21
On the dynamical symmetry of the quantum Kepler problem G.F. Torres del Castillo Departamento de F ´ ısica Matem´ atica, Instituto de Ciencias Universidad Aut´ onoma de Puebla, 72570 Puebla, Pue., exic o J.L. Calvario Ac ´ ocal Inst itut o de F ´ ısica Universidad Aut´ onoma de Puebla, Apartado postal J48, 72570 Puebla, Pue., exico Abstract. Using the fact that the Schr¨ odinger equation for the two-dimensional Kepler problem with negative energy is equivalent to an integral equation on the unit sphere in the three-dimensional space, the eigenfunctions and the generators of a dynamical symmetry group for this problem are obtained from the usual spherical harmonics and the angular momentum operators on the spher e. It is shown that if the spherical harmonics are eigenfunctions of  L y , instead of  L z , the corresponding Rev. Mex . F´ ı s.  44, 344–352 (1998) 1

On the Dynamical Symmetry of the Quantum Kepler Problem

  • Upload
    weiss13

  • View
    214

  • Download
    0

Embed Size (px)

Citation preview

Page 1: On the Dynamical Symmetry of the Quantum Kepler Problem

8/11/2019 On the Dynamical Symmetry of the Quantum Kepler Problem

http://slidepdf.com/reader/full/on-the-dynamical-symmetry-of-the-quantum-kepler-problem 1/21

On the dynamical symmetry of the quantum

Kepler problem∗

G.F. Torres del Castillo

Departamento de Fısica Matem´ atica, Instituto de Ciencias

Universidad Aut´ onoma de Puebla, 72570 Puebla, Pue., Mexico

J.L. Calvario Acocal

Instituto de Fısica

Universidad Aut´ onoma de Puebla, Apartado postal J48,

72570 Puebla, Pue., Mexico

Abstract. Using the fact that the Schrodinger equation for the two-dimensional

Kepler problem with negative energy is equivalent to an integral equation on the

unit sphere in the three-dimensional space, the eigenfunctions and the generators of

a dynamical symmetry group for this problem are obtained from the usual spherical

harmonics and the angular momentum operators on the sphere. It is shown that if

the spherical harmonics are eigenfunctions of Ly, instead of Lz, the corresponding

∗Rev. Mex. Fıs. 44, 344–352 (1998)

1

Page 2: On the Dynamical Symmetry of the Quantum Kepler Problem

8/11/2019 On the Dynamical Symmetry of the Quantum Kepler Problem

http://slidepdf.com/reader/full/on-the-dynamical-symmetry-of-the-quantum-kepler-problem 2/21

eigenfunctions of the Schrodinger equation are separable in parabolic coordinates. It

is also shown that in the case of zero energy, the Schrodinger equation for the Kepler

problem in two or three dimensions is equivalent to an integral equation on the two-

or three-dimensional Euclidean space, respectively.

Resumen. Usando el hecho de que la ecuacion de Schrodinger para el problema

de Kepler en dos dimensiones con energıa negativa equivale a una ecuacion integral

sobre la esfera de radio 1 en el espacio tridimensional, se obtienen las eigenfunciones

y los generadores de un grupo de simetrıa dinamica para este problema a partir

de los armonicos esfericos usuales y los operadores de momento angular sobre la

esfera. Se muestra que si los armonicos esfericos son eigenfunciones de Ly, en lugar de

Lz, las eigenfunciones correspondientes de la ecuacion de Schrodinger son separables

en coordenadas parabolicas. Se muestra tambien que en el caso de energıa cero,

la ecuacion de Schrodinger para el problema de Kepler en dos o tres dimensiones

equivale a una ecuacion integral sobre el espacio Euclideano de dimension dos o tres,

respectivamente.

PACS: 03.65.-w; 02.20.-a; 02.30.Gp

1. Introduction

It is a well known fact that in the Kepler problem (in classical or quantum mechanics),

besides the angular momentum, there exists another conserved vector, known as the

Hermann–Bernoulli–Laplace–Runge–Lenz (HBLRL) vector; whereas the conservation

of the angular momentum follows from the invariance of the potential under rotations,

the conservation of the HBLRL vector is associated with a “hidden symmetry”, that

is, with transformations that mix the position and momentum variables, leaving the

2

Page 3: On the Dynamical Symmetry of the Quantum Kepler Problem

8/11/2019 On the Dynamical Symmetry of the Quantum Kepler Problem

http://slidepdf.com/reader/full/on-the-dynamical-symmetry-of-the-quantum-kepler-problem 3/21

Hamiltonian invariant (see, e.g., Refs. 1–5).

In the case where the energy is negative, the commutation relations (or the Pois-

son brackets) between the components of the angular momentum and of the HBLRL

vector coincide with those of a basis for the generators of rotations in the four-

dimensional Euclidean space (or of the rotations in three dimensions, if one considers

the Kepler problem in two dimensions) (see, e.g., Refs. 3–10). Fock [1] showed that

the corresponding Schrodinger equation possesses, in effect, a symmetry group iso-

morphic to the group of rotations in four dimensions by transforming this equation

into an equation on the unit sphere in the four-dimensional space, where the sym-

metry becomes obvious. This transformation allows to find the energy levels and to

express the eigenfunctions in terms of hyperspherical surface harmonics [1,3,4].

