IOP PUBLISHING

JOURNAL OF PHYSICS A: MATHEMATICAL AND THEORETICAL

J. Phys. A: Math. Theor. 42 (2009) 285301 (18pp)

doi:10.1088/1751-8113/42/28/285301

A generalized quantum nonlinear oscillator B Midya and B Roy Physics & Applied Mathematics Unit, Indian Statistical Institute, Kolkata 700 108, India E-mail: [email protected] and [email protected]

Received 14 January 2009, in final form 11 May 2009 Published 19 June 2009 Online at stacks.iop.org/JPhysA/42/285301 Abstract We examine various generalizations, e.g. exactly solvable, quasi-exactly solvable and non-Hermitian variants, of a quantum nonlinear oscillator. For all these cases, the same mass function has been used and it has also been shown that the new exactly solvable potentials possess shape invariance symmetry. The solutions are obtained in terms of classical orthogonal polynomials. PACS numbers: 03.65.−w, 03.65.Ge

1. Introduction Recently, there has been a surge of interest in obtaining exact [1] and quasi-exact solutions [2] to the position-dependent mass Schr¨odinger equation (PDMSE) for various potentials and mass functions by using various methods such as Lie algebraic techniques [3], supersymmetric quantum mechanics (factorization method) [4, 5], the shape invariance approach [6], point canonical transformation [7], path integral formalism [8], the transfer matrix method [9], etc. Apart from the intrinsic interest, the motivation behind this issue arises because of the relevance of position-dependent mass in describing the physics of many microstructures of current interest, such as compositionally graded crystals [10], quantum dots [11], 3 He clusters [12], metal clusters [13], etc. The concept of position-dependent mass comes from the effective mass approximation [14] which is a useful tool for studying the motion of carrier electrons in pure crystals and also for the virtual-crystal approximation in the treatment of homogeneous alloys (where the actual potential is approximated by a periodic potential) as well as in graded mixed semiconductors (where the potential is not periodic). Attention to the effective mass approach stems from the extraordinary development in crystallographic growth techniques, which allow for the production of a non-uniform semiconductor specimen with abrupt heterojunctions. In these mesoscopic materials, the effective mass of the charge carriers is position dependent. Consequently, the study of the effective mass Schr¨odinger equation becomes relevant for a deeper understanding of the non-trivial quantum effects observed on these nanostructures. The position-dependent (effective) mass is also used in the construction of pseudo-potentials, which have a significant computational advantage in the quantum Monte 1751-8113/09/285301+18$30.00 © 2009 IOP Publishing Ltd Printed in the UK

1

J. Phys. A: Math. Theor. 42 (2009) 285301

B Midya and B Roy

Carlo method [15]. It has also been found that such equations appear in very different areas. For example, it has been shown that a constant mass Schr¨odinger equation in curved space and equations based on deformed commutation relations can be interpreted in terms of PDMSE in the flat space [16] and a PT symmetric cubic anharmonic oscillator [17]. The nonlinear differential equation (1 + λx 2 )x¨ − (λx)x˙ 2 + α 2 x = 0,

λ > 0,

(1)

was studied by Mathews and Lakshmanan in [18, 19] as an example of a nonlinear oscillator, and it was shown that the solution of (1) is x = Asin(ωt + φ)

(2)

with the following additional restriction linking frequency and amplitude: α2 . (3) 1 + λA2 Furthermore, (1) can be obtained from the Lagrangian [18] 1 1 L= (x˙ 2 − α 2 x 2 ) (4) 2 (1 + λx 2 ) so that both the kinetic and the potential terms depend on the same parameter λ. So this nonlinear oscillator must be considered as a particular case of a system with a positiondependent effective mass. Recently in a series of papers [20, 21], this particular nonlinear system has been generalized to higher dimensions and various properties of this system have been studied. The classical Hamiltonian corresponding to the λ-dependent oscillator is given by [18, 21]       √ x2 1 1 2 , P Px + g = 1 + λx 2 px , g = mα 2 , (5) H = x 2m 2 1 + λx 2 ω2 =

, L the Lagrangian and m px being the canonically conjugate momentum defined by px = ∂L ∂ x˙ the mass.   It has been shown in [21] that in the space L2 (, dμ) where dμ = √ 1 2 dx, the 1+λx √ d is skew self-adjoint. Therefore, in contrast to the naive differential operator 1 + λx 2 dx expectation of ordering ambiguities, the transition from the classical system to the quantum one is given by defining the momentum operator √ d (6) Px = −i 1 + λx 2 dx so that  √  √ d d 2 2 2 2 (1 + λx )px → − . 1 + λx 1 + λx dx dx Therefore, the quantum version of the Hamiltonian (5) with h ¯ = 1 becomes [21]     2 1 d 1 1 1 2 d ˆ H =− , (7) (1 + λx ) 2 − λx + g 2m dx 2m dx 2 1 + λx 2 where g = α(mα + λ). It is to be noted that in [21], the value of the parameter g has been slightly modified from that given in equation (5). It may be pointed out that this λ-dependent system can be considered as a deformation of the standard harmonic oscillator in the sense that for λ → 0, all the characteristics of the linear oscillator are recovered. In [21], PDMSE corresponding to this nonlinear oscillator has been solved exactly as a Sturm–Liouville problem, and λ-dependent eigenvalues and eigenfunctions were obtained for 2

J. Phys. A: Math. Theor. 42 (2009) 285301

B Midya and B Roy

both λ > 0 and λ < 0. The λ-dependent wavefunctions were shown to be related to a family of orthogonal polynomials that can be considered as λ-deformations of the standard Hermite polynomials. Also, the Schr¨odinger factorization formalism, intertwining method and shape invariance approach were discussed with reference to this particular quantum Hamiltonian. The existence of a λ-dependent Rodrigues formula, a generating function and λ-dependent recursion relations were obtained. In this paper our objective is to re-examine this problem and obtain a closed-form expression for  the normalization constant, modified generating function and recursion relations  for  = αλ -deformed Hermite polynomials. A relation between the -deformed Hermite polynomials and Jacobi polynomials will also be obtained. We shall also obtain a number of exactly solvable, quasi-exactly solvable and non-Hermitian potentials corresponding to the same mass function m(x) = (1 + λx 2 )−1 . It will be seen that some of these potentials are generalizations of the nonlinear oscillator potentials while the others are of different types. It will be shown that these exactly solvable potentials are shape invariant. Moreover, these potentials can also be complexified and by doing so we shall also obtain a number of exactly solvable non-Hermitian potentials within the framework of PDMSE. As a method of obtaining these results, we shall use point canonical transformation consisting of a change of coordinate only. The organization of the paper is as follows: in section 2, we shall obtain exactly solvable potentials and a relation between -deformed Hermite polynomials and Jacobi polynomials; in section 3, it is shown that the exactly solvable potentials are shape invariant; in section 4, we obtain exactly solvable non-Hermitian potentials; section 5 deals with complex quasi-exactly solvable potentials and finally section 6 is devoted to a discussion. 2. Exactly solvable potentials for the mass m(x) =



