The Supersymmetric Dirac Equation
The Supersymmetric Dirac Equation
DIRAC EQUATION
The Application to Hydrogenic Atoms
This page intentionally left blank
THE SUPERSYMMETRIC
DIRAC EQUATION
The Application to Hydrogenic Atoms
Allen Hirshfeld
Technical University of Dortmund, Germany
Distributed by
World Scientific Publishing Co. Pte. Ltd.
5 Toh Tuck Link, Singapore 596224
USA office: 27 Warren Street, Suite 401-402, Hackensack, NJ 07601
UK office: 57 Shelton Street, Covent Garden, London WC2H 9HE
Paul Dirac photograph by A. Börtzells Tryckeri, courtesy of AIP Emilio Segre Visual Archives,
E. Scott Bar and Weber Collections.
For photocopying of material in this volume, please pay a copying fee through the Copyright
Clearance Center, Inc., 222 Rosewood Drive, Danvers, MA 01923, USA. In this case permission to
photocopy is not required from the publisher.
ISBN-13 978-1-84816-797-1
ISBN-10 1-84816-797-0
Printed in Singapore.
I wish to dedicate this book to the memory of my father, Dr. Martin A. Hirshfeld,
who taught me many things, among them the meaning of proof.
v
August 27, 2011 9:21 9in x 6in The Supersymmetric Dirac Equation b1217-fm
Preface
One morning, when I was an undergraduate, over a cup of coffee in the cafeteria
of the Weizman Institute, where he was working on his PhD in physics, Charles
Robinson opened my eyes to the wonders of the Dirac equation, and sowed the
seed for this book.
I met Paul Dirac in the early 1960s, on the occasion of his visit to the Physics
Department of the Tel-Aviv University. We talked about my physics project at that
time (I was working on my Master’s thesis), and I attended his talk, where he
explained the difficulties in the foundations of quantum field theory. Afterwards
he autographed his book for me.
I am very grateful to Jens Peder Dahl, Professor Emeritus of the Chemistry
Department of the Technical University of Denmark, for the fruitful and
encouraging correspondence on his work.
I wish to thank Dirk Fischer for his assistance with the technical aspects of
the preparation of the manuscript, and especially for his help in the preparation of
the figures.
vii
August 27, 2011 9:21 9in x 6in The Supersymmetric Dirac Equation b1217-fm
Contents
Preface vii
1. Introduction 1
ix
August 27, 2011 9:21 9in x 6in The Supersymmetric Dirac Equation b1217-fm
Contents xi
Appendices 171
Appendix A. The Confluent Hypergeometric Function . . . . . . . . 171
Appendix B. Orthogonality Relations of Hypergeometric
Functions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 175
Appendix C. More Integrals Involving Hypergeometric
Functions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 179
Appendix D. Normalization for the Γ-Induced Scheme . . . . . . . 182
Appendix E. Normalization for the ∆-Induced Scheme . . . . . . . 185
Bibliography 195
Index 199
August 27, 2011 9:21 9in x 6in The Supersymmetric Dirac Equation b1217-fm
List of Figures
xiii
August 27, 2011 9:21 9in x 6in The Supersymmetric Dirac Equation b1217-fm
Chapter 1
Introduction
“He succeeded in combining quantum mechanics and the theory of relativity, . . . and
introduced the idea of an antiparticle . . . Unfortunately, only the expert can appreciate
Dirac’s awesome work. To this day, that work has not lost any of its splendor.”
Martinus Veltman, Facts and Mysteries in Elementary Particle Physics, World Scientific
Publishing, Singapore, 2003.
The sentiment expressed in this quotation is certainly true. But let us pause to
consider the phrase “only the expert can appreciate Dirac’s awesome work . . . ”. We
expect a student beginning to study quantum field theory to understand the working
of relativistic quantum theory as it applies to free particles. The next interesting
case, that of a particle in a Coulomb field, is given at best a somewhat cursory
treatment in most texts. But it was precisely this problem that was the test of the
new theory — could it explain the fine structure in the spectrum of hydrogen? The
fact of this success and the detailed structure that it predicts were the hallmark of
the theory for many years. It was the work devoted to testing this structure, with the
highest possible precision, which revealed the need for the radiative corrections of
quantum electrodynamics and provided us with the most accurate evaluation of the
fine structure, which we possess to the present day, and one of the most impressive
successes in all of physics.
It is thus highly desirable that the modern student of relativistic quantum
mechanics should have an appreciation of this decisive episode in the physics
of the twentieth century. The advances of the theory in recent years have led to
certain simplifications with respect to Dirac’s time. In particular, the discovery that
supersymmetry plays a central role in the structure of the theory of a particle in
a Coulomb field means that while gaining a mastery of this topic, students are
at the same time gaining experience that will be of importance in appreciating
the questions with which their future work will be concerned, namely, does
supersymmetry play a role in the structure of the universe? Even if this question is,
contrary to expectation, answered in the negative, the knowledge of supersymmetry
1
August 27, 2011 9:21 9in x 6in The Supersymmetric Dirac Equation b1217-ch01
Introduction 3
The beginning of the “new quantum theory” was marked, as a matter of fact, by
Pauli’s [Pau26] group theoretical treatment of the hydrogen atom. Pauli’s analysis
is based on the existence of the Laplace vector A as a conserved quantity of the
motion of a point particle moving under the influence of an inverse-quadratic force.
Its three components form, together with the three components of the orbital angular
momentum, the generators of the Lie algebra so(4). The quadratic Casimir operator
of this algebra then yields a formula for the spectrum of the bound-state energy
levels of hydrogenic atoms. The advantage of this approach over Schrödinger’s
[Sch26] method of solving a differential equation is that it automatically yields
an understanding of the degeneracy of the spectral lines, which is treated in the
Schrödinger approach as an “accidental” symmetry.
A full analysis of the problem of an electron with spin in a Coulomb field
involves the two-component Pauli formalism. It is only necessary to replace p by
(σ · p) to get the correct gyromagnetic ratio. The inclusion of the spin variables
provides a formalism in which quadratic operators of the theory factorize. The
effectiveness of such factorizations was already noted by Hull and Infeld [HI51],
whose work was in a sense a precursor of the theory of supersymmetry, which arose
originally in the quantum field theory of elementary particles. The Schrödinger
equation for the radial component of the wave function is seen to be the product of
two linear differential operators, whose eigenvalues are related to the magnitude of
the operator Anr = (σ ·A), in a way that is the quantum version of the eccentricity
formula e = |A|. Couching the results of the treatment of spin in a supersymmetric
language leads to an understanding of the full degeneracy of the hydrogen spectrum
as a consequence of the SO(4) × S(2) supersymmetry group.
When we ask for an understanding of the fine structure of the spectrum we
move on to the treatment of the relativistic spectrum. Today we undertake a study of
the solutions of the Dirac equation describing an electron in a Coulomb potential.
In this book we first follow in the footsteps of Sommerfeld, who undertook a
treatment of the relativistic problem, although he had at his disposal neither the
modern quantum theory nor the notion of electron spin. It is all the more astonishing
that he obtained a formula for the spectrum of hydrogen that agrees with the results
of the Dirac theory. The Sommerfeld theory not only gives the correct formula for
the energy of the hydrogen orbits, it also describes their relativistic eccentricity.
One of the purposes of this book is to explain this success, based on the work of
Biedenharn [Bie83].
When we go over to the relativistic problem the SO(4) symmetry breaks down,
but a remnant of the Laplace vector survives, namely the Johnson–Lippmann
operator A [JL50]. The Johnson–Lippmann operator is, in the non-relativistic
August 27, 2011 9:21 9in x 6in The Supersymmetric Dirac Equation b1217-ch01
approximation, the operator Anr . In this way the degeneracy of the spectrum
reduces to that of SO(3) × S(2), with SO(3) the symmetry generated by the total
angular momentum J = L + S, where S is the spin vector that generates the SO(3)
rotation group, and S(2) is the supersymmetry group.
By describing the Dirac theory in a supersymmetric framework we are able to
derive the solutions directly, without resorting to the heuristic methods employed
formerly. Here we follow the work of Dahl and Jørgensen [DJ95]. The solutions
of the Dirac equation are obtained by investigating two different extensions of
the solution space. We describe the relation of the solutions to the second-order
Kramer’s equation, which was used formerly to construct solutions of the Dirac
equation.
Finally, the non-relativistic approximations of the solutions found for the Dirac
equation are shown to reduce to the solutions of the Pauli equation. This justifies
the choice of sign for the solutions, and allows a check on the various normalization
factors used in this book.
The Appendices deal with some of the features of the confluent hypergeometric
functions, and their relation to the wave functions of the Pauli and Dirac equations.
The SO(3) × S(2) symmetry of the Dirac theory is broken by the radiative
corrections of quantum field theory, notably in the Lamb shift [LR47], which
describes the energy difference between the 2s 12 and 2p 12 levels of hydrogen.
Because of this, the Dirac theory of bound states belongs in a course of study after
non-relativistic quantum mechanics and before the introduction to the quantum
theory of interacting fields. It is for such students that the present book is intended.
A note on units
This book uses the system of units in which and the velocity of light are set to
unity: = c = 1, where = h/(2π), and h is Planck’s constant. The Einstein
convention is used throughout, according to which an index that appears in an
expression twice is summed over. The fine structure constant is α = e2 /4π, where
e is the charge of the electron.
August 27, 2011 9:21 9in x 6in The Supersymmetric Dirac Equation b1217-ch02
Chapter 2
The Classical Kepler Problem
“It is not eighteen months since I first caught a glimpse of the light, three months since
the dawn, very few days since the unveiled Sun, most admirable to gaze upon, burst upon
me. Nothing restrains me; I shall indulge my sacred fury; I shall triumph over mankind by
the honest confession that I have stolen the golden vases of the Egyptians to build up a
tabernacle for my God far from the confines of Egypt. If you forgive me, I rejoice; if you are
angry, I can bear it; the die is cast, the book is written, to be read either now or by posterity,
I care not which; it may well wait a century for a reader, as God himself has waited six
thousand years for someone to behold his work.”
We are all familiar with the modern understanding of our solar system. The
Sun sits at the center of a ring of essentially concentric orbits, each planet following
an elliptical path around the Sun, which sits not at the center but at one focus of
the ellipse. It is the picture shown in Figure 2.1. No one has ever seen this picture.
Even today there are no satellites sitting at that vantage point tracking the orbits
and plotting their paths, much less in the seventeenth century when this picture was
first proposed by Johannes Kepler. Kepler and the astronomers before and since
him have constructed their understanding of the solar system by watching the Sun
and planets, observable as points of light, move across the sky. We only see this
two-dimensional projection of their paths.
Kepler’s “unveiled Sun” was the realization that the observational data of
Tycho Brahe, his mentor, could be explained by the following three laws:
Kepler’s laws of motion were explained by Newton. Newton was able to show
that they could be deduced from his universal law of gravitation, involving a central
5
August 27, 2011 9:21 9in x 6in The Supersymmetric Dirac Equation b1217-ch02
force that falls off as the square of the distance. Taking the mass of the central
attractor as infinite and the planet as a point particle leads to the problem of a point
mass in a central potential field. The motion takes place in a fixed plane because of
the conservation of angular momentum. The fact that the major axis of the ellipse
remains stationary for a 1/r potential was known, by the time of Laplace, to be due
to the conservation of the Laplace (Runge–Lenz) vector [Gol75], whose magnitude
is equal to the eccentricity of the ellipse. With the model of an atom as a miniature
solar system these concepts became relevant to the motion of electrons about the
nucleus.
dr dr dr̂ dr dθ
v= = r̂ + r = r̂ + r θ̂, (2.1)
dt dt dt dt dt
where r = rr̂ is the position vector and r̂ is a unit vector on the direction of r.
θ̂ is a unit vector perpendicular to r̂:
The acceleration is
dv d2 r dr dθ d2 θ dθ dθ
a= = 2 r̂ + 2 θ̂ + r 2 θ̂ + r − r̂
dt dt dt dt dt dt dt
2
d2 r dθ 1 d 2 dθ
= − r r̂ + r θ̂. (2.3)
dt2 dt r dt dt
Equation (2.3) tells us that the acceleration is entirely radial, that is, parallel
to r, if and only if
d dθ
r2 = 0, (2.4)
dt dt
which is equivalent to saying that r(t)2 dθ/dt is a constant independent of t.
Combining this with the next lemma gives a proof of Kepler’s second law:
Lemma. If the position of a particle over time is described by the vector function
r(t), then the rate at which the radial vector sweeps out area is given by
dA r(t)2 dθ
= . (2.5)
dt 2 dt
Proof. Given a circle with center at the origin and radius r, the area swept out by
the radius as it moves through an angle of ∆θ is given by (r 2 /2)∆θ. It follows
that if ∆A is the area swept out by the radial vector from time s to time t, and if
the distance from the origin stays constant during this time interval, then
r2
∆A = ∆θ, (2.6)
2
where ∆θ = θ(t) − θ(s). If r does not stay constant, then we can find two points in
the interval [s, t], call them t1 and t2 , where r takes on its minimum and maximum
values, respectively, over this interval: r(t1 ) ≤ r ≤ r(t2 ). It follows that
r(t1 )2 r(t2 )2
∆θ ≤ ∆A ≤ ∆θ. (2.7)
2 2
We now divide by ∆t = t − s:
r(t1 )2 ∆θ ∆A r(t2 )2 ∆θ
≤ ≤ , (2.8)
2 ∆t ∆t 2 ∆t
and take the limit as s approaches t. This forces t1 and t2 to also approach t and
yields
r(t)2 dθ dA r(t)2 dθ
≤ ≤ . (2.9)
2 dt dt 2 dt
August 27, 2011 9:21 9in x 6in The Supersymmetric Dirac Equation b1217-ch02
Lemma. Let r(t) be the position of a particle at time t, and v(t) its velocity. The
momentum of the particle is p = mv, where m is its mass. If the particle is moving
under the influence of a central force:
dp
F = = f (r)r̂, (2.10)
dt
where f is an arbitrary function of r, then the orbital angular momentum
L=r×p (2.11)
L · r = (r × p) · r = (r × r) · p = 0, (2.16)
L2 = L · (r × p) = (p × L) · r. (2.20)
Now
d dp k 2 dθ
(p × L) = ×L= − r̂ × mr r̂ × θ̂
dt dt r2 dt
dθ dθ d
= −mk [r̂ × (r̂ × θ̂)] = mk θ̂ = (mk r̂). (2.21)
dt dt dt
This means that the time derivative of p × L − mk r̂ is zero, so
1
A= (p × L) − r̂ (2.22)
mk
is a constant vector, and
or
L2
|r| + r · A = . (2.25)
mk
Equation (2.19) follows from the equalities |r| = r and r · A = re cos θ.
An ellipse with foci F1 and F2 can be defined as the locus of points P for
which |F1 P | + |F2 P | = 2a, with a the semimajor axis.
The point P is described
by the vector function r(t) = r(t)r̂(t), with r(t) = x(t)2 + y(t)2 and r̂(t) =
(cos θ(t), sin θ(t)). See Figure 2.2.
August 27, 2011 9:21 9in x 6in The Supersymmetric Dirac Equation b1217-ch02
a
F1 F2
b
Lemma. The general equation of an ellipse with one focus at the origin and
semimajor axis along the x-axis is given in polar coordinates by
where e and c are constants, 0 < e < 1, c > 0. The semimajor axis is
c
a= , (2.27)
1 − e2
and the semiminor axis is
c
b= √ . (2.28)
1 − e2
The eccentricity of the ellipse is
e= 1 − (b/a)2 . (2.29)
r = c − er cos θ. (2.30)
x2 + y 2 = c2 − 2ecx + e2 x2 , (2.31)
or
or
c2
(1 − e2 )(x + ea)2 + y 2 = , (2.34)
1 − e2
which may be rewritten as
(x + ea)2 y2
+ = 1. (2.35)
a2 b2
This is the equation of an ellipse in Cartesian coordinates. The center of the ellipse
is at (−ae, 0), which is a distance ae from the origin. The foci are bisected by a
line that
intersects the ellipse at two points. At these points the distance from the
foci is (ae)2 + b2 = a, and the sum of the distances from the foci, which is the
same for all points of the ellipse, is 2a.
The perigee, which is the point of closest approach, corresponds to θ = 0, so
L2
rmin = a(1 − e) = . (2.36)
mk(1 + e)
L2
rmax = a(1 + e) = . (2.37)
mk(1 − e)
A · L = 0, (2.38)
so A lies in the plane of motion of r. As a matter of fact, A lies along the line
joining the two foci.
We express the eccentricity in terms of the constants of the motion, the energy
E, and the angular momentum L. We use the identities
(a) (p × L) · (p × L) = p2 L2 ,
(b) (p × L) · r = L · (r × p) = L2 . (2.39)
We then have
p2 L2 2L2 2L2 p2 k 2L2 E
|A| = 2 2 −
2
+1 = − +1 = +1 = e2 , (2.40)
m k mkr mk2 2m r mk2
August 27, 2011 9:21 9in x 6in The Supersymmetric Dirac Equation b1217-ch02
where
p2 k
E= − (2.41)
2m r
is the constant energy of a particle in a fixed orbit.
From the equation for the perigee we have
L2
rmin = a(1 − e) = , (2.42)
mk(1 + e)
or
L2
1 − e2 = . (2.43)
mka
Comparing this to Eq. (2.40):
2L2 E
1 − e2 = − , (2.44)
mk 2
we get
k
E=− . (2.45)
2a
The energy of a particle in orbit depends only on the semimajor axis of the ellipse.
We can now derive Kepler’s third law. By Kepler’s first law a particle
√ in orbit
moves on an ellipse. The semiminor axis of the ellipse is b = a 1 − e2 . By
Kepler’s second law the area of the ellipse is
τ
dA L
A= dt = τ, (2.46)
0 dt 2m
The square of the period of the orbit is therefore proportional to the cube of a,
which is the mean distance.
August 27, 2011 9:21 9in x 6in The Supersymmetric Dirac Equation b1217-ch02
Notes on Chapter 2
The conservation of the Laplace vector and the interpretation of its length as
eccentricity are first mentioned by Jacob Hermann, in a letter to Bernoulli. It
was then used by Laplace in his treatise on mechanics. It was rediscovered by
Hamilton in the nineteenth century, and included in textbooks by Gibbs and Runge.
Pauli learned about it from his teacher Lenz, and after Pauli’s quantum mechanical
treatment of the hydrogen atom (Chapter 8) it is commonly referred to as the
Runge–Lenz vector. The early history of the Laplace vector was traced by Goldstein
[Gol75] [Gol76]. See also the Wikipedia page at http://en.wikipedia.org/wiki/
Laplace–Runge–Lenz_vector. The presentation here follows Bressoud [Bre91].
The origin of the Laplace vector as the boost-generator of the relativistic extension
of the theory has been proposed by Dahl [Dah95].
August 27, 2011 9:21 9in x 6in The Supersymmetric Dirac Equation b1217-ch02
Chapter 3
Symmetry of the Classical Problem
“Symmetry, as wide or as narrow as you may wish to define it, is one idea by which man
through the ages has tried to comprehend and create order, beauty, and perfection.”
15
August 27, 2011 9:21 9in x 6in The Supersymmetric Dirac Equation b1217-ch03
for all n, so
−1
et(RAR )
= Re tA R−1 . (3.5)
In particular, the tangent vector to RetA R−1 , which goes through the identity I at
t = 0, is RAR −1 .