When the energy is positive, the commutation relations between the components

of the angular momentum and of the HBLRL vector correspond to those of a set of

generators of the Lorentz group and the Schrodinger equation can be transformed into

an equation on a two-sheeted hyperboloid in Minkowski space where the global action

of the Lorentz group can be seen explicitly [1,4]. Finally, when the energy is equal

to zero, the commutation relations of the angular momentum and the HBLRL vector

coincide with those of a set of generators of the Euclidean group in three dimensions

and in Ref. 4 it is shown that the Schr odinger equation can be transformed into an

equation on a paraboloid. However, since the rotations in four dimensions are the

isometries of the unit sphere and the Lorentz transformations are the isometries of an

hyperboloid in Minkowski space, one would expect that in the case of zero energy the

Schrodinger equation could be transformed into an equation on the Euclidean space,

where the symmetry of the Schrodinger equation would be manifest.

In this paper we use the fact that the the Schr odinger equation for the two-

3

Page 4: On the Dynamical Symmetry of the Quantum Kepler Problem

8/11/2019 On the Dynamical Symmetry of the Quantum Kepler Problem

http://slidepdf.com/reader/full/on-the-dynamical-symmetry-of-the-quantum-kepler-problem 4/21

dimensional Kepler problem with negative energy can be transformed into an equa-

tion on the unit sphere in three-dimensional space to find the energy levels and the

eigenfunctions explicitly, obtaining a relationship between the generating functions

of the associated Legendre functions and of the associated Laguerre polynomials. We

show that the HBLRL vector (in two dimensions) can be derived from the expres-

sions for the usual angular momentum operators and that the eigenfunctions of one of

the components of the HBLRL vector are the separable solutions of the Schrodinger

equation for the 1/r potential in parabolic coordinates. The analogue of this result

in classical mechanics is given in the Appendix. We also consider the Schrodinger

equation for the Kepler problem in two and three dimensions with zero energy, show-

ing that this equation can be transformed into one on the two- or three-dimensional

Euclidean space, respectively, whose solutions can be easily obtained.

2. The Schrodinger equation for the bound states of the

two-dimensional Kepler problem

In this section we give a treatment of the Schrodinger equation for the Kepler problem

with negative energy in two dimensions parallel to that given in Refs. 3 and 4 for the

three-dimensional case, following a simpler procedure.

4

Page 5: On the Dynamical Symmetry of the Quantum Kepler Problem

8/11/2019 On the Dynamical Symmetry of the Quantum Kepler Problem

http://slidepdf.com/reader/full/on-the-dynamical-symmetry-of-the-quantum-kepler-problem 5/21

By expressing the solution of the Schrodinger equation for the two-dimensional Kepler

problem,

− h2

2M ∇2ψ − k

rψ = Eψ, (1)

as a Fourier transform,

ψ(r) = 1

2πh Φ(p)eip·r/hd2p, (2)

using the fact that

(1/r)ei(p−p)·r/hd2r = 2πh/|p − p|, one obtains the integral

equation

( p2 − 2ME )Φ(p) = Mk

πh

Φ(p)

|p− p|d2p, (3)

where p ≡ |p|. Throughout this section, we shall consider bound states only, for

which E < 0. Then, by means of the stereographic projection, the vector p can be

replaced by a unit vector n = (nx, ny, nz) according to [1,3]

p = ( px, py) = p0 (nx, ny)1 − nz

, (4)

where

p0 ≡√ −2ME, (5)

or, equivalently,

n = (nx, ny, nz) = (2 p0 px, 2 p0 py, p2 − p2

0)

p2 + p20

. (6)

Under the correspondence between p and n given by Eqs. (4) and (6), the plane is

mapped onto the unit sphere and making use of the spherical coordinates θ , φ, of n,

from Eq. (4) we find that

p = p0

1 − cos θ(sin θ cos φ, sin θ sin φ) = p0 cot(θ/2)(cos φ, sin φ), (7)

5

Page 6: On the Dynamical Symmetry of the Quantum Kepler Problem

8/11/2019 On the Dynamical Symmetry of the Quantum Kepler Problem

http://slidepdf.com/reader/full/on-the-dynamical-symmetry-of-the-quantum-kepler-problem 6/21

therefore,

p = p0 cot(θ/2), d2p = 14 p20 csc4(θ/2)dΩ, (8)

where dΩ = sin θdθdφ is the solid angle element, and

|p− p| = p0|n− n|

(1 − nz)1/2(1 − nz)1/2

= 12 p0 csc(θ/2) csc(θ/2)|n− n|, (9)

where n is the unit vector corresponding to p according to Eq. (6). Substituting

Eqs. (5), (8) and (9) into Eq. (3) one gets

csc3(θ/2)Φ(n) = Mk

2πhp0

csc3(θ/2)Φ(n)

|n− n| dΩ,

hence, by defining

Φ(n) ≡ 2−3/2 p0 csc3(θ/2)Φ(n) = p0

p2 + p2

0

2 p20

3/2Φ(p), (10)

one arrives at the integral equation

Φ(n) = Mk

2πhp0

Φ(n)

|n

−n

|dΩ. (11)