1 1+λx2



Here we shall obtain exact to PDMSE for a number of potentials with the same   1 solutions mass function m(x) = 1+λx 2 . For this purpose, we first write PDMSE corresponding to the Hamiltonian given in equation (7) with m = 1 and λ > 0 as    2 1 dψ g 2 d ψ −(1 + λx ) 2 − λx ψ = Eψ − dx dx λ 1 + λx 2 g E = 2e − , λ

(8) (9)

where e is the energy for the Hamiltonian (7). Now expanding (1 + λx 2 )−1 for |x| < √1λ , we can write the potential of equation (8) as g (10) V (x) = − + gx 2 − λO(x 3 ). λ It is clear from (10) that the term (− gλ ) in equation (9) cancels from both sides of equation (8), so that the new eigenvalues (9) are actually the old eigenvalues e of the Hamiltonian (7). Also, as λ → 0, the potential and the eigenvalues of equation (8) reduce to those of a linear harmonic oscillator. Now generalizing the potential of equation (8) as below, the corresponding PDMSE reads as  √

√ 2 √ B 2 − A2 − A λ λx dψ 2 d ψ 2 + + B(2A + λ) − (1 + λx ) 2 − λx + A ψ = Eψ. dx dx 1 + λx 2 1 + λx 2 (11) 3

J. Phys. A: Math. Theor. 42 (2009) 285301

B Midya and B Roy

It is seen from (11) that if we √put B = 0, then the potential reduces to that of the nonlinear oscillator with gλ = A2 + A λ. It is to be noted that this generalization should correctly reproduce the λ → 0 limit, in which case equation (11) reduces to the Schr¨odinger equation for a linear harmonic oscillator. In the appendix we have shown that in the√ limit λ → 0 and for A = √αλ (which is one of the solution of the quadratic equation A2 + A λ = gλ ), B = 0 the potential of equation (11), the energy eigenvalues (18) and the wavefunction given in (19) reduce to those of a linear harmonic oscillator. This particular generalization is made so that it corresponds to the hyperbolic Scarf II potential [22] in the constant mass case. In order to solve (11), we now perform a transformation involving a change of variable given by √ 1 dx = √ sinh−1 ( λx), (12) z= √ F (x) λ where F (x) = 1 + λx 2 ,

λ > 0.

(13)

Under transformation (12), equation (11) reduces to a Schr¨odinger equation d2 ψ + V (z)ψ(z) = Eψ(z), dz2 where the potential V (z) is given by √ √ √ √ √ V (z) = (B 2 − A2 − A λ) sech2 (z λ) + B(2A + λ) tanh(z λ) sech(z λ) + A2 . −

Potential (15) is a standard solvable potential and the solutions are given by [22] √ √ √ s −1 (ir−s− 12 ,−ir−s− 12 ) ψn (z) = Nn i n (1 + sinh2 (z λ))− 2 e−rtan (sinh(z λ)) Pn (i sinh(z λ),

(14)

(15)

(16)

(α,β) B √ and Pn (x) is the Jacobi polynomial λ

where Nn is the normalization constant, s = √Aλ , r = [24]. The normalization constants Nn , n = 0, 1, 2, . . . , are given by [23] √    1/2  λn!(s − n) s − ir − n + 12 s + ir − n + 12 Nn = . π 2−2s (2s − n + 1) The eigenvalues En are given by √ √ En = n λ(2A − n λ),

n = 0, 1, 2, . . . < s.

(17)

(18)

Subsequently by performing the inverse of transformation (12), we find the solution to PDMSE (11) as √    1/2  λn!(s − n) s − ir − n + 12 s + ir − n + 12 ψn (x) = π 2−2s (2s − n + 1)  (19)  √ √ A s −1 (ir−s− 12 ,−ir−s− 12 ) . i n (1 + λx 2 )− 2 e−rtan (x λ) Pn (ix λ), n = 0, 1, 2, · · · < s = √ λ At this point, it is natural to ask the following question: are there other solvable potentials  1 corresponding to the mass function m(x) = 1+λx 2 ? The answer to this question is in the affirmative. The procedure to obtain these potentials is similar and so instead of treating each case separately, we have presented the potentials and the corresponding solutions in table 1. The first two and the last two potentials in table 1 are actually the generalizations of the nonlinear oscillator potential. Although the other two potentials in the table are not generalizations of the nonlinear oscillator potential, nevertheless they are exactly solvable potentials with the same mass function. 4

s − n + a, s2 = s − n − a, s3 = a − n − s, s4 = −(s

V (x)

W (x) √ λx A √1+λx 2



B 2 −A2 −A λ + 2 1+λx √ √λx λ) 1+λx 2

A2 +

B2 A2

2B



B(2A +

1 + B √1+λx 2

√ √ λx , B 1+λx 2



λx A √1+λx + 2

+

B A

< A2

√ B2 1+λx 2 − 2B √ A2 √ λx A(A− λ) , B > A2 2 λx

+

B A

−A



2 1+λx √ x λ

√ 0x λ∞

√ A2 +B 2 +A λ − B(2A 2 λx √ 2 λ) 1+λx + λx 2 2



+

A ,A < B √ 0x λ∞

A2 +B 2 −A |λ| − B(2A − 1+λx 2 √ √ x |λ| 2 |λ|) 1+λx 2 −A

s

i n (1 + λx 2 )− 2 e−rtan

A

1+λx √ x λ

2

− B x √1 λ

2



A2 + BA2 − (A + √ 2 n λ)2 − (A+nB√λ)2 √ √ n λ(2A − n λ)

√ |λ| A √x1+λx 2

1 − B √1+λx 2

√ √ n |λ|(2A + n |λ|)





|λ| A √x1+λx − 2

B A

B2 A2

− A2 + (A + √ n |λ|)2 − B√2 (A+n |λ|)2

√1 |λ|

−1 (x



λ)

√ (ir−s− 12 ,−ir−s− 12 ) Pn (ix λ)

√1 |λ|

x |λ| 2B 1+λx 2 −

x

ψn (x)

2

A2 + BA2 − (A − √ 2 n λ)2 − (A−nB√λ)2

=

B √ ,r λ 1

=

B λ ,a

 s21  s22 √ √ √x λ √x λ 1 + 2 2 1+λx  1+λx √ λ Pn(s1 ,s2 ) √x1+λx 2

1−

2 1+λx √ x λ

 s23 √  s24 2 1+λx √ −1 + 1 x λ

 2 √ Pn(s3 ,s4 ) 1+λx x λ √ r−s ( 1√ + λx 2 − 1)( 2 ) r+s ( 1 + λx 2 + 1)−( 2 )

Pn

√ ( 1 + λx 2 )

√ s  −r  (1 − x |λ|)( 2 ) (1 + √ ( r  +s  ) x |λ|) 2  (s  −r  − 12 ,s  +r  − 12 )  √ x |λ| Pn

2 −( s 2+n ) √ λx − 1 e−a |λ|x 2 1+λx

 √   |λ| Pn(−s −n−ia,−s −n+ia) −i √x1+λx 2

ai , i = 0, 1, . . . √ (A − i λ, B)

R(ai ) √ √ λ[2A − (2i + 1) λ]

√ (A − i λ, B)

√ A2 − [A − (i + 1) λ]2 2 2 B √ + BA2 − [A−(i+1) λ]2

√ (A + i λ, B)

√ A2 − [A + (i + 1) λ]2

√ (A − i λ, B)

B √ + BA2 − [A+(i+1) λ]2 √ √ λ[2A − (2i + 1) λ]

√ (A + i |λ|, B)

√  √  |λ| 2A + (2i + 1) |λ|

√ (A + i |λ|, B)

B2 A2

2



=

r1 s−n , s1

=

2

B 2√ [A+(i+1) |λ|]2

√ −A2 +[A+(i +1) |λ|]2

5

B Midya and B Roy

1 √ − |λ|

En √ √ n λ(2A − n λ)

A √ ,r λ

The first four entries correspond to λ > 0 and the last two correspond to λ < 0.

