† , one can perform the same computations and use (400), (406), (403), (404), (410), (411), (414) and (415) to prove that (395)-(397) also hold. z Computing ∆(A)∆(A† ) and ∆(A† )∆(A) we obtain: ˜ α˜1 N , q α˜2 N ) ⊗ A) ∆(A)∆(A† ) = (αA ⊗ Ψ(pα1 N , q α2 N ) + α ˜ Ψ(p ˜ β˜1 N , q β˜2 N ) ⊗ A† ) ×(βA† ⊗ Ψ(pβ1 N , q β2 N ) + β˜Ψ(p = αβAA† ⊗ Ψ(pα1 N , q α2 N )Ψ(pβ1 N , q β2 N ) ˜ Ψ(p ˜ β˜1 N , q β˜2 N ) ⊗ Ψ(pα1 N , q α2 N )A† +αβA ˜ α˜1 N , q α˜2 N )A† ⊗ AΨ(pβ1 N , q β2N ) +αβ ˜ Ψ(p ˜ α˜1 N , q α˜2 N )Ψ(p ˜ β˜1 N , q β˜2 N ) ⊗ AA† +α ˜ β˜Ψ(p
and ˜ β˜1 N , q β˜2N ) ⊗ A† ) ∆(A† )∆(A) = (βA† ⊗ Ψ(pβ1 N , q β2 N ) + β˜Ψ(p ˜ α˜1 N , q α˜2 N ) ⊗ A) ×(αA ⊗ Ψ(pα1 N , q α2 N ) + α ˜ Ψ(p = αβA† A ⊗ Ψ(pβ1 N , q β2 N )Ψ(pα1 N , q α2 N ) ˜ α˜1 N , q α˜2 N ) ⊗ Ψ(pβ1 N , q β2N )A +αβA ˜ † Ψ(p ˜ β˜1 N , q β˜2N )A ⊗ A† Ψ(pα1 N , q α2 N ) +αβ˜Ψ(p ˜ α˜1 N , q α˜2 N ) ⊗ A† A ˜ β˜1 N , q β˜2N )Ψ(p +α ˜ β˜Ψ(p respectively. Thus, ∆(A)∆(A† ) − γ∆(A† )∆(A) = αβ(AA† − γA† A) ⊗ Ψ(pα1 N , q α2 N )Ψ(pβ1 N , q β2 N ) ˜ α˜1 N , q α˜2 N )Ψ(p ˜ β˜1 N , q β˜2 N ) ⊗ (AA† − γA† A) +α ˜ β˜Ψ(p ˜ Ψ(p ˜ β˜1 N , q β˜2 N ) ⊗ Ψ(pα1 N , q α2 N )A† +αβ(A ˜ β˜1 N , q β˜2N )A ⊗ A† Ψ(pα1 N , q α2 N )) −γ Ψ(p ˜ α˜1 N , q α˜2 N )A† ⊗ AΨ(pβ1 N , q β2 N ) +αβ( ˜ Ψ(p †˜ −γA Ψ(pα˜1 N , q α˜2 N ) ⊗ Ψ(pβ1 N , q β2N )A). Therefore, [A, A† ]γ = AA† − γA† A implies ∆([A, A† ]γ ) = [A, A† ]γ ⊗ Ψ(pα1 N , q α2 N )Ψ(pβ1 N , q β2N ) ˜ α˜1 N , q α˜2 N )Ψ(p ˜ β˜1 N , q β˜2 N ) ⊗ [A, A† ]γ +Ψ(p provided that ˜
˜
˜
˜
˜ β1 N , q β2 N ) ⊗ Ψ(pα1 N , q α2 N )A† = γ Ψ(p ˜ β1 N , q β2 N )A ⊗ A† Ψ(pα1 N , q α2 N ) AΨ(p ˜ α˜1 N , q α˜2 N )A† ⊗ AΨ(pβ1 N , q β2 N ) = γA† Ψ(p ˜ α˜1 N , q α˜2 N ) ⊗ Ψ(pβ1 N , q β2N )A, Ψ(p 70
(421)
αβ = 1,
and α ˜ β˜ =
or ˜ β˜1 N , q β˜2 N ) ⊗ Ψ(pα1 N , q α2 N ) = γ Ψ(p ˜ β˜1 (N −1) , q β˜2 (N −1) ) ⊗ Ψ(pα1 (N −1) , q α2 (N −1) ), Ψ(p ˜ α˜1 (N +1) , q α˜2 (N +1) ) ⊗ Ψ(pβ1 (N +1) , q β2 (N +1) ) = γ Ψ(p ˜ α˜1 N , q α˜2 N ) ⊗ Ψ(pβ1 N , q β2N ), Ψ(p and α ˜ β˜ = 1,
αβ = 1,
which are (418), (419) and (406), respectively.
5.4
✷
Relevant particular cases
Let us now apply the above general formalism to particular deformed algebras, well spred in the literature. 5.4.1
Jagannathan-Srinivasa deformation
, we obtain the Jagannathan-Srinivasa (p, q)- factors A. Taking R(x, y) = x−y p−q and (p, q)-factorials [n]p,q =
pn − q n , p−q
and [n]!p,q =
(
1 for
n=0 for
((p,q);(p,q))n (p−q)n
n ≥ 1,
(422)
respectively. Referring the readers to [57] for details on (p, q)-calculus, let us restrict the present description to some new relevant properties. Proposition 5.6 If n and m are nonnegative integers, then [n]p,q =
n−1 X
pn−1−k q k ,
k=0
[n + m]p,q = = [−m]p,q = [n − m]p,q = = [n]p,q =
q m [n]p,q + pn [m]p,q pm [n]p,q + q n [m]p,q , −q −m p−m [m]p,q , q −m [n]p,q − q −m pn−m [m]p,q p−m [n]p,q − q n−m p−m [m]p,q , [2]p,q [n − 1]p,q − pq[n − 2]p,q . 71
(423)
Proposition 5.7 The (p, q)−binomial coefficients ((p, q); (p, q))n n , 0 ≤ k ≤ n; n ∈ N, = k p,q ((p, q); (p, q))k ((p, q); (p, q))n−k
(424)
where ((p, q); (p, q))m = (p − q)(p2 − q 2 ) · · · (pm − q m ), m ∈ N satisfy the following identities n n (425) = n − k p,q k p,q n n k(n−k) k(n−k) , =p = p n − k q/p k q/p n n n+1 n+1−k k , (426) +q = p k − 1 p,q k p,q k p,q n n n+1 n+1−k k (427) +p = p k − 1 p,q k p,q k p,q n−1 n n −(p − q ) k − 1 p,q with
n k
= q/p
(q/p; q/p)n , (q/p; q/p)k (q/p; q/p)n−k
where (q/p; q/p)n = (1 − q/p)(1 − q 2 /p2 ) · · · (1 − q n /pn ) and the (p, q)-shifted factorial ((a, b); (p, q))n ≡ (a − b)(ap − bq) · · · (apn−1 − bq n−1 ) n X n (−1)k p(n−k)(n−k−1)/2 q k(k−1)/2 an−k bk . = k p,q k=0
Proposition 5.8 If the quantities x, y, a and b are such that xy = qyx, ba = pab, [i, j] = 0 for i ∈ {a, b} and j ∈ {x, y}, and, moreover, p and q commute with each element of the set {a, b, x, y}, then n X n n an−k bk y k xn−k . (428) (ax + by) = k p,q k=0
The latter result is a generalization of noncommutative form of the q-binomial theorem [41], which can be obtained setting a, b and p equal to 1, i.e. n X n n y k xn−k , (429) (x + y) = k q k=0
72
where
n k
= (q; q)n /(q; q)k (q; q)n−k ,
q 2
with (q; q)n = (1 − q)(1 − q ) · · · (1 − q n ). Proof of Proposition 5.8 A proof has been proposed in [57]. Here we provide another one by induction on n. Indeed, the result is true for n = 1. Suppose it remains valid for all n ≤ m and prove that this is also true for n = m + 1: (ax + by)m+1 = (ax + by)m (ax + by) m X m am−k bk y k xm−k (ax + by) = k p,q k=0 m X m pk am+1−k bk y k xm+1−k = k p,q k=0 m X m q m−k am−k bk+1 y k+1xm−k + k p,q k=0 m X m m+1 m+1 pk am+1−k bk y k xm+1−k =a x + k p,q k=1 m−1 X m q m−k am−k bk+1 y k+1xm−k + bm+1 y m+1 + k p,q k=0
= am+1 xm+1 + bm+1 y m+1 ! m X m m am+1−k bk y k xm+1−k q m+1−k pk + k − 1 p,q k p,q k=1 m X m+1 m+1 m+1 am+1−k bk y k xm+1−k =a x + k p,q k=1
=
+bm+1 y m+1 , m+1 X m+1 k=0
k
am+1−k bk y k xm+1−k
p,q
where the use of (426) has been made. Hence the result is true for all n ∈ N. ✷ The R(p, q)−derivative is thus reduced to the (p, q)−derivative [57] 1 ∂p,q = (P − Q), (430) (p − q)z namely, for f ∈ O(DR ),
∂p,q f (z) =
f (pz) − f (qz) . z(p − q) 73
(431)
The associated algebra Ap,q , generated by {1, A, A† , N}, satisfies the relations: A A† − pA† A = q N , A A† − qA† A = pN [N, A† ] = A† [N, A] = −A,
(432) (433)
and its realization on O(DR ), engendered by {1, z, z∂z , ∂p,q }, satisfies the relations z ∂p,q − p ∂p,q z = q z∂z z ∂p,q − q ∂p,q z = pz∂z [z∂z , z] = z [z∂z , ∂p,q ] = −∂p,q .
(434)
Therefore, the differential operator dp,q is then given by dp,q = (dz)
1 (P − Q) (p − q)z
(435)
with the following properties: dp,q 1 = 0,
dp,q z = (dz),
dp,q (z∂z ) = (dz)(z∂z + 1)∂p,q
2 dp,q ∂p,q = (dz)∂p,q
and
d2p,q
(436)
= 0.
The differential of f ∈ O(DR ) is then dp,q f (z) = (dz)
f (pz) − f (qz) (p − q)z
(437)
affording the Leibniz rule f (pz) − f (qz) g(qz) (p − q)z g(pz) − g(qz) +(dz)f (pz) (p − q)z = {dp,q f (z)} · g(qz) + f (pz) · dp,q g(z)
dp,q (f g)(z) = (dz)
(438)
or, equivalently, f (pz) − f (qz) g(pz) (p − q)z g(pz) − g(qz) +(dz)f (qz) (p − q)z = {dp,q f (z)} · g(pz) + f (qz) · dp,q g(z).
dp,q (f g)(z) = (dz)
74
(439)
The (p, q)−integration is obtained from (388) as follows: ∞ X Qν p−q zf (z) = (p − q) zf (z) Ip,q f (z) = P −Q P ν+1 ν=0 ∞ X
= (p − q)z
(440)
f (zq ν /pν+1 )q ν /pν+1 .