The constant factors included in the definition (10) are chosen in such a way that Φ

is dimensionless and Φ is normalized over the sphere if and only if ψ is normalized

over the plane [3]. Since the distance between points on the sphere and the solid

angle element dΩ are invariant under rotations of the sphere, Eq. (11) is explicitly

invariant under these transformations, thus showing that the rotation group SO(3) is

a symmetry group of the original equation (1), for E < 0. Substituting Eqs. (7), (8)

and (10) into Eq. (2) one obtains the wave function ψ(r) in terms of the solution of the integral equation (11)

ψ(r) = p0

2√

2πh

Φ(θ, φ) csc(θ/2)eip0 cot(θ/2)(x cosφ+y sinφ)/hdΩ. (12)

6

Page 7: On the Dynamical Symmetry of the Quantum Kepler Problem

8/11/2019 On the Dynamical Symmetry of the Quantum Kepler Problem

http://slidepdf.com/reader/full/on-the-dynamical-symmetry-of-the-quantum-kepler-problem 7/21

The integral equation (11) can be easily solved using the fact that the spherical

harmonics form a complete set for the functions defined on the sphere, therefore the

function Φ can be expanded in the form

Φ(θ, φ) =∞l=0

lm=−l

almY lm(θ, φ). (13)

Substituting Eq. (13) into Eq. (11), making use of the expansion

1

|n− n| =∞l=0

lm=−l

2l + 1Y ∗lm(θ, φ)Y lm(θ, φ),

where θ, φ and θ, φ are the spherical coordinates of n and n, respectively, and of

the orthonormality of the spherical harmonics one obtains

∞l=0

lm=−l

1 − 2Mk

hp0(2l + 1)

almY lm(θ, φ) = 0,

which implies that, in order to have a nontrivial solution, 2Mk = hp0(2l + 1), for

some l; hence, according to Eq. (5),

E = − 2Mk2

h2(2l + 1)2, (14)

(cf. Ref. 10). The 2l + 1 coefficients alm, (m = −l, −l + 1, . . . , l) are arbitrary andalm = 0 for all l = l. Thus, the degeneracy of the energy level (14) is 2l + 1; all the

spherical harmonics of degree l are solutions of Eq. (11), corresponding to the energy

(14) and the solutions of the homogeneous integral equation (11) are precisely the

eigenfunctions of L2, the square angular momentum operator.

According to the preceding results, the solutions of the Schrodinger equation (1), for

E < 0, are given by Eq. (12),

ψ(x, y) = p0√

2πh

Φ(θ, φ)cos(θ/2)eip0 cot(θ/2)(x cosφ+y sinφ)/hdθdφ, (15)

7

Page 8: On the Dynamical Symmetry of the Quantum Kepler Problem

8/11/2019 On the Dynamical Symmetry of the Quantum Kepler Problem

http://slidepdf.com/reader/full/on-the-dynamical-symmetry-of-the-quantum-kepler-problem 8/21

where Φ(θ, φ) is an eigenfunction of L2. The rotational symmetry of the Hamiltonian

(1) suggests the use of polar coordinates, in terms of which Eq. (15) takes the form

ψ(r, ϕ) = p0√

2πh

Φ(θ, φ)cos(θ/2)eip0 cot(θ/2)r cos(ϕ−φ)/hdθdφ

= p0√

2πh

Φ(θ, φ)

∞m=−∞

im

eim(ϕ−φ)J m

( p0r/h)cot(θ/2)

cos(θ/2)dθdφ,(16)

where we have made use of the expansion eix sin θ = ∞

m=−∞ eimθJ m(x) (see, e.g.,

Ref. 11, Sec. 4). Taking Φ(θ, φ) as a spherical harmonic, Y lm(θ, φ), which are the

normalized separable eigenfunctions of L2 in spherical coordinates, from Eq. (16) we

obtain

ψlm(r, ϕ) = p0√

2πh

(−1)m

2l + 1

(l − m)!

(l + m)!

1/2P ml (cos θ)eimφ

×∞

m=−∞

im

eim(ϕ−φ)J m

( p0r/h)cot(θ/2)

cos(θ/2)dθdφ

= p0

h

2l + 1

(l − m)!

(l + m)!