Suppose A is fixed and R(s) is a curve of matrices that passes through I at
s = 0 and is differentiable in s. Then R(s) is invertible for s close to the identity,
and differentiating
R(s)R(s)−1 = I (3.6)
at s = 0 gives
d −1 d
R (s)|s=0 = − R(s)|s=0 = −B, (3.7)
ds ds
and
d
[R(s)AR(s)−1 ]s=0 = BA − AB , (3.8)
ds
by Leibniz’s rule. Define the Lie bracket [ , ] by
[B, A] = BA − AB . (3.9)
B + B T = 0, (3.11)
so B is antisymmetric. The dimension of the tangent space, and hence the manifold
O(n), is thus n(n − 1)/2. O(3) is a three-dimensional subgroup of the nine-
dimensional group of all invertible 3 × 3 matrices. O(4) is six-dimensional.
Conversely, let A be an antisymmetric matrix.
T T
etA = etA , (3.12)
But it is more instructive to derive this from the statements above. Indeed, let B ∈ G.
Then R(s) = esB is a curve of matrices in G. Hence for each fixed s, conjugation
by R(s) carries G into itself, and carries I into itself, and hence must carry the
tangent space G into itself. Hence if A ∈ G, the curve C(s) = R(s)AR(s)−1 lies
in G; C(s) ∈ G for all s. But then C(s + h) − C(s) ∈ G for any s and h. Dividing
by h and passing to the limit implies C (s) ∈ G. At s = 0 this tells us that for
B ∈ G and A ∈ G also [B, A] ∈ G.
August 27, 2011 9:21 9in x 6in The Supersymmetric Dirac Equation b1217-ch03
The corresponding facts are true in general: If G is a linear Lie group then
its tangent space G at I is the space spanned by A ∈ G, and the curve etA is a
one-parameter subgroup of G. We say that G is the Lie subalgebra of the space of
all matrices.
More generally, a linear space G with a bracket mapping G ×G → G is called a
Lie algebra if the bracket is bilinear, antisymmetric, and satisfies the Jacobi identity.
A group G that is a finite-dimensional manifold, where the group multiplication
is consistent with the manifold structure in an obvious way, is called a Lie group.
The tangent space to G at I (in the sense of differentiable manifolds) then inherits
the structure of a Lie algebra. Given any finite-dimensional Lie algebra G over
the real numbers, there always exists a Lie group G having G as its Lie algebra.
All such G will be “locally isomorphic”, but may differ in their global structure.
There is a unique G that is connected and simply connected having G as its Lie
algebra.
o(3)
Let δi denote infinitesimal rotations about the xi -axis, i = 1, 2, 3, so that
0 0 0 0 0 1 0 −1 0
δ1 = 0 0 −1, δ2 = 0 0 0, δ3 = 1 0 0. (3.17)
0 1 0 −1 0 0 0 0 0
Then
The δi are clearly linearly independent and so form a basis of o(3). So if we write
the most general element of o(3) as a = a1 δi + a2 δ2 + a3 δ3 then the preceding
equations for the brackets of the basis elements show that the Lie product in o(3)
August 27, 2011 9:21 9in x 6in The Supersymmetric Dirac Equation b1217-ch03
[a, b] = a × b. (3.20)
Notice that this operation can also be described by thinking of a as a matrix and b
as a vector and multiplying the vector b by the matrix a. For example, the fact that
[δ1 , δ2 ] = δ3 can be expressed as
0 0 0 0 0
0 0 −1 1 = 0. (3.21)
0 1 0 0 1
e(3)
Let E(3) denote the group of all Euclidean motions in three-dimensional space. So
E(3) consists of all transformations of R3 of the form
r → Rr + v, (3.22)
with a ∈ o(3) and b ∈ R3 . But recall from Eq. (3.20) that we have an identification
of o(3) with R3 . Therefore we can identify the six-dimensional algebra e(3) as a
vector space with
If [ , ]o(3) denotes the bracket in the Lie algebra o(3) then the Lie bracket in e(3),
which is just the commutator of 4 × 4 matrices restricted to matrices in e(3) as
August 27, 2011 9:21 9in x 6in The Supersymmetric Dirac Equation b1217-ch03
above, is
The bracket of two elements of the form (a, 0) is just the o(3) bracket in the a
position, that is
o(4)
This is the algebra of all antisymmetric 4 × 4 matrices, so a basis is given by
0 0 0 0 0 0 1 0 0 −1 0 0
0 0 −1 0 0 0 0
δ1 = , δ2 = 0 0 0 0, δ3 = 1 ,
0 1 0 0 −1 0 0 0 0 0 0 0
0 0 0 0 0 0 0 0 0 0 0 0
0 0 0 1 0 0 0 0 0 0 0 0
0 0 0 0 0 0 0 1 0 0
ω1 = , ω3 = 0 0 ,
0 0 0 0, ω2 = 0 0 0 0 0 0 0 1
−1 0 0 0 0 −1 0 0 0 0 −1 0
(3.31)
with a ∈ o(3) and b ∈ R3 . Once again, this six-dimensional Lie algebra can be
identified as a vector space with o(3)⊕ o(3). The bracket of two as is the same as
before, as is the bracket of an a and a b. What is new is the bracket of two bs. In
the algebra o(4) we have
o(3,1)
This is the algebra of 4 × 4 matrices of the form
a b
, (3.34)
bT 0
with a ∈ o(3) and b ∈ R3 . It can be shown that this is the Lie algebra of the Lorentz
group — the group of all linear transformations of R4 that preserve the quadratic
form
Once again the bracket between two as and an a and a b is as before. But now
In all three cases, e(3), o(4), o(3, 1), the Lie algebra decomposes as a vector
space direct sum
G = k + p, (3.37)
with
In these three cases k = p = o(3) as vector spaces, and the [k, k] and [k, p] brackets
are the o(3) brackets. The bracket of p back into k is given by λ[ , ]o(3) , where
λ = 1 for o(4), λ = 0 for e(3), and λ = −1 for o(3, 1).
It is notable that in the o(4) case we have k = p as vector spaces, and when we
make this vector space identification all three brackets are the same: The subspaces
and
are both subalgebras isomorphic to k and we have the Lie algebra direct sum
decomposition
G = ∆ ⊕ ∆T . (3.41)
Thus o(4) is Lie algebra isomorphic to o(3)⊕ o(3). This identification in Lie theory
has remarkable physical consequences as will become apparent in Chapter 8.
August 27, 2011 9:21 9in x 6in The Supersymmetric Dirac Equation b1217-ch03
The first three axioms are just the axioms for a Lie algebra: The space of functions
on phase space form a Lie algebra under the Poisson bracket. Axiom (d) ties the
Poisson bracket in with usual multiplication. It says that the Poisson bracket acts
as a derivation. A commutative algebra A having a bracket satisfying the above
axioms is called a Poisson algebra.
For example, the phase space X for a single classical particle moving in
ordinary three-dimensional space is the six-dimensional space consisting of all
possible positions and momenta. So X is six-dimensional, with coordinates
(x1 , x2 , x3 , p1 , p2 , p3 ). The Poisson brackets of these coordinate functions are
When these bracket relations are realized we say that the coordinates p1 , p2 , p3 are
the canonically conjugate variables to x1 , x2 , x3 . Using the axioms we find
and, by iteration,
Therefore
∂g
{xi , g} = , (3.45)
∂pi
and similarly
∂g
{pi , g} = − , (3.46)
∂xi
August 27, 2011 9:21 9in x 6in The Supersymmetric Dirac Equation b1217-ch03
So if we identify the function xk with the kth basis element in R3 then the first
equation in (3.54) identifies the bracket with σ(M ) with the matrix Mij and so
σ defines an isomorphism between the Lie algebra gl(3, R) and the subalgebra of
A consisting of all homogeneous polynomials of degree two, which are of degree
one in x and degree one in p. For example, if L3 = x1 p2 − x2 p1 then
and
The first of these equations says that L3 acts on r as an infinitesimal rotation about
the x3 -axis. Let o(3)∈ gl(3, R) be the Lie algebra of infinitesimal rotations and let
L denote the restriction of σ to o(3). Recall that δ3 denotes infinitesimal rotation
about the x3 -axis (with similar notation for δ1 and δ2 ) and then L3 = L(δ3 ).
Let us use × to denote the vector product on R3 . We have made the
identification of o(3) with R3 so that the Lie bracket is given by ×. That is, the
elements of o(3) are regarded as vectors and the Lie bracket becomes the vector
product. The elements of δi then become the standard basis vectors. If we form the
vector-valued function
L=r×p (3.57)
L(ξ) = L · ξ, (3.58)
under this time evolution is thus given by {H, f }. We say that H is invariant under
the action σ of the Lie algebra G if
{σ(ξ), H} = 0 (3.60)
Thus invariance of the Hamiltonian under the full translation group implies
conservation of the vector-valued function p. This is the principle of the
conservation of linear momentum.
Thus invariance of the Hamiltonian under the full group of Euclidean motions
containing all translations and rotations implies conservation of both linear and
angular momentum. One of Galileo’s principal revolutionary achievements was to
replace the geocentric Aristotelian theory, which has an O(3) symmetry centered
at the origin of the Earth, with the larger symmetry group E(3). Indeed, it is easy
to check that the map
(a, b) → a · L + b · p (3.62)
we have
1 1
{H, Ai } = − ijk {1/r, pj }Lk − pj {pj , xi /r}
m m
1
= (ijk xj Lk /r 3 + pi /r − xi pj xj /r3 ), (3.64)
m
Using Eq. (3.46), since {Li , f (r)} = 0 for f an arbitrary function of r. The first
term is
1 1
ijk ksn xj xs pn /r3 = (δis δjn − δin δjs )xj xs pn /r 3 , (3.65)
m m
which cancels the second and third terms. We are using
Li = ijk xj pk . (3.66)
We want to see if these six quantities close to a Lie algebra. We still have
1
{Li , Aj } = jsn {Li , ps Ln } − {Li , xj /r}
mk
1 1
= jsn ({Li , ps }Ln + ps {Li , Ln }) − {Li , xj }
mk r
(p × L)k xk
= ijk − = ijk Ak , (3.69)
mk r
below:
and
(mk){xi /r, Aj } = jsn {xi /r, ps Ln } = jsn ({xi , ps Ln }/r + {1/r, ps Ln }xi )
= jsn (ps {xi , Ln }/r + {xi , ps }Ln /r + {1/r, ps }Ln xi )
= jsn (−nik ps xk /r + δis Ln /r − xs xi Ln /r 3 )
= −(δji δsk − δjk δsi )ps xk /r + jin Ln /r − jsn xs xi Ln /r 3
= pi xj /r − δij xm pm /r − ijm Lm /r − jmn xi xm Ln /r3 .
(3.71)
The latter yields, remembering that r · L = 0, and using Exercises 3.1 and 3.2,
and
Also
and
−2H
{Ai , Aj } = ijk Lk . (3.79)
mk2
The occurrence of H(r, p) in the last equation prevents the quantities Li and
Ai from closing to a Lie algebra on the whole phase space. But when the motion
is in a fixed orbit H(r, p) = E, with E a constant energy. We can then divide the
phase space into three sets, the two open regions where H > 0 and H < 0, and
their common boundary where H = 0. In the region where H > 0 we can rescale
A by 2H/mk 2 to get an o(3, 1) algebra. In the region where H < 0 we can rescale
by −2H/mk2 to get an o(4) algebra.
That is, we define
A = N A, (3.80)
where
2H
N −2 = − , (3.81)
mk 2
to get
This shows that the quantities Li and Ai close to form an o(4) Lie algebra.
August 27, 2011 9:21 9in x 6in The Supersymmetric Dirac Equation b1217-ch03
We can use the Lie algebra isomorphism of o(4) with o(3)⊕ o(3) to form the
direct sum of two o(3) algebras. Define
Mi = Li + Ai , Ni = Li − Ai . (3.83)
Notes on Chapter 3
This review of Hamiltonian mechanics is standard, and can be found in any
good text on classical mechanics. It is included here to make the treatment self-
contained. The same is true for the relation between Lie groups and Lie algebras.
More information can be found in Weaver and Sattinger [WS10], or Marsden and
Ratiu [MR99]. We follow here the treatment of Guillemin and Sternberg [GS90].
The calculation of the Poisson bracket between the components of the Laplace
vector is the classical version of the calculation of the corresponding commutator
by Sudbery [Sud86].
August 27, 2011 9:21 9in x 6in The Supersymmetric Dirac Equation b1217-ch03
Chapter 4
From Solar Systems to Atoms
Lord Rutherford, quoted in Rutherford and the Nature of the Atom, E. N. da C. Andrade,
Heinemann, London, 1964.
31
August 27, 2011 9:21 9in x 6in The Supersymmetric Dirac Equation b1217-ch04
(p × L)
A=− − r̂. (4.1)
Zαm
The scattering process is described as in Figure 4.1. Long before the collision
takes place the initial impulse of a particle in the beam is pi = mv0 ei , where v0
is the velocity of the incoming particle and ei is a unit vector in the direction of
the incoming velocity. The final velocity is pf = mv0 ef , with the same velocity
v0 (the process is elastic) and ef is a unit vector in the direction of the outgoing
velocity. Li = Lf = mv0 bk̂, where L is the conserved angular momentum, and
k̂ is a unit vector pointing up through the page, perpendicular to the scattering
plane. b is the impact parameter, the distance that the incoming particle would
pass by the point target in the absence of interaction. The initial incoming direction
is r̂i = −ei , and the final outgoing direction r̂f = ef .
The initial Laplace vector is
mv02
Ai = − bni + ei , (4.2)
Zα
e
f
n
f
b
v
f
ni
ei
m
r
b
v Θ
c
Figure 4.1. The incoming and outgoing directions and the scattering angle.
August 27, 2011 9:21 9in x 6in The Supersymmetric Dirac Equation b1217-ch04
where ni is the unit vector in Figure 4.1 perpendicular to ei and in the scattering
plane. The final Laplace vector is
mv02
Af = − bnf − ef , (4.3)
Zα
With (ef ·ei ) = cos θ, where θ is the scattering angle, and (nf ·ei ) = cos (π/2 + θ)
= − sin θ, this leads to
and this enables us to calculate the scattering angle for any impact parameter.
b Θ
db dΘ
section is
dσ b db
= . (4.7)
dΩ sin θ dθ
The result for the Thompson atom, which involves multiple scattering,
decreases much more rapidly then the characteristic sin−4 (θ/2) of Rutherford’s
formula. This is the reason for Rutherford’s surprise, noted earlier. Together with his
associates, Geiger and Marsden, he verified the formula in great detail, including
its proportionality to Z 2 , and the inverse squared proportionality to the kinetic
energy of the incoming particles. The formula was also found to be accurate down
to small distances, and the finite size of the nucleus, at which deviations from the
formula were observed, was around 10−14 m = 10 Fermi for an aluminium foil.
The success of the formula proved the correctness of the assumptions on which it
is based, and established the solar system model for atoms.
Notes on Chapter 4
Rutherford scattering is analyzed using the Laplace vector by Basano and Bianchi
[BB80], and in textbooks on classical mechanics; see, for example, Poole, Goldstein
and Safko [PGS01].
August 27, 2011 9:21 9in x 6in The Supersymmetric Dirac Equation b1217-ch05
Chapter 5
The Bohr Model
“The spectrum of the hydrogen atom has proved to be the Rosetta stone of modern physics.”
Theodor W. Hänsch, Arthur L. Schawlow, and George W. Series, in Rev. Mod. Phys. 54,
p. 697 (1982).
The first postulate bases the model on the picture introduced by Rutherford.
The second postulate introduces quantization of orbital angular momentum of an
atomic electron moving under the influence of an inverse square (Coulomb) force:
L = N, N = 1, 2, 3, . . . (5.2)
N=1
N=2
N=3
postulating that this particular feature of the classical theory is not valid for the case
of an atomic electron. The postulate was based on the fact that atoms are observed
by experiment to be stable — even though this is not predicted by classical theory.
The fourth postulate, 2πν = Ei − Ef , is really just Einstein’s postulate that the
frequency of a photon of electromagnetic radiation is equal to the energy carried
by the photon divided by Planck’s constant; see Figure 5.2.
Zα mv 2
2
= , (5.3)
r r
where α = e2 /4π is the fine structure constant, v is the speed of the electron in
its orbit, and r is the radius of the orbit. The left-hand side of this equation is the
Coulomb force acting on the electron, and the right-hand side is ma, where a is
the centripetal acceleration of the electron in its circular orbit.
Now, the orbital angular momentum of the electron, L = mvr, is a constant,
because the force acting on the electron is entirely in the radial direction. Applying
August 27, 2011 9:21 9in x 6in The Supersymmetric Dirac Equation b1217-ch05
Figure 5.3. The electron in the ground state of the Bohr atom. (a) Classical picture; (b) Quantum
Mechanical picture.
mvr = N, N = 1, 2, 3, . . . (5.4)
Figure 5.4. The probability of finding the electron at a given distance from the nucleus.
in the Bohr model. On the other hand, Eq. (5.7) shows that for large values of Z
the electron velocity becomes relativistic, and the model cannot be applied.
Next we calculate the total energy of an atomic electron moving in one of
the allowed orbits. Let us define the potential energy to be zero when the electron
is at rest infinitely distant from the nucleus. Then the potential energy V at any
finite distance r can be obtained by integrating the work that would be done by the
Coulomb force acting from r to ∞. Thus
∞
Zα Zα
V = 2
dr = − . (5.8)
r r r
The potential energy is negative because the Coulomb force is attractive; it takes
work to move the electron from r to ∞ against this force. The kinetic energy of
the electron, Ekin , can be evaluated, with the aid of Eq. (5.3), to be
1 Zα
Ekin = mv 2 = . (5.9)
2 2r
The total energy of the electron, E, is then
Zα
E = Ekin + V = − = −Ekin , (5.10)
2r
August 27, 2011 9:21 9in x 6in The Supersymmetric Dirac Equation b1217-ch05
E [eV]
0 N=∞
-0.85 N=4
-1.51 N=3
-3.39 N=2
-13.6 N=1
in agreement with the virial theorem. Using Eq. (5.6) for r in the preceding equation,
we have
m (Zα)2
E=− , N = 1, 2, 3, . . . (5.11)
2 N2
We see that the quantization of the orbital angular momentum of the electron leads
to a quantization of its total energy.
The information contained in Eq. (5.11) is presented as an energy-level
diagram in Figure 5.5. The energy of each level, as evaluated from Eq. (5.11) is
shown on the left, in terms of electron volts, and the quantum number of the level is
shown on the right. The diagram is so constructed that the distance from any level
to the level of zero energy is proportional to the energy of that level. Note that the
lowest allowed total energy occurs for the smallest quantum number N = 1. As
N increases, the total energy of the quantum states becomes less negative, with E
approaching zero as N approaches infinity. Since the state of lowest total energy
is the ground state, the normal state of the electron in a one-electron atom is the
state for which N = 1.
The energy binding the electron to the nucleus may be calculated from
Eq. (5.11). The binding energy is numerically equal to the energy of the ground
state, corresponding to N = 1. This yields, with Z = 1, E = −(m/2)α2 =
−13.6 eV , which agrees very well with the experimentally observed binding
energy for hydrogen. Next we calculate the frequency ν of the electromagnetic
radiation emitted when the electron makes a transition from the quantum state Ni
August 27, 2011 9:21 9in x 6in The Supersymmetric Dirac Equation b1217-ch05
to the quantum state Nf , that is, when an electron initially moving in an orbit
characterized by the quantum number Ni discontinuously changes its motion so
that it moves in an orbit characterized by quantum number Nf . Using Bohr’s fourth
postulate, Eq. (5.2), and Eq. (5.11), we have
Ei − E f Z 2 α2 m 1 1 1 1
ν= = − 2 = R∞ Z 2 − 2 , (5.12)
2π 4π Nf2 Ni Nf2 Ni
with
α2 m
R∞ = , (5.13)
4π
and where Ni and Nf are integers.