ν=0
Setting p = 1, one recovers the q−derivative and q−integral of Jackson [74]. B. Tacking R(x, y) = given by
x−y a x− qb y p
, where a, b ∈ C with a 6= b, the R(p, q)-factors are
R(pn , q n ) = [n]a,b p,q =
pn − q n , apn−1 − bq n−1
The R(p, q)-factorials become ( 1 for n = 0 [n]!a,b ((p,q);(p,q))n p,q = for ((a,b);(p,q))n
n = 0, 1, · · · .
n ≥ 1,
(441)
(442)
The derivative is now given by ∂R(p,q) = ∂p,q
p−q P −Q 1 P −Q = a , b a P − Q pP − qQ z p P − qb Q
(443)
so that for f ∈ O(DR ) we have 1 P −Q f (z) (444) z ap P − qb Q ∞ p X 1 (bp/aq)ν (Q/P )ν f (z) (P − Q) = z aP ν=0 ∞ p X = (bp/aq)ν (Q/P )ν − (Q/P )ν+1 f (z) az ν=0 ∞ p X = (bp/aq)ν f ((q/p)ν z) − f (q/p)ν+1 z . az ν=0
∂R(p,q) f (z) =
Moreover, IR(p,q) =
a P p
75
− qb Q
P −Q
z.
(445)
Applying this to f ∈ O(DR ) we obtain IR(p,q) f (z) =
a P p
− qb Q
zf (z) P −Q X ∞ b 1 a (Q/P )ν zf (z) P− Q = p q P ν=0 X ∞ a b = − (Q/P ) (Q/P )ν zf (z) p q ν=0 ∞ X = (a/p) (Q/P )ν − (b/q) (Q/P )ν+1 zf (z)
(446)
ν=0
=
∞ X
[(a/p)(q/p)ν zf ((q/q)ν z)
ν=0
−(b/q)(q/p)ν+1 zf (q/p)ν+1 ∞ X = (z/p) (q/p)ν af ((q/q)ν z) − bf (q/p)ν+1 . ν=0
Let us display the Hopf algebra structure of Jagannathan-Srinivasa algebra according to the theorem 5.5. To this end notice first that the Leibniz rule of the derivative is given by ∂p,q (f g)(z) = (∂p,q f (z))pz∂z g(z) + (q z∂z f (z))∂p, qg(z)
(447)
˜ from which we deduce Ψ(x, y) = x and Ψ(x, y) = y. So, α2 = 0, α ˜ 1 = 0, β2 = 0 ˜ and β1 = 0. Equations (405)-(407) yield ˜ α = pα1 τ , α ˜ = q α˜2 τ , β = pβ1 τ , β˜ = q β2 τ τ (α1 + β1 ) = 0 and τ (α ˜ 2 + β˜2 ) = 0,
(448) (449)
while equations (412)-(415) are reduced to s1 pα1 (τ +1) pα1 N = q −α˜2 τ q −α˜2 N , ˜
˜
p−α1 τ p−α1 N = s1 q α˜2 (τ −1) q α˜2 N , ˜
˜
s˜1 pβ1 (τ +1) p−β1 N = q 3β2 τ q β2 N .
p−β1 τ p−β1 N = s˜1 q β2 (τ +1) q β2 N , Therefore, α1 = α ˜ 2 = β1 = β˜2 = 0,
α=α ˜ = β = β˜ = s1 = s˜1 = 1
(450)
and equations (408)-(411) are satisfied and (418)-(419) yield γ = 1. Thus, the coproduct, the counit and the antipode are given by ∆(A) = A ⊗ 1 + 1 ⊗ A,
(451) 76
∆(A† ) = A† ⊗ 1 + 1 ⊗ A† , ∆(N) = N ⊗ 1 + 1 ⊗ N + τ 1 ⊗ 1, ∆(1) = 1 ⊗ 1, ǫ(A) = 0, ǫ(A† ) = 0, ǫ(N) = −τ, ǫ(1) = 1 S(A) = −A, S(A† ) = −A† , S(N) = −N − 2τ 1,
(452) (453) (454) (455) S(1) = 1 (456)
respectively, where τ is a real number, usually set equal to 0. 5.4.2
Chakrabarty and Jagannathan deformation
The algebra of Chakrabarty and Jagannathan [24] is obtained by taking R(x, y) 1−xy = (p−1 . Indeed, the R(p, q)-factors and R(p, q)-factorials are reduced to (p−1 , q)−q)x factors and (p−1 , q)-factorials, namely [n]p−1 ,q = and [n]!p−1 ,q =
(
1 for
p−n − q n , p−1 − q n=0
((p−1 ,q);(p−1 ,q))n (p−1 −q)n
for
n ≥ 1,
(457)
respectively. The properties of this deformation can be readily recovered from the previous section 5.4.1 by replacing the parameter p by p−1 . The R(p, q)−derivative is also reduced to (p−1 , q)−derivative. Indeed, p − q 1 − PQ P − Q (p−1 − q)P 1 = (P −1 − Q) =: ∂p−1 ,q . −1 (p − q)z
∂R(p,q) = ∂p,q
Therefore, for f ∈ O(DR )
∂p−1 ,q f (z) =
f (p−1 z) − f (qz) z(p−1 − q)
(458)
(459)
and the differential of f ∈ O(DR ) is given by dp−1 ,q f (z) = (dz) Computing the Leibniz rule we get
f (p−1 z) − f (qz) . z(p−1 − q)
f (p−1 z) − f (qz) g(qz) z(p−1 − q) g(p−1 z) − g(qz) +(dz)f (p−1 z) z(p−1 − q)
dp−1 ,q (f g)(z) = (dz)
77
(460)
(461)
= {dp−1 ,q f (z)} · g(qz) + f (p−1 z) · dp−1 ,q g(z) or, equivalently, f (p−1 z) − f (qz) −1 g(p z) z(p−1 − q) g(p−1z) − g(qz) +(dz)f (qz) z(p−1 − q) = {dp−1,q f (z)} · g(p−1z) + f (qz) · dp−1 ,q g(z).
(462)
dp−1 ,q (f g)(z) = (dz)
We obtain from (388) the action of the (p−1 , q)-integration on f ∈ O(DR ) as follows: ∞ X p−1 − q −1 Ip−1 ,q f (z) = zf (z) = (p − q) Qν P ν+1 zf (z) (463) P −1 − Q ν=0 = (1 − pq)z
∞ X
f (zq ν pν+1 )(pq)ν .
ν=0
The same Hopf algebra structure as that of Jagannathan-Srinivasa is also obtained for Chakrabarty and Jagannathan deformation. 5.4.3
Generalized q-Quesne deformation
The generalized Quesne algebra [53, 95] is found by taking R(x, y) = Indeed, the (p, q)-Quesne factors and factorials are given by [n]Q p,q = and [n]Q p,q ! =
(
1 for
xy−1 . (q−p−1 )y
pn − q −n , q − p−1 n=0
((p,q −1 );(p,q −1 ))n (q−p−1 )n
for
(464)
n ≥ 1,
respectively. There follow some relevant new properties: Proposition 5.9 If n and m are nonnegative integers, then −m m [−m]Q q [m]Q p,q = −p p,q ,
(465)
−m n Q m Q −n [n + m]Q [n]Q [m]Q p,q = q p,q + p [m]p,q = p [n]p,q + q p,q ,
[n −
m]Q p,q
= q
[n]Q p,q =
m
[n]Q p,q
−p
n−m m
q
[m]Q p,q
=p
−m
[n]Q p,q
+p
−m m−n
q
(466) [m]Q p,q , (467)
−1
q−p Q −1 Q [2]Q p,q [n − 1]p,q − pq [n − 2]p,q . −1 p−q 78
(468)
Proof: We obtain Eqs.(465) and (466) applying the relations p−m − q m = −p−m q m (pm − q −m ) and pn+m − q −n−m = q −m (pn − q −n ) + pn (pm − q −m ) = pm (pn − q −n ) + q −n (pm − q −m ), respectively. Eq.(467) follows combining Eqs.(465) and (466). Note that [n]p,q−1 =
q − p−1 pn − q −n pn − q −n = = [n]Q p,q , p − q −1 p − q −1 q − p−1
n = 1, 2, · · · (469)
which, combined with the following identity [n]p,q−1 = [2]p,q−1[n − 1]p,q−1 − pq −1 [n − 2]p,q−1 , gives Eq.(468).
✷
Proposition 5.10 The (p, q)-Quesne binomial coefficients Q ((p, q −1 ); (p, q −1 ))n n , = k p,q ((p, q −1 ); (p, q −1))k ((p, q −1); (p, q −1 ))n−k where
0 ≤ k ≤ n, n ∈ N, satisfy the following properties: Q Q n n n k(n−k) = =p k 1/qp k p,q n − k p,q n , = pk(n−k) n − k 1/qp Q Q Q n+1 n n k −n−1+k = p +q , k k p,q k − 1 p,q p,q Q Q Q n+1 n n k n+1−k = p +p k k p,q k − 1 p,q p,q Q n−1 −(pn − q −n ) . k − 1 p,q
Proof: This is direct using the Proposition 5.6 and Q n n . = k p,q k p,q−1
(470)
(471)
(472)
(473) ✷
79
Proposition 5.11 If the quantities x, y, a and b are such that xy = q −1 yx, ba = pab, [i, j] = 0 for i ∈ {a, b} and j ∈ {x, y}, and, moreover, p and q commute with each element of the set {a, b, x, y}, then Q n X n an−k bk y k xn−k . (ax + by) = k p,q n
(474)
k=0
Proof: By induction on n. Indeed, the result is true for n = 1. Suppose it remains valid for n ≤ m and prove that this is also true for n = m + 1 : (ax + by)m+1 = (ax + by)m (ax + by) Q m X m am−k bk y k xm−k (ax + by) = k p,q k=0 Q m X m pk am+1−k bk y k xm+1−k = k p,q k=0 Q m X m q −m+k am−k bk+1 y k+1xm−k + k p,q k=0 Q m X m m+1 m+1 pk am+1−k bk y k xm+1−k = a x + k p,q k=1 m−1 X m Q + q −m+k am−k bk+1 y k+1xm−k + bm+1 y m+1 k p,q k=0 Q m X m k m+1 m+1 p = a x + k p,q k=1 Q ! m +q −m−1+k am+1−k bk y k xm+1−k + bm+1 y m+1 k − 1 p,q Q m X m+1 m+1 m+1 am+1−k bk y k xm+1−k = a x + k p,q k=1
=
+bm+1 y m+1 m+1 X m + 1 Q k=0
k
am+1−k bk y k xm+1−k ,
p,q
where the use of (472) has been made. Therefore the result is true for all n ∈ N.✷ The (p, q)−Quesne derivative is given by Q ∂p,q = ∂p,q
p − q PQ − 1 1 = (P − Q−1 ). −1 P − Q (q − p )Q (q − p−1 )z 80
(475)
Therefore, for f ∈ O(DR ) Q ∂p,q f (z) =
f (pz) − f (q −1 z) z(q − p−1 )
(476)
and the differential is given by dQ p,q f (z) = (dz)
f (pz) − f (q −1 z) z(q − p−1 )
(477)
leading to the Leibniz rule f (pz) − f (q −1 z) −1 g(q z) z(q − p−1 ) g(pz) − g(q −1z) +(dz)f (pz) z(q − p−1 ) Q = {dp,q f (z)} · g(q −1 z) + f (pz) · dQ p,q g(z)
dQ p,q (f g)(z) = (dz)
(478)
or, equivalently, f (pz) − f (q −1 z) g(pz) z(q − p−1 ) g(pz) − g(q −1 z) +(dz)f (q −1 z) z(q − p−1 ) −1 Q = {dQ p,q f (z)} · g(pz) + f (q z) · dp,q g(z).
dQ p,q (f g)(z) = (dz)
(479)
The action of the (p, q)−Quesne integration on f ∈ O(DR ) is obtained from (388) as follows: Q Ip,q f (z)
∞ X q − p−1 −1 = zf (z) = (p − q) P ν Qν+1 zf (z) −1 P −Q ν=0
= (p−1 − q)z
∞ X
(480)
f (zpν q ν+1 )pν q ν+1 .