1/2(−i)m

× π

0

P ml (cos θ)J m( p0r/h)cot(θ/2) cos(θ/2)dθ eimϕ, (17)

which shows that the separable eigenfunctions of L2 in spherical coordinates corre-

spond to separable eigenfunctions of the Hamiltonian in polar coordinates. Denoting

by I lm the integral between braces in Eq. (17), and introducing an auxiliary parameter

t we have

∞l=0

(2l + 1)I lmtl = π0

∞l=0

(2l + 1)P ml (cos θ)tlJ m

( p0r/h)cot(θ/2)

cos(θ/2)dθ

therefore, making use of the recurrence relation (2l+1) sin θP ml (cos θ) = P

m+1

l+1 (cos θ)−P m+1l−1 (cos θ) and of the generating function of the associated Legendre functions,

(2m)!(1 − x2)m/2

2mm!(1 − 2tx + t2)m+1/2 =

∞k=0

P mm+k(x)tk

8

Page 9: On the Dynamical Symmetry of the Quantum Kepler Problem

8/11/2019 On the Dynamical Symmetry of the Quantum Kepler Problem

http://slidepdf.com/reader/full/on-the-dynamical-symmetry-of-the-quantum-kepler-problem 9/21

(see, e.g., Ref. 12), for m ≥ 0, we find

∞l=0

(2l + 1)I lmtl = π

0

(1 − t2)tm

sin θ

(2m + 2)!(1 − cos2 θ)(m+1)/2

2m+1(m + 1)!(1 − 2t cos θ + t2)m+3/2

×J m

( p0r/h)cot(θ/2)

cos(θ/2)dθ.

Replacing the variable θ by µ ≡ cot(θ/2) one finds that

∞l=0

(2l + 1)I lmtl = ∞0

(1 − t2)tm

(1 − t)2m+3

(2m + 2)!

(m + 1)!

µm+1J m

( p0r/h)µ

µ2 +

1+t1−t

2m+3/2dµ.

The last integral can be evaluated by first differentiating with respect to s the equation ∞0

e−xsxmJ m(x)dx = (2m)!

2mm!(s2 + 1)m+1/2

(see, e.g., Ref. 11, Sec. 15), which yields

∞0

e−xsxm+1J m(x)dx = (2m + 1)!s

2mm!(s2 + 1)m+3/2. (18)

Using now the fact that ∞0

f (x)J n(xy)xdx = g(y) amounts to ∞0

g(x)J n(xy)xdx =

f (y) (see, e.g., Ref. 11, Sec. 13), from Eq. (18) one gets ∞0

(2m + 1)!

2mm!

sxm+1J m(xy)

(x2 + s2)m+3/2dx = yme−ys

thus∞l=0

(2l + 1)I lmtl = 2m+1tm( p0r/h)me− p0r/h e−(2 p0r/h)t/(1−t)

(1 − t)2m+1

and recalling thate−xz/(1−z)

(1

−z )k+1

=∞

n=0

Lkn(x)z n,

where Lkn denote the associated Laguerre polynomials (see, e.g., Ref. 13), it follows

that∞l=0

(2l + 1)I lmtl = 2m+1e− p0r/h( p0r/h)mtm∞k=0

L2mk (2 p0r/h)tk

9

Page 10: On the Dynamical Symmetry of the Quantum Kepler Problem

8/11/2019 On the Dynamical Symmetry of the Quantum Kepler Problem

http://slidepdf.com/reader/full/on-the-dynamical-symmetry-of-the-quantum-kepler-problem 10/21

therefore

I lm = 2m+1

2l + 1 e− p0r/h( p0r/h)mL2ml−m(2 p0r/h)

and the normalized eigenfunctions of the Hamiltonian, for m ≥ 0, are given by

ψlm(r, ϕ) = p0

h

2l + 1

(l − m)!

(l + m)!

1/2(−i)m

2m+1

2l + 1e− p0r/h( p0r/h)mL2m

l−m(2 p0r/h). (19)

Using the relations P −ml = (−1)m

(l − m)!

(l + m)!P ml and J −m = (−1)mJ m, from Eq. (17) it

follows that

ψl,−m = ψ∗lm. (20)

As we have shown, Eq. (15) gives a correspondence between the solutions of the

integral equation (11) and those of the Schrodinger equation (1). As remarked above,

Eq. (11) is explicitly invariant under the rotations of the sphere and, as is well known,

a set of generators of these rotations are the angular momentum operators

Lx = ih

sin φ

∂θ + cot θ cos φ

∂φ

,

Ly = ih

− cos φ

∂θ + cot θ sin φ

∂φ

, (21)

Lz = −ih ∂

∂φ,

where the ˆ indicates that these operators act on functions defined on the sphere.

Then, by means of the correspondence (12) we can find the operators on the wave

functions that correspond to the generators of rotations (21).

From Eqs. (2), (7) and (10) it follows that the function Φ on the sphere corre-

sponding to a given wave function ψ is

Φ(n) = p0

4√

2πh

1

sin3(θ/2)

ψ(r)e−ip0 cot(θ/2)(x cosφ+y sinφ)/hd2r. (22)

10

Page 11: On the Dynamical Symmetry of the Quantum Kepler Problem

8/11/2019 On the Dynamical Symmetry of the Quantum Kepler Problem

http://slidepdf.com/reader/full/on-the-dynamical-symmetry-of-the-quantum-kepler-problem 11/21

By applying, for example, Lx to both sides of the last equation one obtains

LxΦ = ip0

8√

2πh

1

sin3(θ/2)

ψ(r)

−3cot(θ/2)sin φ +

ip0

h cot2(θ/2)(2x sin φ cos φ

+ y sin2 φ − y cos2 φ) + ip0

h y

e−ip0 cot(θ/2)(x cosφ+y sinφ)/hd2r,

which can be rewritten as

h

8√

2π sin3(θ/2)