The essential predictions of the Bohr model are contained in Eqs. (5.11) and
(5.12). We discuss the emission of electromagnetic radiation by a one-electron
Bohr atom in terms of these equations:
According to the Bohr model, each of the five known series of the hydrogen
spectrum arises from a subset of transitions in which the electron goes to a certain
final quantum state Nf . For the Lyman series Nf = 1, for the Barmer series
Nf = 2, for the Paschen series Nf = 3, for the Brakett series Nf = 4, and for the
Pfund series Nf = 5. The wavelength of the lines of all these series are fitted very
accurately by Eq. (5.12) by using the appropriate value of Nf . This was a great
triumph of the Bohr model. The success of the model was particularly impressive
August 27, 2011 9:21 9in x 6in The Supersymmetric Dirac Equation b1217-ch05
because the Lyman, Brakett and Pfund series had not been discovered at the time the
model was developed by Bohr. The existence of these series was predicted, and the
series were soon found experimentally by the persons after whom they were named.
The model worked equally well when applied to the case of one-electron atoms
with Z = 2, i.e., singly ionized helium atoms He+. Such atoms can be produced
by passing a spark through normal helium gas. They make their presence apparent
by emitting a simpler spectrum than that emitted by normal helium atoms. In fact,
the spectrum of He+ is exactly the same as the hydrogen spectrum except that the
reciprocal wavelengths of all the lines are almost exactly four times as great. This is
explained very easily, in terms of the Bohr model, by setting Z 2 = 4 in Eq. (5.11).
The properties of the absorption spectrum of one-electron atoms are also easy
to understand in terms of the Bohr model. Since the atomic electron must have a
total energy exactly equal to the energy of one of the allowed energy states, the
atom can only absorb discrete amounts of energy from the incident electromagnetic
radiation. This fact leads to the idea that we consider the incident radiation to be a
beam of photons, and that only those photons can be absorbed whose frequency is
given by E = 2πν, where E is one of the discrete amounts of energy which can
be absorbed by the atom. The process of absorbing electromagnetic radiation is
then just the inverse of the normal emission process, and the lines of the absorption
spectrum will have exactly the same wavelengths as the lines of the emission
spectrum. Normally the atom is always initially in the ground state N = 1, so
only absorbtion processes from N = 1 to N > 1 can occur. Thus, the only
absorption lines which correspond (for hydrogen) to the Lyman series will normally
be observed. However, if the gas containing the absorbing atoms is at a very high
temperature, then, owing to collisions, some of the atoms will initially be in the
first excited state N = 2, and absorption lines corresponding to the Balmer series
will be observed. Balmer absorbtion lines are actually observed in the hydrogen
gas of some stellar atmospheres. This gives us a way of estimating the temperature
of the surface of a star.
difficult to show that in such a system the electron moves relative to the nucleus as
though the nucleus were fixed and the mass of the electron were slightly reduced to
the value µ, the reduced mass of the system. The equations of motion of the system
are the same as those we have considered if we simply substitute µ for m, where
mM
µ= , (5.14)
m+M
is less than m by a factor 1/(1 + m/M ). Here M is the mass of the nucleus.
To handle this situation Bohr modified his second postulate, Eq. (5.4), to
require
µvr = N, N = 1, 2, 3, . . . (5.15)
Using µ instead of m in this equation takes into account the angular momentum
of the nucleus as well as that of the electron. All the equations are identical with
those derived before, except that the electron mass m is replaced by the reduced
mass µ. In particular, the formula for the frequencies of the spectral lines becomes
1 1
ν = RM Z 2 − 2 , (5.16)
Nf2 Ni
where
µ
RM = R∞ . (5.17)
m
The quantity RM is the Rydberg constant for a nucleus of mass M . For the most
extreme case of hydrogen RM is less than R∞ by about one part in 2000. If we
evaluate RH from Eq. (5.17) we find that the Bohr model, corrected for the finite
nuclear mass, agrees with the experimental data to within three parts in 100,000.
Atoms that follow the same pattern as hydrogen, but with different values
of the parameters, are called hydrogenic atoms. They include systems in which
an electron or the proton are replaced by an exotic particle (see below). They also
include Rydberg atoms, an ordinary atom in which one electron has been elevated to
a very high quantum state, and highly ionized atoms; that is, atoms whose electrons
have been stripped away, leaving only one electron in an orbit around the nucleus.
In Rydberg atoms the methods developed here serve only as approximations, in
which the interactions between the lower-lying electrons are neglected.
An example of an exotic atom is positronium, which consists of a positron
and an electron, revolving about their common center of mass, which lies halfway
between them. The reduced mass is µ = m/2. The corresponding Rydberg constant
is RM = R∞ /2. The frequencies of the emitted spectral lines of positronium are
given by
R∞ 1 1
ν= − 2 . (5.18)
2 Nf2 Ni
August 27, 2011 9:21 9in x 6in The Supersymmetric Dirac Equation b1217-ch05
The frequencies of the emitted lines are then half those of a hydrogen atom
(with infinitely heavy nucleus), Z being equal to one for positronium and for
hydrogen. The radius of the ground state of positronium is
1
rpos = = 2a0 . (5.19)
αµ
Hence, for any quantum state N the radius of the electron is twice as great in the
positronium atom as in the hydrogen atom (with infinitely heavy nucleus).
A muonic atom contains a nucleus of charge Zα and a negative muon, µ− ,
moving about it. The µ− is an elementary particle with charge −e and a mass that
is about 207 times as large as an electron mass. Such an atom is formed when a
proton, or some other nucleus, captures a µ− . The reduced mass of the system is
µ = 186m. Then, from Eq. (5.6), with N = Z = 1, and µ = 186m, we obtain
1
r1 = = 2.8 × 10−3 Å. (5.20)
186mα
Therefore the µ− is much closer to the nuclear (proton) surface than is an electron in
a hydrogen atom. It is this feature that makes muonic atoms interesting, information
about nuclear properties being revealed from their study. The binding energy of a
muonic atom with Z = 1 is calculated from Eq. (5.11), with N = 1 and µ = 186 m;
the result is E = −2530 eV . The wavelength of the first line in the Lyman series is
1
ν = RM 1 − , (5.21)
4
with RM = (µ/m)R∞ , so λ = 1/ν = 6.5 Å, so that the Lyman lines lie in the
x-ray part of the spectrum. X-ray techniques are therefore necessary to study the
spectrum of muonic atoms.
Ordinary hydrogen contains about one part in 6000 of deuterium, or heavy
hydrogen. This is a hydrogen atom whose nucleus contains a proton and a neutron.
The spectrum would be identical to that of hydrogen were it not for the correction
of finite nuclear mass. For a normal hydrogen atom
109737 cm−1
RH = 1
= 109678 cm−1 . (5.22)
1 + 1836
109737cm−1
RD = = 109707 cm−1 . (5.23)
1
1 + 2×1836
August 27, 2011 9:21 9in x 6in The Supersymmetric Dirac Equation b1217-ch05
Hence, RD is a bit larger than RH , so that the spectral lines of the deuterium atom
are shifted to slightly lesser wavelengths compared to hydrogen. Indeed, deuterium
was discovered in 1932 by H. C. Urey following the observation of these shifted
spectral lines. By increasing the concentration of the heavy isotope above its normal
value in a hydrogen discharge tube, one can enhance the intensity of the deuterium
lines which, ordinarily, are difficult to detect. One then readily observes pairs of
hydrogen lines; the shorter wavelength members of the pair correspond exactly to
those predicted from RD . The resolution is easily obtained, the Hα -line pair being
separated by about 1.8 Å, for example, several thousand times greater than the
minimal resolvable separation.
Notes on Chapter 5
Information on Niels Bohr and the Bohr model can be found in Hydrogen: The
Essential Element, by John Rigden [Rig03]. Also in that book are further details
on hydrogenic atoms. Further recommended reading is The Infancy of Atomic
Physics: Hercules in His Cradle, by Alex Keller [Kel83]. The presentation here
follows Eisberg and Resnick [ER74].
August 27, 2011 9:21 9in x 6in The Supersymmetric Dirac Equation b1217-ch05
Chapter 6
Interpretation of the Quantum Rules
47
August 27, 2011 9:21 9in x 6in The Supersymmetric Dirac Equation b1217-ch06
or
p2x x2
+ = 1. (6.3)
2mE 2E/k
The quantization integral px dx is most easily evaluated, for the relation between
px and x that is imposed by this equation, if we consider a geometric interpretation.
The relation between px and x is the equation of an ellipse. Any instantaneous state
of motion of the oscillator is represented by some point in a plot of this equation on a
two-dimensional space having coordinates px and x. Such a space (the p − q plane)
is the phase space, and the plot is a phase diagram of the linear oscillator. During
one cycle of oscillation the point representing the position and the momentum of
the particle travels once around the ellipse. The semiaxes a and b of the ellipse
p2x /b2 + x2 /a2 = 1 are seen, by comparison with Eq. (6.3), to be
√
b = 2mE and a = 2E/k. (6.4)
Now the area of an ellipse is πab. The value of the integral px dx is just equal to
that area. Thus we obtain
px dx = πab. (6.5)
In our case
2πE
px dx = , (6.6)
k/m
but
k/m = 2πν, (6.7)
E = 2πN ν, (6.10)
Note that the allowed states of oscillation are represented by a series of ellipses
in phase space, the area enclosed between successive ellipses always being 2π =
2π = h. We find that the classical situation corresponds to 2π → 0, allvalues
of E and hence all ellipses being allowed if that were true. The quantity px dx
is called a phase integral; in classical physics it is the integral of the dynamical
quantity called the action over one oscillation of the motion. Hence, the Planck
energy quantization condition is equivalent to the quantization of the action.
The full quantum mechanical treatment of course gives E = 2πν(N + 12 ), so
that the Sommerfeld–Wilson quantization condition does not give the zero-point
energy correctly.
We can also deduce the Bohr quantization of angular momentum from the
Sommerfeld–Wilson rule, Eq. (6.1). An electron moving in a circular orbit of
radius r has an angular momentum, mvr = L, which is constant. The angular
coordinate is θ, which is a periodic function of time. That is, θ versus t is a saw-
tooth function, increasing linearly from zero to 2π rad in one period and repeating
this pattern in each succeeding period. The quantization rule
pq dq = 2πNq , (6.11)
and
2π
Ldθ = L dθ = 2πL, (6.13)
0
The hypothesis of de Broglie [Bro24] was that the dual, that is the wave-
particle, behavior of radiation applies equally well to matter. Just as a photon has a
light wave associated with it that governs its motion, so a material particle (e.g., an
electron) has an associated matter wave that governs its motion. Since the universe
is composed entirely of matter and radiation, de Broglie’s suggestion is essentially
a statement about a grand symmetry of nature. Indeed, he proposed that the wave
aspects of matter are related to its particle aspects in exactly the same quantitative
way that is the case for radiation. According to de Broglie, for matter and radiation
alike the total energy of an entity is related to the frequency ν of the wave associated
with its motion by the equation
E = 2πν, (6.14)
and the momentum p of the entity is related to the wavelength λ of the associated
wave by the equation
p = 2π/λ. (6.15)
Here the particle concepts, energy E and momentum p, are connected through
Planck’s constant h = 2π = 2π, to the wave concepts, frequency ν and
wavelength λ. Equation (6.15), in the following form, is called de Broglie’s relation:
2π
λ= . (6.16)
p
It predicts the de Broglie wavelength λ of a matter wave associated with the motion
of a material particle having a momentum p.
The wave nature of light propagation is not revealed in experiments in
geometrical optics, for the important dimensions of the apparatus used there are
very large compared to the wavelength of light. Geometrical optics involves ray
propagation, which is similar to the trajectory motion of classical particles. When
the characteristic dimension of an optical apparatus becomes comparable to, or
smaller than, the wavelength of the light going through it, we are in the domain
of physical optics, where diffraction effects are observed, and the wave nature of
light propagation becomes apparent.
To observe wavelike aspects in the motion of matter, therefore, we need
systems with apertures or obstacles of suitable small dimensions. Using apparatus
with characteristic dimensions of 1 Å, wavelike aspects of the motion of electrons
should be apparent.
Elsasser pointed out, in 1926, that the wave nature of matter might be tested
in the same way that the wave nature of x-rays was first tested, namely by allowing
a beam of electrons of appropriate energy to fall on a crystalline solid. The atoms
August 27, 2011 9:21 9in x 6in The Supersymmetric Dirac Equation b1217-ch06
2πr = N λ, N = 1, 2, 3, . . . (6.20)
Thus the allowed orbits are those in which the circumference of the orbit can
contain exactly an integral number of de Broglie wavelengths.
Imagine the electron moving in a circular orbit at constant speed. It has a wave
associated with its motion. The wave, of wavelength λ, is wrapped repeatedly
around the circular orbit. The resultant wave that is produced will have zero
intensity at any point unless the wave at each traversal is exactly in phase at that
point with the wave in other traversals. If the wave in each traversal is exactly
in phase, they join on perfectly in orbits that accommodate integral numbers of
de Broglie wavelengths, as illustrated in Figure 6.1. But the condition that this
happens is just the condition that Eq. (6.20) be satisfied.
This wave picture gives no suggestion of progressive motion. Rather, it
suggests standing waves, as in a stretched string of a given length. In a stretched
string only certain wavelengths, or frequencies of vibration, are permitted. Once
such modes are excited, the vibration goes on indefinitely if there is no damping.
To get standing waves, however, we need oppositely directed traveling waves
of equal amplitude. For the atom this requirement is satisfied by the fact that
the electron can traverse an orbit in either direction and still have the same
magnitude of angular momentum required by Bohr. The de Broglie standing wave
August 27, 2011 9:21 9in x 6in The Supersymmetric Dirac Equation b1217-ch06
Notes on Chapter 6
The presentation here follows Eisberg and Resnick [ER74].
August 27, 2011 9:21 9in x 6in The Supersymmetric Dirac Equation b1217-ch07
Chapter 7
Sommerfeld’s Model for Non-Relativistic Electrons
53
August 27, 2011 9:21 9in x 6in The Supersymmetric Dirac Equation b1217-ch07
One differentiates Eq. (7.8) with respect to θ and obtains the linear differential
equation:
d2 s
+ s − C = 0, (7.9)
dθ2
where we have used the abbreviation
Zαm
C= . (7.10)
p2θ
To evaluate the integral in Eq. (7.15) we use the expression Eq. (7.5) for pr
and obtain
2
pr dr e2 sin2 θ
pr dr = 2 dθ = pθ dθ. (7.16)
r dθ (1 + e cos θ)2
Thus, the radial quantization condition, Eq. (7.15), takes the form
θ=2π
1 e2 sin2 θ dθ nr
= . (7.17)
2π θ=0 (1 + e cos θ)2 nθ
We then determine H from the energy equation, Eq. (7.8), and knowledge of
the orbit s = s(θ). Eq. (7.8) becomes
n2θ 2 2
H + Zα(A cos θ + C) = (A sin θ + A2 cos2 θ + C 2 + 2AC cos θ)
2m
n2θ 2
= (A + C 2 + 2AC cos θ). (7.23)
2m
Comparing the coefficients of cos θ yields
n2θ
Zα = C, (7.24)
m
which holds because of Eq. (7.10). The constant term yields
n2θ 2 n2
H + ZαC = (A + C 2 ) = θ (e2 + 1)C 2 . (7.25)
2m 2m
This leads to
n2θ m
(Zα)2
H= 2 − (1 − e2 ) C 2 − ZαC = − . (7.26)
2m 2 (nθ + nr )2
For a given orbit the energy is constant
m
(Zα)2
H=E=− . (7.27)
2 (nθ + nr )2
For the circular orbits e = 0, and from Eq. (7.20) it follows that nr = 0. Thus, for
Bohr’s circular orbits
m
(Zα)2
E=− . (7.28)
2 n2θ
Defining N = nθ + nr we find that there are multiple orbits for N > 1, all of the
same energy:
m (Zα)2
EN = − . (7.29)
2 N2
August 27, 2011 9:21 9in x 6in The Supersymmetric Dirac Equation b1217-ch07
N = 1, 2, 3, 4, . . . (7.30)
nθ = 1, 2, 3, . . . , N. (7.31)
The integer N is called the principal quantum number, and nθ is called the
azimuthal quantum number. nθ = 0 is not allowed; for nr = 0 it would correspond
to an orbit going through the force center.
This degeneracy in the total energy of an electron, due to orbits of very different
shape but common N , is the result of a very delicate balance between potential and
kinetic energy, which is characteristic of treating the inverse square Coulomb force
by the methods of classical mechanics. Exactly the same phenomena is found in
planetary or satellite motion, which is governed by the inverse square gravitational
force. For instance, a satellite may be launched into any one of a whole family of
elliptical orbits, all of which correspond to the same state of the same total energy
and have the same semimajor axis. Of course, there is no quantization of the orbit
parameters in these macroscopic cases, but as far as degeneracy is concerned they
are completely analogous to the case of the hydrogen atom.
To get the complete degeneracy of the electron, neglecting the spin, we turn
to a group-theoretical analysis in the next chapter.
Notes on Chapter 7
The history of atomic models is told in Keller [Kel83]. The importance of hydrogen
in this context is emphasized by Rigden [Rig03]. This chapter develops a non-
relativistic version of Sommerfeld’s work [Som16], following Biedenharn [Bie83].
See also Sommerfeld’s book [Som24].
August 27, 2011 9:21 9in x 6in The Supersymmetric Dirac Equation b1217-ch08
Chapter 8
Quantum Mechanics of Hydrogenic Atoms
Heisenberg [ Hei25], on the basis of the spectral data of hydrogen, wrote a paper
in which he formulated for the first time the canonical commutation relations
between the basic variables of position and momenta, [xi , pj ] = iδij . He gave his
manuscript to Born, who submitted it to the Zeitschrift für Physik. Born recognized
that Heisenberg’s equations could be expressed in matrix form. The formulation of
the matrix formalism of quantum mechanics was published in two papers in 1925,
by Born and Jordan [BJ25], and by Born, Heisenberg and Jordan [MBJ26].
The immediate question was whether the new theory was capable of explaining
the spectrum of hydrogen, as the Bohr model had done. The answer was given by
Pauli [Pau26], and independently by Dirac [Dir26], towards the end of 1925. Pauli
obtained the same formula that Bohr had obtained in 1913, only this time the route
to the formula was a coherent theory — the new theory of quantum mechanics.
In the following we shall show, following Pauli, that the basic commutation
relations, and a general knowledge of the symmetry group SO(4), lead to the Balmer
formula, and a first estimate of the degeneracy of the energy levels.
The classical Kepler problem is characterized by six conserved quantities: the
three components of the orbital angular momentum L = (L1 , L2 , L3 ) and the
three components of the Laplace vector, A = (A1 , A2 , A3 ), satisfying the two
constraints A·L = 0 and |A|2 = 2EL2 /((Zα)2 m) + 1. We shall see that these
relations all have their counterparts in the quantum mechanical case.
8.1 Quantization
In general, we go from a classical to a quantum system by following the Dirac
quantization procedure [Dir53], i.e., we replace observable quantities by Hermitian
operators, and Poisson brackets of observables by commutators, according to the
rule {, } → i[, ].