ν=0
The same structure of Hopf algebra as for Jagannathan-Srinivasa is also found for the generalized q−Quesne deformation. 5.4.4
(p, q; µ, ν, h)−deformation
The deformed Hounkonnou-Ngompe [52] algebra is obtained by taking R(x, y) = h(p, q) 81
y ν xy − 1 , xµ (q − p−1 )y
such that 0 < pq < 1, pµ < q ν−1 , p > 1, and h(p, q) is a well behaved real and non-negative function of deformation parameters p and q such that h(p, q) → 1 as (p, q) → (1, 1). Here the R(p, q)−factors become (p, q; µ, ν, h)-factors, namely [n]µ,ν p,q,h = h(p, q)
q νn pn − q −n . pµn q − p−1
(481)
Proposition 5.12 The (p, q; µ, ν, h)−factors verify the following properties, for m, n ∈ N: [−m]µ,ν p,q,h = −
q −2νm+m [m]µ,ν p,q,h , −2µm+m p
(482)
q νn q νm−m µ,ν [n] + [m]µ,ν p,q,h p,q,h pµm pµn−n q νm q νn−n = µm−m [n]µ,ν + [m]µ,ν p,q,h p,q,h , p pµn
[n + m]µ,ν p,q,h =
q ν(n−2m)+m q −νm+m µ,ν [n] − [m]µ,ν p,q,h p,q,h p−µm pµ(n−2m)−n+m q −νm q ν(n−2m)−n+m µ,ν = −µm+m [n]p,q,h − µ(n−2m)+m [m]µ,ν p,q,h , p p
[n − m]µ,ν p,q,h =
[n]µ,ν p,q,h =
(483)
(484)
q − p−1 q −ν 1 q 2ν−1 µ,ν µ,ν [2] [n − 1] − [n − 2]µ,ν p,q,h p,q,h p,q,h . (485) −1 −µ 2ν−1 p − q p h(p, q) p
Proof: This is direct using the Proposition 5.9 and the fact that [n]µ,ν p,q,h = h(p, q)
q νn Q [n] . pµn p,q,h
(486) ✷
Proposition 5.13 The (p, q, µ, ν, h)− binomial coefficients
n k
µ,ν
p,q,h
[n]µ,ν q νk(n−k) p,q,h ! := µ,ν = [k]p,q,h![n − k]µ,ν pµk(n−k) p,q,h !
82
n k
Q
p,q
,
(487)
where 0 ≤ k ≤ n, n k n+1 k n+1 k
n ∈ N, satisfy the following properties: µ,ν µ,ν n = , (488) n − k p,q,h p,q,h µ,ν µ,ν µ,ν q (ν−1)(n+1−k) q νk n n + , (489) = (µ−1)k k p,q,h k − 1 p,q,h p pµ(n+1−k) p,q,h µ,ν µ,ν µ,ν q νk q ν(n+1−k) n n = (µ−1)k + (490) k p,q,h p(µ−1)(n+1−k) k − 1 p,q,h p p,q,h µ,ν νn n−1 n −n q . −(p − q ) µn k − 1 p,q,h p
Proof: This is direct using the Proposition 5.10 and the fact that n [n]µ,ν p,q,h ! = h (p, q)
q n(n+1)/2 Q [n] !, pn(n+1)/2 p,q
(491)
where the use of Eq.(486) has been made. Proposition 5.14 If the quantities x, y, a and b are such that xy =
✷ q ν−1 yx, pµ
qν ab, [i, j] = 0 for i ∈ {a, b} and j ∈ {x, y}, and, moreover, p and q pµ−1 commute with each element of the set {a, b, x, y}, then ba =
µ,ν n X n an−k bk y k xn−k . (ax + by) = k p,q,h n
(492)
k=0
Proof: By induction on n. Indeed, the result is true for n = 1. Suppose it remains valid for n ≤ m and prove that this is also true for n = m + 1 : (ax + by)m+1 = (ax + by)m (ax + by) µ,ν m X m am−k bk y k xm−k (ax + by) = k p,q,h k=0 µ,ν m X q νk m+1−k k k m+1−k m a b y x = k p,q,h q (µ−1)k k=0 µ,ν (ν−1)(m−k) m X q m + am−k bk+1 y k+1xm−k k p,q,h pµ(m−k) k=0 µ,ν m X q νk m+1−k k k m+1−k m m+1 m+1 a b y x = a x + k p,q,h q (µ−1)k k=1
83
+
m−1 X k=0
+bm+1 y m+1 m+1 m+1
= a
x
+
q
µ,ν
q (ν−1)(m−k) m−k k+1 k+1 m−k a b y x µ(m−k) p,q,h p
m k
m+1 m+1
+b
y
(ν−1)(m+1−k)
pµ(m+1−k)
+
m X
k=1 µ,ν
m k−1
q νk q (µ−1)k !
m k
µ,ν
p,q,h
am+1−k bk y k xm+1−k
p,q,h
µ,ν m X m+1 m+1 m+1 = a x + am+1−k bk y k xm+1−k k p,q,h k=1 m+1 m+1
+b
=
m+1 X k=0
y µ,ν m+1 am+1−k bk y k xm+1−k , k p,q,h
where the use of (489) has been made. Therefore, the result is true for all n ∈ N. ✷ The R(p, q)-derivative is then reduced to the (p, q; µ, ν, h)−derivative, given by p−q Qν P Q − 1 h(p, q) µ P −Q P (q − p−1 )Q h(p, q) Qν µ,ν (P − Q−1 ) ≡ ∂p,q,h . = (q − p−1 )z P µ
∂R(p,q) = ∂p,q
(493)
Therefore the (p, q; µ, ν, h)−derivative and the (p, q; µ, ν, h)−differential of f ∈ O(DR ) are given by µ,ν ∂p,q,h
f (zq ν /pµ−1 ) − f (zq ν−1 /pµ ) = h(p, q) z(q − p−1 )
(494)
and dµ,ν p,q,h f (z) = (dz)h(p, q)
f (zq ν /pµ−1 ) − f (zq ν−1 /pµ ) z(q − p−1 )
(495)
respectively, with the Leibniz rule f (zq ν /pµ−1 ) − f (zq ν−1 /pµ ) g(zq ν−1 /pµ ) −1 z(q − p ) g(zq ν /pµ−1 ) − g(zq ν−1 /pµ ) + (dz)f (zq ν /pµ−1 )h(p, q) z(q − p−1 ) µ,ν ν−1 µ = {dp,q,hf (z)} · g(zq /p ) + f (zq ν /pµ−1 ) · dµ,ν p,q,h g(z)
dµ,ν p,q,h (f g)(z) = (dz)h(p, q)
84
which is equivalent to f (zq ν /pµ−1 ) − f (zq ν−1 /pµ ) g(zq ν /pµ−1 ) −1 z(q − p ) g(zq ν /pµ−1 ) − g(zq ν−1 /pµ ) + (dz)f (zq ν−1 /pµ )h(p, q) z(q − p−1 ) µ,ν ν µ−1 = {dp,q,hf (z)} · g(zq /p ) + f (zq ν−1 /pµ ) · dµ,ν p,q,h g(z).
dµ,ν p,q,h (f g)(z) = (dz)h(p, q)
From (388) we obtain the action of the (p, q, µ, ν, h)-integration on f ∈ O(DR ) as follows: q − p−1 P µ /Qν p−1 − q P µ /Qν−1 zf (z) = zf (z) h(p, q) P − Q−1 h(p, q) 1 − P Q ∞ p−1 − q P µ X j j P Q zf (z) = h(p, q) Qν−1 j=0 ∞ p−1 − q X j+µ j+1−ν P Q zf (z) = h(p, q) j=0 ∞ z(p−1 − q) pµ X = f (zpj+µ q j+1−ν ). ν−1 h(p, q) q j=0
µ,ν Ip,q,h f (z) =
(496)
From the derivative Leibniz rule µ,ν ∂p,q,h (f (z)g(z)) =
q νz∂z µ,ν ∂p,q,h f (z) (µ−1)z∂z g(z) p (ν−1)z∂z q µ,ν f (z) ∂p,q,h (g(z)), + µz∂ z p
(497)
˜ one deduces Ψ(x, y) = x−(µ−1) y ν and Ψ(x, y) = y −µxν−1 . Hence, (405)-(407) yield α = p−α1 τ (µ−1) q α2 τ ν , α ˜ = p−α˜1 τ µ q α˜2 τ (ν−1) , (498) ˜ ˜ (499) β = p−β1 τ (µ−1) q β2 τ ν , β˜ = p−β1 τ µ q β2 τ (ν−1) , (α1 + β1 )τ = 0, (α2 + β2 )τ = 0, (α ˜ 1 + β˜1 )τ = 0, (α ˜ 2 + β˜2 )τ = 0. (500) Equations (408)-(411) are of course satisfied and (418)-(419) give ˜
˜
γ = p−[β1 µ+α1 (µ−1)] q β2 (ν−1)+α2 ν and γ = p−[α˜1 µ+β1 (µ−1)] q α˜2 (ν−1)+β2 ν (501) From equations (412)-(415) we deduce s1 = pα1 (µ−1) q −α2 ν , s1 = pβ1 (µ−1) q −β2ν (502) µ−1 ν µ−1 ˜ ν α ˜ 1 = −α1 , α ˜ 2 = −α2 , β˜1 = −β1 , β2 = −β2 (503) µ ν−1 µ ν −1 85
so that α ˜ = α−1 ,
β˜ = β −1 ,
γ = 1.
(504)
Setting κα =
q α2 ν pα1 (µ−1)
and κβ =
q β2 ν pβ1 (µ−1)
, with α1 , α2 , β1 , β1 ∈ R,
(505)
we get ˜ α˜1 N , q α˜2 N ) = κ−N , Ψ(pα1 N , q α2 N ) = κN α , Ψ(p α β1 N β2 N N ˜ β˜1 N β˜2 N −N Ψ(p , q ) = κβ , Ψ(p , q ) = κ .
(506) (507)
β
The remaining conditions are (α1 + β1 )τ = 0 and (α2 + β2 )τ = 0 which hold if τ = 0 or β1 = −α1 and β2 = −α2 . Suppose τ = 0. Then α = α ˜ = β = β˜ = 1. Then, the coproduct, the counit, the antipode are defined as follows: −N ∆(A) = A ⊗ κN α + κα ⊗ A, −N ∆(A† ) = A† ⊗ κN ⊗ A† , β + κβ ∆(N) = N ⊗ 1 + 1 ⊗ N, ∆(1) = 1 ⊗ 1, ǫ(A) = 0, ǫ(A† ) = 0, ǫ(N) = 0, ǫ(1) = 1, † S(A) = −κ−1 S(A† ) = −κ−1 S(N) = −N, α A, β A ,
(508) (509) (510) (511) (512) S(1) = 1. (513)
Suppose now τ 6= 0 that means β1 = −α1 and β2 = −α2 . So, −1 ˜1 = κα . α = κτα , β = κ−τ α , s1 = κα , s
Thus, the coproduct, the counit, the antipode are defined as follows: ∆(A) = A ⊗ καN +τ + κα−N −τ ⊗ A, (514) † † −N −τ N +τ † ∆(A ) = A ⊗ κα + κα ⊗ A , (515) ∆(N) = N ⊗ 1 + 1 ⊗ N + τ 1 ⊗ 1, (516) ∆(1) = 1 ⊗ 1, (517) † ǫ(A) = 0, ǫ(A ) = 0, ǫ(N) = −τ, ǫ(1) = 1, (518) −1 † † S(A) = −κα A, S(A ) = −κα A , S(N) = −N − 2τ 1, S(1) = 1. (519)
86
6
R(p, q)-deformed Rogers-Szeg¨ o polynomials: associated quantum algebras, deformed Hermite polynomials and relevant properties
This section addresses a new characterization of R(p, q)−deformed RogersSzeg¨o polynomials by providing their three-term recursion relation and the associated quantum algebra built with corresponding creation and annihilation operators. The whole construction is performed in a unified way, generalizing all known relevant results which are straightforwardly derived as particular cases. Continuous R(p, q)−deformed Hermite polynomials and their recursion relation are also deduced. Novel relations are provided and discussed. The present investigation aims at giving a new realization of the previous generalized deformed quantum algebras and an explicit definition of the R(p, q)−RogersSzeg¨o polynomials, together with their three-term recursion relation and the deformed difference equation giving rise to the creation and annihilation operators.