ψ(r)

3

∂y + 2x

∂x

∂y + y

∂ 2

∂y2 − ∂ 2

∂x2

− p2

0

h2 y

e−ip·r/hd2r

= h

8√ 2π sin3

(θ/2) e−ip·r/h

∂y

+ x ∂

∂x

∂y

+ y ∂ 2

∂y2

∂ 2

∂x2 −

p20

h2 yψ(r)d2r,

where we have integrated by parts. Now, assuming that ψ satisfies Eq. (1), the last

term can be replaced according to p20ψ = h2∇2ψ + (2Mk/r)ψ, hence,

LxΦ = p0

4√

2πh sin3(θ/2)

e−ip·r/h 1

p0

h2

2

∂y + 2x

∂x

∂y − 2y

∂ 2

∂x2

− M ky

r

ψ(r)d2r,

which is of the form (22), with ψ replaced by 1 p0

h2

2

∂ ∂y

+ 2x ∂ ∂x

∂ ∂y

− 2y ∂ 2

∂x2

− Mky

r

ψ =

1 p0

h2

2

x ∂ ∂y

− y ∂ ∂x

∂ ∂x

+ h2

2∂ ∂x

x ∂ ∂y

− y ∂ ∂x

− Mky

r

ψ. Thus, under the correspondence

given by Eq. (22), restricted to the solutions of Eq. (1) for a fixed value of E , ˆLx

corresponds to the operator 1 p0

h2

2

x ∂ ∂y

− y ∂ ∂x

∂ ∂x

+ h2

2∂ ∂x

x ∂ ∂y

− y ∂ ∂x

− Mky

r

, which,

apart from the constant factor (1/p0), coincides with Ay, one of the cartesian com-

ponents of the HBLRL vector

A ≡ 1

2(p× L− L× p) − M kr

r , (23)

where L = r × p and p = −ih∇ (cf. Refs. 3 and 14).

In a similar manner, one finds that the operators Ly and Lz correspond to−

(1/p0)Ax

and Lz, respectively. Thus, one concludes that the components of the HBLRL vec-

tor (23) are associated with the SO(3) symmetry of Eq. (1) and that the operators

(1/p0)Ay, −(1/p0)Ax and Lz obey the same commutation relations as Lx, Ly and Lz.

11

Page 12: On the Dynamical Symmetry of the Quantum Kepler Problem

8/11/2019 On the Dynamical Symmetry of the Quantum Kepler Problem

http://slidepdf.com/reader/full/on-the-dynamical-symmetry-of-the-quantum-kepler-problem 12/21

The Schrodinger equation (1) is also known to be separable in parabolic coordinates

(see, e.g., Ref. 6). In fact, in terms of the parabolic coordinates u and v defined by

x = 12

(u2 − v2), y = uv, Eq. (1) takes the form

− h2

2M

1

u2 + v2

∂ 2ψ

∂u2 +

∂ 2ψ

∂v2

− 2k

u2 + v2ψ = Eψ. (24)

Substituting ψ = U (u)V (v) into Eq. (24), one obtains the separated equations

− h2

2M d2

U du2 − Eu2U = (k + λ)U, − h

2

2M d2

V dv2 − Ev2V = (k − λ)V, (25)

where λ is a separation constant. Each of these separated equations, for E < 0, has

the form of the Schrodinger equation for a harmonic oscillator (cf. Refs. 6, 7 and 15);

making use of the dimensionless variables ξ ≡

p0/h u and η ≡

p0/h v, from Eqs.

(25) we get

−d2U

dξ 2 + ξ 2U =

2M (k + λ)

hp0U, −d2V

dη2 + η2V =

2M (k − λ)

hp0V,

therefore, in order to have well-behaved solutions of Eqs. (25),

M (k + λ)

hp0= n1 + 1

2,

M (k − λ)

hp0= n2 + 1

2, (26)

where n1 and n2 are non-negative integers, and

ψ = Ce−ξ2/2H n1(ξ )e−η2/2H n2(η) = Ce− p0u2/2hH n1(

p0/h u)e− p0v

2/2hH n2(

p0/h v),

(27)

where the H n are Hermite polynomials and C is a normalization constant [6].

Since H n(−x) = (−1)nH n(x) and the couples (u, v) and (−u, −v) correspond to

the same point (x, y), in order to have a single-valued wave function it is necessary

12

Page 13: On the Dynamical Symmetry of the Quantum Kepler Problem

8/11/2019 On the Dynamical Symmetry of the Quantum Kepler Problem

http://slidepdf.com/reader/full/on-the-dynamical-symmetry-of-the-quantum-kepler-problem 13/21

Page 14: On the Dynamical Symmetry of the Quantum Kepler Problem

8/11/2019 On the Dynamical Symmetry of the Quantum Kepler Problem

http://slidepdf.com/reader/full/on-the-dynamical-symmetry-of-the-quantum-kepler-problem 14/21

It should be remarked that, as we have seen in the preceding paragraph, the

conservation of the HBLRL vector follows from the separability of the Schrodinger

equation (1) in parabolic coordinates. (As shown in the Appendix, in a similar

manner, the conservation of the classical HBLRL vector follows from the separability

of the Hamilton–Jacobi equation for the two-dimensional Kepler problem.) In this

context, the HBLRL vector arises in a very natural way (compare with the derivation

of the classical HBLRL vector given, e.g., in Ref. 16).