59
August 27, 2011 9:21 9in x 6in The Supersymmetric Dirac Equation b1217-ch08
For the basic observables, position xi and momentum pj , this means that
{xi , pj } = δij is replaced by [xi , pj ] = iδij . More complex commutators are
reduced by analogs of the rules (a)–(d) of Section 3.3, e.g.,
Li = ijk xj pk . (8.4)
(p × L)† = −L × p, (8.7)
(a) (p × L)·(p × L) = p2 L2 ,
(b) (p × L)·p = 2ip2 ,
(c) p·(p × L) = 0,
(d) (p × L)·r = L2 + 2i(p·r),
(e) r·(p × L) = L2 . (8.16)
Li = imn xm pn , (8.21)
and similarly
so
In general, we say that a vector operator v, whose components satisfy with the
components of the angular momentum
1 1
(Zαm)[pi , Aj ] = jsn [pi , ps Ln ] − [pi , Ls pn ] − Zαm[pi , xj /r]
2 2
= i(pi pj − p2 δij ) + i(Zαm)(δij /r − xi xj /r 3 ). (8.26)
August 27, 2011 9:21 9in x 6in The Supersymmetric Dirac Equation b1217-ch08
To calculate [pi , 1/r] we use the analog of the rule (3.46), generalized to the
quantum context, [pi , g] = −i∂g/∂xi , for g(x) a function of x.
We evaluate the terms separately. The first two expressions are identical to those
in the classical evaluation:
= i(δjs δkn − δjn δks )Ln (ps pj − p2 δsj + (Zαm)δsj /r − (Zαm)xs xj /r3 )
1
A = (N 2 ) 2 A. (8.36)
which are equivalent to the original ones since (8.38) represents a real linear
transformation of the original generators. The new commutation relations are
simpler since M and N are two commuting angular momentum vectors. Therefore
so(4) is the direct sum of so(3) ⊕ so(3) subalgebras. The first is generated by M
and the second is generated by N.
The so(4) algebra has two Casimir operators; these are quadratic operators
commuting with all the generators:
1
C1 = L2 + A2 , C2 = (L·A + A·L). (8.40)
2
or substituting (8.43)
4M 2 + 1 = N 2 . (8.45)
N 2 → N 2. (8.47)
cannot be explained on the basis of SO(3) alone, since SO(3) does not guarantee that
the different values of the angular momentum l all have the same energy. However,
it is SO(4) that is the geometrical symmetry group, not SO(3). In fact it follows
from j1 = j2 that the degeneracy of EN is N 2 , since there are N 2 = (2j1 + 1)2
basis vectors in the direct product space corresponding to the energy EN .
It must be emphasized that N 2 does not give the full degeneracy of the levels,
which is experimentally 2N 2 . The further factor of two comes from the spin of
the electron, which has so far not entered our considerations. Another indication
of this fact is discussed below.
Classically the eccentricity e is the length of the vector A. We might attempt to
take this relation over to quantum mechanics. From Eqs. (8.36), (8.42), and (8.47)
we find that
N = 1).
This result for the eccentricity is strictly applicable to the orbits of spinless
particles. For the electron, which has spin- 12 , it is not applicable. In Chapter 10 we
shall find the value of the eccentricity for the orbits of the spinning electron.
Notes on Chapter 8
This chapter is a review of Pauli’s paper [Pau26]. If you need to review the
derivation of quantum mechanical commutation relations, or the application of
group-theoretical symmetry considerations, an excellent place to look is the very
lucid book by Adams [Ada94].
August 27, 2011 9:21 9in x 6in The Supersymmetric Dirac Equation b1217-ch09
Chapter 9
The Schrödinger Equation
and the Confluent Hypergeometric Functions
In late 1925 Schrödinger read de Broglie’s paper in which it was proposed that
particles have an associated wavelength. Schrödinger went to work attempting to
generalize de Broglie’s concept of matter waves. The first sentence of Schrödinger’s
classic paper [Sch26] reads as follows: “In this paper I wish to consider, first,
the simplest case of the hydrogen atom, and show that the customary quantum
conditions can be replaced by another postulate, in which the concept of ‘whole
numbers’, merely as such, is not introduced.” In modern language Schrödinger is
saying that he can dispense with the use of quantum conditions; they are replaced
by the requirement that the argument of the hypergeometric function describing
the solution of the wave equation must be a nonzero negative integer if the wave
function is to fall off at infinity (see below).
The mathematics employed in the paper was familiar to physicists — it
involved the solutions of differential equations. Schrödinger set up a wave equation
in a form appropriate for the hydrogen atom — a negatively charged electron
orbiting a positively charged nucleus. The solutions are normalizable only for
discrete values of the energy. In 1926 Born suggested the interpretation of the
solution of the wave equation: The magnitude of the quantity |ψ|2 , where ψ is
the solution of the wave equation, is a measure of the probability density of the
electron. Schrödinger’s wave equation gives the maximum magnitude of |ψ|2 in
the ground state of the atom at a distance of 0.529 Å from the nucleus, in agreement
with the Bohr radius, and the known size of the atom, whose diameter is about 1 Å.
69
August 27, 2011 9:21 9in x 6in The Supersymmetric Dirac Equation b1217-ch09
wave function:
1 2 Zα
Hψ(r) = − ∇ − ψ(r) = Eψ(r). (9.1)
2m r
This equation is most naturally solved by transforming from Cartesian to spherical
polar coordinates, in which system it is completely separable. We use the identity
L2 = r2 p2 − r(r · p) · p. (9.2)
∂2 2 ∂ 1 ∂ 2 ∂
p2r = − 2
− =− 2 r , (9.7)
∂r r ∂r r ∂r ∂r
and
L2
p2 = p2r + , (9.8)
r2
as the quantum-mechanical counterpart to Eq. (7.4).
August 27, 2011 9:21 9in x 6in The Supersymmetric Dirac Equation b1217-ch09
where the Y (θ, φ) are spherical harmonics and R(r), the radial part of the wave
function, satisfies
2Zαm l(l + 1)
−2m(H − E) = −pr + 2mE −
2
− R
r r2
1 d 2 dR 2Zαm l(l + 1)
= 2 r + 2mE + − R = 0,
r dr dr r r2
(9.10)
subject to the conditions that limr→0 rR(r) = 0 and rR(r) be square integrable.
We introduce the dimensionless variable
Zαm
x=2 r, (9.11)
λ
and obtain
d2 R dR
x2 + 2x + [−µx2 + λx − l(l + 1)]R = 0. (9.12)
dx2 dx
We have defined the parameter µ by
−E λ2
µ= . (9.13)
2m (Zα)2
The parameter λ > 0 is yet to be determined. For bound states (E < 0) we also
have µ > 0.
Next we factor out the behavior near x = 0 and x = ∞, and look for solutions
of the type
√
R(x) = xl e− µx
u(x). (9.14)
1 F1 (l + 1 − λ; 2l + 2; x) −−−−→ ex , (9.19)
x→∞
the function u(x) diverges for large x unless l + 1 − λ is a negative integer or zero,
in which case the confluent hypergeometric function is a polynomial. This means
that λ must be a positive integer, which we denote by N ,
λ = N, N = 1, 2, 3, . . . (9.20)
√
Hence, with µ = 1/2, Eq. (9.13) becomes
1 E N2
µ= =− (9.21)
4 2m (Zα)2
and we find again that the energy spectrum is discrete, with
m (Zα)2
EN = − , N = 1, 2, 3, . . . (9.22)
2 N2
The variable x is now
Zαm
x=2 r. (9.23)
N
The spherical harmonics are normalized to unity, so the normalization of the wave
function (9.9) is
|R(r)|2 r2 dr = |u(r)|2 dr = 1, (9.24)
with
|u(x)|2 dx = 1. (9.26)
The radial solution, which is regular at the origin and falls off exponentially at
infinity, is
3
u(r) 2Zαm 2 uN l (x)
RN l (r) = = (9.27)
r N x
where
uN l (x) = AN l xl+1 e− 2 x 1 F1 (−N + l + 1; 2l + 2; x),
1
(9.28)
and AN l is the normalization constant. The normalization constant is calculated in
Appendix B:
1 (N + l)!
AN l = . (9.29)
(2l + 1)! (N − l − 1)!(2N )
To obtain a polynomial solution we must have N ≥ l + 1.
then u(x) is given by Eq. (9.31). The solution that is regular at the origin is
dq q −x
Lq (x) = q!1 F1 (−q, 1; x) = ex (x e ), (9.35)
dxq
see Appendix A. This function is the Laguerre polynomial of order q.
From Eq. (9.31) we see that the polynomial solution to Eq. (9.30) is obtained
by differentiating Lq (x) p times. We denote these solutions by
dp Lq (x)
Lpq (x) = . (9.36)
dxp
These functions are known as the associated Laguerre polynomials. It is a
polynomial of degree q − p. By carrying out the differentiation explicitly, we
find
∞
q!(−q)p (−q + p)k k
Lpq (x) = x , (9.37)
p! k!(p + 1)k
k=0
[Γ(q + 1)]2
Lpq (x) = (−1)p 1 F1 (−q + p, p + 1; x). (9.38)
Γ(p + 1)Γ(q − p + 1)
From Eq. (9.28) we see that the radial functions for bound states in a Coulomb
potential are
32
2Zαm
xl e− 2 x L2l+1
1
RN l (r) = NN l N +1 (x), (9.39)
N
Notes on Chapter 9
The analysis of the Schrödinger equation is conventional, e.g., Mertzbacher
[Mer61]. The hypergeometric functions and the special functions which arise in
different physical contexts are reviewed, for example, in the book by Seaborn
[Sea80].
August 27, 2011 9:21 9in x 6in The Supersymmetric Dirac Equation b1217-ch09
Chapter 10
Non-Relativistic Hydrogenic Atoms with Spin
In 1922 Stern and Gerlach passed a beam of silver atoms through an inhomogeneous
magnetic field, and found that it was split into two components. Goudsmit and
Uhlenbeck proposed in 1925, on the basis of atomic spectra, that the electron was
characterized by an additional quantum number with value 1/2, which would also
explain the results of the Stern–Gerlach experiment. Pauli [Pau26] developed a
two-component formalism, which took the spin of the electron into account. The
two-component states transform as spinors, see [BJ53] or [FK05].
Our treatment of spin below has the benefit of hindsight. We present in this
chapter a complete theory of the non-relativistic hydrogenic atom with a spinning
electron. We thereby correct the long-held assumption that this theory is somehow
incomplete, needing the insight provided by the Dirac theory to understand it.
This is because the spin is a completely non-relativistic phenomena, as has been
emphasized by Lévy-Leblond [LL74]. The Dirac theory is presented in this book
as an extension of the successful two-component Pauli formalism.
We use the operator Knr to characterize the states, which was introduced by
Biedenharn and Louck [BL81] in 1981 as the non-relativistic form of an operator
that was used by Dirac [Dir53]. We then show how we are led to a factorized form
of the Pauli Hamiltonian, which allows a solution of the problem analogously to
the harmonic oscillator. This method, which was developed for the treatment of a
great variety of problems by Infeld [Inf41], and Hull and Infeld [HI51] in 1951, then
suggests a supersymmetric interpretation, which is developed in the next chapter.
77
August 27, 2011 9:21 9in x 6in The Supersymmetric Dirac Equation b1217-ch10
where
ψ+
ψ= . (10.2)
ψ−
ψ+ is the spin-up component, ψ− the spin-down component. If the degeneracy of
the spin-up levels and the degeneracy of the spin-down levels is N 2 , the degeneracy
of ψ is 2N 2 . The two-component Hamiltonian HP is
p2
HP = I2 , (10.3)
2m
where I2 is the unit 2 × 2 matrix, and HP is called the Pauli Hamiltonian. This just
says that spin-up particles and spin-down particles have exactly the same dynamics.
If p is considered as a Hermitian operator in the two-dimensional complex
space then it may be expanded in terms of a basis of this space,
p = pi σi . (10.4)
If we are to have the usual relation between the square of this vector and the scalar
vector norm we must have
(σ·p)(σ·p) = (σ1 )2 p21 + (σ2 )2 p22 + (σ3 )2 p33
+{σ1 , σ2 }p1 p2 + {σ1 , σ3 }p1 p3 + {σ2 , σ3 }p2 p3
= p21 + p22 + p23 = p2 . (10.5)
This holds if
{σi , σj } = σi σj + σj σi = 2δij , (10.6)
where juxtaposition signifies matrix multiplication. Matrices which satisfy such a
relation are said to form a Clifford algebra. Since space is three-dimensional, there
are exactly three such matrices.
Let
σ1 σ2 σ3 = X. (10.7)
Then
X 2 = σ1 σ2 σ3 σ1 σ2 σ3 = σ2 σ3 σ2 σ3 = −I2 , (10.8)
and
X = σ1 σ2 σ3 = iI2 . (10.9)
August 27, 2011 9:21 9in x 6in The Supersymmetric Dirac Equation b1217-ch10
and in general
σi σj = iijk σk . (10.11)
Then
[σi , σj ] = σi σj − σj σi = 2iijk σk . (10.12)
where A and B are any two vectors which commute with σ. With the help of this
identity the Hamiltonian can be written as
(σ·p)(σ·p)
HP = . (10.16)
2m
The advantage of writing the Hamiltonian as in Eq. (10.16) can be seen when
we consider an electron in a magnetic field. The presence of the magnetic field can
be accounted for by the rule of minimal substitution:
p → π = p − eA, (10.17)
where A is the vector potential (in this paragraph only), which satisfies ∇ ×A = B,
and B is the magnetic field. The Hamiltonian becomes
1 π2 i
HP → Hmag = (σ·π)(σ·π) = + σ·π × π. (10.18)
2m 2m 2m
We now have
and
1 1
σ·π × π = mij σm [πi , πj ] = mij ijk σm (ie)Bk = (ie)σk Bk . (10.20)
2 2
Therefore
π2 eσ π2 e π2
Hmag = − ·B = −g S ·B = − µ·B, (10.21)
2m m 2 2m m 2m
where Si denote the spin matrices, Si = 12 σi , and the magnetic moment is µ.
µB = e/(2m) is called the Bohr magneton, and g is the gyromagnetic ratio, here
g = 2, with
µ = gµB S. (10.22)
so that the Si are the generators of an su(2)-algebra, and S2 = s(s + 1), where
s = 1/2 is the spin of the electron.
This is in agreement with the experimental results for the interaction of the
spin of the electron with a magnetic field (up to corrections due to quantum
electrodynamics). The “classical” value of g is g = 1, where
e
µ=g L, (10.24)
2m
and L is the orbital angular momentum. This is nowadays recognized as the value
appropriate for spin-one photons. In fact, it is conjectured that g = 1/s in general.
The conjecture has been proved for s ≤ 2, and for any half-integral s ([LL74] and
references therein).
This method of introducing a two-component description for particles with
spin one-half, and the fact that this automatically leads to a value g = 2, was
apparently first noticed by Feynman [Sak67]. It copies Dirac’s method for the
relativistic case, but works just as well for non-relativistic fermions, and could
have been introduced in this case. It brings out the fact that spin is inherently a
non-relativistic phenomena [LL74].
The Clifford algebra structure of the Pauli matrices allows us to factorize
quadratic operators. Consider the quadratic operator L2 = L21 + L22 + L23 . We can
factorize this with the help of the spin operator σ. One finds
which is easily verified from Eq. (10.15) and the equation L × L = iL. The total
angular momentum is J = L + 12 σ. It factorizes as
2 3 1
J = σ·L + σ·L + . (10.26)
2 2
We now introduce, following Biedenharn and Louck [ BL81], the non-
relativistic analog of Dirac’s operator [Dir53]:
and
1 1
2
J = Knr + Knr − 2
= Knr − 1/4. (10.29)
2 2
Since Knr commutes with J2 and L2 this yields, when applied to a simul-
taneous eigenvector, Knr → κ, L2 → l(κ)(l(κ) + 1), J2 → j(κ)(j(κ) + 1) with
κ, κ positive
l(κ) = (10.30)
|κ| − 1, κ negative
and
v × L + L × v = 2iv. (10.34)
August 27, 2011 9:21 9in x 6in The Supersymmetric Dirac Equation b1217-ch10
j/m m= 1
2
m = − 12
1 l+m+ 12 l−m+ 12
j =l+ 2 2l+1 2l+1
l−m+ 12 l+m+ 12
j =l− 1
2 − 2l+1 2l+1
We note that (σ·r̂) is a scalar operator with respect to rotations, so that (σ·r̂)χκm
has the same j and m as χκm . But since (σ·r̂) anticommutes with Knr (see theorem
of Section 10.2), the result of (σ·r̂)χκm must be proportional to χ−κm . That is,
(σ·r̂)χκm = cχ−κ
m . (10.47)
August 27, 2011 9:21 9in x 6in The Supersymmetric Dirac Equation b1217-ch10
(σ·r̂)(σ·r̂) = 1, (10.48)
we have |c|2 = 1 and c = ±1. The sign is fixed by choosing r̂ = eˆ3 = (0, 0, 1),
which is equivalent to taking θ = 0. From Eq. (10.42)
2l + 1
Yml (ê3 ) = δm0 , (10.49)
4π
so we have
2l(κ) + 1
χκm (r̂) = C(l(κ) 12 j(κ); 0m) ξ m . (10.50)
4π
Equation (10.46) becomes
2m( 2lA + 1) C lA 12 jA ; 0m = c 2lB + 1 C lB 12 jB ; 0m , (10.51)
(σ·r̂)χ−κ
m = −χm .
κ
(10.55)
an expression for the eccentricity, which is a key concept of the book, again for
the relativistic as well as the non-relativistic cases.
Using the definition
and the fact that (σ·A) anticommutes with Knr , one finds the result that
|A|2 + L2 + 1 = N 2 . (10.58)
(σ·A)2 = N 2 − Knr
2
. (10.61)
Hence
and
α(E, κ) = N 2 − κ2 . (10.63)
1
For A = (N 2 ) 2 A we get
(σ·A)ψN κm = − 1 − κ2 /N 2 ψN,−κ,m . (10.64)
We shall see, in Section 10.7, that this is the expression for the eccentricity.
August 27, 2011 9:21 9in x 6in The Supersymmetric Dirac Equation b1217-ch10
A−
l+1 RN,l+1 (r) = gl+1 RN l (r), A+
l+1 RN l = gl+1 RN,l+1 (r), (10.72)
with
12
1 1 l Zαm
gl = Zαm − 2 , A∓
l = ±ipr + − , (10.73)
l2 N r l
and A+ = (A− )† .
Applying the raising operator to RN,l (r), and then the lowering operator to
the resulting RN,l+1 (r), yields
A− + 2
l+1 Al+1 RN l = g(l+1) RN l , (10.74)
or
d (l + 2) Zαm d l Zαm
+ − − + − RN l (r)
dr r l+1 dr r l+1
1 1
= (Zαm)2 − RN l (r). (10.75)
(l + 1)2 N2
This can be simplified to
1 d 2 d l(l + 1) 2Zαm
r − + − 2mEN RN l (r) = 0, (10.76)
r2 dr dr r2 r
with EN given by Eq. (8.48). We have recovered the Schrödinger equation for the
radial part of the Pauli wave function. But since it has now been factorized into the
product of two factors, which raise and lower l without changing the energy, we
can now explain the degeneracy in l, and the degeneracy 2N 2 of an N -multiplet.
The tower of states terminates when A+ l+1 RN l = 0, that is, when gl+1 = 0.
From Eq. (10.73) we see that this implies N = l+1. From the equation A+ N RN l = 0
we further have
d N −1 1
− + − RN,N −1 (r) = 0, (10.77)
dr r N
and hence
N −1
2(Zαm)r
e−
Zαm
RN,N −1 (r) = NN,N −1 N r . (10.78)
N
The radial wave function is nodeless. These orbits are called “circular”. For the
ground state (N = 1) we have
All other eigenfunctions in the N-multiplet can be generated from the explicit
result for the eigenfunction with l = N − 1. Taking κ = N − 1 (since l(N − 1) =
N − 1) we can use this same eigenfunction in Eq. (10.70), which now functions
as a lowering operator, generating the eigenfunction for l = N − 2. Iterating this
procedure yields all the eigenfunctions in an N-multiplet.