6.1
R(p, q)-Rogers-Szeg¨ o polynomials and related quantum algebras
This section aims at providing realizations of (R, p, q)−deformed quantum algebras induced by (R, p, q)−Rogers-Szeg¨o polynomials. We first define the latter and their three-term recursion relation, and then following the procedure elaborated in [40, 58], we prove that every sequence of these polynomials forms a basis for the corresponding deformed quantum algebra. Indeed, Galetti in [40], upon recalling the technique of construction of raising and lowering operators which satisfy an algebra akin to the usual harmonic oscillator algebra, by using the three-term recursion relation and the differentiation expression of Hermite polynomials, has shown that a similar procedure can be carried out to construct a q-deformed harmonic oscillator algebra, with the help of relations controlling the Rogers-Szeg¨o polynomials. Following this author, Jagannathan and Sridhar in [58] adapted the same approach to construct a Bargman-Fock realization of the harmonic oscillator as well as realizations of qand (p, q)- deformed harmonic oscillators based on Rogers-Szeg¨o polynomials. As matter of clarity, this section is stratified as follows. We first develop the synoptic schemes of known different generalizations and then display the formalism of R(p, q)-Rogers-Szeg¨o polynomials.
87
6.1.1
Hermite polynomials and harmonic oscillator approach
The Hermite polynomials are defined as orthogonal polynomials satisfying the three-term recursion relation Hn+1 (z) = 2zHn (z) − 2nHn−1 (z)
(520)
and the differentiation relation d Hn (z) = 2nHn−1 (z). dz
(521)
Inserting Eq. (521) in Eq. (520), one gets d Hn+1 (z) = 2z − Hn (z) dz
(522)
which includes the introduction of a raising operator (see [40] and references therein), defined as a ˆ+ = 2z −
d dz
(523)
such that the set of Hermite polynomials can be generated by the application of this operator to the first polynomial H0 (z) = 1, i.e., Hn (z) = a ˆn+ H0 (z).
(524)
From Eq.(521), one defines the lowering operator a ˆ− as a ˆ− Hn (z) =
1 d Hn (z) = nHn−1 (z). 2 dz
(525)
Furthermore one constructs a number operator in the form n ˆ=a ˆ+ aˆ− .
(526)
One can readily check that these operators satisfy the canonical commutation relations [ˆa− , aˆ+ ] = 1,
[ˆ n, a ˆ− ] = −ˆa− ,
[ˆ n, a ˆ+ ] = a ˆ+ ,
(527)
although the operators a ˆ− and a ˆ+ are not the usual creation and annihilation operators associated with the quantum mechanics harmonic oscillator. Thus, we see that one can obtain raising, lowering and number operators from the two basic relations satisfied by the Hermite polynomials, i. e. the three-term recursion 88
relation and the differentiation relation, respectively, so that they satisfy the well known commutation relations. On the other hand, if one considers the usual Hilbert space spanned by the vectors |ni, generated from the vacuum |0i by the raising operator a ˆ+ , then together with the lowering operator a ˆ− , the following relations hold aˆ− a ˆ+ − a ˆ+ a ˆ+ = 1, h0|0i = 1, |ni = a ˆn+ |0i, aˆ− |0i = 0.
(528)
In particular, the next expressions, established using the previous equations, are in order: a ˆ+ |ni = |n + 1i, a ˆ− |ni = |n − 1i, hm|ni = n!δmn .
(529)
Now, on the other hand, examining the procedure given in [58], the authors considered the sequence of polynomials 1 ψn (z) = √ hn (z), (530) n! where n X n n zk , (531) hn (z) = (1 + z) = k k=0
obeying the relations
√ d ψn (z) = nψn−1 (z), dz √ n + 1ψn+1 (z), (1 + z)ψn (z) = d (1 + z) ψn (z) = nψn (z), dz
(532) (533) (534)
d ((1 + z)ψn (z)) = (n + 1)ψn (z). (535) dz Here equations (533) and (534) are the recursion relation and the differential equation for polynomials ψn (z), respectively. By analogy to the work done by Galleti, Jagannathan and Sridhar proposed the following relations: d d n ˆ = (1 + z) , (536) a ˆ+ = (1 + z), a ˆ− = , dz dz for creation (or raising), annihilation (or lowering) and number operators, respectively, and found that the set {ψn (z) |n = 0, 1, 2, · · ·} forms a basis for the Bargman-Fock realization of the harmonic oscillator (527). 89
6.1.2
Rogers-Szeg¨ o polynomials and q-deformed harmonic oscillator
Here in analogous way as Jagannathan and Sridhar [58], we perform a construction of the creation, annihilation and number operators from the three-term recurrence relation and the q−difference equation founding the Rogers-Szeg¨o polynomials. This procedure a little differs from that used by Galetti [40] to obtain raising, lowering and number operators. The Rogers-Szeg¨o polynomials are defined as n X n z k , n = 0, 1, 2 · · · (537) Hn (z; q) = k q k=0
and satisfy a three-term recursion relation Hn+1 (z; q) = (1 + z)Hn (z; q) − z(1 − q n )Hn−1 (z; q)
(538)
as well as the q-difference equation ∂q Hn (z; q) = [n]q Hn−1 (z; q).
(539)
In the limit case q → 1, the Rogers-Szeg¨o polynomial of degree n (n = 0, 1, 2, · · ·) well converges to n X n hn (z) = zk k k=0
as required. Defining n 1 X n 1 zk , Hn (z) = p ψn (z; q) = p k [n]!q [n]!q k=0 q
n = 0, 1, 2 · · · ,
(540)
one can straightforwardly infer that
q ∂q ψn (z; q) = [n]q ψn−1 (z; q)
(541)
∂qn+1 ψn (z; q) = 0 and ∂qm ψn (z; q) 6= 0 for any m < n + 1.
(542)
with the property that for n = 0, 1, 2, · · ·
It follows from Eqs. (538) and (540) that the polynomials {ψn (z; q) | n = 0, 1, 2, · · ·} satisfy the following three-term recursion relation q q [n + 1]q ψn+1 (z; q) = (1 + z)ψn (z; q) − z(1 − q) [n]q ψn−1 (z; q) (543) 90
and the q−difference equation ((1 + z) − (1 − q)z ∂q ) ψn (z; q) =
q
[n + 1]q ψn+1 (z; q)
(544)
obtained from Eq.(541). Hence, it is natural to formally define the number operator N as Nψn (z; q) = nψn (z; q)
(545)
determined for the creation and annihilation operators expressed as A† = 1 + z − (1 − q)z ∂q
and A = ∂q
(546)
respectively. Indeed, the proofs of the following relations are immediate: Nψn (z; q) = nψn (z; q), q † A ψn (z; q) = [n + 1]q ψn+1 (z; q), q Aψn (z; q) = [n]q ψn−1 (z; q),
(547) (548) (549)
A† Aψn (z; q) = [n]q ψn (z; q) = [N]q ψn (z; q), AA† ψn (z; q) = [n + 1]q ψn (z; q) = [N + 1]q ψn (z; q).
(550) (551)
Therefore, one concludes that the set of polynomials {ψn (z; q) |n = 0, 1, 2, · · ·} provides a basis for a realization of the q-deformed harmonic oscillator algebra given by AA† − qA† A = 1, 6.1.3
[N, A] = −A,
[N, A† ] = A†
(552)
R(p, q)−generalized Rogers-Szeg¨ o polynomials and quantum algebras
We can now supply the general procedure for constructing the recursion relation for the R(p, q)−Rogers-Szeg¨o polynomials and the related R(p, q)-difference equation that allow to define the creation, annihilation and number operators for a given (R, p, q)−deformed quantum algebra. This is summarized as follows. Theorem 6.1 If φi (x, y) (i = 1, 2, 3) are functions satisfying: φi (p, q) 6= 0 for i = 1, 2, 3, φi (P, Q)z k = φki (p, q)z k for z ∈ C, k = 0, 1, 2, · · ·
i = 1, 2
(553) (554)
and if, moreover, the following relation between R(p, q)−binomial coefficients holds n n n+1 n+1−k k + φ2 (p, q) = φ1 (p, q) k − 1 R(p,q) k R(p,q) k R(p,q) 91
n
n
−φ3 (p, q)R(p , q )
n−1 k−1
(555)
R(p,q)
for 1 ≤ k ≤ n, then the R(p, q)−Rogers-Szeg¨o polynomials defined as n X n zk , Hn (z; R(p, q)) = k R(p,q) k=0
n = 0, 1, 2 · · ·
(556)
satisfy the three-term recursion relation Hn+1 (z; R(p, q)) = Hn (φ1 (p, q)z : R(p, q)) +zφn2 (p, q)Hn zφ−1 2 (p, q); R(p, q) −zφ3 (p, q)R(pn , q n )Hn−1 (z; R(p, q))
(557)
and R(p, q)−difference equation ∂R(p,q) Hn (z; R(p, q)) = R(pn , q n )Hn−1 (z; R(p, q)).
(558)
Proof: Multiplying the two sides of the relation (555) by z k and adding for k = 1 to n we get n n X X n+1 n k k z = φ1 (p, q) zk k k R(p,q) R(p,q) k=1 k=1 n X n n+1−k (p, q) zk + φ2 k − 1 R(p,q) k=1 n X n−1 n n z k . (559) −φ3 (p, q)R(p , q ) k − 1 R(p,q) k=1
After a short computation and using the condition (555) we get Eq.(557). Then there immediately results the proof of Eq.(558). Setting 1 ψn (z; R(p, q)) = p Hn (z; R(p, q)), (560) R!(pn , q n )
and using the equations (557) and (558) yield the three-term recursion relation φ1 (P, Q) + zφn2 (p, q)φ−1 (P, Q) − zφ3 (p, q)∂R(p,q) ψn (z; R(p, q)) = 2 p R(pn+1 , q n+1 ) ψn+1 (z; R(p, q)) (561)
and R(p, q)−difference equation
∂R(p,q) ψn (z; R(p, q)) =
p
R(pn , q n ) ψn−1 (z; R(p, q)) 92
(562)
for the polynomials ψn (z; R(p, q)) with the virtue that for n = 0, 1, 2, · · · n+1 m ∂R(p,q) ψn (z; R(p, q)) = 0 and ∂R(p,q) ψn (z; R(p, q)) 6= 0 for m < n + 1.