(r−s− 12 ,−r−s− 12 )



√ A(A− |λ|) − 1+λx 2 2 A2 + BA2

=

√B . |λ|

+A

√ A(A+ λ) 1+λx 2

x

=

√A &r  |λ|

2

A2 +

1 √ − |λ|

+ n + a), s 

J. Phys. A: Math. Theor. 42 (2009) 285301

Table 1. Exactly solvable shape invariant potentials V (x), superpotential W (x), energy eigenvalue En and wavefunctions ψn (x), where s =

J. Phys. A: Math. Theor. 42 (2009) 285301

B Midya and B Roy

2.1. Relation between the -deformed Hermite polynomial and Jacobi polynomial, generating function, recursion relation Here we shall obtain a correspondence between the -deformed Hermite polynomials [21] and Jacobi polynomials. We recall that the Hamiltonian for the nonlinear oscillator is given by [21]   2 d g x2 1 1 2 d . + H = − (1 + λx ) 2 − λx 2 dx 2 dx 2 1 + λx 2 After introducing adimensional variables (y, ) as was done in [21] √ λ y = αx, = , α the Schr¨odinger equation H ψ = ψ reduces to    1 y2 d2 1 d 1+ − (1 + y 2 ) 2 − y ψ = ψ. + 2 dy 2 dy 2 1 + y 2

(20)

(21)

The eigenvalues and eigenfunctions for  < 0 are [21] 1

ψm (y, ) = Hm (y, )(1 − ||y 2 ) (2||)   m = m + 12 − 12 m2 , m = 0, 1, 2, . . . ,

(22)

where Hm (y, ) is the -deformed Hermite polynomial whose Rodrigues formula and generating function are given in (27). For  > 0, ψm (y, ) = Hm (y, )(1 + y 2 )− 2 (23)   m = m + 12 − 12 m2 , m = 0, 1, 2 · · · , N ,   where N denotes the greatest integer lower than m = 1 . On the other hand, putting B = 0 and A = √αλ in solution (19) of equation (11), the eigenfunctions of equation (21) can be written in terms of the Jacobi polynomial as √ 1 1 (− 1 − 1 ,− 1 − 1 ) ψn (y) = Nn (1 + y 2 )− 2 Pn 2  2  (iy ), n = 0, 1, 2 . . . < ( > 0). (24)  For  < 0, putting B = 0, A = √α|λ| in the wavefunction of the fifth entry of table 1 and using (20), we obtain 1 (− 1 − 1 ,− 1 − 1 )  ψn (y) = Nn (1 + y 2 )− 2 Pn 2  2  (y ||), n = 0, 1, 2 . . . ( < 0). (25) 1

Comparing equations (22) and (25) and also equations (23) and (24), it is possible to derive a relation between the -deformed Hermite polynomial Hn (y, ) and the Jacobi polynomial (α,β) Pn (x) as n  √ 1 1 (− 1 − 1 ,− 1 − 1 ) Hn (y, ), ∀ . (26) Pn 2  2  (iy ) = √ n! 2i  The Rodrigues formula and the generating function for the -deformed Hermite polynomial Hn (y, ) are given by [21] 1 1 dn  + −( 1 + 1 )  zy = 1 + y 2 , Hn (y, ) = (−1)n zy 2 n zyn zy  2 , dy (27) 2 1 F (t, y, ) = (1 + (2ty − t )) . It was shown [21] that the polynomials obtained from the generating function F (t, y, ) with those obtained from the Rodrigues formula are essentially the same and only differ in 6

J. Phys. A: Math. Theor. 42 (2009) 285301

B Midya and B Roy

the values of the global multiplicative coefficients. We have observed that if the generating function F (t, y, ) is taken as  1 ∞  1 − n tn 2 1 1  Hn (y, ) , (28) (1 + (2ty − t )) = 1 n 2 2− n n! n=0 where (a)n represents the P¨ochhammer symbol given by (a)n = (a+n) , then the polynomials

(a) obtained from the above relation are exactly similar to those obtained from the Rodrigues formula given in equation (27). Correspondingly, the recursion relations are obtained as ((2n + 1) − 2)[2(1 − n)y Hn (y, ) + ((2n − 1) − 2)nHn−1 (y, )] = (n − 2)Hn+1 (y, )

(29)

and ((n − 2) − 2)[2((2n − 1) − 2)nHn (y, ) − ((n − 1) − 2)Hn (y, )]  = n((2n − 1) − 2)[2((n − 2) − 2)y Hn−1 (y, )  − (n − 1)((2n − 3) − 2)Hn−2 (y, )],

(30)

where the ‘prime’ denotes differentiation with respect to y. For  → 0, equations (29) and (30) give the recursion relations for the Hermite polynomial [24]. 3. Shape invariance approach to supersymmetric PDMSE The supersymmetric approach to PDMSE [5] may be discussed either by reducing PDMSE to a constant mass Schr¨odinger equation or by starting with modified intertwining operators consisting of first-order differential operators. Here, we shall be following the latter approach. Thus, we consider operators of the form   d 1 . (31) −i A = Px − iW (x), A† = Px + iW (x), Px = √ dx m(x) We now consider the supercharges Q, Q† defined by     0 A† 0 0 . Q= , Q† = 0 0 A 0 The supersymmetric Hamiltonian is then obtained as   †  PDM 0 AA H− PDM † = = {Q, Q } = H 0 H+PDM 0

(32)  0 , AA†

where the component Hamiltonians are given by    d 1 d2 m W 2 + W H±PDM = − + ± . √ m(x) dx 2 2m2 dx m

(33)

(34)

The Hamiltonians H±PDM are supersymmetric partners and the potentials are W  (x) . V±PDM = W 2 (x) ± √ m(x) It can be easily seen that the following commutation and anticommutation relations

(35)

2

Q2 = Q† = [Q, H PDM ] = [Q† , H PDM ] = 0 {Q, Q† } = {Q† , Q} = 0

(36) 7

J. Phys. A: Math. Theor. 42 (2009) 285301

B Midya and B Roy

together with equation (33) complete the standard supersymmetry algebra [22,  25]. For unbroken supersymmetry (SUSY), the ground state of H− has zero energy E0(−) = 0   provided the ground-state wavefunction ψ0(−) (z) given by Aψ0(−) = 0  x   ψ0(−) (x) = N0 exp − m(y)W (y) dy (37) is normalizable. In this case it can be shown that, apart from the ground state of H− , the partner Hamiltonians H± have identical bound-state spectra. In particular, they satisfy (−) En+1 = En(+) ,

n = 0, 1, 2 . . . .