We close this section pointing out that from the commutation relations of the

angular momentum operators (21), [Li, L j ] = ihεijk Lk, it follows, in the usual way,

that the action of Lz ± iLx on an eigenfunction of L2 and Ly yields another eigen-

function of L2 and Ly with the same eigenvalue of L2 and the eigenvalue of Ly shifted

by ±h. Therefore, owing to the results of the preceding subsection, the operators

Lz ± (i/p0)Ay (which correspond to Lz ± iLx) raise or lower the eigenvalue my of the

wave functions

ψlmy= Ce− p0u

2/2hH l+my( p0/h u)e− p0v

2/2hH l−my( p0/h v) (31)

[see Eqs. (27) and (28)] by one unit. In fact, a straightforward computation, using

again the relation p20ψ = h2∇2ψ + (2Mk/r)ψ, gives the simple expression

Lz ± i

p0Ay = ± h

i

p0

2h u ∓

h

2 p0

∂u

p0

2h v ±

h

2 p0

∂v

. (32)

The operators in the right-hand side of the last equation can be recognized as raising

or lowering operators corresponding to the linear harmonic oscillators described by

Eqs. (25). Letting a1 ≡ p0/2h u +

h/2 p0 ∂/∂u, a

1 ≡ p0/2h u −

h/2 p0 ∂/∂u,a2 ≡

p0/2h v +

h/2 p0 ∂/∂v, a†

2 ≡

p0/2h v −

h/2 p0 ∂/∂v, we have

Lz + i

p0Ay = −iha†

1a2, Lz − i

p0Ay = iha1a†

2, − 1

p0Ax =

h

2(a†

1a1−a†2a2). (33)

14

Page 15: On the Dynamical Symmetry of the Quantum Kepler Problem

8/11/2019 On the Dynamical Symmetry of the Quantum Kepler Problem

http://slidepdf.com/reader/full/on-the-dynamical-symmetry-of-the-quantum-kepler-problem 15/21

These equations correspond to the well-known Schwinger’s realization of the Lie al-

gebra of the rotation group in terms of creation and annihilation operators.

3. The quantum-mechanical Kepler problem with zero energy

When E = 0, Eq. (3) gives

p2Φ(p) = Mk

πh

Φ(p)

|p− p|d2p. (34)

Making now the change of variable

p = 2Mk

h

q

q 2, (35)

where q is a dimensionless vector in the plane and q = |q|, we get

p = 2Mkhq

, d2p = 4M 2k2

h2q 4 d2q, |p− p| = 2Mk

h|q− q|

qq ,

and from Eq. (34) it follows that

Φ(q) = 1

Φ(q)

|q− q|d2q, (36)

where

Φ(q) ≡ 2Mk

hq 3 Φ(q) =

h2 p3

4M 2k2Φ(p). (37)

Equation (36) is explicitly invariant under SE(2), the group of rigid transformations

of the Euclidean plane onto itself that do not change the orientation.

15

Page 16: On the Dynamical Symmetry of the Quantum Kepler Problem

8/11/2019 On the Dynamical Symmetry of the Quantum Kepler Problem

http://slidepdf.com/reader/full/on-the-dynamical-symmetry-of-the-quantum-kepler-problem 16/21

The homogeneous integral equation (36) can be easily solved. Substituting Φ(q) =

f (s)eis·qd2s into Eq. (36) we get

f (s)eis·qd2s = 1

f (s)eis·q

d2sd2q

|q− q| =

f (s)eis·q

s d2s

hence, f (s) = 0 only if |s| = 1; thus, Φ(q) = eis·q is a solution of Eq. (36) for any (con-

stant) unit vector s. These solutions are separable in cartesian coordinates: eis·q =

eisxxeisyy, with q = (x, y) and s = (sx, sy). Since eis·q = ∞

m=−∞ ime−imαeimθJ m(ρ),

where (ρ, θ) are the polar coordinates of q and s = (cos α, sin α), it follows that

Φ(ρ, θ) = J m(ρ)eimθ

is a separable solution in polar coordinates of Eq. (36), for anyintegral value of m. Hence, the degeneracy of the energy level E = 0 is infinite. It

may be noticed that the solutions to Eq. (36) coincide with those of the Helmholtz

equation ∇2Φ = −Φ. A somewhat similar result holds in the case of the integral

equation (11), which is equivalent to the eigenvalue equation L2Φ = l(l + 1)h2Φ,

since −h−2L2 is the Laplace operator of the sphere; however, in the latter case, all

non-negative integral values of l are allowed, while in the case of the Laplace operator

of the plane, only one of its eigenvalues is relevant, which is related to the fact thatE only takes the value zero.