(563)
Now, formally defining the number operator N as Nψn (z; R(p, q)) = nψn (z; R(p, q)),
(564)
and the raising and lowering operators by −1 A† = φ1 (P, Q) + zφN 2 (p, q)φ2 (P, Q) − zφ3 (p, q)∂R(p,q) and A = ∂R(p,q) ,
(565)
respectively, the set of polynomials {ψn (z; R, p, q) | n = 0, 1, 2, · · ·} provides a basis for a realization of R(p, q)−deformed quantum algebra AR(p,q) satisfying the commutation relations (369). Provided the above formulated theorem, we can now show how the realizations in terms of Rogers-Szeg¨o polynomials can be derived for different known deformations simply by determining the functions φi (i = 1, 2, 3) that satisfy the relations (553)-(555).
6.2
Continuous R(p, q)−Hermite polynomials
We exploit here the peculiar relation established in the theory of q−deformation between Rogers-Szeg¨o polynomials and Hermite polynomials [55, 57, 73, 74] and given by n X n in θ −2i θ ei(n−2k)θ , n = 0, 1, 2, · · · , (566) Hn (cos θ; q) = e Hn (e ; q) = k q k=0
where Hn and Hn stand for the Hermite and Rogers-Szeg¨o polynomials, respectively. Is also of interest the property that all the q−Hermite polynomials can be explicitly recovered from the initial one H0 (cos θ; q) = 1, using the three-term recurrence relation Hn+1 (cos θ; q) = 2 cos θHn (cos θ; q) − (1 − q n )Hn−1 (cos θ; q)
(567)
with H−1 (cos θ; q) = 0. In the same way we define the R(p, q)-Hermite polynomials through the R(p, q)Rogers-Szeg¨o polynomials as Hn (cos θ; R(p, q)) = ein θ Hn (e−2i θ ; R(p, q)), Then the next statement is true. 93
n = 0, 1, 2, · · · .
(568)
Proposition 6.2 Under the hypotheses of the theorem 6.1, the continuous R(p, q)Hermite polynomials satisfy the following three-term recursion relation n
Hn+1 (cos θ; R(p, q)) = ei θ φ12 (p, q)φ1 (P, Q)Hn (cos θ; R(p, q)) n +e−i θ φ22 (p, q)φ−1 2 (P, Q)Hn (cos θ; R(p, q)) n n −φ3 (p, q)R(p , q )Hn−1 (cos θ; R(p, q)).
(569)
Proof: Multiplying the two sides of the three-term recursion relation (557) by ei(n+1)θ , we obtain, for z = e−2iθ , ei(n+1)θ Hn+1 (e−2iθ ; R(p, q)) = ei(n+1)θ Hn φ1 (p, q)e−2iθ ; R(p, q) −2iθ + ei(n−1)θ φn2 (p, q)Hn φ−1 ; R(p, q) 2 (p, q)e − ei(n−1)θ φ3 (p, q)R(pn , q n )Hn−1 e−2iθ ; R(p, q) = eiθ einθ φ1 (P, Q)Hn e−2iθ ; R(p, q) −2iθ + e−iθ φn2 (p, q)einθ φ−1 ; R(p, q) 2 (P, Q)Hn e − φ3 (p, q)R(pn , q n )ei(n−1)θ Hn−1 e−2iθ ; R(p, q) .
The required result follows from the use of the equalities n einθ φ1 (P, Q)Hn e−2iθ ; R(p, q) = φ12 (p, q)φ1 (P, Q)einθ Hn e−2iθ ; R(p, q) , (570)
−n inθ −2iθ Hn e−2iθ ; R(p, q) einθ φ−1 ; R(p, q) = φ2 2 (P, Q)φ−1 2 (P, Q)e 2 (P, Q)Hn e (571)
with φj (P, Q)e−2ikθ = φkj (p, q)e−2ikθ , j = 1, 2, k = 0, 1, 2, · · · .
(572)
6.3
Relevant particular cases
The following pertinent cases deserve to be raised, as their derivation from the previous general theory appeals concrete expressions for the deformed function R(p, q). 6.3.1
R(x, y) =
x−y p−q
In this case, the R(p, q)−factors are simply given by [n]p,q = R(pn , q n ) =
pn − q n , p−q 94
n = 0, 1, 2, · · ·
with the R(p, q)−factorials defined by 1n for n = 0 Y pk − q k ((p, q); (p, q))n [n]!p,q = = p−q (p − q)n k=1
for
n ≥ 1.
(573)
They correspond to the Jagannathan-Srinivasa (p, q)-numbers and (p, q)-factorials [57, 58]. There result the following relevant properties. Proposition 6.3 If n and m are nonnegative integers, then [n]p,q =
n−1 X
pn−1−k q k ,
k=0
[n + m]p,q [−m]p,q [n − m]p,q [n]p,q
= = = =
q m [n]p,q + pn [m]p,q = pm [n]p,q + q n [m]p,q , −q −m p−m [m]p,q , q −m [n]p,q − q −m pn−m [m]p,q = p−m [n]p,q − q n−m p−m [m]p,q , [2]p,q [n − 1]p,q − pq[n − 2]p,q . (574)
Proposition 6.4 The (p, q)−binomial coefficients [n]!p,q ((p, q); (p, q))n n ≡ = , k p,q [k]!p,q [n − k]!p,q ((p, q); (p, q))k ((p, q); (p, q))n−k where 0 ≤ k ≤ n, m ∈ N, satisfy the n = k p,q n+1 = k p,q n+1 = k p,q
(575)
n ∈ N, and ((p, q); (p, q))m = (p − q)(p2 − q 2 ) · · · (pm − q m ), following identities: n n n k(n−k) k(n−k) , (576) =p =p n − k q/p k q/p n − k p,q n n n+1−k k , (577) +q p k − 1 p,q k p,q n−1 n n n n n+1−k k (578) − (p − q ) +p p k − 1 p,q k − 1 p,q k p,q
with
n k
= q/p
(q/p; q/p)n , (q/p; q/p)k (q/p; q/p)n−k
where (q/p; q/p)n = (1 − q/p)(1 − q 2 /p2 ) · · · (1 − q n /pn ) and the (p, q)−shifted factorial ((a, b); (p, q))n ≡ (a − b)(ap − bq) · · · (apn−1 − bq n−1 ) n X n (−1)k p(n−k)(n−k−1)/2 q k(k−1)/2 an−k bk . (579) = k p,q k=0
95
The algebra Ap,q , generated by {1, A, A† , N}, associated with (p, q)− Janagathan - Srinivasa deformation, satisfies the following commutation relations [57, 58]: A A† − pA† A = q N , [N, A† ] = A† ,
A A† − qA† A = pN [N, A] = −A.
(580)
The (p, q)-Rogers-Szeg¨o polynomials studied in [58] appear as a particular case obtained by choosing φ1 (x, y) = φ2 (x, y) = φ(x, y) = x and φ3 (x, y) = x − y. Indeed, φ(p, q) = p 6= 0, φ3 (p, q) = p − q 6= 0, φ(P, Q)z k = φk1 (p, q)z k and Eq.(578) shows that n−1 n n n+1 n+1−k k . − (p − q)[n]p,q +p =p k − 1 p,q k − 1 p,q k p,q k p,q Hence, the hypotheses of the above theorem are satisfied and, therefore, the (p, q)Rogers-Szeg¨o polynomials n X n z k n = 0, 1, 2, · · · (581) Hn (z; p, q) = k p,q k=0
satisfy the three-term recursion relation Hn+1 (z; p, q) = Hn (pz; p, q) + zpn Hn (p−1 z; p, q) −z(pn − q n )Hn−1 (z; p, q)
(582)
and (p, q)−difference equation ∂p,q Hn (z; p, q) = [n]p,q Hn−1 (z; p, q).
(583)
Finally, the set of polynomials 1 ψn (z; p, q) = p Hn (z; p, q), [n]!p,q
n = 0, 1, 2, · · ·
(584)
forms a basis for a realization of the (p, q)−deformed harmonic oscillator and quantum algebra Ap,q satisfying the commutation relations (580) with the number operator N defined as Nψn (z; p, q) = nψn (z; p, q),
(585)
relating the annihilation and creation operators given by A = ∂p,q
and A† = P + zpN P −1 − z(p − q)∂p,q 96
(586)
respectively. Naturally, setting p = 1 one recovers the results of the subsection 6.1.2. The continuous (p, q)−Hermite polynomials have been already suggested in [57] without any further details. In the above achieved generalization, these polynomials are given by Hn (cos θ; p, q) = einθ Hn (e−2iθ ; p, q) n X n ei(n−2k)θ , = k p,q k=0
n = 0, 1, 2, · · · .
(587)
Since for the (p, q)−deformation φ1 (x, y) = φ2 (x, y) = x and φ3 (x, y) = x − y, from the Proposition 6.2 we deduce that the corresponding sequence of continuous (p, q)−polynomials satisfies the three-term recursion relation n
Hn+1 (cos θ; p, q) = p 2 (eiθ P + e−iθ P −1)Hn (cos θ; p, q) −(pn − q n )Hn−1 (cos θ; p, q),
(588)
with P eiθ = p−1/2 eiθ . This relation turns to be the well-known three-term recursion relation (567) for continuous q−Hermite polynomials in the limit p → 1. As matter of illustration, let us explicitly compute the first three polynomials using the relation (588), with H−1 (cos θ; p, q) = 0 and H0 (cos θ; p, q) = 1: H1 (cos θ; p, q) = p0 (eiθP + e−iθP −1)1 − (p0 − q 0 )0 = eiθ + e−iθ = 2 cos θ 1 1 = eiθ + e−iθ . 0 p,q 1 p,q 1
H2 (cos θ; p, q) = p 2 (eiθ P + e−iθ P −1)(eiθ + e−iθ ) − (p − q)1 −2iθ = e2iθ + + p+ q= 2 cos 2θ+ p+ q e 2 2 2 e−iθ . e0iθ + e2iθ + = 2 p,q 1 p,q 0 p,q
H3 (cos θ; p, q) = p(eiθ P + e−iθ P −1 )(e2iθ + e−2iθ + p + q) −(p2 − q 2 )(eiθ + e−iθ ) = e3iθ + e−3iθ + (p2 + pq + q 2 )(eiθ + e−iθ ) = 2 cos3θ + 2(p2 + pq+ q 2 ) cosθ 3 3 3 3 −iθ iθ 3iθ e−3iθ . e + e + e + = 3 p,q 2 p,q 1 p,q 0 p,q 6.3.2
R(x, y) =
1−xy (p−1 −q)x
The R(p, q)-factors and R(p, q)−factorials are reduced to (p−1 , q)-numbers and (p−1 , q)-factorials, namely, [n]p−1 ,q =
p−n − q n , p−1 − q 97
and [n]!p−1 ,q =
(
1 for
n=0
((p−1 ,q);(p−1 ,q))n (p−1 −q)n
for
n ≥ 1,
(589)
respectively, which exactly reproduce the (p, q)-numbers and (p, q)−factorials introduced by Chakrabarty and Jagannathan [24]. The other properties can be recovered similarly to those of section 6.3.1 replacing the parameter p by p−1 . The R(p, q)−derivative is also reduced to (p−1 , q)−derivative. Indeed, p − q 1 − PQ P − Q (p−1 − q)P 1 = (P −1 − Q) ≡ ∂p−1 ,q −1 (p − q)z
∂R(p,q) = ∂p,q
(590)
obtained by a simple replacement of the dilatation operator P by P −1 . The algebra Ap−1 ,q , generated by {1, A, A† , N}, associated with (p, q)− Chakrabarty and Jagannathan deformation satisfies the following commutation relations: A A† − p−1 A† A = q N , [N, A† ] = A†
A A† − qA† A = p−N [N, A] = −A.