To summarize: The properties of the operator Anr are equivalent to a definition
of the radial eigenfunctions; moreover, the eigenvalue conditions deducible from
Anr → 0 are equivalent to the eigenvalue condition (obtained from analysis) for the
confluent hypergeometric function 1 F1 (a, b; z). The eigenvalue condition referred
to above is the condition that the confluent hypergeometric function reduces to a
polynomial for the parameter a being zero or a negative integer.
we have
e= 1 − κ2 /N 2 , (10.81)
with κ = ±1, ±2, . . ., ±(N − 1), −N . This corresponds to the operator equation
1
eop = [1 − Knr
2
/N 2 ] 2 . (10.82)
Notes on Chapter 10
This chapter follows essentially the work of Biedenharn and Louck [Bie62].
The importance of Clifford algebras for the description of particles with spin is
especially emphasized in geometrical algebra; see Doran and Lasenby [DL03].
Spinors are discussed, at an elementary level, by Bade and Jehle [BJ53], and in the
framework of geometric algebra, by Doran and Lasenby [DL03] and Francis and
Kosowsky [FK05]. The factorization method was developed by Infeld [Inf41] and
Hull and Infeld [HI51].
August 27, 2011 9:21 9in x 6in The Supersymmetric Dirac Equation b1217-ch11
Chapter 11
Elements of Supersymmetric Quantum Mechanics
The theory of supersymmetry has its origins in quantum field theories [SDB04]
and elementary particle physics [WZ74], but it is also a valuable tool for describing
the solutions of many ordinary quantum mechanical systems [Wit90], [RLB90]. It
is used, for example, to provide a simple proof of the positive-mass conjecture in
general relativity [Wit81], and in the proof of various index theorems in differential
geometry [Tak08]. The following is a summary of the most pertinent definitions
that we need in our discussion of supersymmetry in non-relativistic hydrogenic
atoms and, later, in the Dirac–Kepler problem.
V = V1 ⊕ V2 , (11.1)
such that the vectors of V1 are even with respect to some parity P, while those of
V2 are odd. A general vector in V may then be written as the sum of an even and
an odd vector:
|ν1 |ν1 0
|ν = = + . (11.2)
|ν2 0 |ν2
91
August 27, 2011 9:21 9in x 6in The Supersymmetric Dirac Equation b1217-ch11
The element Ω11 effects an even-even transition. Similarly, the element Ω22 is
odd-odd. The element Ω12 describes the transformation of an odd state into an
even one, the element Ω21 the transformation of an even state into an odd one. The
linear operators form a graded algebra. The even and the odd operators are called
homogeneous.
The set of linear operators define a superalgebra if it is closed under a
generalized commutator, which, for two homogeneous operators Ω1 and Ω2 , is
the ordinary commutator between Ω1 and Ω2 unless both elements are odd. In
the latter case, it is the anticommutator between Ω1 and Ω2 . The definition of the
generalized commutator is extended to arbitrary operators by requiring it to be a
bilinear operation.
Consider the so-called s(2) superalgebra defined by the supercharges
1 0 A+ i 0 A+
Q1 = √ − , Q2 = √ − , (11.4)
2 A 0 2 −A 0
where the operators A+ and A− are the Hermitian adjoints of each other. The
operators
1 0 0
Q = √ (Q1 + iQ2 ) = (11.6)
2 A− 0
and
1 0 A+
Q† = √ (Q1 − iQ2 ) = (11.7)
2 0 0
are the supersymmetry generators. The parity operator that is responsible for the
grading is represented by the operator
1 0
P= . (11.8)
0 −1
Therefore
[Q1 , H] = [Q1 , Q21 + Q22 ] = [Q1 , Q22 ] = −Q2 {Q1 , Q2 } + {Q1 , Q2 }Q2 = 0.
(11.11)
Similarly [Q2 , H] = 0.
Thus, the supersymmetric Hamiltonian is the sum of the squares of the
supersymmetric charges, or the anticommutator of the supersymmetry generators.
These are relations that characterize supersymmetric quantum mechanics, with the
proviso that H+ and H− be closely related to the Hamiltonians of actual physical
systems. We say that a system which satisfies these relations is an s(2) superalgebra,
and we speak of an S(2) supersymmetry.
Assume now that the vector |ν defined by Eq. (11.2) is the eigenvector of the
supersymmetric Hamiltonian with eigenvalue E:
|ν1 H+ 0 |ν1 |ν1
H = =E . (11.12)
|ν2 0 H− |ν2 |ν2
|A− |ν1 |2 = ν1 |A+ A− |ν1 = A− ν1 |A− ν1 = Eν1 |ν1 ,
|A+ |ν2 |2 = ν2 |A− A+ |ν2 = A+ ν2 |A+ ν2 = Eν2 |ν2 , (11.13)
where we have used that A+ and A− are the Hermitian adjoints of each other.
These relations show that E is positive or zero. They also show that A− |ν1 and
A+ |ν2 vanish if and only if E is zero.
Therefore the vectors
+
|ν1 0 † 0 A |ν2
Q = , Q = (11.14)
0 A− |ν1 |ν2 0
and H− with positive energy are paired. They are transformed into each other by
the operators A− and A+ . Thus the levels of the super-Hamiltonian are doubly
degenerate. If, however, either H+ or H− has an eigenstate with energy zero, then
there is no partner state and, hence, no degeneracy.
We have seen that an eigenstate of H with energy zero is annihilated by Q and
Q† . If, conversely, an eigenstate of H is annihilated by Q and Q† , then its energy
is zero. This follows from Eq. (11.10), according to which {Q, Q† } = H.
For supersymmetry to be an exact symmetry, one requires that H have a
supersymmetrically invariant ground state, i.e., a state that is annihilated by Q and
Q† . Such a ground state has zero energy, and either |ν1 or |ν2 is zero.
We have thus seen that if supersymmetry is an exact symmetry, then the two
Hamiltonians H+ and H− have the same set of eigenvalues, except that one of
them has a zero eigenvalue that is not shared by the other Hamiltonian.
This completes our summary of elementary supersymmetric quantum
mechanics. We shall now turn to its application to non-relativistic hydrogenic atoms
with spin.
1
Hmag = (π 2 + πy2 ) − gµB BS3 , (11.16)
2m x
Inserting this in the foregoing equations, then the comparison with the formula for
H, Eq. (11.16) — and the demand that g = 2 — leads to the equation Q21 = Q22 =
2 H. It follows that H commutes both with Q1 and Q2 . Together with Eq. (11.20)
1
In terms of Q, Q† we have
Thus we have the supersymmetry algebra s(2). We may say that the requirement
g = 2 is equivalent to demanding that the Pauli Hamiltonian, together with
August 27, 2011 9:21 9in x 6in The Supersymmetric Dirac Equation b1217-ch11
When Q acts on the vector |ν then we have (σ · A) acting on the positive κ
component, so that (σ·A) lowers l(κ) by one unit, whereas Q† acts on the negative
κ component, and raises l(κ) by one unit.
We also have
|ν1 † |ν1 2 |ν1 |ν1
H = {Q, Q } = (σ·A) =E , (11.26)
|ν2 |ν2 ν2 |ν2
with
κ2
E= 1− 2 , (11.27)
N
0.0
E [eV]
-13.6
l=0 l=1 l=1 l=2
j=1/2 j=3/2
The constructed S(2) supersymmetry has implications for the spectrum. The
consequences of a (super-) symmetry group of the Hamiltonian is that the energy
eigenspaces consist of irreducible representations of the group. The irreducible
representations of S(2) are either one- or two-dimensional. States within the
subspace Hκ , with fixed j and m, transform irreducibly under the superalgebra
s(2). The multiplets have dimensionality two, unless E = 0, in which case they are
one-dimensional.
We can now interpret the spectrum of a non-relativistic spin- 12 particle in a
Coulomb field (see Figure 11.1):
1. The degeneracies of levels with fixed j and m but with l(κ) = j ± 12 are a
consequence of the s(2) supersymmetry algebra, constructed above. The level
of degeneracy is two, except for states with l = j − 12 and
m (Zα)2
E=− , (11.28)
2 (l + 1)2
Notes on Chapter 11
The relation of the factorization method to supersymmetry has been emphasized
by Stahlhofen and Bleuler [SB89]. The application of supersymmetry to the non-
relativistic hydrogen spectrum was performed by Tangerman and Tjon [TT93]. The
various applications of supersymmetry to quantum mechanics are reviewed in Roy,
Lahiri and Bagchi [RLB90] and Khare, Cooper and Sukhatma [KCS95].
August 27, 2011 9:21 9in x 6in The Supersymmetric Dirac Equation b1217-ch12
Chapter 12
Sommerfeld’s Derivation of the Relativistic
Energy Level Formula
99
August 27, 2011 9:21 9in x 6in The Supersymmetric Dirac Equation b1217-ch12
p2θ
p2 = p2r + , (12.2)
r2
where the radial momentum is
mṙ
pr = √ , (12.3)
1 − v2
and the angular momentum is
mr2 θ̇
pθ = √ . (12.4)
1 − v2
From angular momentum conservation, the component pθ is a constant of the
motion. We define as before the variable s = 1/r, and note that
pr ds
=− . (12.5)
pθ dθ
Introducing these variables into Eq. (12.1), we find that
H + Zαs
2 p 2 ds 2
θ
=1+ + s2 . (12.6)
m m dθ
As before, one differentiates Eq. (12.6) with respect to θ and obtains the linear
differential equation
d2 s
+ γ 2 (s − C) = 0, (12.7)
dθ 2
where we have used the abbreviations
(Zα)2
γ2 = 1 − (12.8)
p2θ
August 27, 2011 9:21 9in x 6in The Supersymmetric Dirac Equation b1217-ch12
and
ZαH
C= . (12.9)
γ 2 p2θ
From Eq. (12.7) we obtain the general solution for the relativistic orbits:
1
s= = A cos (γθ) + B sin (γθ) + C, (12.10)
r
where A, B are the integration constants and C is given in Eq. (12.9). Defining the
distance of closest approach (perihelion) to occur at θ = 0 makes B = 0, yielding
the simpler equation
1
= A cos (γθ) + C. (12.11)
r
This result shows that the classical Kepler orbits have the form of conic
sections — just as in the non-relativistic case — but that the correct angle variable
(in terms of which the orbit is actually a conic section) is not θ but ψ = γθ. Thus,
for elliptic orbits (bound states) to move from one perihelion (ψ = 0) to the next
(ψ = 2π) requires θ = 2π/γ. Since the parameter γ is less than unity, classical
bound state relativistic Kepler motion takes place in an ellipse whose perihelion is
advancing (that is, moving in the same sense as the orbit is transversed).
We now express the elliptic orbits in geometrical terms, using the eccentricity e.
For ψ = 0 one has the perihelion distance r = a(1−e) and, for ψ = π, the aphelion
distance r = a(1 + e). The orbit equation, Eq. (12.11), is then
1 1 1 + e cos ψ
= . (12.12)
r a 1 − e2
Except for the occurrence of ψ = γθ in place of θ, one sees that the orbits are
exactly of non-relativistic form. In other words, in a moving reference frame, such
that the perihelion appears fixed, the orbit is non-relativistic in form.
It is now easy to derive Sommerfeld’s formula by applying the Sommerfeld–
Wilson quantization rules to the phase integrals for pθ and pr :
θ=2π
pθ dθ = 2πnθ , (12.13)
θ=0
so that we obtain
2
1 dr e2 γ sin2 ψ
pr dr = pθ dθ = pθ dψ, (12.16)
r dθ (1 + e cos ψ)2
upon using the orbit Eq. (12.12). Thus, the radial quantization condition,
Eq. (12.14), takes the form
ψ=2π 2 2
1 e sin ψ dψ nr
2
= . (12.17)
2π ψ=0 (1 + e cos ψ) γn θ
Notice that the left-hand side of Eq. (12.17) is identical to that for non-
relativistic motion, since the integral involves only the geometry of the ellipse.
It has the value
ψ=2π 2 2
1 e sin ψ dψ
= (1 − e2 )− 2 − 1.
1
(12.18)
2π ψ=0 (1 + e cos ψ)2
By contrast, the right-hand side of Eq. (12.17) differs from the non-relativistic case
solely by the appearance of the factor γ. We may phrase this more suggestively
by saying that the sole effect of relativity in the radial quantization condition
consists in replacing the orbital angular momentum, nθ , by the “irrational angular
momentum” parameter, γnθ .
A = eC. (12.20)
August 27, 2011 9:21 9in x 6in The Supersymmetric Dirac Equation b1217-ch12
H 2 + (Zα)2 C 2 + 2H(Zα)C
+ [2(Zα)2 AC + 2H(Zα)A] cos (γθ) + (Zα)2 A2 cos2 (γθ)
(Zα)2
= m + nθ A 1 −
2 2 2
+ C 2 + 2AC cos (γθ)
n2θ
A2 (Zα)2 2
+ cos (γθ) . (12.21)
n2θ
The coefficients of the terms quadratic in cos (γθ) agree, they are (Zα)2 A2 . The
coefficients of the terms linear in cos (γθ) agree, because
(Zα)2
HZα = n2θ 1 − C = n2θ γ 2 C. (12.22)
n2θ
The constant terms yield
2 (Zα)4 2(Zα)2 (Zα)2
H 1+ 4 4 + 2 2 = m2 + H 2 2 4 γ 2 e2 + 1 , (12.23)
nθ γ nθ γ nθ γ
or
(Zα)2 m 2
1+ [(1 − γ 2
) + 2γ 2
− γ 2 2
e − 1] = . (12.24)
n2θ γ 4 H
This simplifies to
(Zα)2 2 m 2
1+ 2 γ (1 − e 2
) = , (12.25)
nθ γ 4 H
and, inserting (1 − e2 ) from Eq. (12.18):
(Zα)2 m 2
1+ = . (12.26)
(nθ γ + nr )2 H
Equivalently, for orbits for which H = E = constant,
− 12
(Zα)2
E =m 1+ . (12.27)
(nr + γnθ )2
This is the Sommerfeld formula for the energy levels of the bound states in the
relativistic regime.
For the ground state, nθ = 1, the Lorentz transformation factor γ would appear
in a transformation from the rest frame to a frame moving, at a given instant, in a
direction tangent to the orbit with a velocity v = Zα. Such a Lorentz transformation
gives rise to a time dilation between the fixed and transformed systems by the factor
August 27, 2011 9:21 9in x 6in The Supersymmetric Dirac Equation b1217-ch12
Chapter 13
The Dirac Equation
We are now ready to tackle the relativistic problem of an electron moving rapidly
in a Coulomb potential, using the full power of modern quantum mechanics. The
appropriate formalism was developed by Dirac [Dir28]. In the present chapter we
review the standard approach to this problem: The separation of variables in
spherical polar coordinates and the derivation of differential equations for the
radial components of the wave function. We emphasize that the treatment is a
direct generalization of the treatment of a non-relativistic spin- 12 particle, with
the three-dimensional Clifford algebra replaced by the four-dimensional Clifford
algebra, corresponding to a replacement of the two-component Pauli spinors by
four-component Dirac spinors ([BJ53], [FK05]).
{γ µ , γ ν } = γ µ γ ν + γ ν γ µ = 2g µν , (13.1)
with g µν the Minkowsky metric: gµν = diag(1, −1, −1, −1). The γ µ are the
generators of a four-dimensional Clifford algebra in Minkowski space, in the same
way as the σi are the generators of a three-dimensional Clifford algebra in Euclidian
space, {σi , σj } = 2δij .
The equation of motion for a free particle with mass m is given by
105
August 27, 2011 9:21 9in x 6in The Supersymmetric Dirac Equation b1217-ch13
where we identify the parameter m with the mass. Choosing the minus sign would
yield
which is the Dirac equation for a particle with negative mass −m. We shall consider
the Eq. (13.5) to be unphysical. It will nevertheless be used in connection with the
unphysical Hamiltonian H̄ of Eq. (15.1) below.
The gamma matrices are given (in the Dirac representation) by
I2 0 0 σ
γ0 = , γ= , (13.6)
0 −I2 −σ 0
with I2 the 2 × 2 unity matrix, and σ the Pauli matrices. We shall also need the
matrix
0 I2
γ5 = iγ 0 γ 1 γ 2 γ 3 = , (13.7)
I2 0
with i, j = 1, 2, 3, and
(αi )2 = β 2 = I4 . (13.10)
August 27, 2011 9:21 9in x 6in The Supersymmetric Dirac Equation b1217-ch13
For the stationary states ψ(t, r) = ψ(r)e−iEt , where E is the energy of the state,
and Eq. (13.12) reduces to
with
p = −i∇. (13.16)
We have the relation between the spin matrices and the quantities α:
α = γ5 Σ = Σγ5 . (13.18)
which holds for any two vectors A and B which commute with Σ. This is a
generalization of the identity (10.15).
The Dirac operator is a constant of the motion. Indeed, it commutes with the
Hamiltonian for an arbitrary central potential V(r).
There is a relation between the orbital angular momentum squared and the
Dirac operator squared:
P = βI, (13.33)
where I is the inversion operator which inverts the position and momenta,
The total angular momentum also commutes with the parity operation:
To specify the parity of the states is superfluous, since its eigenvalues are known
when those of {H, J2 , J3 , K} are given. We denote the eigenvectors, which we
assume are normalized to unity, by
κ = −σ·L − 1, (13.43)
j3
where Yjl (r̂) is a normalized spin-angular function formed by the combination
of the Pauli spinors with the spherical harmonics of order l. The radial functions
RA and RB depend on κ. The factor i is inserted to make RA and RB real for
bound-state solutions.
so
∂ Σ·L
Σ·p = i(Σ·r̂) − + . (13.50)
∂r r
Now rewrite Σ·L in terms of the Dirac operator K,
∂ βK + 1
Σ·p = −i(Σ·r̂) + . (13.51)
∂r r
This is the relativistic generalization of Eq. (10.66). Multiply this equation from
the left by γ5 and get, by Eq. (13.18):
∂ βK + 1
α·p = −i(α· r̂) + . (13.52)
∂r r
We obtain for the Hamiltonian the expression
∂ βK + 1
H = −i(α· r̂) + + βm + V (r). (13.53)
∂r r
With this Hamiltonian the equation for the bound states becomes
0 σ·r̂ ∂ 1 0 −σ·r̂ κ m 0
i − − −i + + V (r)
σ·r̂ 0 ∂r r σ·r̂ 0 r 0 −m
RA YA RA Y A
× =E . (13.54)
iRB YB iRB YB
where
Zα
na = √ . (13.61)
1 − 2
Since in the non-relativistic limit
1 (Zα)2
→1− + O((Zα)4 ) (13.62)
2 N2
we have
na → N (13.63)
in this limit. For this reason na is called the “apparent principal quantum number”.
In terms of the functions uA (x), uB (x) we find the coupled differential
equations
d κ 1 1 − Zα
− uB (x) = − uA (x),
dx x 2 1+ x
(13.64)
d κ 1 1 + Zα
+ uA (x) = + uB (x).
dx x 2 1− x
Notes on Chapter 13
This is the work of Dirac [Dir28], referred to in Chapter 1. The considerations here
are by now completely standard; see the texts by Bjorken and Drell [BD64], Sakurai
[Sak67], or Rose [Ros61]. The Dirac equation is here motivated as a generalization
of the Pauli equation in Chapter 10.
August 27, 2011 9:21 9in x 6in The Supersymmetric Dirac Equation b1217-ch14
Chapter 14
The Primary Supersymmetry of the Dirac Equation
thus replaces the key equation of our non-relativistic analysis of the spectrum of the
hydrogen atom. This equation expresses the primary supersymmetry of the Dirac
equation. The supersymmetry of the Dirac equation in a Coulomb field may be seen
just by looking at the pattern formed by the lines of the spectrum in Figure 14.1.