The eigenfunctions of H± corresponding to the same eigenvalue are related by 1  (−) = En(+) 2 ψn(+) (x) Aψn+1  1 (−)  A† ψn(+) (x) = En(+) 2 ψn+1 (x).

(38)

(39)

It may be noted here that the superpotential W (x) and therefore the factorization of the Hamiltonian could be generated from the ground-state solution of the equation. In a remarkable paper [26], Gendenshtein explored the relationship between SUSY and solvable potentials. The pair of potentials V± (x, a0 ), a0 being a set of parameters, is called shape invariant if it satisfies the relationship [5, 22] V+ (x, a0 ) = W 2 (x, a0 ) + W  (x, a0 ) = W 2 (x, a1 ) − W  (x, a1 ) + R(a0 ) (40) = V (x, a1 ) + R(a0 ), where a1 is some function of a0 and R(a0 ) is independent of x. When SUSY is unbroken, the energy spectrum of any shape-invariant potential is given by [22] n−1  En(−) = R(ai ), E0(−) = 0. (41) i=0

We are now going to study the factorization and the shape invariance property of the potentials for PDMSE. As an example, let us consider the generalized nonlinear oscillator of section 2. For this, it is now necessary to choose the superpotential W (x) so that H− can be identified with the Hamiltonian of equation (11). In this case, we choose the superpotential to be √ λx 1 W = A√ +B√ . (42) 2 1 + λx 1 + λx 2 Therefore, the Hamiltonians H−PDM and H+PDM can be factorized as H−PDM = A† A = −(1 + λx 2 )

√   √ √ B 2 − A2 − A λ λx d2 d + A2 + − λx + B(2A + λ) dx 2 dx 1 + λx 2 1 + λx 2

H+PDM = AA†

(43)

√   √ √ λx d2 d B 2 − A2 + A λ + A2 . = −(1 + λx ) 2 − λx + B(2A − λ) + dx dx 1 + λx 2 1 + λx 2 These two Hamiltonians are related by √ √ √ (44) H+PDM (x; A, B) = H−PDM (x; A − λ, B) + λ(2A − λ) 2

so that they satisfy the shape invariance condition H+PDM (x, a0 ) = H−PDM (x, a1 ) + R(a0 ), √ √ √ where {a0 } = (A, B), {a1 } = (A − λ, B) and R(a0 ) = λ(2A − λ). 8

(45)

J. Phys. A: Math. Theor. 42 (2009) 285301

B Midya and B Roy

The ground state ψ0 (x, a0 ) of the Hamiltonian H−PDM is found by solving Aψ0 (x, a0 ) = 0 and has a zero energy, i.e. H−PDM (x, a0 )ψ0 (x, a0 ) = 0.

(46)

Now using (45) we can see that ψ0 (x, a1 ) is an eigenstate of H+PDM with the energy E1 = R(a0 ), because H+PDM (x, a0 )ψ0 (x, a1 ) = H−PDM (x, a1 )ψ0 (x, a1 ) + R(a0 )ψ0 (x, a1 ) = R(a0 )ψ0 (x, a1 ) (using (46)). Next, using the intertwining relation equation (45), we see that

H−PDM (x, a0 )A† (x, a0 )



= A

(x, a0 )H+PDM (x, a0 )

(47) and

H−PDM (x, a0 )A† (x, a0 )ψ0 (x, a1 ) = A† (x, a0 )H+PDM (x, a0 )ψ0 (x, a1 ) = A† [H−PDM (x, a1 ) + R(a0 )]ψ0 (x, a1 )

(48)

and hence using (46), we arrive at H−PDM (x, a0 )A† (x, a0 )ψ0 (x, a1 ) = R(a1 )A† (x, a0 )ψ0 (x, a1 ).

(49)

This indicates that A† (x, a0 )ψ0 (x, a1 ) is an eigenstate of H−PDM with an energy E1 = R(a0 ). Now iterating this process, we will find the sequence of energies for H−PDM as En(−) =

n−1 

√ √ R(ai ) = n λ(2A − n λ),

E0(−) = 0,

(50)

i=0

with corresponding eigenfunctions being ψn (x, a0 ) = A† (x, a0 )A† (x, a1 ) . . . A† (x, an−1 )ψ0 (x, an ),

(51)

where

√ √ √ √ ai = f (ai−1 ) = f (f (. . . (f (a0 ))) = (A − i λ, B) and R(ai ) = λ[2(A − i λ) − λ].    i times We have found a number of other potentials which are shape invariant for the same mass function. For all these potentials, the energy, wavefunctions and other parameters related to the shape invariance property are given in table 1. 3.1. Shape invariance approach to PDMSE with broken supersymmetry When  supersymmetry is broken, neither of the wavefunctions ψ0(±) (x) ≈ x√ m(y)W (y) dy] is normalizable and in this case all the energy values are degenerate, exp[± i.e. H+ and H− have identical energy eigenvalues [22, 26] En(−) = En(+)

(52)

with ground-state energies greater than zero. So far as we know, little attention has been paid till now to study problems involving broken SUSY in the case of PDMSE. Broken supersymmetric shape invariant systems in the case of a constant mass Schr¨odinger equation have been discussed in [27]. Below, we illustrate the two-step procedure discussed in [28] for obtaining the energy spectra in PDMSE when the SUSY is broken. For this, we consider the superpotential as √  1 1 + λx 2 x B , 0 < x < √ , λ < 0. −√ (53) W (x, A, B) = A |λ| √ 2 x |λ| |λ| 1 + λx 9

J. Phys. A: Math. Theor. 42 (2009) 285301

B Midya and B Roy

Then the supersymmetric partner potentials are obtained using (35) as √ √ A(A − |λ|) B(B − |λ|) − − (A + B)2 V− (x, A, B) = 1 + λx 2 λx 2 √ √ A(A + |λ|) B(B + |λ|) V+ (x, A, B) = − − (A + B)2 . 1 + λx 2 λx 2 The ground-state wavefunction is obtained from (37) as ψ0(−) ∼ x

√B |λ|

(1 + λx 2 ) 2

√A |λ|

.

(54)

(55) ψ0(−)

For A > 0, B > 0 the ground-state wavefunction is normalizable which means that the SUSY is unbroken. But for A > 0, B < 0 and A < 0, B > 0, neither of ψ0(±) is normalizable. Hence, SUSY is broken in both cases. We shall discuss the case A > 0, B < 0. In this case, the eigenstates of V± (x, A, B) are related by ψn(+) (x, a0 ) = A(x, a0 )ψn(−) (x, a0 ) ψn(−) (x, a0 ) = A† (x, a0 )ψn(+) (x, a0 ), En(−) (a0 ) = En(+) (a0 ).