From Eqs. (2), (35) and (37) it follows that the solutions of the Schrodinger

equation (1) with E = 0 are related with the functions Φ through

Φ(ρ, θ) = Mk

πh2ρ3

ψ(x, y)e−2iMk(x cos θ+y sin θ)/h2ρd2r, (38)

where, again, ρ, θ are the polar coordinates of q. On the other hand, the generators

of the rigid motions of the plane can be chosen as

P 1 = −ih

cos θ

∂ρ − sin θ

ρ

∂θ

,

P 2 = −ih

sin θ

∂ρ +

cos θ

ρ

∂θ

, (39)

16

Page 17: On the Dynamical Symmetry of the Quantum Kepler Problem

8/11/2019 On the Dynamical Symmetry of the Quantum Kepler Problem

http://slidepdf.com/reader/full/on-the-dynamical-symmetry-of-the-quantum-kepler-problem 17/21

Lz =

−ih

∂θ

.

P 1 and P 2 generate translations in the x and y directions, respectively, and Lz gen-

erates rotations about the origin. The commutation relations of the operators (39)

are

[ P 1, P 2] = 0, [Lz, P 1] = ih P 2, [Lz, P 2] = −ih P 1. (40)

The operators on the wave functions corresponding to the generators of the symmetry

transformations (39) can be obtained following the same steps as in Sec. 2.3, making

use of Eq. (38). One finds, for instance,

P 1Φ = Mk

πh2ρ3

e−ip·r/h h3

2Mk

− ∂

∂x + x

∂ 2

∂y2 − ∂ 2

∂x2

− 2y

∂x

∂y

ψ(x, y)d2r

hence, recalling that in the present case h2∇2ψ + (2Mk/r)ψ = 0, the operator corre-

sponding to the generator of translations P 1 is

h3

2Mk

− ∂

∂x + 2x

∂ 2

∂y2 +

2Mk

h2r x − 2y

∂x

∂y

= hMk

h

2

2

x ∂ ∂y − y ∂ ∂x

∂ ∂y + ∂ ∂y

x ∂ ∂y − y ∂ ∂x

+ M kxr

= − hMk Ax

[see Eq. (23)]. In a similar way, one finds that the operators corresponding to P 2 and

Lz are (−h/Mk)Ay and Lz, respectively. Thus, Ax, Ay and Lz generate symmetry

transformations of the Schrodinger equation (1), as in the case where E is negative.

The fact that Lz is the operator corresponding to Lz under the correspondence

(38) implies that, under this relationship, the separable solutions in polar coordinates

of the Schrodinger equation with E = 0 correspond to the separable solutions in polar

coordinates of Eq. (36) (which are of the form J m(ρ)eimθ). On the other hand, since

Eq. (30) holds for any value of E , the separable solutions in parabolic coordinates of

the Schrodinger equation with E = 0 are eigenfunctions of Ax and, therefore, they

17

Page 18: On the Dynamical Symmetry of the Quantum Kepler Problem

8/11/2019 On the Dynamical Symmetry of the Quantum Kepler Problem

http://slidepdf.com/reader/full/on-the-dynamical-symmetry-of-the-quantum-kepler-problem 18/21

correspond to eigenfunctions of P 1, which are the separable solutions in cartesian

coordinates of Eq. (36).

Considering now the Schrodinger equation (1) in three dimensions, writing

ψ(r) = 1

(2πh)3/2

Φ(p)eip·r/hd3p,

one obtains the integral equation

( p2 − 2ME )Φ(p) = Mk

π2h

Φ(p)

|p− p|2 d3p, (41)

which takes the place of Eq. (3). Making E = 0 and p = 2Mkq/(hq 2), where q is a

dimensionless vector in three dimensions, one finds that Eq. (41) is equivalent to

Φ(q) = 1

2π2

Φ(q)

|q− q|2 d3q, (42)

where

Φ(q) ≡

2Mkh

3/2 Φ(q)

q 4 =

h

2Mk

5/2

p4Φ(p).

Equation (42) is manifestly invariant under the rigid transformations of the three-

dimensional Euclidean space, which means that this group of transformations consti-

tute a symmetry group of the three-dimensional Kepler problem with zero energy.

It can be easily seen that the functions Φ = eis·q, where s is a constant vector

with |s| = 1, and Φ = jl(ρ)Y lm(θ, φ) are separable solutions of Eq. (42) in cartesian

coordinates and spherical coordinates, respectively, and these functions are also solu-

tions of the Helmholtz equation ∇2Φ = −Φ; the existence of these sets of separable

solutions corresponds to the separability of the Schrodinger equation (1) in parabolic

and spherical coordinates.

18

Page 19: On the Dynamical Symmetry of the Quantum Kepler Problem

8/11/2019 On the Dynamical Symmetry of the Quantum Kepler Problem

http://slidepdf.com/reader/full/on-the-dynamical-symmetry-of-the-quantum-kepler-problem 19/21

4. Concluding remarks

The cases considered in this paper, as well as those treated in Refs. 3 and 4, show the

usefulness of exhibiting the underlying symmetry of the quantum Kepler problem,

which allows to change the Schrodinger equation by a simpler condition.