115
August 27, 2011 9:21 9in x 6in The Supersymmetric Dirac Equation b1217-ch14
E [eV]
κ = -1 κ=1 κ = -2 κ=2
j=1/2 j=3/2
In order that the K-odd operator A be a constant of the motion we must have
[A , H] = 0. To calculate this commutator we need the following terms:
Exercise 14.3 Prove [Kγ5 f (r), H] = iK(Σ· r̂)f (r) + 2mKγ5 βf (r).
2i
[A , H] = − x1 βKγ5 + x2 V (r)K(Σ· r̂)
r
− x3 f (r)K(Σ· r̂) − 2ix3 mβKγ5 f (r). (14.9)
We group diagonal and anti-diagonal operators separately and set Eq. (14.9) to
zero. This yields
x
+ mx3 f (r) + K(Σ· r̂)[x2 V (r) − x3 f (r)] = 0.
1
−2iβKγ5 (14.10)
r
This is fulfilled if the diagonal and the anti-diagonal terms vanish separately:
x1
= −x3 mf (r), x2 V (r) = x3 f (r). (14.11)
r
We integrate the second equation over the interval (r, ∞), with the requirement
that f (r) and V (r) tend to zero when r → ∞, and find
AK + KA = 0. (14.25)
So the parity operator P anticommutes with the complete expression for the
Johnson–Lippmann operator A,
AP + P A = 0. (14.28)
14.3 Eccentricity
We now derive an expression for the magnitude of the Johnson–Lippmann operator
in terms of the eigenvalues of the Dirac operator. We evaluate the operator A2 as
A† A, and derive the following expression for A2 :
i −i
A2 = (H − βm)γ5 K − Σ·r̂ Kγ5 (H − βm) − Σ·r̂
Zαm Zαm
1
= K 2 (H − βm)2 + 1
(Zαm)2
i
− [(Σ·r̂)Kγ5 (H − βm) − (H − βm)γ5 K(Σ·r̂)]. (14.29)
Zαm
Exercise 14.4 Prove K 2 (H − βm)2 = K 2 H 2 − m2 + 2mβ Zα
r .
Exercise 14.5 Prove (Σ · r̂)Kγ5 (H −βm)−(H −βm)γ5 K(Σ · r̂) = −2iβK 2 /r.
κ2
e2 = 1 − (1 − 2 ), (14.31)
(Zα)2
Pκ = K/|κ|, (14.40)
where K is the Dirac operator. It is represented by the matrix specified in Eq. (11.8).
The functions ψ(r; , |κ|) and ψ(r; , −|κ|) are transformed into each other
by the operator A, and they are both eigenfunctions of A2 with eigenvalue e2 .
Thus, ψ(r; ) is an eigenvector of the supersymmetric Hamiltonian. The relation
(14.33) shows that there is a one-to-one correspondence between the values of
and e in a given pair of stacks, i.e., for a fixed value of j and, hence, of |κ|. The
supersymmetric Hamiltonian H is thus closely tied to the Hamiltonian H, as it
should be.
Having completely identified the exact supersymmetry of the Dirac–Kepler
problem, we must determine the ψ(r; ) corresponding to the bottom level, for
August 27, 2011 9:21 9in x 6in The Supersymmetric Dirac Equation b1217-ch14
which the eigenvalue of H, and, hence, e is zero. We ask, accordingly, for a Dirac
spinor ψ0 (r; , |κ|) such that
In the notation of Eq. (13.48), we must determine radial functions RA (r) and
RB (r) satisfying the equation
RA (r)YA 0
A = . (14.42)
iRB (r)YB 0
which holds for any pair of radial functions. We use the matrix form of γ5 ,
Eq. (13.1), and the fact that ψ0 (r; , κ) is an eigenfunction of K with eigenvalue
κ. Then, using the explicit form of A, Eq. (14.18), in Eq. (14.42):
−i RA (r))YA
Aψ0 (r; , κ) = Kγ5 (H − βm) − Σ·r̂)
Zαm iRB (r)YB
κ ∂ βκ 1 RA (r)YB
=− + +
Zαm ∂r r r iRB (r)YA
κ −RB (r)YB RA (r)YB 0
− + = . (14.45)
mr iRA (r)YA iRB (r)YA 0
which, as we shall see below, is consistent in this case with (13.60) (because
|κ| = na ), we get instead
1 d
− Sign(κ)x + x + κ + 1 RA (x) = ZαRB (x),
2 dx
1 d
− Sign(κ)x + x − κ + 1 RB (x) = −ZαRA (x). (14.48)
2 dx
Inserting one of these equations into the other shows that both RA (x) and RB (x)
satisfy the equation
2
d 1
x − Sign(κ)x − γ 2 |κ|2 R(x) = 0, (14.49)
dx 2
with
(Zα)2
γ2 = 1 − . (14.50)
|κ|2
The second solution is physically unacceptable because of its behavior at the origin.
The first solution behaves properly at the origin, but it is only normalizable for
negative values of κ. Hence, we conclude that Eq. (14.41) only has a proper solution
for negative values of κ. Identifying the solution R(x) with RA (x) and inserting
August 27, 2011 9:21 9in x 6in The Supersymmetric Dirac Equation b1217-ch14
κ + γ|κ| Zα
RB (x) = RA (x) = RA (x). (14.54)
Zα κ − γ|κ|
The ψ0 (r; ) corresponding to the lowest energy of a double stack has now
been found to have the form
0
ψ0 (r; ) = , (14.55)
ψ0 (r; , κ)
with
YA
ψ0 (r; , κ) = R0 (x) κ+γ|κ| , (14.56)
i Zα
YB
and
R0 (x) = (const)xγ|κ|−1 e− 2 .
x
(14.57)
κ is negative and the functions YA and YB must be chosen according to the rules for
parallel coupling, described in Section 13.4. Thus, we have identified the ground
state of the supersymmetric Hamiltonian. The corresponding eigenvector satisfies
the equations
with γ as in Eq. (14.50). We also note that the relation na = |κ| implies that the
definitions (13.60) and (14.47) of the variable x coincide when e = 0.
Such states correspond to the “circular” orbits of the Bohr model. They are non-
degenerate under supersymmetry and hence correspond to the eigenvalue e = 0
for the operator A, which is the relativistic generalization of Anr = (σ·A). Thus
we have established the result that the states corresponding to the circular orbits
are those with vanishing eccentricity.
August 27, 2011 9:21 9in x 6in The Supersymmetric Dirac Equation b1217-ch14
for +κ and −κ separately. In fact, it suffices to solve the equations for one sign.
The solution corresponding to the other sign may then be obtained by Eq. (14.34).
Instead of solving Eq. (14.61) for a given value of κ, we may, of course,
focus on the radial equations for the same κ-value, and this is what we shall
do in the following chapter. In either case it is advantageous to exploit the way
supersymmetry reflects itself in an extended function space.
Notes on Chapter 14
The Johnson–Lippmann operator was announced as a conserved quantity of the
Dirac equation by Johnson and Lippmann [JL50]. The realization that it reduces to
the Laplace vector in the non-relativistic limit is due to Biedenharn [Bie83]. That
the Johnson–Lippmann operator generates a supersymmetry group was realized
by Nieto, Kostelecky and Traux [NKT85]. The parallelism with the non-relativistic
case was pointed out by Dahl and Jørgensen [DJ95].
August 27, 2011 9:22 9in x 6in The Supersymmetric Dirac Equation b1217-ch15
Chapter 15
Extending the Solution Space
H̄ = γ5 Hγ5 . (15.2)
ψ̄ = γ5 ψ (15.3)
H̄ ψ̄ = E ψ̄, (15.4)
127
August 27, 2011 9:22 9in x 6in The Supersymmetric Dirac Equation b1217-ch15
j, m and −κ. Thus, the sign of κ is reversed. The ψ̄-functions differing merely in
the sign of κ are transformed into each other by the operator Ā = γ5 Aγ5 in a
similar way as in Eq. (14.34).
It follows that the stationary bound states of the negative-mass Hamiltonian
define a supersymmetric structure that is, in all details, the same as the
supersymmetric structure defined by the positive-mass Hamiltonian, except that
the roles of κ and −κ are reversed.
If we now group the positive-mass states with a given κ-value together with the
negative-mass states with the same κ-value, then we obviously encounter a double
stack of exactly the same appearance as before. The question is then: Is it possible
to introduce a parity operator and generators such that this double stack can be
properly described by supersymmetry? It turns out that this is indeed the case.
When we group the positive- and negative-mass solutions together then the
states with the same κ-value form a new supersymmetric structure. The appearance
of this structure is a direct consequence of the supersymmetric structure that
combines states with opposite values of κ in the original Dirac–Kepler problem.
The latter structure was based on the presence of the Johnson–Lippmann operator.
We expect, accordingly, that this operator also plays a central role for the description
of the new structure.
There is a rather direct way of carrying a positive-mass state with a given
κ-value into a negative-mass state with the same κ-value, namely to apply the
operators A and γ5 in succession. Diagonalizing the operator γ5 A will therefore
separate the double stack into two columns with different eigenvalues of γ5 A. These
columns are not affected if the operator is multiplied by an arbitrary constant and
a diagonal operator added so as to obtain an operator whose eigenvalues merely
differ in sign. Such an operator defines a grading of the function space and, hence,
a center for the description of the supersymmetry.
Using the explicit form of the operator A, Eq. (14.18), we find its action on
Dirac spinors:
RA (r) −
κ
(1 − )RB (r) YB
RA (r)YA Zα
A =
κ . (15.5)
iRB (r)YB i RB (r) − (1 + )RA (r) YA
Zα
Hence
i −RB (r) +
κ
(1 − )RA (r) YA
RA (r)YA Zα
−iZαγ5 A = iZα .
iRB (r)YB κ
−RA (r) + (1 + )RB (r) YB
Zα
(15.7)
Since the spinors before and after the transformation have the same value of κ,
they also have the same spin and angular variable dependence. The transformation
can then be expressed as a transformation of the radial functions, i.e., as a linear
transformation in the space of functions VR spanned by RA (r) and RB (r). That is
RA κ( − 1) Zα RA
−iZαγ5 A = . (15.8)
RB −Zα κ( + 1) RB
The eigenvalues of this matrix are, by the Cayley characteristic equation, the roots of
κ( − 1) − λ Zα
det = 0, (15.9)
−Zα κ( + 1) − λ
i.e., λ = κ ± γ|κ|. The grading may therefore be defined by
KH
Γ=− − iZαγ5 A, (15.10)
m
which has the eigenvalues ±γ|κ|.
with
and
−1
γ|κ| 0
D ΓD = . (15.15)
0 −γ|κ|
where γ is as in Eq. (14.50). The grading operator is Pκ = Γ/γ|κ|. The new basis
vectors are
(Zα)c1 (Zα)c2
(w1 , w2 ) = D(uA , uB ) = (uA , uB ) .
−(γ|κ| + κ)c1 (γ|κ| − κ)c2
(15.17)
The vector (w1 , 0) is an eigenvector of Γ, referred to the new basis, corresponding
to the eigenvalue γ|κ|, the vector (0, w2 ) is an eigenvector of Γ corresponding to
the eigenvalue −γ|κ|. The inverse of Eq. (15.17) is
−1 1 (γ|κ| − κ)c2 −(Zα)c2
(uA , uB ) = D (w1 , w2 ) = (w1 , w2 ) .
det (γ|κ| + κ)c1 (Zα)c2
(15.18)
(15.20)
From this we see that if we put
c1 (κ + γ|κ|)(κ − γ|κ|)
= , (15.21)
c2 (κ − γ|κ|)(κ + γ|κ|)
and
c2 1 (κ − γ|κ|)2 (1 − ) − (Zα)2 (1 + ))
√
c1 4(Zα)γ|κ| 1 − 2
c2 1 (κ − γ|κ|)(κ + γ|κ|) −1 (κ)2 − γ 2 |κ|2
=− √ = ,
c1 2(Zα)γ|κ| 1 − 2 2γ|κ| 1 − 2
(15.24)
as well as
(κ)2 − γ 2 |κ|2
= ena , (15.25)
1 − 2
August 27, 2011 9:22 9in x 6in The Supersymmetric Dirac Equation b1217-ch15
or
The integral of a function times its derivative vanishes for acceptable functions.
Since the remaining integrals are equal, we find
1 1
w1 w1 = w2 w2 . (15.32)
x x
It is clear that these operators are the Hermitian conjugates of each other. The
equations may then be written
ena ena
b− w 1 = − w2 , b+ w2 = − w1 . (15.34)
2γ|κ| 2γ|κ|
The operator b+ is the analog of A+ : it transforms an even state into an odd state.
The operator b− is the analog of A− , it transforms an odd state into an even state.
The functions w1 and w2 take the place of ν1 and ν2 , respectively. The equations
of (15.34) imply that
2 2
ena ena
b+ b− w1 = w1 , b− b+ w2 = w2 . (15.35)
2γ|κ| 2γ|κ|
becomes γ, as shown in Eq. (14.59). Hence na = γ|κ|. Inserting this result in
(15.26) yields
dw1 1 γ|κ| dw2 1 γ|κ|
− − − w1 = 0, − − w2 = 0. (15.37)
dx 2 x dx 2 x
The solutions of these equations are
This gives
d2 w1 γ|κ|(γ|κ| − 1) na 1
+ − + − w1 = 0,
dx2 x2 x 4
(15.43)
d2 w2 γ|κ|(γ|κ| + 1) na 1
+ − + − w2 = 0.
dx2 x2 x 4
August 27, 2011 9:22 9in x 6in The Supersymmetric Dirac Equation b1217-ch15
(15.45)
2
x ddxG22 + (2γ|κ| + 2 − x) dG
dx + (na − γ|κ| − 1)G2 = 0.
2
For nr = 0, we have reproduced the result (15.39). For the other solutions,
we must determine the relative values of N1 and N2 . This is done in Appendix D.
The result is
N1 2γ|κ|(2γ|κ| + 1)
=− . (15.52)
N2 ena
Our final task is to express the original radial functions uA (x) and uB (x) in
terms of w1 (x) and w2 (x). Let us put
1 1
(c1 , c2 ) = , , (15.53)
(κ + γ|κ|)(κ − γ|κ|) (κ − γ|κ|)(κ + γ|κ|)
in accordance with Eq. (15.21). We get then, by Eq. (15.18),
uA (x) = Sign(κ)
2γ|κ| − (κ − γ|κ|)(κ − γ|κ|)w1 (x)
+ (κ + γ|κ|)(κ + γ|κ|)w2 (x) ,
(15.54)
uB (x) = 2γ|κ|1
− (κ + γ|κ|)(κ − γ|κ|)w1 (x)
+ (κ − γ|κ|)(κ + γ|κ|)w2 (x) ,
× (κ + γ|κ|)(κ + γ|κ|)1 F1 (−nr + 1, 2γ|κ| + 2; x) ,
(15.55)
uB (x) = N xγ|κ| e−x/2 (κ + γ|κ|)(κ − γ|κ|)1 F1 (−nr , 2γ|κ|; x)
+ ena x
2γ|κ|(2γ|κ|+1) (κ − γ|κ|)(κ + γ|κ|)1
× F1 (−nr + 1, 2γ|κ| + 2; x) .
The overall normalization constant corresponding to the normalization
condition
[u2A (x) + u2B (x)]dx = 1, (15.56)
Notes on Chapter 15
This type of extension has been widely discussed in the literature. Sukumar
[Suk85a; Suk85b] found the supersymmetric structure of the equations in this
decomposition, and utilized it to solve them. It was also used by Jarvis and
Stedman [JS86], who solved the second-order Kramer’s equation, rather than
the first-order Dirac equation (see Chapter 17). The same equation was solved
by Swainson and Drake [SD91], and by Goodman and Ignjatovi¢ [GI97], in an
article that prompted an avalanche of papers tracing earlier work on this subject
([Sta98], [Wal79], [Kol66], [Won86], [WY82], [AB78], [Bie83]). This method
leads to a different relative normalization of the solutions than that used here. The
uncoupled differential equations had been previously obtained by Infeld [Inf41].
The uncoupled differential equations of (15.43) may be said to arise from the
coupled equations of (15.26) by factorization, according to the general method of
Hull and Infeld [HI51]. The advantage of the method used here, which utilizes
the primary supersymmetry discussed in Chapter 14, is that it leads directly to the
solutions of the Dirac equation, and makes clearer the nature of the coefficients
appearing.
August 27, 2011 9:22 9in x 6in The Supersymmetric Dirac Equation b1217-ch15
Chapter 16
A Different Extension of the Solution Space
ψ̃ = (Σ·r̂)ψ, (16.2)
where ψ describes a stationary bound state of the original Hamiltonian. Like βγ5 ,
the operator Σ·r̂ commutes with J and anticommutes with K. The function ψ̃ bears,
therefore, a similar relation to ψ as does the function ψ̄. Hence, the stationary bound
states of the Hamiltonian H̃ also define a supersymmetric structure that is, in all
details, the same as the supersymmetric structure defined by the positive-mass
Hamiltonian, except that the roles of κ and −κ are reversed.
In a similar way to what we did for the negative-mass states, we may now
group the positive-mass states with a given κ-value together with the new ψ̃ states
with the same κ-value and obtain a double stack of exactly the same appearance
as previously. Again, we may ask if it is possible to introduce a parity operator and
generators such that this double stack can be properly described by supersymmetry.
We now explain how this comes about.
139
August 27, 2011 9:22 9in x 6in The Supersymmetric Dirac Equation b1217-ch16
∆ = 1 + (Σ·r̂)A (16.9)
Now replace the equations of (16.23) with the equations obtained from them
by multiplying them from the left by x. This gives
du1 x
−x + − na u1 = −ena u2 ,
dx 2
(16.26)
du2 x
x + − na u2 = −ena u1 .
dx 2
Let us multiply the first of these equations from the left by u2 , the second by u1 ,
and integrate over x. A comparison of the resulting equations gives
d
d
u2
x
u1 = u1
−x
u2 . (16.27)
dx dx
This relation justifies the designation normal modes for the functions u1 (x) and
u2 (x).
August 27, 2011 9:22 9in x 6in The Supersymmetric Dirac Equation b1217-ch16
a+ a− u1 = (ena )2 u1 ,
(16.31)
a− a+ u2 = (ena )2 u2 .
Hence, the analog of Eq. (16.30) becomes
a+ a− 0 u1 2 u1
= (ena ) . (16.32)
0 a− a+ u2 u2
Thus, the quantity (ena )2 replaces the quantity E as the eigenvalue of the
supersymmetric Hamiltonian. It is positive or zero, as it should be, and it is a
well-defined function of the physical energy. It becomes zero when e = 0.
The function u2 (x) equals x times the R0 (x) of Eq. (14.57), as we would expect.
From Eq. (16.18) we find
uB 1− 1−γ
=− =−
uA 1+ 1+γ
1 − γ2 Zα
=− =−
1+γ |κ|(1 + γ)
−|κ| + γ|κ| κ + γ|κ|
= = , (16.36)
Zα Zα
since κ is negative, in accordance with Eq. (14.56).
The substitutions
u1 (x) = xγ|κ| e−x/2 F1 (x),
(16.39)
u2 (x) = xγ|κ| e−x/2 F2 (x),
introduce new radial functions, proportional to F1 (x) and F2 (x), that must satisfy
the equations
d2 F1 dF1
x + (2γ|κ| + 1 − x) + (na − γ|κ| − 1)F1 = 0,
dx2 dx
(16.40)
d2 F2 dF2
x 2 + (2γ|κ| + 1 − x) + (na − γ|κ|)F2 = 0.
dx dx
The regular solutions of these equations are confluent hypergeometric functions;
hence, we get
u1 (x) = N1 xγ|κ| e−x/2 1 F1 (−nr + 1, 2γ|κ| + 1; x),
(16.41)
u2 (x) = N2 xγ|κ| e−x/2 1 F1 (−nr , 2γ|κ| + 1; x),
where nr is the radial quantum number:
1√ na − γ|κ|
uA (x) = 1 + u1 (x) − u2 (x) ,
2 na + γ|κ|
(16.45)
1√ na − γ|κ|
uB (x) = 1 − u1 (x) + u2 (x) .