(56)

Now we can show that the potentials in equation (54) are shape invariant by two different relations between the parameters. Step 1. The potentials of equation (54) are shape invariant if we change A → A + √ √ |λ| and B → B + |λ|. The shape invariant condition is given by    V+ (x, A, B) = V− (x, A + |λ|, B + |λ|) + (A + B + 2 |λ|)2 − (A + B)2 . (57) Now for B < − √1|λ| , it is seen that the superpotential (53) resulting from the change of

parameters as above falls in the class of a broken SUSY problem for which E0(−) = 0. Though the potentials of equation (54) are shape invariant but we are unable to determine the spectra for these potentials because of the absence of a zero √ energy ground state. Another way of parameterizations A → A + |λ| and B → −B gives us   V+ (x, A, B) = V− (x, A + |λ|, −B) + (A − B + |λ|)2 − (A + B)2 , (58) which √ shows that V− and V+ are shape invariant. This change of parameters (A → A + |λ| and B → −B) leads to a system with unbroken SUSY since √ the parameter B |λ|, −B) is zero. changes sign. Hence, the ground-state energy of the potential V√ − (x, A + From relation (58), we observe that V+ (x, A, B) and V− (A + |λ|, −B) differ only by a constant; hence, we have  ψ+ (x, A, B) = ψ− (x, A + |λ|, −B) (59)   En(+) (A, B) = En(−) (x, A + |λ|, −B) + (A − B + |λ|)2 − (A + B)2 . Thus, if we can √ evaluate the spectrum and energy eigenfunctions of unbroken SUSY H−PDM (x, A + |λ|, −B), then we can determine the spectrum and eigenfunctions H+PDM (x, A, B) with broken SUSY. In the second step, we will do this. Step 2. With the help of shape invariant formalism in the case of unbroken √ SUSY for PDMSE (see section 3), we obtain a spectrum and eigenfunctions for V− (x, A + |λ|, −B) as     En(−) (A + |λ|, −B) = (A − B + |λ| + 2n |λ|)2 − (A − B + |λ|)2 (60)  ( √B − 12 , √A|λ| − 12 ) √B √A ψn(−) (x, A + |λ|, −B) ∝ x |λ| (1 + λx 2 ) 2 |λ| Pn |λ| (1 + 2λx 2 ). 10

J. Phys. A: Math. Theor. 42 (2009) 285301

B Midya and B Roy

Now using (60), (59) and (56), we obtain a spectrum and eigenfunctions for V − (x, A, B) with broken SUSY as   En(−) (A, B) = (A − B + |λ| + 2n |λ|)2 − (A + B)2 (61) 1−B ( 12 − √B|λ| , √A|λ| − 12 ) √A √ ψn(−) (x, A, B) ∝ x |λ| (1 + λx 2 ) 2 |λ| Pn (1 + 2λx 2 ). A similar approach can√be applied in the case of A < 0 and B > 0. In this case, we change (A, B) into (−A, B + |λ|) and the shape invariance condition is   V+ (x, A, B) = V− (x, −A, B + |λ|) + (B − A + |λ|)2 − (A + B)2 , (62) and En(−) (A, B) = (B − A + ψn(−) (x, A, B) ∝ x

1−A √ |λ|



 |λ| + 2n |λ|)2 − (A + B)2

(1 + λx 2 ) 2

√B |λ|

( √B|λ| − 12 , √A|λ| − 12 )

Pn

(63)

(1 + 2λx 2 ).

4. Exactly solvable PT symmetric potentials in PDMSE Here we shall find exactly solvable complex potentials, some of which are related to the nonlinear oscillator potential, within the framework of PDMSE. Before we consider any particular potential, let us note that a quantum mechanical Hamiltonian H is said to be PT symmetric [28] if PT H = HPT ,

(64)

where P is the parity operator acting as spatial reflection and T stands for time reversal, acting as the complex conjugation operator. Their action on the position and momentum operators are given by P : x → −x,

p → −p,

T : x → x,

p → −p,

i → −i.

(65)

For a constant mass Schr¨odinger Hamiltonian, the condition for PT symmetry reduces to V (x) = V ∗ (−x). However in the case of position-dependent mass, an additional condition is required. To see this we note that in the present case, the Hamiltonian is of the form d2 1 m (x) d + V (x). (66) − 2m(x) dx 2 2m2 (x) dx From (65), it follows that the conditions for the Hamiltonian (66) to be PT symmetric are H =−

m(x) = m(−x),

V (x) = V ∗ (−x).

(67) 2 −1

It may be pointed out that here we are working with a mass profile m(x) = (1 + λx ) which is an even function and consequently satisfies the first condition of (67). To generate non-Hermitian interaction in the present case, we introduce a complex coupling constant. As an example, let us first consider the potential appearing in (11). It can be seen from (18) that the energy for this potential does not depend on one of the potential parameters, namely B. Thus, we consider the complex potential  √

√ √ B 2 − A2 − A λ λx 2 V (x) = + iB(2A + λ) +A . (68) 1 + λx 2 1 + λx 2 From (68), it can be easily verified that V (x) = V ∗ (−x) so that the Hamiltonian (66) with this potential is PT symmetric. In this case, the spectrum is real and given by (18). Proceeding in a similar way, we have obtained the spectrum of a number of PT symmetric potentials and the results are given in table 2. Incidentally, all the potentials in table 2 are shape invariant and the results can also be obtained algebraically. 11

V (x)

W (x)

−B 2 −A2 −A



λ

1+λx 2

√ √λx 2 + iB(2A + λ) 1+λx 2 +A

√ λx A √1+λx 2

1 + iB √1+λx 2

r1 s−n , s1

= s − n + a, s2 = s − n − a, s3 = a − n − s, s4 = −(s + n + a), s  =

En √ √ n λ(2A − n λ)

√A &r  |λ|

ψn (x) s

i n (1 + λx 2 )− 2 e−rtan

−1 (x

√ λ)

√  ix λ

 s22 √ λ 1 + √x1+λx 2 

(ir−s− 12 ,−ir−s− 12 )

Pn A2 −

B2 A2

A2 −

B2 A2

√ A(A+ λ) 1+λx 2







λx + i2B √1+λx , B < A2 2

2

1+λx − 2iB √ + λx √ 0x λ∞

√ A2 −B 2 +A λ λx 2

− iB(2A + λ)

√ 0x λ∞

√ A2 −B 2 −A |λ| 1+λx 2 1 √ − |λ|

√ 1+λx 2 2 A ,A λx 2





i BA − A √


x |λ| 2 |λ|) 1+λx 2 − A

√ − (A − n λ)2 +

2 1+λx √ x λ

A2 −

B2 A2

√ − (A + n λ)2 +

A

1+λx √ x λ

− iB x √1 λ

√ √ n λ(2A − n λ)



2

2 B√ (A−n λ)2

2 B√ (A+n λ)2



√ √ n λ(2A + n |λ|)