(591)
Hence, the (p−1 , q)−Rogers-Szeg¨o polynomials n X n zk Hn (z; p , q) = k p−1 ,q −1
k=0
n = 0, 1, 2, · · ·
(592)
obey the three-term recursion relation Hn+1 (z; p−1 , q) = Hn (p−1 z; p−1 , q) + zp−n Hn (pz; p−1 , q) −z(p−n − q n )Hn−1 (z; p−1 , q)
(593)
and (p−1 , q)−difference equation ∂p−1 ,q Hn (z; p, q) = [n]p−1 ,q Hn−1 (z; p, q).
(594)
Finally, the set of polynomials 1 ψn (z; p−1 , q) = p Hn (z; p−1 , q), [n]!p−1 ,q 98
n = 0, 1, 2, · · ·
(595)
forms a basis for a realization of the (p−1 , q)−deformed harmonic oscillator and quantum algebra Ap−1 ,q generating the commutation relations (591) with the number operator N formally defined as Nψn (z; p−1 , q) = nψn (z; p−1 , q),
(596)
and the annihilation and creation operators given by A = ∂p−1 ,q
and A† = P −1 + zp−N P − z(p−1 − q)∂p−1 ,q ,
(597)
respectively. Naturally, setting p = 1 permits to recover the results of the subsection 6.1.2. 6.3.3
R(x, y) =
xy−1 (q−p−1 )y
In this case, the R(p, q)−factors and R(p, q)−factorials are reduced to [n]Q p,q =
pn − q −n , q − p−1
and [n]!Q p,q
=
(
1 for
n=0
((p,q −1 );(p,q −1 ))n (q−p−1 )n
for
n ≥ 1,
(598)
introduced in our previous work [53], generalizing the q−Quesne algebra [95]. Then follow some remarkable properties: Proposition 6.5 If n and m are nonnegative integers, then −m m [−m]Q q [m]Q p,q = −p p,q ,
[n +
[n −
m]Q p,q m]Q p,q
=
=
[n]Q p,q =
(599)
−m
n Q m Q −n q [n]Q [m]Q p,q + p [m]p,q = p [n]p,q + q p,q , −m m−n −m Q n−m m Q m Q [m]Q q [m]p,q = p [n]p,q + p q q [n]p,q − p p,q , q − p−1 Q −1 Q [2]p,q [n − 1]Q p,q − pq [n − 2]p,q . −1
p−q
(600) (601) (602)
Proof: Eqs.(599) and (600) are immediate by the application of the relations p−m − q m = −p−m q m (pm − q −m ) and pn+m − q −n−m = q −m (pn − q −n ) + pn (pm − q −m ) = pm (pn − q −n ) + q −n (pm − q −m ), respectively, while Eq.(601) results from the combination of Eqs.(599) and (600). Finally, the relation [n]p,q−1 =
pn − q −n q − p−1 pn − q −n q − p−1 Q = = [n] , n = 1, 2, · · · (603) p − q −1 p − q −1 q − p−1 p − q −1 p,q 99
cumulatively taken with the identity [n]p,q−1 = [2]p,q−1[n − 1]p,q−1 − pq −1 [n − 2]p,q−1 gives Eq.(602).
✷
Proposition 6.6 The (p, q)−Quesne binomial coefficients where
n k
Q
((p, q −1 ); (p, q −1 ))n , ((p, q −1 ); (p, q −1))k ((p, q −1); (p, q −1 ))n−k
=
p,q
(604)
0 ≤ k ≤ n; n ∈ N, satisfy the following properties n k
Q
=
p,q
n n−k
Q
=p
k(n−k)
p,q
n+1 k
Q
n+1 k
Q
=p
k
p,q
= p
p,q
n k
Q
n k
k
n k
+q
=p 1/qp
−n−1+k
p,q
Q
k(n−k)
+p
n+1−k
p,q
−(pn − q −n )
n k−1
n n−k
Q
n k−1
,
,
(605)
1/qp
(606)
p,q
Q
Qp,q n−1 . k − 1 p,q
(607)
Proof: It is straightforward, using the Proposition 6.3 and
n k
Q
p,q
=
n k
.
(608)
p,q −1
✷ † Finally, the algebra AQ , generated by {1, A, A , N}, associated with (p, q)− p,q Quesne deformation satisfies the following commutation relations: p−1 A A† − A† A = q −N −1 , [N, A† ] = A† ,
qA A† − A† A = pN +1 [N, A] = −A.
(609)
The (p, q)−Rogers-Szeg¨o polynomials corresponding to the Quesne deformation [53] are deduced from our generalization by choosing φ1 (x, y) = φ2 (x, y) = φ(x, y) = x and φ3 (x, y) = y − x−1 . Indeed, it is worthy of attention that we get 100
in this case φ(p, q) = p 6= 0, φ3 (p, q) = q − p−1 6= 0, φ(P, Q)z k = φk1 (p, q)z k and from Eq.(607)
n+1 k
Q
p,q
=p
k
n k
Q
+p
n+1−k
p,q
n k−1
Q
p,q
−1
− (q − p )[n]p,q
n−1 k−1
Q
.
p,q
Hence, the hypotheses of the theorem are satisfied and, therefore, the (p, q)− Rogers - Szeg¨o polynomials HnQ (z; p, q)
Q n X n zk , = k p,q k=0
n = 0, 1, 2, · · ·
(610)
satisfy the three-term recursion relation Q Hn+1 (z; p, q) = HnQ (pz; p, q) + zpn HnQ (p−1 z; p, q) Q − z(pn − q −n )Hn−1 (z; p, q)
(611)
and the (p, q)−difference equation Q Q ∂p,q HnQ (z; p, q) = [n]Q p,q Hn−1 (z; p, q).
(612)
Thus, the set of polynomials 1 HnQ (z; p, q), ψnQ (z; p, q) = q Q [n]!p,q
n = 0, 1, 2, · · ·
(613)
forms a basis for a realization of the (p, q)− Quesne deformed harmonic oscillator and quantum algebra AQ p,q engendering the commutation relations (609) with the number operator N formally defined as NψnQ (z; p, q) = nψnQ (z; p, q),
(614)
and the annihilation and creation operators given by Q A = ∂p,q
and A† = P + zpN P −1 − z(q − p−1 )∂p,q ,
(615)
respectively. Naturally, setting p = 1 gives the Rogers-Szeg¨o polynomials associated with the q−Quesne deformation [95]. The continuous (p, q)−Hermite polynomials corresponding to the (p, q)− generalization of Quesne deformation [53] can be defined as follows: inθ Q −2iθ HQ Hn (e ; p, q) n (cos θ; p, q) = e
101
Q n X n ei(n−2k)θ , = k p,q k=0
n = 0, 1, 2, · · · .
(616)
Since for the (p, q)-generalization of Quesne deformation [53] φ1 (x, y) = φ2 (x, y) = x and φ3 (x, y) = y − x−1 , from the Proposition 6.2 we deduce that the corresponding sequence of continuous (p, q)−Hermite polynomials satisfies the three-term recurrence relation n
iθ −iθ −1 2 HQ P )HQ n+1 (cos θ; p, q) = p (e P + e n (cos θ; p, q) Q n −n −(p − q )Hn−1 (cos θ; p, q).
6.3.4
ν
µ
R(x, y) = h(p, q)y /x
xy − 1 (q − p−1 )y
(617)
Here 0 < pq < 1 , pµ < q ν−1 , p > 1 , h is a well behaved real and non-negative function of deformation parameters p and q such that h(p, q) → 1 as (p, q) → (1, 1). The R(p, q)−factors become (p, q; µ, ν, h)-numbers introduced in our previous work [52] and defined by [n]µ,ν p,q,h = h(p, q)
q νn pn − q −n . pµn q − p−1
(618)
Proposition 6.7 The (p, q; µ, ν, h)−numbers verify the following properties, for m, n ∈ N: [−m]µ,ν p,q,h = −
q −2νm+m [m]µ,ν p,q,h , p−2µm+m
q νn q νm−m µ,ν [n] + [m]µ,ν p,q,h p,q,h pµm pµn−n νn−n νm q q [m]µ,ν = µm−m [n]µ,ν p,q,h + p,q,h , p pµn
(619)
[n + m]µ,ν p,q,h =
q −νm+m µ,ν q ν(n−2m)+m [n] − [m]µ,ν p,q,h p,q,h −µm µ(n−2m)−n+m p p q ν(n−2m)−n+m q −νm − [m]µ,ν = −µm+m [n]µ,ν p,q,h p,q,h , p pµ(n−2m)+m
(620)
[n − m]µ,ν p,q,h =
[n]µ,ν p,q,h =
q − p−1 q −ν 1 q 2ν−1 µ,ν µ,ν [2] [n − 1] − [n − 2]µ,ν p,q,h p,q,h . p − q −1 p−µ h(p, q) p,q,h p2ν−1 102
(621)
(622)
Proof: It is direct using the Proposition 6.5 and the fact that [n]µ,ν p,q,h = h(p, q)
q νn Q [n] . pµn p,q
(623) ✷
Proposition 6.8 The (p, q, µ, ν, h)− binomial coefficients Q µ,ν [n]!µ,ν q νk(n−k) n n p,q,h , := µ,ν = µk(n−k) k p,q k p,q,h [k]!µ,ν p p,q,h [n − k]!p,q,h
(624)
where 0 ≤ k ≤ n; n ∈ N, satisfy the following properties µ,ν µ,ν n n = , k p,q,h n − k p,q,h
n+1 k
µ,ν
n+1 k
=
p,q,h
µ,ν
p,q,h
q νk p(µ−1)k
=
n k
q νk p(µ−1)k
µ,ν
q (ν−1)(n+1−k) + pµ(n+1−k) p,q,h
n k
µ,ν
(625)
n k−1
µ,ν
,
µ,ν n + (µ−1)(n+1−k) k − 1 p,q,h p p,q,h µ,ν νn n−1 n −n q −(p − q ) µn . k − 1 p,q,h p q ν(n+1−k)
(626)
p,q,h
(627)
Proof: There follow from the Proposition 6.6 and the fact that n [n]!µ,ν p,q,h = h (p, q)
q n(n+1)/2 [n]!Q p,q , pn(n+1)/2
(628)
where use of Eq.(623) has been made. ✷ † The algebra Aµ,ν , generated by {1, A, A , N}, associated with (p, q, µ, ν, h)p,q,h deformation, satisfies the following commutation relations: ν−1 N +1 q qν † −1 † , p A A − µ A A = h(p, q) p pµ ν N +1 qν † q † qA A − µ A A = h(p, q) µ−1 p p [N, A† ] = A† , [N, A] = −A. (629) The (p, q, µ, ν, h)-Rogers-Szeg¨o [52] polynomials are deduced from the above general construction by setting φ1 (x, y) = x1−µ y ν , φ2 (x, y) = x−µ y ν−1 and φ3 (x, y) = 103
y−x−1 . h(p,q)
Indeed, φi (p, q) 6= 0 for i = 1, 2, 3; φi (P, Q)z k = φi (p, q)k z k for i = 1, 2 and the property (627) furnishes µ,ν µ,ν µ,ν q ν(n+1−k) q νk n n n+1 + = (µ−1)k k p,q,h p(µ−1)(n+1−k) k − 1 p,q,h k p p,q,h µ,ν q − p−1 µ,ν n−1 − . [n] h(p, q) p,q,h k − 1 p,q,h Therefore, the (p, q, µ, ν, h)-Rogers-Szeg¨o polynomials are defined as follows: µ,ν n X n z k , n = 0, 1, 2 · · · (630) Hn (z; p, q, µ, ν, h) = k p,q,h k=0
with the three-term recursion relation ν q z : p, q, µ, ν, h Hn+1 (z; p, q, µ, ν, h) = Hn pµ−1 ν q (ν−1)n p +z µn Hn z; p, q, µ, ν, h p q ν−1 q νn −z µn (pn − q −n )Hn−1 (z; p, q, µ, ν, h) p
(631)
and (p, q, µ, ν, h)−difference equation µ,ν ∂p,q,h Hn (z; p, q, µ, ν, h) = [n]µ,ν p,q,h Hn−1 (z; p, q, µ, ν, h).