2 na + γ|κ|
N√
uA (x) = 1 + xγ|κ| e−x/2 [nr 1 F1 (−nr + 1, 2γ|κ| + 1; x)
2
− (na − γ|κ|)1 F1 (−nr , 2γ|κ| + 1; x)],
(16.46)
N√
uB (x) = 1 − xγ|κ| e−x/2 [nr 1 F1 (−nr + 1, 2γ|κ| + 1; x)
2
+ (na − γ|κ|)1 F1 (−nr , 2γ|κ| + 1; x)].
1/2
1 Γ(nr + 2γ|κ| + 1)
N = , (16.47)
Γ(2γ|κ| + 1) na (na − κ)nr !
Notes on Chapter 16
The standard method of solving the radial equations (13.64) is by introducing a
normal-mode representation. This method was introduced by Gordon [Gor28].
It is also used by Bethe and Salpeter [BS57]. We now understand that in these
works one is actually diagonalizing the operator ∆, without paying attention to
the underlying supersymmetry. With proper attention paid to this symmetry the
procedure is simplified and the algebra becomes more transparent. The decoupled
equations (16.38) may be said to arise from the coupled equations (16.23) by
factorization.
August 27, 2011 9:22 9in x 6in The Supersymmetric Dirac Equation b1217-ch16
Chapter 17
The Relation of the Solutions to Kramer’s Equation
The solutions of the Dirac equation have been widely investigated in the framework
of a second-order differential equation called Kramer’s equation. The solutions of
Kramer’s equation are superpositions of the solutions of Dirac’s equation and the
solutions of an unphysical equation. When we have found the solutions of Kramer’s
equation we still have to project out the solutions of Dirac’s equation [MG58]. This
used to involve additional work. But in the present approach it turns out that this
work is already done.
In this chapter we work out the relations between the solutions given here and
those obtained by the solution of Kramer’s equation. We find that, when Kramer’s
equation is expressed in terms of the quantity Γ, it reduces to the differential
equations for w1 , w2 . When it is expressed in terms of ∆, it reduces to the
equations for u1 , u2 . So we can also find the solutions to the Dirac equation,
uA , uB , by solving Kramer’s equation for w1 , w2 , or u1 , u2 , and then using the
known relations between w1 , w2 and uA , uB , or between u1 , u2 and uA , uB . In
the following sections we shall see how this works out in detail.
Ω11 Ω12
Ω(F1 , F2 ) = (F1 , F2 ) , (17.1)
Ω21 Ω22
S = c1 F1 + c2 F2 (17.2)
149
August 27, 2011 9:22 9in x 6in The Supersymmetric Dirac Equation b1217-ch17
S = a1 F1 + a2 ΩF1
S = b1 F2 + b2 ΩF2 . (17.4)
c1 = a1 + a2 Ω11 ,
c2 = a2 Ω12 . (17.6)
This gives
c2 Ω11
a1 = c1 − ,
Ω12
c2
a2 = , (17.7)
Ω12
and, hence,
c1
S = a2 Ω12 − Ω11 F1 + ΩF1 . (17.8)
c2
so that
c1 = b2 Ω21 ,
c2 = b1 + b2 Ω22 , (17.10)
August 27, 2011 9:22 9in x 6in The Supersymmetric Dirac Equation b1217-ch17
or
c1 Ω22
b1 = c2 − ,
Ω21
c1
b2 = , (17.11)
Ω21
and
c2
S = b2 Ω21 − Ω22 F2 + ΩF2 . (17.12)
c1
Now, Eq. (17.3) allows us to write
c2 c1
λ = Ω11 + Ω21 = Ω22 + Ω12 . (17.13)
c1 c2
We conclude, therefore, that if the matrix is traceless,
By taking the difference between the two expressions for S, we finally get
Hence, (a2 , −b2 ) is an eigenvector of the matrix in Eq. (17.3) with eigenvalue −λ.
That −λ is an eigenvalue when λ is, is a consequence of the assumption that the
matrix is traceless, Eq. (17.14).
problems of the kind discussed in the previous sections, the salient feature of these
problems being that the two-dimensional matrix to be diagonalized has zero trace.
The radial eigenfunctions of the operator Γ may be read off from the equations
of (15.18):
w1 uA c1 Zα −c1 (κ + γ|κ|) uA
=D = . (17.17)
w2 uB c2 (Zα) c2 (−κ + γ|κ|) uB
Each of them parallels the function S in Eq. (17.2). That function may also be
expressed in the form (17.4), with factors a2 and b2 as described in the comments to
Eq. (17.16); after identifying the vectors F1 and F2 of the previous section with the
functions uA and uB we find that we may write the radial function corresponding
to the eigenvalue γ2 = −γ|κ| as
Therefore
1
γ|κ|uA (x) + ΓuA (x) = w1 (x),
Zαc2
1
γ|κ|uB (x) + ΓuB (x) = w1 (x). (17.21)
(κ − γ|κ|)c2
w2 = c1 (Zα)(−γ|κ|uA + ΓuA ), or
w2 = c1 (κ + γ|κ|)(−γ|κ|uB + ΓuB ); (17.23)
To determine the spinors ψδ1 and ψδ2 , we exploit the expression (16.12),
which allows us to write for the radial function corresponding to the eigenvalue
δ1 = κ/na the expressions
d2 κ
u1 = − √ uA + ∆uA , or
1 + na
d2 κ
u1 = − √ uB + ∆uB . (17.25)
1 − na
This gives that
√
1 + YA
(3)
xψδ1 (x) = N u1 (x) √ . (17.26)
i 1 − YB
Finally, we get for the radial function corresponding to the eigenvalue λ2 = −κ/na
the expressions
d1 κ
u2 = √ − uA + ∆uA , or
1+ na
d1 κ
u2 = − − uB + ∆uB , (17.27)
1− na
and, hence,
√
1 + YA
(4)
xψδ2 (x) = N u2 (x) √ . (17.28)
i 1 − YB
state, which has no supersymmetric partner. In this case, and in this case
alone,
(1) (κ − γ|κ|) YA (1 ) YA
xψγ1 (x) = N w1 (x) =N w1 (x) ,
iZα YB i κ+γ|κ|
Zα
YB
(17.29)
where w1 (x) = xR0 (x), and we have used Eq. (14.54). Similarly,
√
1 + YA YA
xψδ2 (x) = N u2 (x) √
(4)
=N (4 )
u2 (x) κ+γ|κ| ,
i 1 − YB i Zα YB
(17.30)
where u2 (x) = xR0 (x), and we have used (16.36).
The eigenspinors that we have found above are, with respect to supersymmetric
quantum mechanics, the parallels of the eigenfunctions of the Johnson–
Lippmann operator A. Thus, A turns the supersymmetric partners |, j, mj , κ
and |, j, mj , −κ > into each other (Eq. (14.34)). It has the eigenvalues ±e and
the eigenfunctions (14.36). Similarly, the operator Γ turns the spinors ψ and ψ̄
into each other. It has the eigenvalues ±γ|κ| and the eigenfunctions (17.29) and
(17.22). Finally, the operator ∆ turns the spinors ψ and ψ̃ into each other. It has
the eigenvalues ±γ|κ| and the eigenfunctions (17.22) and (17.24).
Zα
O+ = β(H − E) = m + βγ5 Σ·p − β E + ,
r
Zα
O− = β(H̄ − E) = −m + βγ5 Σ·p − β E + . (17.31)
r
August 27, 2011 9:22 9in x 6in The Supersymmetric Dirac Equation b1217-ch17
The operators O+ and O− commute. Both ψ and ψ̄ will therefore satisfy the
equation
2
Zα
O− O+ ψ = −m2 + E + + βγ5 (Σ·p)βγ5 (Σ·p)
r
Zα Zα
−βγ5 (Σ·p)β E + −β E+ βγ5 (Σ·p) ψ = 0.
r r
(17.32)
Now
and
Zα Zα
−βγ5 (Σ·p)β E + −β E+ βγ5 (Σ·p)
r r
Zα Zα
= γ5 (Σ·p) E + − γ5 (Σ·p) E +
r r
1 i
= Zαγ5 , (Σ·p) = 2 Zαγ5 (Σ·r̂). (17.34)
r r
1 ∂ 2 ∂ K 2 − Z 2 α2 βK 2ZαE
−(m2 − E 2 ) + 2 r − − 2 +
r ∂r ∂r r2 r r
i 1 ∂ 2 ∂
+ 2 Zαγ5 Σ·r̂ ψ = −(m2 − E 2 ) + 2 r
r r ∂r ∂r
K 2 − Z 2 α2 Γ 2ZαE
− + 2+ ψ = 0, (17.37)
r2 r r
where
The operator Γ agrees with the expression we have been using, Eq. (15.10):
KH
Γ=− − iZαγ5 A, (17.39)
m
because
KH −i
Γ=− − iZαγ5 Kγ5 (H − βm) − Σ·r̂
m Zαm
KH H
=− − γ5 Kγ5 − β + iZαγ5 (Σ·r̂)
m m
KH H
=− +K − β + iZαγ5 (Σ·r̂)
m m
Γ2 = K 2 − (Zα)2 . (17.42)
August 27, 2011 9:22 9in x 6in The Supersymmetric Dirac Equation b1217-ch17
Inserting Eq. (17.42) into Eq. (17.37) yields Kramer’s equation in the
form
1 ∂ 2 ∂ Γ(Γ − 1) 2ZαE
r − + + (E − m ) ψ = 0.
2 2
(17.43)
r2 ∂r ∂r r2 r
We can rewrite the kinetic energy term with the identity
1 ∂ 2 ∂ ∂2 ∂ 1 ∂2
2
r = 2 + 2r = r. (17.44)
r ∂r ∂r ∂r ∂r r ∂r 2
Going over to the variable x, defined by Eq. (13.60), Eq. (17.37) then becomes
∂2 na 1
x−1 2 x − x−2 Γ(Γ − 1) + − ψ = 0. (17.45)
∂x x 4
Assume now that ψ is an eigenfuntion of Γ with eigenvalue +γ|κ|. We may then
replace the operator Γ in Eq. (17.46) by its eigenvalue γ|κ|, by which the whole
operator in front of xψ becomes independent of spin and angular coordinates.
Multiplying on the left by x we get:
2
∂ γ|κ|(γ|κ| + 1) na 1
+ + − xψ = 0. (17.46)
∂x2 x2 x 4
The equation may now be factorized, in accordance with Eq. (17.22), and we obtain
a radial function that is seen to satisfy the upper equation in (15.43):
d2 w1 γ|κ|(γ|κ| − 1) na 1
+ − + − w1 = 0. (17.47)
dx2 x2 x 4
Similarly, the eigenvalue −γ|κ| gives a radial function that satisfies the lower
equation in (15.43):
d2 w2 γ|κ|(γ|κ| + 1) na 1
+ − + − w2 = 0. (17.48)
dx2 x2 x 4
Thus, solutions of Kramer’s equation are solutions of the differential equations for
w1 (x) and w2 (x). These are standard differential equations, which have explicit
solutions in terms of confluent hypergeometric functions. These are connected to
the radial solutions of the Dirac equation by the known matrix of Eq. (15.17).
The fact that the differential equations (15.43) may be derived from Kramer’s
equation seems to give the functions w1 (x) and w2 (x) an elevated status with
respect to the Dirac equation. The functions u1 (x) and u2 (x) that arise from
diagonalizing the operator ∆ lead, however, to simpler expressions for uA (x) and
uB (x) and, unlike w1 (x) and w2 (x), the functions u1 (x) and u2 (x) are mutually
orthogonal. Accordingly, it is of interest to note that the second-order differential
August 27, 2011 9:22 9in x 6in The Supersymmetric Dirac Equation b1217-ch17
equations for u1 (x) and u2 (x) may also be derived from Kramer’s equation. To
show this, we insert the form of the Hamiltonian into expression (15.10) for Γ. We
make the assumption that κ is a good quantum number and therefore replace K by
κ in expression (17.38) for Γ. We may then show that Γ can be written
∂ Zαm
Γ= r− r[1 + (Σ·r̂)A]. (17.49)
∂r κ
Indeed
−i
(Σ·r̂)A = (Σ·r̂)Kγ5 (H − βm) − 1, (17.50)
Zαm
and therefore
−i i
1 + (Σ·r̂)A = (Σ·r̂)Kγ5 (H − βm) = (Σ·r̂)γ5 K(H − βm)
Zαm Zαm
i iκ
= (Σ·r̂)γ5 (H − βm)K = (Σ·r̂)γ5 (H − βm).
Zαm Zαm
(17.51)
Notes on Chapter 17
Kramer’s equation was used by Martin and Glauber [MG58] to obtain solutions
of the Dirac equation, but also a similar idea was behind many of the methods
discussed in the Notes on Chapter 15.
August 27, 2011 9:22 9in x 6in The Supersymmetric Dirac Equation b1217-ch17
Chapter 18
Non-Relativistic Approximation
In this chapter we look at the non-relativistic limit of the theory developed in the
previous chapters. For the energy levels we have of course the Bohr formula, as
well as higher order effects. We are interested here in the reduction of the wave
functions. It is of course well known that in the non-relativistic limit the Dirac wave
functions reduce to the Pauli wave functions. Here we gain further insight into the
separation of the Dirac radial functions into the two terms w1 and w2 . It turns out
that for negative values of κ the wave function uA reduces to the two-component
Pauli wave function uN l , and for positive values to the Pauli wave function uN,l+1 .
Since the Dirac wave function is degenerate with respect to positive and negative
values of κ, due to the supersymmetry, it is clear that both components uN l and
uN,l+1 will be present in the relativistic case. In the ∆-scheme, with its separation
of the Dirac of the wave function into u1 and u2 , we find the same result after the
application of recursion relations of the hypergeometric functions.
The results of the computations in this chapter are not surprising. The explicit
demonstration is nevertheless instructive, and verifies many of the calculations
performed in the book. In particular we verify the choice of the overall phase,
Sign(κ), in Eq. (15.55), which ensures the equivalence of the expressions for the
solutions of the radial Dirac equation, which were derived independently. Hence,
our choice of phase is in coincidence with [DJ95].
Relativistic effects become important with increasing velocity. Let v be a
typical velocity. For hydrogenic atoms the Bohr formula for the lowest state,
N = 1, is
m
EBohr = EN − m = − (Zα)2
2
= kinetic energy + potential energy = −kinetic energy, (18.1)
in accordance with the virial theorem [HGS01]: potential energy =−2 × kinetic
energy. So the velocity v can be replaced by the more convenient parameter
ζ = Zα.
161
August 27, 2011 9:22 9in x 6in The Supersymmetric Dirac Equation b1217-ch18
ζ2
− 1 ≈ Bohr = − . (18.4)
2N 2
The terms of order O(ζ 4 ) may be calculated in first-order perturbation theory from
the Pauli formalism, for the relativistic mass-correction, the Thomas precession
[ Jac98], and the so-called Darwin term [ Dar28], which is a result of the
zitterbewegung [Ros61].
The apparent principal quantum number na has the non-relativistic
approximation
ζ
na = √ ≈ N. (18.5)
1 − 2
We then have
mζ
x≈2 r. (18.6)
N
From Eq. (15.47) the radial quantum number is nr = na − γ|κ| ≈ N − |κ|.
August 27, 2011 9:22 9in x 6in The Supersymmetric Dirac Equation b1217-ch18
ena x
+ (κ + γ|κ|)(κ − γ|κ|)1 F1 (−nr + 1, 2γ|κ|+2; x) ,
2γ|κ|(2γ|κ| + 1)
γ|κ| −x/2
uB (x) = N x e (κ + γ|κ|)(κ − γ|κ|)1 F1 (−nr , 2γ|κ|; x)
ena x
+ (κ − γ|κ|)(κ + γ|κ|)1 F1 (−nr + 1, 2γ|κ|+2; x) .
2γ|κ|(2γ|κ| + 1)
(18.9)
since
1 (N + ¯l − 1)!(N + ¯l)(N − ¯l)
(2l)!(2¯l + 1)
¯ (N − ¯l)!(2N )
1 (N + ¯l)!
= ¯ = AN l̄ , (18.13)
(2l + 1)! (N − ¯l − 1)!(2N )
from Eq. (9.29). The small component is uB (x) = O(ζ).
August 27, 2011 9:22 9in x 6in The Supersymmetric Dirac Equation b1217-ch18
In both cases the relativistic state reduces in the non-relativistic limit to the Pauli
state.
N√
uA (x) = 1 + xγ|κ| e−x/2
2
× [nr 1 F1 (−nr + 1, 2γ|κ| + 1; x) − (na − κ)1 F1 (−nr , 2γ|κ| + 1; x)],
N√
uB (x) = 1 − xγ|κ| e−x/2
2
× [nr 1 F1 (−nr + 1, 2γ|κ| + 1; x) + (na − κ)1 F1 (−nr , 2γ|κ| + 1; x)].
(18.18)
August 27, 2011 9:22 9in x 6in The Supersymmetric Dirac Equation b1217-ch18
N
uA (x) ≈ √ x|κ| e−x/2 nr [1 F1 (−nr + 1, 2|κ| + 1; x) − 1 F1 (−nr , 2|κ| + 1; x)],
2
uB (x) ≈ 0. (18.19)
and
nr 1 (N + ¯l)! 1 (N + ¯l)!
= ¯ = AN l̄ . (18.23)
2¯l + 1 (2¯l)! (2N ) nr nr ! (2l + 1)! (2N ) n̄r !
Hence
N
uA (x) ≈ √ x|κ| e−x/2
2
× [nr 1 F1 (−nr + 1, 2|κ| + 1; x) − (nr + 2|κ|)1 F1 (−nr , 2|κ| + 1; x)].
(18.25)
yields
N
uA (x) ≈ − √ x|κ| e−x/2 2|κ|1 F1 (−nr , 2|κ|; x). (18.27)
2
Now use |κ| = l + 1.
N
uA (x) ≈ − √ xl+1 e−x/2 (2l + 2)1 F1 (−N + l + 1, 2l + 2; x)
2
= −AN l xl+1 e−x/2 1 F1 (−N + l + 1, 2l + 2; x) = −uN l (x), (18.28)
since
N 1 (N + l + 1)!
(2l + 2) √ ≈
2 (2l + 1)! 2N (N + l + 1) (N − l − 1)!
1 (N + l)!
= = AN l , (18.29)
2l + 1 2N (N − l + 1)!
from
1 (N + l + 1)!
N ≈ , (18.30)
(2l + 2)! N (N + l + 1)(N − l − 1)!
see Eq. (E.12).
We have regained the wave functions of Eqs. (18.16) and (18.17).
August 27, 2011 9:22 9in x 6in The Supersymmetric Dirac Equation b1217-ch19
Chapter 19
Conclusions
“There’s a reason physicists are so successful at what they do, and that is they study the
hydrogen atom and the helium ion and then they stop.”
Attributed to Richard Feynman by J. Rigden, in Hydrogen: The Essential Element, Harvard
University Press, Cambridge, MA, 2002.
One of the central concepts in this book is the Laplace vector, which gives
us the connection between an attractive central force, which falls off inversely
proportional to the distance squared, and the elliptical orbits that a particle moving
under the influence of this force follows. It governs the motion of a planet about the
Sun, and of an electron about the nucleus. It is valid in the classical domain, and, in a
suitably modified form, also in the quantum domain. With a further generalization, a
remnant of the Laplace vector even appears as a conserved quantity in the relativistic
version of this problem at the quantum level.