√ 2 − BA2 − A2 + (A + n |λ|)2 +

|λ| 1 A √x1+λx − iB √1+λx 2 2

√1 |λ| √

x |λ| 2 − 2iB 1+λx 2 − A −

x

> A2

B2 A2

√1 |λ|

B2 A2

|λ| A √x1+λx − i BA 2

2

B√ (A+n |λ|)2

 s21 √ √x λ 2 1+λx

√ λ Pn(s1 ,s2 ) √x1+λx 2

√  s23 √ 2 2 1+λx 1+λx √ √ −1 x λ x λ 1−

 s24 +1

 2 √ Pn(s3 ,s4 ) 1+λx x λ √ √ r−s r+s ( 1 + λx 2 − 1)( 2 ) ( 1 + λx 2 + 1)−( 2 ) (r−s− 12 ,−r−s− 12 ) √ Pn ( 1 + λx 2 ) √ ( s  −r  ) √ r  +s  (1 − x |λ|) 2 (1 + x |λ|)( 2 ) (s  −r  − 12 ,s  +r  − 12 ) √ Pn (x |λ|)

2 −( s 2+n ) √ λx −1 e−a |λ|x 1+λx 2

 √   |λ| Pn(−s −n−ia,−s −n+ia) −i √x1+λx 2

B Midya and B Roy

1 √ − |λ|

− iB(2A −

x

√ A(A− |λ|) 1+λx 2

√ A(A− λ) ,B 2 λx

A2 −



λx A √1+λx + i BA 2

= i √B|λ| .

J. Phys. A: Math. Theor. 42 (2009) 285301

12 Table 2. Exactly solvable PT symmetric potentials, where s = √A , r = i √B , r1 = i Bλ , a = λ λ The first four entries correspond to λ > 0 and the last two correspond to λ < 0.

J. Phys. A: Math. Theor. 42 (2009) 285301

B Midya and B Roy

5. Quasi-exactly solvable PT symmetric potentials in PDMSE The complex sextic potential in the constant mass Schr¨odinger equation has been discussed in [29]. By using transformation (12) for λ > 0, we obtain the corresponding quasi-exactly solvable potentials in PDMSE. For λ > 0, the potential is taken as V (x) =

6  ck k=1

λ

k 2

√ (sinh(x λ))−k ,

(69)

where for V (x) to be PT symmetric, c1 , c3 , c5 are purely imaginary and c2 , c4 , c6 are real. Following [29], the ansatz for the wavefunction is taken as ⎞ ⎛ 4  √ b j λ))−j ⎠ , (70) ψ(x) = f (x) exp ⎝− j (sinh(x 2 λ j =1 where f (x) is some polynomial function of x. We shall focus on the following choices of f (x): (a)f (x) = 1

√ (sinh(x λ))−1 (b)f (x) = + a0 √ λ √ √ (sinh(x λ))−1 (sinh(x λ))−2 (c)f (x) = + a0 . + a1 √ λ λ For complex potentials, a0 is purely imaginary in (b), but in (c) a1 is purely imaginary, but a0 is real. Without going into the details of calculation, which are quite straightforward, let us summarize our results. Case 1. f (x) = 1 In this case, the relation between the parameters ci and bi is found to be c1 = −3b3 + 2b1 b2 , c2 = −6b4 + 3b1 b3 + 2b22 , c4 = 8b2 b4 + 92 b32 , c5 = 12b3 b4 , c6 = 8b42 ,

c3 = 4b1 b4 + 6b2 b3 (71)

and E = b2 − 12 b12 .

(72)

Without loss of generality, we can choose c6 = 12 which fixes the leading coefficient of V (x). It gives b4 = ± 14 . Taking the positive sign to ensure the normalizability of the wavefunction, we obtain √ √ √  b3 (sinh(x λ))−3 b1 (sinh(x λ))−1 b2 (sinh(x λ))−2 − ψ(x) = exp − − √ √ λ λ λ λ √ −4  (sinh(x λ)) . (73) − 4λ2 Now if b1 and b3 are purely imaginary, then c1 , c3 , c5 are also purely imaginary. In that case, V (x) in equation (69) and ψ(x) in equation (70) are PT symmetric and E is real. 13

J. Phys. A: Math. Theor. 42 (2009) 285301

B Midya and B Roy

√ −1 (sinh(x √ λ)) λ

Case 2. f (x) = + a0 , where a0 is purely imaginary. In this case, the wavefunction is of the form

√ (sinh(x λ))−1 ψ(x) = + a0 √ λ

(74) √ √ √ √ b1 (sinh(x λ))−1 b3 (sinh(x λ))−3 b2 (sinh(x λ))−2 (sinh(x λ))−1 exp − − . − − √ √ λ 4λ2 λ λ λ In this case, the relation between the parameters is given by c1 = −6b3 + 2b1 b2 + a0 , c4 = 2b2 +

9 2 b , 2 3

c2 = − 52 + 3b1 b3 + 2b22 ,

c5 = 3b3 ,

c6 =

1 . 2

c3 = b1 + 6b2 b3

(75)

a0 satisfies the condition a03 − 3b3 a02 + 2b2 a0 − b1 = 0.

(76)

The energy is given by E = − 12 b12 + 3b2 − 3a0 b3 + a02 .

(77)

We now consider two special cases. (a) b1 = b3 = 0 and a02 < 0. √ In this case, c1 is purely imaginary and c3 = c5 = 0. Moreover, c1 = a0 = ±i 2b2 . So we get two different complex potentials corresponding to the above two values of c1 with the same real energy eigenvalues. The potential, energy values and the eigenfunctions are given by √ √ 1 (sinh(x λ))−6 2b2 V (x) = + 2 (sinh(x λ))−4 3 2 λ  2 5 λ √ √ −2 i 2b2 √ 2b2 − 2 (sinh(x λ)) ± √ (sinh(x λ))−1 + λ λ E = b2 > 0 (78)

√ −1    √ √ (sinh(x λ) 1 b2 ψ(x) = ± i 2b2 exp − (sinh(x λ))−2 − 2 (sinh(x λ))−4 . √ λ 4λ λ It can be easily seen from the above equations that the potential is PT symmetric, while the wavefunction is odd under PT symmetry. (b) b1 = 0, b3 = 0 Then from (76), we get    a0 = 12 3b3 ± 9b32 − 8b2 .

(79)

So in order to make a0 imaginary, we must have 9b32 − 8b2 < 0 or b32 = −|b3 |2  89 b2 . In this case also there exist two different complex potentials corresponding to two values of b3 with the same real energy eigenvalues E = 3b2 − 3a0 b3 + a02 . √

−2



−1

sinh(x λ)) sinh(x λ) Case 3. f (x) = ( + a1 ( √λ ) + a0 , where a1 is imaginary and a0 is real. λ In this case, the relation between the parameters is given by    2  a0 = 12 2b2 − b32 ± 2b2 − 3b32 + 2 . (80) a1 = 2b3 ,

14

J. Phys. A: Math. Theor. 42 (2009) 285301

B Midya and B Roy

The wavefunction, energy and the potential are of the form  √ √     (sinh(x λ))−2 (sinh(x λ))−1 1  2 2 2 + 2b3 ψ± (x) = + 2b2 − b3 ± 2b2 − 3b3 + 2 √ λ 2 λ √ −2 √ −1    (sinh(x λ)) (sinh(x λ)) exp −2b3 b2 − b32 − b2 √ λ λ