(632)
Hence, the set of polynomials 1 ψn (z; p, q, µ, ν, h) = q Hn (z; p, q, µ, ν, h), [n]!µ,ν p,q,h
n = 0, 1, 2, · · ·
(633)
forms a basis for a realization of the (p, q, µ, ν, h)−deformed algebra Ap,q,µ,ν,h satisfying the commutation relations (629) with the number operator N formally defined as NψnQ (z; p, q, µ, ν, h) = nψn (z; p, q, µ, ν, h), together with the annihilation and the creation operators given by ν−1 N q Pµ (q − p−1 ) µ,ν Qν µ,ν † − z ∂ , A = ∂p,q,h and A = µ−1 + z P pµ Qν−1 h(p, q) p,q,h respectively. 104
(634)
(635)
The continuous (p, q, µ, ν, h)−Hermite polynomials [52] can be now deduced as: Hn (cos θ; p, q, µ, ν, h) = einθ Hn (e−2iθ ; p, q, µ, ν, h) µ,ν n X n = ei(n−2k)θ , n = 0, 1, 2, · · · . k p,q,h
(636)
k=0
Since for the (p, q, µ, ν, h)−deformation φ1 (x, y) = x1−µ y ν , φ2 (x, y) = x−µ y ν−1 −1 , from the Proposition 6.2 the corresponding sequence of and φ3 (x, y) = y−x h(p,q) continuous (p, q, µ, ν, h)−Hermite polynomials satisfies the three-term recursion relation n
Hn+1 (cos θ; p, q, µ, ν, h) =
7
qν 2
Qν Hn (cos θ; p, q, µ, ν, h) n p(µ−1) 2 P µ−1 n q (ν−1) 2 Q−(ν−1) + µn Hn (cos θ; p, q, µ, ν, h) P −µ p 2 q νn −(pn − q −n ) µn Hn−1 (cos θ; p, q, µ, ν, h). (637) p
Concluding remark
We have first deformed the Heisenberg algebra with the set of parameters {q, l, λ} to generate a new family of generalized coherent states respecting the Klauder criteria. In this framework, the matrix elements of relevant operators have been exactly computed and investigated from functional analysis point of view. Then, relevant statistical properties have been examined. Besides, a proof on the sub-Poissonian character of the statistics of the main deformed states has been provided. This property has been finally used to determine the induced generalized metric, characterizing the geometry of the considered system. Next, a unified method of defining structure functions from commutation relations of deformed single-mode oscillator algebras has been presented. A natural approach to building coherent states associated to deformed algebras has been then deduced. Known deformed algebras have been given as illustration and such mathematical properties as continuity in the label, normalizability and resolution of the identity of their corresponding coherent states have been discussed. Besides, we have generalized a class of two - parameter deformed Heisenberg algebras related to meromorphic functions. There have been probed relevant families of coherent states maps and their corresponding hypergeometric series. The latter constitutes a generalization of known hypergemotric series. Moreover, a R(p, q)-binomial theorem, generalizing the (p, q)-binomial theorem given in [57] has been deduced. We have also defined the R(p, q)-trigonometric, hyperbolic and (p, q)-Bessel functions, including their main relevant properties. 105
Then, we have provided a new noncommutative algebra related to the R(p, q)deformation and shown that the notions of differentiation and integration can be extended to it, thus generalizing well known q or/and (p, q)-differential and integration calculi [22, 31, 74]. Besides, we have performed a general procedure of constructing the Hopf algebra structure compatible with the R(p, q)-algebra. As illustration, relevant examples have been given. Finally, we have defined and discussed a general formalism for constructing R(p, q)− deformed Rogers-Szeg¨o polynomials. The displayed approach not only provides novel relations, but also generalizes well known standard and deformed Rogers-Szeg¨o polynomials. A full characterization of the latter, including the data on the three-term recursion relations and difference equation, has been provided. We have succeeded in elaborating a new realization of R(p, q)−deformed quantum algebra generalizing the construction of q−deformed harmonic oscillator creation and annihilation operators performed in [40, 58]. The continuous R(p, q)−Hermite polynomials have been also investigated in detail and relevant particular cases and examples have been exhibited.
Acknowledgements This work is partially supported by the Abdus Salam International Centre for Theoretical Physics (ICTP, Trieste, Italy) through the Office of External Activities (OEA) - Prj-15. The ICMPA is in partnership with the Daniel Iagolnitzer Foundation (DIF), France.
References [1] E. Abe, Hopf algebras (Cambridge University Press, Cambridge, 1977). [2] N.I. Akhiezer, The Classical Moment Problem and Some Related Questions in Analysis (Olivier and Boyd, London, 1965). [3] S. T. Ali, J.-P. Antoine, J.-P. Gazeau, Coherent States, Wavelets and their Generalizations (Springer-Verlag, New-York, 1999). [4] S. T. Ali, J.-P. Antoine, J.-P. Gazeau and U.A. Mueller, Coherent states and their generalizations: A mathematical overview, Rev. Math. Phys.7 (1995), 1013-1104. [5] C. Aragone, E. Chalbaud, S. Salamo, Intelligent spin states, J. Math. Phys. 17 (1976), 1963-1971.
106
[6] C. Aragone, G. Guerri, S. Salamo, and J.L. Tani, Intelligent spin states, J. Phys. A: Math. Nucl. Gen. 7 (1974), L149-L151. [7] M. Arik and D.D. Coon, Hilbert spaces of analytic functions and generated coherent states, J. Math. Phys. 17 (1976), 424-427. [8] N.M. Atakishiyev, M.K. Atakishiyeva, A q-analogue of the Euler gamma integral, Theoret. Math. Phys. 129 (2001), 1325-1334. [9] E. Balo¨ıtcha, M.N. Hounkonnou and E.B. Ngompe Nkouankam, Unified (p, q; α, β, ν; γ)-deformation: Irreducible representations and induced deformed harmonic oscillator, J. Math. Phys. 53 (2012), 013504-013514. [10] A. O. Barut and L. Girardello, New coherent states associated with noncompact groups, Commun. Math. Phys. 21 (1971), 41-55. [11] Ju. M. Berezanski´ı, Expansions in Eigenfunctions of Selfadjoint Operators, (Amer. Math. Soc., Providence, Rhode Island, 1968). [12] L. C. Biedenharn, The quantum group SUq (2) and a q-analogue of the boson operators, J. Phys. A 22 (1989), L873-L878. [13] V.V. Borzov, E.V. Damaskinsky and S.V. Yegorov, Somme remarks on the representations of the generalized deformed algebra, arXiv:q-alg/9509022. [14] T. Brzezi´ nski, J.L. Egusquinza, A.J. Macfarlane, Generalised Harmonic Oscillator Systems and Their Fock Space Description, Phys. Lett. B 311 (1993), 202-206. [15] J.D. Bukweli Kyemba and M.N. Hounkonnou, Characterization of (R, p, q)deformed Rogers-Szeg¨o polynomials: associated quantum algebras, deformed Hermite polynomials and relevant properties, J. Phys. A: Math. Theor. 45 (2012), 225204. [16] J.D. Bukweli Kyemba and M.N. Hounkonnou, On generalized oscillator algebras and their associated coherent states, submitted. [17] J.D. Bukweli Kyemba and M.N. Hounkonnou, (q; l, λ)-deformed Heisenberg algebra: coherent states, their statistics and geometry, Afr. Diaspora J. Math. 14 (2012), 38-56. [18] J.D. Bukweli Kyemba and M.N. Hounkonnou, R(p, q)-calculus: differentiation, integration and Hopf algebras, submitted.
107
[19] I.M. Burban, Unified (q; α, β, γ; ν)-deformation of one-parametric q-deformed oscillator algebras, Phys. Let. A 42 (2009), 056201, arxiv.org/pdf/0806.0613 [20] I. M. Burban, Two-parameter deformation of oscillator algebra, Phys. Lett. B 319 (1993), 485-489. [21] I. M. Burban, On (p, q; α, β, l)-deformed oscillator and its generalized quantum Heisenberg-Weyl algebra, Phys. Lett. A 366 (2007), 308-314. [22] I. M. Burban and A. U. Klimyk, P, Q−differentiation, P, Q−integration, and P, Q− hypergeometric functions related to quantum groups. Integral Transforms and Special Functions. 2(1994), 15-36. [23] B. L. Cerchiai, R. Hinterding, J. Madore and J. Wess, The geometry of q-deformed phase space, Eur. Phys. J. C 8 (1999), 533-546. [24] R. Chakrabarti and R. Jagannathan, A (p, q)−oscillator realisation of twoparameter quantum algebras. J. Phys. A: Math. Gen. 24 (1991), L711-L718. [25] V. Chari and A. Pressley, A guide to Quantum groups (Cambridge, Cambridge University Press, 1994). [26] W.-S. Chung, K.-S Chung, S.-T. Nam and C.-I. Um, Generalized deformed algebra, Phys.Lett.A, 183 (1993), 363-370. [27] S.M. Dancoff, Non-adiabatic meson theory of nuclear forces, Phys. Rev. 78 (1950), 382-385. [28] M. Daoud, Photon-added coherent states for exactly solvable Hamiltonians, Phys. Letters A 305 (2002), 135-143. [29] V. G. Drinfel’d, Hopf Algebras and the quantum Yang-Baxter equation, Dokl. Akad. Nauk SSSR, 283 (1985), 1060-1064. [30] V.G. Drinfel’d, Quantum Group in Proc. Int. Congr. Math., Berkeley (1986), pp. 798-820. [31] A. Dobrogowska and A. Odzijewicz, Second order q−difference equations solvable by factorization method, J. Comput. Appl. Math. 193 (2006), 319346. [32] M. El Baz and Y. Hassouni, Deformed exterior algebra, Quons and their coherent states, Int. Jour. Mod. Phys. 18 (2003), preprint: IC/2002/106.
108
[33] L. D. Faddeev and L. A. Takhtajan, The Quantum Method of the Inverse Problem and the Heisenberg XY Z model, Russ. Math. Surveys 34 (1979), 11-68. [34] L. D. Faddeev and N. Yu. Reshetkhin, Hamiltonian structures for integrable models field theory, Teor. Mat. Fiz., 56 (1983), 323-343. [35] L.D. Faddeev, N.Yu. Reshitikhin and L.A. Takhtajan, Quantization of Lie groups and Lie algebras, Leningrad Math J. 1 (1990), 193-225. [36] Ph. Feinsilver, Commutators, anti-commutators and Eulerian calculus, Rocky Mountain J. Math., 12 (1982), 171-183. [37] Ph. Feinsilver, Discrete analogues of the Heisenberg-Weyl algebra, Mn. Math., 104 (1987), 89-108. [38] R. Floreanini, L. Lapointe and L. Vinet, A note on (p, q)−oscillators and bibasic hypergeometric functions, J. Phys. A: Math. Gen. 26 (1993), L611L614. [39] R. Floreanini, L. Lapointe and L. Vinet, A quantum algebra approach to basic multivariable special functions, J. Phys. A: Math. Gen. 27 (1994), 6781-6797. [40] D. Galetti, A realization of the q-deformed harmonic oscillator: Rogers-Szeg¨o and Stieltjes-Wigert polynomials, Braz. J. Phys. 33 (2003), 148-157. [41] G. Gasper and M. Rahman, Basic Hypergeometric Series (Cambridge, Cambridge University Press, 1990). [42] G. Gasper and M. Rahman, Basic Hypergeometric Series (Cambridge University Press, Cambridge, 2004). [43] J.-P. Gazeau, Coherent States in Quantum Physics (WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim, 2009). [44] J.-P. Gazeau and J.R. Klauder, Generalized Coherent States for Arbitrary Quantum Systems (Academic Press, New-York, 2000). [45] M.I. Gelfand, M. I. Graev and V. S. Retakh, (r, s)−Hypergeometric functions of one variable Russian Acad. Sci. Dokl. Math 48 (1994), 591-596. [46] R. Gilmore, Geometry of symmetrized states Ann. Phys. (NY) 74 (1972), pp 391-463.