The results of Sec. 2 show, among other things, that the usual spherical harmonics

are related with the associated Laguerre polynomials and the Hermite polynomials,

which form bases for representations of the rotation group; these results also show

one of the many connections between the Kepler problem with negative energy and

the isotropic oscillator (cf. also Ref. 9 and the references cited therein).

Acknowledgments

This work was supported in part by CONACYT. The authors would like to thank

Prof. K.B. Wolf for correspondence relating to the topic of this paper and Profs. H.V.

McIntosh and E.G. Kalnins for useful comments.

Appendix

The separability of the Schrodinger equation and of the Hamilton–Jacobi equation

for the two-dimensional Kepler problem in polar coordinates is a consequence of the

invariance of the Hamiltonian under rotations about the center of force (as in the

case of any central potential) which, in turn, is equivalent to the conservation of

the angular momentum Lz. The 1/r potential is distinguished by the existence of

another conserved vector —the HBLRL vector— which turns out to be related with

the separability of the above-mentioned equations in parabolic coordinates.

19

Page 20: On the Dynamical Symmetry of the Quantum Kepler Problem

8/11/2019 On the Dynamical Symmetry of the Quantum Kepler Problem

http://slidepdf.com/reader/full/on-the-dynamical-symmetry-of-the-quantum-kepler-problem 20/21

The Hamilton–Jacobi equation for the two-dimensional Kepler problem written

in the parabolic coordinates u, v , defined by x = 12(u2 − v2), y = uv, is

1

2M

1

u2 + v2

∂S

∂u

2

+

∂S

∂v

2 − 2k

u2 + v2 +

∂ S

∂t = 0. (A.1)

Even though both coordinates are non-ignorable, Eq. (A.1) admits separable solutions

of the form

S (u,v,t) = −Et + f (u) + g(v). (A.2)

In fact, substituting Eq. (A.2) into Eq. (A.1) one obtains the separated equations

1

2M

df

du

2

− k − Eu2 = λ, 1

2M

dg

dv

2

− k − Ev2 = −λ, (A.3)

where λ is a separation constant. Multiplying the first equation in (A.3) by v2 and

the second one by u2 and subtracting one finds that

1

2M

v2

df

du

2

− u2

dg

dv

2 + k(u2 − v2) = λ(u2 + v2), (A.4)

therefore, since u2 + v2 = 2r, df/du = ∂S/∂u = pu = px(∂x/∂u) + py(∂y/∂u) =

upx + vpy and dg/dv = ∂S/∂v = pv = −vpx + upy, from Eq. (A.4) it follows that

λ = − 1

M

(xpy − ypx) py − M kx

r

= −Ax

M ,

where A ≡ p × L − Mkr/r now denotes the classical HBLRL vector (cf. Eq. (30)).

From Eqs. (A.3) it follows that, if E ≤ 0, then −k ≤ λ ≤ k (cf. Eqs. (26)).

In a similar way one finds that Ay corresponds to a separation constant using the

coordinate system y = 12

(v2 − u2), x = uv.

References

1. V. Fock, Z. Physik 98 (1935) 145.

20

Page 21: On the Dynamical Symmetry of the Quantum Kepler Problem

8/11/2019 On the Dynamical Symmetry of the Quantum Kepler Problem

http://slidepdf.com/reader/full/on-the-dynamical-symmetry-of-the-quantum-kepler-problem 21/21

2. V. Bargmann, Z. Physik 99 (1936) 576.

3. M. Bander and C. Itzykson, Rev. Mod. Phys. 38 (1966) 330.

4. M. Bander and C. Itzykson, Rev. Mod. Phys. 38 (1966) 346.

5. H.V. McIntosh, in Group theory and its applications , Vol. 2, E.M. Loebl (ed.),

(Academic Press, N.Y., 1971).

6. A. Cisneros and H.V. McIntosh, J. Math. Phys. 10 (1969) 277.

7. A. Cisneros and H.V. McIntosh, J. Math. Phys. 11 (1970) 870.

8. M.J. Englefield, Group Theory and the Coulomb Problem , (Wiley, N.Y., 1972).

9. O.L. de Lange and R.E. Raab, Operator Methods in Quantum Mechanics , (Ox-

ford University Press, Oxford, 1991).

10. G.F. Torres del Castillo and J.L. Calvario Acocal, Rev. Mex. Fıs. 43 (1997) 649.

11. H. Hochstadt, The Functions of Mathematical Physics , (Wiley-Interscience, New

York, 1971), reprinted by Dover (New York, 1986), Chap. 8.

12. G. Arfken, Mathematical Methods for Physicists , 3rd ed. (Academic Press, San

Diego, 1985), Chap. 12.

13. H. Hochstadt, op. cit., Chap. 2.

14. K.B. Wolf, Supl. Nuovo Cimento 5 (1967) 1041.

15. M. Moshinsky, T.H. Seligman and K.B. Wolf, J. Math. Phys. 13 (1972) 901.

16. H. Goldstein, Classical Mechanics , 2nd ed. (Addison-Wesley, Reading, Mass.,

1980).

21