√ √ −3 (sinh(x λ)) 1 (sinh(x λ))−4 (81) − b3 − √ 4 λ2 λ λ     2 2 2 2 2 (82) E± = −2b3 b2 − b3 + 3b2 − b3 ± 2b2 − 3b32 + 2   √ −6 √ √ −5 2b2 + 92 1 3b3 V (x) = 3 (sinh(x λ)) + √ (sinh(x λ)) + (sinh(x λ))−4 2 2 2λ λ λ λ   2  7 2 4 √ √   2 b + 3b 2b3 2 b3 − 3b3 − 2 2 2 −3 + √ 4b2 − b3 (sinh(x λ)) + (sinh(x λ))−2 λ λ λ  2  2 √ b3 4b2 − 4b2 b3 − 7 + (83) (sinh(x λ))−1 . √ λ Results (80)–(83) are valid for both real and purely imaginary bi . When bi are purely imaginary, the potential and wavefunction are PT symmetric while for real bi PT symmetry is broken. In particular, when b3 is purely imaginary we have a complex PT symmetric two-parameter family of potentials corresponding to two values of a0 with two distinct real eigenvalues. 6. Discussion We have studied various exactly solvable as well as quasi-exactly solvable and   non-Hermitian 1 generalizations of the quantum nonlinear oscillator with the mass function 1+λx 2 . We have also obtained a closed form normalization constant for the eigenfunctions of a quantum nonlinear oscillator. A relationship between the λ-deformed Hermite polynomial and Jacobi polynomial has also been found. By exploiting the supersymmetry of PDMSE, we have obtained some shape invariant potentials corresponding to this particular mass function. We have considered the shape invariance approach to PDMSE with broken supersymmetry as well. As for the future work, we feel that it would be interesting to examine the Lie algebraic symmetry of the exactly solvable potentials. In view of the fact that in the present case transformation (12) is invertible, it seems promising to study whether or not the Lie algebraic symmetry of the constant mass system can be transported back to the non-constant mass case. Another interesting area of investigation would be to study the classical analogs of some of the models (especially the PT symmetric ones) considered here. Acknowledgments The authors would like to thank the referees for constructive criticisms which have helped us to improve the manuscript. Appendix For B = 0, A =

√α , λ

the potential of equation (11) and its energy eigenvalues (18) reduce to 15

J. Phys. A: Math. Theor. 42 (2009) 285301

For |x| <

√1 , λ

B Midya and B Roy

  α2 α2 V (x) = − − α (1 + λx 2 )−1 + λ λ

(A.1)

En = 2nα − nλ.

(A.2)

potential (A.1) can be written as   α2 α2 V (x) = − − α (1 − λx 2 + λ2 x 4 − λ3 x 6 + · · ·) + λ λ = α 2 x 2 − λ(α 2 x 4 − λα 2 x 6 + · · ·) + λ(αx 2 − λαx 4 + · · ·) − α.

(A.3)

For λ → 0, the potential reduces to V (x) = α 2 x 2 − α.

(A.4)

It is clear from (A.4) and (A.2) that for λ → 0, potential (11) and the energy eigenvalues (18) reduce to those of a simple harmonic oscillator. For A = √αλ , B = 0 and using relation (26), the expression for the wavefunction (19) is   √ λ  2 − 2α λ , (A.5) αx, ψn (x) = Nn (1 + λx ) Hn α where

1 α  n2 Nn 2n n! λ      1/2   α n αλ − n αλ − n + 12 αλ − n + 12 = .   2α 1 π n!22n− λ λn− 2 2α −n+1 λ

Nn =

(A.6)

Now for λ → 0, the λ-deformed Hermite polynomial becomes the conventional Hermite polynomial Hn [21]. Consequently, at the λ → 0 limit the unnormalized wavefunction given in equation (A.5) reduces to √ αx 2 ψn (x) ∝ e− 2 Hn ( αx). (A.7) √ 1 Using the asymptotic formula (az + b) ∼ 2π e−az (az)az+b− 2 (see 6.1.39 of [24]) in (A.6), we have √

1/2 α − √nλα  Nn = √ n . (A.8) π 2 n! Therefore from equations (A.7) and (A.8), it follows that for λ → 0 the wavefunction given in equation (19) reduces to that of a simple harmonic oscillator. References [1] Dekar L et al 1998 J. Math. Phys. 39 2551 Dekar L et al 1999 Phys. Rev. A 59 107 Bagchi B et al 2004 Czech. J. Phys. 54 1019 Bagchi B et al 2004 Mod. Phys. Lett. A 19 2765 Yu J, Dong S H and Sun G H 2004 Phys. Lett. A 322 290 Yu J and Dong S H 2004 Phys. Lett. A 325 194 Ganguly A et al 2006 Phys. Lett. A 360 228 Alhaidari A D 2002 Phys. Rev. A 66 042116 Dong S H and Lozada-Cassou M 2005 Phys. Lett. A 337 313 Chen G and Chen Z 2004 Phys. Lett. A 331 312 16

J. Phys. A: Math. Theor. 42 (2009) 285301

[2] [3]

[4]

[5] [6]

[7]

[8]

[9] [10] [11] [12] [13] [14]

[15] [16] [17] [18] [19] [20]

[21] [22] [23] [24]