109
[47] R. Gilmore, On properties of coherent states, Rev. Mex. Fis. 23 (1974), 143-187. [48] R.J. Glauber, Photon correlations, Phys. Rev. Letters 10 (1963), 84-86. [49] R.J. Glauber, Coherent and incoherent states of the radiation field, Phys. Rev. 131 (1963), 2766-2788. [50] A. Horzela and F.H. Szafraniec, A measure-free approach to coherent states, J. Phys. A Math. Theor. 45 (1012), 244018. [51] M.N. Hounkonnou and J.D. Bukweli Kyemba, Generalized (R, p, q)deformed Heisenberg algebras: coherent states and special functions, J. Math. Phys. 51 (2010), 063518. [52] M. N. Hounkonnou and E. B. Ngompe Nkouankam, New (p, q, µ, ν, f )deformed states. J. Phys. A: Math. Theor. 40 (2007), 12113-12130. [53] M .N Hounkonnou and E. B. Ngompe Nkouankam, On (p, q, µ, ν, φ1, φ2 ) generalized oscillator algebra and related bibasic hypergeometric functions, J. Phys. A: Math. Theor. 40 (2007), 8835-8843. [54] M. N. Hounkonnou and K. Sodoga, Generalized coherent states for associated hypergeometric-type functions, J. Phys. A: Math. Gen. 38 (2005), 7851-7857. [55] E. H. Ismail Moudard, Classical and Quantum Orthogonal Polynomials in one Variable. Enc. Math Appl. Vol. 98 (Cambridge, Cambridge University Press, 2005). [56] F.H. Jackson, The application of basic numbers to Bessel’s and Legendre’s functions II in Proc. London Math. Soc.(2), 3 (1905), pp 192-220. [57] R. Jagannathan and K. Srinivasa Rao, Two-parameter quantum algebras, twin-basic numbers, and associated generalized hypergeometric series, arXiv:math/0602613. [58] R. Jagannathan and R. Sridhar, (p, q)-Rogers-Szeg¨o polynomials and the (p, q)-oscillator, arXiv:1005.4309v1 [math.QA] [59] A.D. Jannussis, L.C. Papaloucas and P.D. Siafarikas, Eigenfunctions and Eigenvalues of the Q-Differential Operators, Hadronic J. 3 (1980), 16221632. [60] A.D. Jannussis, G. Bbodimas, D. Sourlas and V. Zisis, Remarks on the qquantization, Lett. Nuovo Cim. 30 ( 1981), 123-127. 110
[61] A. Jannussis, G. Bbodimas, D. Sourlas, L.C. Papaloucas and P.D. Siafaricas, Q-algebras without interaction, Lett. Nuovo Cim. 37 ( 1983), 119-123. [62] A. Jannussis, G. Bbodimas, D. Sourlas, L. Papaloucas, P.D. Siafaricas and K. Vlachos, Some properties of Q-analysis and applications to noncanonical mechanics, Hadronic J. 6 ( 1983), 1653-1686. [63] M. Jimbo, A q-difference analogue of U(g) and Yang-Baxter equation, Lett. Math. Phys. 10 (1985), 63-69. [64] M. Jimbo, Quantum R-matrix for the generalized Toda system, Comm. Math. Phys. 102 (1986), 537-547 . [65] G. Junker and P. Roy, Conditionally exactly solvable problems and nonlinear algebras, Phys. Lett. A 232 (1997), 155-161. [66] G. Kalnins, W. Miller and S. Mukhejee, Models of q-algebra representations: matrix elements of the q-oscillator algebra, J. Math. Phys. 34 (1993), 53335356 [67] C. Kassel, Quantum groups (Springer-Verlag, New York, 1995). [68] A.A. Kirillov, Elements of the theory of representation (Springer, Berlin, Heidelberg, New York, 1976). [69] R. J. Klauder, Continuous-representation theory I. Postulates of continuous representation theory J. Math. Phys. 4 (1963), 1055-1058. [70] R. J. Klauder, Continuous-representation theory II. Generalized relation between quantum and classical dynamics J. Math. Phys. 4 (1963), 1058-1073. [71] J.R. Klauder, K.A. Penson and J.-M. Sixderniers, Constructing coherent states through solutions of Stieljes and Hausdorff moment problems, Phys. Rev. A 64 (2001), 013817. [72] J.R. Klauder and B.S. Skagerstam, Coherent states: Applications in Physics and Mathematical Physics (Singapore: World Scientific, 1985) [73] A. Klimyk and K. Schm¨ udgen, Quantum Groups and their Representation, (Berlin Heidelberg, Springer-Verlag, Berlin Heidelberg, 1997). [74] R. Koekoek and R. F. Swarttouw, The Askey-scheme of orthogonal polynomials and its q−analogue, (TUDelft Report No. 98-17, 1998).
111
[75] B. Kostant, Group Representation in Mathematics and Physics, ed. by V. Bargamann, Lecture Notes Phys., Vol 6 (Springer, Berlin, Heidelberg, New York, 1970). [76] K. Kosi´ nski, M. Majewski and P. Ma´slanka, Representations of generalized oscillator algebra, arXiv:q-alg/9501012v1 [77] P.P. Kulish and N.Y. Reshetikhin, Quantum linear problem for the sineGordon equation and higher representations, J. Sov. Math. 23 (1983), 24352445. [78] V.V. Kuryshkin, Annales de la Fondation Louis de Broglie 5 (1980), p. 111. [79] A. J. Macfarlane, On q-analogues of the quantum harmonic oscillator and quantum group SU(2)q , J. Phys. A 22 (1989), 4581-4588. [80] V.I. Man’ko, G. Marmo, E.C.G. Sudarshan and F. Zaccaria, f -oscillators and nonlinear coherent states, Phys. Scr. 55 (1997), 528-541. [81] R. L. de Matos Filho and W. Vogel, Nonlinear coherent states, Phys. Rev. A 54 (1996), 4560-4563. [82] V. Maximov and A. Odzijewicz, The q−deformation of quantum mechanics of one degree of freedom. J. Math. Phys. 36 (1995), 1681-1689. [83] S. Meljanac, M. Milekovi, and S. Pallua, Unified view of deformed singlemode oscillator algebras, Phys. Lett. B 328 (1994), 55-59. [84] M.M. Nieto and L.M. Simmmons, Jr., Coherent states for general potentials, Phys. Rev. Lett. 41 (1978), 207-210. [85] M.M. Nieto and L.M. Simmmons, Jr., Coherent states for general potentials: I, II and III, Phys. Rev. D 20 (1979), 1321-1350. [86] M.M. Nieto, L.M. Simmmons, Jr., and V.P. Gutschick, Coherent states for general potentials: IV, Phys. Rev. D 20 (1981), 391-402. [87] A. Odzijewicz, Quatum algebras and q-special functions related to coherent states maps of the disc. Commun. Math. Phys. 192 (1998), 183-215. [88] V. Pasquier and H. Saleur, Common structures between finite systems and conformal field theories through quantum groups, Nucl. Phys. B 330, (1990),523-556. [89] A. M. Perelomov, Coherent states for arbitrary Lie group, Commun. Math. Phys. 26 (1972), 222-236. 112
[90] A.M. Perelomov, Generalized Coherent States and their Applications (Springer-Verlag, Berlin Heidelberg, 1986). [91] A.D. Polyanin and A. V. Manzhirov, Handbook of integral equations (CRC Press, Boca Raton, 1998). [92] D. Popov, Gazeau-Klauder quasi-coherent states for the Morse oscillator, Phys. Letters A 316 (2003), 369-381. [93] A.P. Prudnikov, Yu A. Brychkov and O.I. Marichev, Integrals and Series, vol 3 (Gordon and Breach, New York 1990). [94] C. Quesne, New q-deformed coherent states with an explicitly known resolution of unity, J. Phys. A: Math. Gen. 35 (2002), 9213-9226. [95] C. Quesne, K.A. Penson and V.M. Tkachuk, Maths-type q-deformed coherent states for q > 1, Phys. Lett. A 313 (2003),29-36. arXiv:quant-ph/0303120v2. [96] V. Sahai, q−Representations of Lie algebras G(a, b), J. Indian Math. Soc. 63 (1997), 83-98. ¨ [97] E. Schr¨odinger, Der stetige Ubergang von der Mikro- zur Makromechanik, Naturwissensehaften, 14 (1926), 664-666. [98] J.A. Shohat and J.D. Tamarkin, The Problem of Moments, (APS, New York, 1943). [99] E. K. Sklyanin, Some Algebraic Structures Connected with the Yang-Baxter Equation. Representations of Quantum Algebras, Funct. Anal. Appl., 17 (1983, 34-48). [100] I.N. Sneddon, The Use of Integral Transforms (McGraw-Hill, New York, 1974). [101] H. S. Snyder, Quantized space-time, Phys. Rev. 71 (1947), 38-41 [102] E.C. G. Sudarshan, Equivalence of semiclassical and quantum mechanical descriptions of statistical light beams, Phys. Rev. Letters 10 (1963), 277279. [103] C.-P. Sun, H.-C. Fu, The q-deformed boson realisation of the quantum group SU(n)q and its representations, J. Phys. A: Math. Gen. 22 (1989), L983L986.
113
[104] I.E. Tamm, The relativistic interaction of elementary particles, J. Phys. UdSSR 9 (1945), 449-465. [105] M.A. Vasiliev, High spin algebras and quantization on the sphere and hyperboloid, Int. J. Mod. Phys. A 6 (1991), 1115. [106] H. J. de Vega, Yang-Baxter Algebras, Integrable Theories and Quantum Groups, Int. J. Mod. Phys. A 4, (1989), 2371. [107] J. Wess, q-deformed phase space and its lattice structure, Int. J. Mod. Phys. A 12 (1997), 4997-5006. [108] E. Witten, Gauge theories, vertex models, and quantum groups, Nucl. Phys. B 330 (1990), 285-346. [109] H. Yan, q-deformed oscillator algebra as a quantum group , J. Phys. A: Math. Gen. 23 (1990), L1155-L1160. [110] L. You-quan and S. Zheng-mao, A deformation of quantum mechanics, J. Phys. A: Math. Gen. 25 (1992), 6779-6788. [111] S. Yu, M. Baker and D. D. Coon, First and Second Factorization in a Dual Multiparticle Theory with Nonlinear Trajectories, Phys. Rev. D 5(1972),3108-3121. [112] W.M. Zhang, D.H. Feng and R.G. Gilmore, Coherent states: theory and some applications, Rev. Mod. Phys. 62 (1990), 867-927.
114