B Midya and B Roy

Jiang L et al 2005 Phys. Lett. A 345 279 Gang C 2004 Phys. Lett. A 329 22 Mustafa O and Mazharimousavi Habib S 2006 Phys. Lett. A 358 259 Cruz y Cruz S et al 2007 Phys. Lett. A 369 400 Lozada-Cassou M et al 2004 Phys. Lett. A 331 45 Koc R et al 2002 J. Phys. A: Math. Gen. 35 L527 Bagchi B et al 2005 Europhys. Lett. 72 155 Roy B and Roy P 2002 J. Phys. A: Math. Gen. 35 3961 Koc R and Koca M 2003 J. Phys A: Math. Gen. 36 8105 Quesne C 2007 SIGMA 3 067 Yahiaoui S A and Bentaiba M 2007 arXiv:0711.2265v1 Yahiaoui S A and Bentaiba M 2008 arXiv:0803.4376v1 G¨onul B et al 2002 Mod. Phys. Lett. A 17 2057 Quesne C 2006 Ann. Phys 321 1221 Ganguly A and Nieto L M 2007 J. Phys. A: Math. Theor. 40 7265 Dutra AdeSouza et al 2003 Europhys. Lett. 62 8 Koc R and Tutunculer H 2003 Ann. Phys., Lpz. 12 684 Tanaka T 2006 J. Phys. A: Math. Gen. 39 219 Milanovic V and Ikonic Z 1999 J. Phys. A: Math. Gen. 32 7001 Plastino A R et al 1999 Phys. Rev. A 60 4398 Samani K and Loran F arXiv:quant-ph/0302191 Bagchi B et al 2005 J. Phys. A: Math. Gen. 38 2929 Yahiaoui S A et al 2007 arXiv:0704.3425v1 Roy B and Roy P arXiv:quant-ph/0106028 Alhaidary A D 2002 Phys. Rev. A 66 042116 Gonul B et al 2002 Mod. Phys. Lett. A 17 2453 Mustafa O and Mazharimousavi Habib S 2006 J. Phys. A: Math. Gen. 39 10537 Moayedi J K et al 2003 J. Mol. Struct. Theochem. 663 15 Yung K C and Lee J H 1994 Phys. Rev. A 50 104 Chetouani L et al 1995 Phys. Rev. A 52 82 Dong S H et al 2003 Int. J. Theor. Phys. 42 2999 Ou Y C et al 2004 J. Phys. A: Math. Gen. 37 4283 Bastard G 1988 Wave Mechanics Applied to Semiconductor Heterostructures (Les Ulis: Les Editions de Physique) Geller M R and Kohn W 1993 Phys. Rev. Lett. 70 3103 Serra L and Lipparini E 1997 Europhys. Lett. 40 667 Barranco M, Pi M, Gatica S M, Hernandez E S and Navarro J 1997 Phys. Rev. B 56 8997 Puente A, Serra L and Casas M 1994 Z. Phys. D 31 283 Preston M A 1965 Physics of the Nucleus (Reading, MA: Addison-Wesley) p 210 Wannier G H 1937 Phys. Rev. 52 191 Slater J C 1949 Phys. Rev 76 1592 Luttinger J M and Kohn W 1955 Phys. Rev. 97 869 Bachlet G B, Ceperley D M and Chiocchetti M G B 1989 Phys. Rev. Lett. 62 2088 Foulkes W M C and Schluter M 1990 Phys. Rev. B 42 505 Quesne C and Tkachuk V M 2004 J. Phys. A: Math. Gen. 37 4267 Bagchi B, Banerjee A, Quesne C and Tkachuk V M 2005 J. Phys. A: Math. Gen. 38 2929 Mostafazadeh A 2005 J. Phys. A: Math. Gen. 40 6557 Mostafazadeh A 2005 J. Phys. A: Math. Gen. 38 8185 (erratum) Mathews P M and Lakshmanan M 1974 Q. Appl. Math. 32 215 Lakshmanan M and Rajasekar S 2003 Nonlinear Dynamics, Integrability, Chaos and Patterns (Advanced Texts in Physics) (Berlin: Springer) Carinena J F et al 2004 Nonlinearity 17 1941 Carinena J F et al 2005 Regul. Chaot. Dyn. 10 423 Carinena J F et al 2007 Phys. Atom. Nucl. 70 505 Carinena J F et al 2007 Ann. Phys. 322 434 Cooper F, Khare A and Sukhatme U 2000 Suoersymmetry in Quantum Mechanics (Singapore: World Scientific) Levai G 2001 Czech. J. Phys. 51 1 Abramowitz M and Stegun I A 1964 Handbook of Mathematical Functions (New York: Dover)

17

J. Phys. A: Math. Theor. 42 (2009) 285301

B Midya and B Roy

[25] Junker G 1996 Supersymmetric Methods in Quantum and Statistical Physics (Berlin: Springer) Bagchi B K 2001 Supersymmetry in Quantum and Classical Mechanics (London: Chapman and Hall) [26] Gendenshtein L E 1983 JETP Lett. 38 356 Gendenshtein L E and Krive I V 1985 Sov. Phys. Usp. 28 645 [27] Dutt R et al 1993 Phys. Lett. A 174 363 Gangopadhyaya A et al 2001 Phys. Lett. A 283 279 [28] Bender C M and Boettcher S 1998 Phys. Rev. Lett. 80 5243 [29] Bagchi B et al 2000 Phys. Lett. A 269 79

18

A generalized quantum nonlinear oscillator

electrons in pure crystals and also for the virtual-crystal approximation in the treatment of .... solvable non-Hermitian potentials within the framework of PDMSE.

249KB Sizes 4 Downloads 256 Views

Recommend Documents

a generalized hybrid real-coded quantum evolutionary ...
International journal of computer science & information Technology (IJCSIT) ...... 1988, Masters degree in Computer Science from Asian Institute of Technology (AIT), ... includes Evolutionary Computations, Intelligent Computer Networking, ...

The Quantum Harmonic Oscillator Ladder Operators - Eli Lansey
Mar 30, 2009 - As a first example, I'll discuss a particular pet-peeve of mine, which is something covered in many introductory quantum mechanics classes: The algebraic solution to quantum (1D) simple harmonic oscillator.1 The one-dimensional, time-i

Delay stage circuitry for a ring oscillator
of Search . ..... phase deadband characteristics to be easily optimized. Another object of the present .... (ASIC), a memory controller, or a graphics engine. Master.

Examples of the Generalized Quantum Permanent ...
Sep 6, 2010 - qubits to the second register, and 1 qubit as ancillary ... all the domain of the generator. ... Emphasizing the domain QRM , the state |ψ1> can.

Phase-shifting oscillator
Flg. 2. The complete circuit for an audio oscillator. wpELESS WORLD FEBRUARY 19E. When we apply to the n.t.c. a sine wav voltage with an r. m.s. value, v,: v, ...

Yig-tuned solid state microwave oscillator
Mar 29, 1971 - device 11, which may be a GaAs bulk effect device or an Im patt diode ... of a dc. blocking capacitor 19, the other side of the capacitor. 19 being ...

Yig-tuned solid state microwave oscillator
Mar 29, 1971 - coupled to one end of a quarter wavelength transmission line. 15 on-the surface of quartz substrate 28 mounted on the base. 23. The other end of the transmission line 15 is electrically connected to an r.f. bypass capacitor 16 which is

Oscillator Model Theory
over the canonical distribution of r, pr, and pθ and substituting into eq 1, we can obtain the low-density transport ... Boltzmann factor. Equation 6 is valid when the ...

A GENERALIZED VECTOR-VALUED TOTAL ...
Digital Signal Processing Group. Pontificia Universidad Católica del Perú. Lima, Peru. Brendt Wohlberg. ∗. T-5 Applied Mathematics and Plasma Physics.

A generalized Tullock contest
Depending on the litigation system, losers have to compensate winners for a portion of their legal ... By simultaneously solving best response functions (8), and accounting for symmetric. Nash equilibrium we obtain the ... In a standard Tullock conte

Quantum criticality as a resource for quantum estimation
1Department of Physics and Astronomy, University of Southern California, Los Angeles, California 90089-0484, USA ..... lar shift of the location of the quantum critical point may be ..... 1 H. Cramer, Mathematical Methods of Statistics Princeton.

Injection-lock a Wien-bridge oscillator - Linear Technology
−15°C), it measured 270Ω. Based on the 5W, 120V spec, it should be about. 2.8k at white hot. That seemed like plenty of impedance range to stabilize a gain of ...

A generalized inquisitive semantics.
the definition of inquisitive semantics can be easily reformulated in such a way ... Recall that a P-index (or a P-valuation) is a map from P to {0, 1}, and we.

A Generalized Complementary Intersection Method ...
The contribution of this paper ... and stochastic collocation method. However, the ... Contributed by the Design Automation Committee of ASME for publication in.

A generalized inquisitive semantics.
the definition of inquisitive semantics can be easily reformulated in such a way ..... The second weak distribution law is ..... mative content in the classical way.