Neglecting for a moment the spin of the electron, we see that the three
conserved components of the Laplace vector A, together with the three conserved
components of the angular momentum vector L, suitably normalised, form a
representation of the closed Lie algebra so(4). This gives a method for computing
the spectrum of a non-relativistic hydrogen-like atom, including the degeneracy of
the spectral lines.
In the classical problem the magnitude of the Laplace vector gives the
eccentricity of the orbits. Sommerfeld constructed a semi-classical model of the
hydrogen atom, which gives the correct positions of the spectral lines, and a
characterization of the orbits according to their eccentricities. The model works for
the non-relativistic hydrogen atom, as well as for the relativistic hydrogen atom.
Even though it neglects the spin of the electron it gives the fine structure of the
spectrum as it results from Dirac’s relativistic theory, and the correct expression
for the eccentricity of the orbits.
When we want to take the spin of the electron into account we are led naturally
to the Pauli formalism, where the wave function of the electron has two components.
167
August 27, 2011 9:22 9in x 6in The Supersymmetric Dirac Equation b1217-ch19
Conclusions 169
2s 29
2p 2p 2p
3/2 3/2
2p
1/2
2s 2s 2p 2s
1/2 1/2 1/2
Appendices
(a)0 = 1,
(a)n = a(a + 1)(a + 2) · · · (a + n − 1), n = 1, 2, 3, . . . . (A.3)
n! = (n − m)!(n − m + 1)m ,
(c − m + 1)m = (−1)m (−c)m , (A.4)
(n + m)! = n!(n + 1)m .
171
August 27, 2011 9:22 9in x 6in The Supersymmetric Dirac Equation b1217-App
is analytic for any real s with z in the interval −1 < z < 1. Its derivatives are
where we have used the identity (−s)k = (−1)k (s − k + 1)k . With s = −1, this
reduces to the geometric series
∞
1
= zk , (A.10)
1−z
k=0
∞
(a + b − n + 1)n n
(1 + x)a+b = x . (A.12)
n=0
n!
August 27, 2011 9:22 9in x 6in The Supersymmetric Dirac Equation b1217-App
Appendices 173
where
n
Un = Sm Tn−m . (A.15)
m=0
∞ n
(a − m + 1)m (b − n + m + 1)n−m n
(1 + x)a+b = x . (A.16)
n=0 m=0
m! (n − m)!
Since this holds for arbitrary x in the interval −1 < x < 1, the terms in different
powers of x are linearly independent. Therefore, the coefficient of each power of
x must vanish separately, that is,
n
(a − m + 1)m (b − n + m + 1)n−m (a + b − n + 1)n
= . (A.18)
m=0
m! (n − m)! n!
∞
(−a)m (−b)n−m (−a − b)n
= . (A.19)
m=0
m! (n − m)! n!
Laguerre polynomials
Consider the function 1 F1 (−q, 1; x). To have a polynomial q must be a positive
integer. Then, in the series
(−q)k
1 F1 (−q, 1; x) = xk (A.20)
k!k!
k=0
k
(−q)k k (q − k + 1)k k (q + 1)m (−k)k−m
= (−1) = (−1) , (A.21)
k! k! m=0
m! (k − m)!
and
to obtain
1 F1 (−q, 1; x)
∞
1
k
1 (−1)m (m + 1) · · · (m + q) m
= xk−m x
q! m=0
(k − m)! m!
k=0
∞
1
k
xk−m (−1)m dq m+q
= x . (A.25)
q! m=0
(k − m)! m! dxq
k=0
Appendices 175
If this differentiation is carried out in Eq. (A.27), the generating function for the
associated Laguerre polynomials is found to be
Lpq (x) ∞
tp
(−1)p e−xt/(1−t) = tq . (A.29)
(1 − t)p+1
q=0
q!
Notes on Appendix A
A good source of information on hypergeometric functions is the book by Seaborn
[Sea80]. The recurrence relations used in the text are given there in terms of the
associated Laguerre polynomials.
This leads to the appropriate orthogonality relation for the Laguerre polynomials.
By substituting from Eq. (B.1) into the integral in Eq. (B.2) and changing the
August 27, 2011 9:22 9in x 6in The Supersymmetric Dirac Equation b1217-App
The form of the integrand in Eq. (B.4) suggests using the generating function
in Eq. (A.29) to write
p+1 −x p up −xu/(1−u) p vp −xv/(1−v)
x e (−1) e (−1) e
(1 − u)p+1 (1 − v)p+1
∞
∞
uj v k
= xp+1 e−x Lpj (x)Lpk (x). (B.5)
j=0 k=0
j!k!
to show that
∞
xp+1 e−x[1+u/(1−u)+v/(1−v)] dx
0
−p−2
u v
= Γ(p + 2) 1 + +
(1 − u) (1 − v)
p+2
(1 − u)(1 − v)
= Γ(p + 2) . (B.8)
(1 − uv)
Appendices 177
to obtain
∞
(p + 2)k
(1 − u)(1 − v)(p + 1)! (uv)k+p
k!
k=0
∞
(p + k + 1)!
= (1 − u)(1 − v) (uv)k+p
k!
k=0
∞
= Lpj (x)Lpk (x)dx. (B.10)
0
Since u and v are arbitrary and independent, the coefficients of uj v k for any given
j and k must be the same on both sides of this equation. In particular if j = k = q,
∞
(q + 1)! q! 1
+ = xp+1 e−x [Lpq (x)]2 dx. (B.11)
(q − p)! (q − p − 1)! (q!)2 0
From this result the value of the integral is seen to be
∞
[Γ(q + 1)]3
xp+1 e−x [Lpq (x)]2 dx = (2q + 1 − p). (B.12)
0 (q − p)!
Another integral we will need for normalizing the Dirac radial
functions is
∞
xp e−x [Lpq (x)]2 dx. (B.13)
0
Γ(p + 1)(uv)p uj v k ∞ p −x p
= x e Lj (x)Lpk (x)dx. (B.15)
(1 − uv)p+1 j!k! 0
j=0 k=0
If the denominator on the left-hand side of Eq. (B.15) is expanded as in Eq. (A.9),
then Eq. (B.15) becomes
∞ ∞ ∞
(p + 1)s uj v k ∞ p −x p
Γ(p + 1) (uv)s+p = x e Lj (x)Lpk (x)dx.
s=0
s! j=0
j!k! 0
k=0
(B.17)
Now, Lpj (x)
= 0 for j < p, therefore, we can change the lower limits on the sums
on the right-hand side of Eq. (B.17) to p without affecting the sums. Next, redefine
the summation index on the left-hand side to be j = s + p. This gives
∞ ∞ ∞
(p + 1)j−p uj v k ∞ p −x p
Γ(p + 1) j
(uv) = x e Lj (x)Lpk (x)dx.
j=p
(j − p)! j=p
j!k! 0
k=p
(B.18)
Now use the definition of the Kronecker delta to rewrite the left-hand side of this
equation. After some rearrangement of terms the equation then reads
∞ ∞
∞
1
j k
u v xp e−x Lpj (x)Lpk (x)dx
j=p k=p
j!k! 0
Γ(p + 1)(p + 1)k−p
− δjk = 0. (B.19)
(k − p)!
The coefficients of uj v k for different combinations of j and k must vanish
separately. Thus,
∞
(k!)2 p!(p + 1)k−p
xp e−x Lpj (x)Lpk (x)dx = δjk . (B.20)
0 (k − p)!
With k − p equal to an integer Γ(p + 1)(p + 1)k−p = Γ(k + 1), from Eq. (A.4).
Therefore, for the more general case where q − p is an integer, but q and p are not
necessarily integers,
∞
[Γ(q + 1)]3
xp e−x [Lpq (x)]2 dx = . (B.21)
0 (q − p)!
Expressed in terms of 1 F1 (−n; 2a; x) the orthogonality relation for the Lpn (x)
becomes
∞
[Γ(2a)]2
= 2(a + n) n!, (B.22)
Γ(2a + n)
where we have used Eq. (9.38).
August 27, 2011 9:22 9in x 6in The Supersymmetric Dirac Equation b1217-App
Appendices 179
Notes on Appendix B
Be careful of different conventions concerning the associated Laguerre functions,
for example Mertzbacher [Mer61] vs Landau and Lipshitz [LL77].
These integrals are special cases of the integral (F.16) discussed by Landau
and Lipshitz, namely
∞
Jµsp (α, α ) = z µ−1+s e−(k+k )z/2 1 F1 (α, µ; kz)1 F1 (α ; µ − p; k z)dz.
0
(C.2)
August 27, 2011 9:22 9in x 6in The Supersymmetric Dirac Equation b1217-App
∞
(a)n (b)n n
2 F1 (a, b, c; z) = z . (C.3)
n=0
n!(c)n
∞
(a1 )n (a2 )n · · · (ap )n n
p Fq (a1 , a2 , . . . , ap , b1 , b2 , . . . , bq ; z) = z , (C.4)
n=0
n!(b 1 )n (b2 )n · · · (bq )n
The indices α and α are negative integers, so that Eq. (C.6) is a polynomial which,
in the limit → 0, is determined by the last term only. This is summarized in the
relation
0 m > 2n2 ,
n1 ! Γ(µ)
m 2 F1 (α, α , µ; −−2 ) −−−→ (−1)n2 m = 2n2
→0
(n1 − n2 )! Γ(µ + n2 )
undefined m < 2n2
(C.7)
where n1 = max(−α, −α ), and n2 = min(−α, −α ).
The integrals in (C.2) are
10
I11 = J2a+2 (1 − n, 1 − n),
10
I22 = J2a (−n, −n), (C.8)
02
I12 = J2a+2 (1 − n, n).
August 27, 2011 9:22 9in x 6in The Supersymmetric Dirac Equation b1217-App
Appendices 181
00
I11 = (2a + 2n)J2a+2 (1 − n, 1 − n)
[Γ(2a + 2)]2
= 2(a + n) (n − 1)!
Γ(2a + n + 1)
[Γ(2a)]2
00
I22 = (2a + 2n)J2a (−n, −n) = 2(a + n) n! (C.10)
Γ(2a + n)
The value of I22 calculated here agrees with Eq. (B.22), calculated by more
elementary methods.
A little more work is required for I12 . The recursion formula, Eq. (F.15),
namely
02
J2a+2 (1 − n, −n) = 2a(2a + 1)[J2a
00
(1 − n, −n))
− 2J2a
00
(−n, −n) + J2a
00
(−1 − n, −n)]. (C.12)
[Γ(2a)]2
I12 = −2(2a + 1) n! (C.13)
Γ(2a + n)
Notes on Appendix C
Appendix F of Landau and Lipshitz [LL77] is an invaluable source of information
on integrals of hypergeometric functions, which often seems to be insufficiently
appreciated. The limit processes necessary for the application of the results of this
appendix to the cases considered follow [GI97].
August 27, 2011 9:22 9in x 6in The Supersymmetric Dirac Equation b1217-App
To determine N1 /N2 insert this into the first of the equations in (15.34), where
−
b is defined in Eq. (15.33):
− d γ|κ| na ena
b w1 = − + − w1 = − w2 . (D.2)
dx x 2γ|κ| 2γ|κ|
First calculate
dw1 γ|κ| 1
= N1 xγ|κ| e−x/2 − 1 F1 (−nr , 2γ|κ|; x)
dx x 2
d1 F1 (−nr , 2γ|κ|; x)
+ . (D.3)
dx
Appendices 183
(nr + 2γ|κ|)
− x1 F1 (−nr + 1, 2γ|κ| + 2; x)
2γ|κ|(2γ|κ| + 1)
This yields
nr (nr + 2γ|κ|)
N1 x1 F1 (−nr + 1, 2γ|κ| + 2; x)
2γ|κ|
2γ|κ|(2γ|κ| + 1)
ena
= −N2 x1 F1 (−nr + 1, 2γ|κ| + 2; x). (D.8)
2γ|κ|
The result is
N1 ena 2γ|κ|(2γ|κ| + 1) 1
=− =− 2γ|κ|(2γ|κ| + 1), (D.9)
N2 nr (nr + 2γ|κ|) ena
2
uA (x) + u2B (x) dx = 1. (D.10)
N 2
2κ(κ − γ|κ|) dx x2γ|κ| e−x [1 F1 (−nr , 2γ|κ|; x)]2
4ena
+ (κ2 − γ 2 |κ|2 )(κ2 2 − γ 2 |κ|2 )
2γ|κ|(2γ|κ| + 1)
(ena )2
+ 2κ(κ + γ|κ|)
(2γ|κ|)2 (2γ|κ| + 1)2
2γ|κ|+2 −x
× dx x 2
e [1 F1 (−nr + 1, 2γ|κ| + 2 : x)] = 1. (D.11)
August 27, 2011 9:22 9in x 6in The Supersymmetric Dirac Equation b1217-App
[ Γ(2γ|κ|) ]2
x2γ|κ| e−x [1 F1 (−nr , 2γ|κ|; x)]2 dx = 2(2γ|κ| + nr ) nr !
Γ(2γ|κ| + nr )
[Γ(2γ|κ| + 2)]2
= 2(2γ|κ| + nr ) (nr − 1)!
Γ(2γ|κ| + nr + 1)
[Γ(2γ|κ|)]2
= −2(2γ|κ|)(2γ|κ| + 1) (nr )! (D.12)
Γ(2γ|κ| + nr )
so that
1 [ Γ(2γ|κ|) ]2
=4 nr !
N 2 Γ(2γ|κ| + nr )
κ(κ − γ|κ|)na − 2(ena )Zα|κ| 2 − γ 2
κ(κ + γ|κ|)(ena )2 na
+ . (D.13)
nr (nr + 2γ|κ|)
Now use
|κ|2 (2 − γ 2 )
nr (nr + 2γ|κ|) = (ena )2 = (D.14)
1 − 2
to reduce this to
1 8na [ Γ(2γ|κ|) ]2 2 2
= γ |κ| nr !, (D.15)
N 2 Γ(2γ|κ| + nr )
which is equivalent to
1 Γ(2γ|κ| + nr )
N = . (D.16)
Γ(2γ|κ| + 1) (2na )nr !
August 27, 2011 9:22 9in x 6in The Supersymmetric Dirac Equation b1217-App
Appendices 185
where nr = na − γ|κ| is, for u2 (x), a positive number or zero, and, for u1 a
positive number (excluding zero).
To determine N1 /N2 insert these solutions into the second of the equations of
(16.23), where a+ is defined in Eq. (16.29):
d x
+
a u2 = x + − na u2 = −ena u1 . (E.2)
dx 2
We calculate
du2
x = N2 xγ|κ| e−x/2
dx
x d1 F1 (−nr , 2γ|κ| + 1; x)
γ|κ| − F
1 1 (−n r , 2γ|κ| + 1; x) + x .
2 dx
(E.3)
d1 F1 (−nr , 2γ|κ| + 1; x)
x
dx
= −nr [1 F1 (−nr + 1, 2γ|κ| + 1; x) − 1 F1 (−nr , 2γ|κ| + 1; x)]. (E.4)
The (x/2) terms cancel, and the other 1 F1 (−nr , 2γ|κ| + 1; x) terms have a
coefficient γ|κ| + nr − na , which vanishes because of (15.47). We are left with
August 27, 2011 9:22 9in x 6in The Supersymmetric Dirac Equation b1217-App
N2 nr = N1 (ena ), or
N1 nr nr nr
= = = . (E.6)
N2 ena na − κ
2 2 (na + κ)(na − κ)
2
uA (x) + u2B (x) dx = 1. (E.7)
N 2
(1 + ) n2r x2γ|κ| e−x [1 F1 (−nr + 1, 2γ|κ| + 1; x)]2 dx
4
2γ|κ| −x
+(na − κ) 2
x 2
e [1 F1 (−nr , 2γ|κ| + 1; x)] dx
N 2
+ (1 − ) n2r x2γ|κ| e−x [1 F1 (−nr + 1, 2γ|κ| + 1; x)]2 dx
4
2γ|κ| −x
+(na − κ) 2
x 2
e [1 F1 (−nr , 2γ|κ| + 1; x)] dx . (E.8)
[ Γ(2γ|κ| + 1) ]2
x2γ|κ| e−x [1 F1 (−nr , 2γ|κ| + 1; x)]2 dx = (nr − 1)!
Γ(nr + 2γ|κ|)
[ Γ(2γ|κ| + 1) ]2
x2γ|κ| e−x [1 F1 (−nr + 1, 2γ|κ| + 1; x)]2 dx = nr !
Γ(nr + 2γ|κ| + 1)
(E.9)
so
N 2 2 [ Γ(2γ|κ| + 1) ]2 [ Γ(2γ|κ| + 1) ]2
nr (nr − 1)! + (na − κ)2 nr !
2 Γ(nr + 2γ|κ|) Γ(nr + 2γ|κ| + 1)
N 2 [ Γ(2γ|κ| + 1) ]2 nr !
= nr (nr + 2γ|κ|) + (na − κ)2 = 1. (E.10)
2Γ(nr + 2γ|κ| + 1)
Now use
Appendices 187
This yields
N 2 [ Γ(2γ|κ| + 1) ]2 nr !
1= 2na (na − κ),
2 Γ(nr + 2γ|κ| + 1)
1 Γ(nr + 2γ|κ| + 1)
N 2 = , (E.12)
[ Γ(2γ|κ| + 1) ]2 na (na − κ)nr !
1/2
1 Γ(nr + 2γ|κ| + 1)
N = .
Γ(2γ|κ| + 1) na (na − κ)nr !
August 27, 2011 9:22 9in x 6in The Supersymmetric Dirac Equation b1217-App
If k = n
(a) (p × L)·(p × L) = p2 L2 ,
(b) (p × L)·p = 2ip2 ,
(c) p·(p × L) = 0,
(d) (p × L)·r = L2 + 2i(p·r),
(e) r·(p × L) = L2 .
189
August 27, 2011 9:22 9in x 6in The Supersymmetric Dirac Equation b1217-solution
Solution:
Solution:
Solution:
∂ 1∂ 2
[r, pr ] = rpr − pr r = (−i) r− r = −i(1 − 2) = i.
∂r r∂
Exercise 9.2 Prove that the quantized version of pr follows from the Weyl–Moyal
quantization prescription applied to p and r̂.
Solution:
1 r r −i ∂ ∇·r r·r ∂ 1
·p + p· = 2 + − 3 = −i + = pr .
2 r r 2 ∂r r r ∂r r
We have
1 ∂
pr ψ = (−i) (rψ),
r ∂r
and
1 ∂
(pr ψ)∗ = (i) (rψ)∗ .
r ∂r
Therefore
[ψ ∗ (pr ψ) − (pr ψ)∗ ψ]d3 r
∗ ∂ ∂ ∗
= −i d cos θ dφ (rψ) (rψ) + (rψ) (rψ)
∂r ∂r
∂ ∂
= −i d cos θ dφ [(rψ)(rψ)∗ ] dr = −i d cos θ dφ |rψ|2 dr.
∂r ∂r
For limr→∞ (rψ) = 0 and limr→0 (rψ) = 0 the condition for Hermiticity is
satisfied.
Exercise 14.1 Prove [Σ·r̂, H] = − 2i
r βKγ5 .
Solution:
r ri
i
[Σ·r̂, H] = [Σ·r̂, α·p] = Σi , αj pj + [Σi , αj pj ] .
r r
August 27, 2011 9:22 9in x 6in The Supersymmetric Dirac Equation b1217-solution
2i 2i 2i 2i
[Σ·r̂, H] = γ5 + γ5 (Σ·L) = γ5 (Σ·L + 1) = − γ5 βK.
r r r r
Solution:
Solution:
[Kγ5 f (r), H]
= K[γ5 f (r), H] + [K, H]γ5 f (r) = Kγ5 [f (r), H] + K[γ5 , H]f (r)
= Kγ5 [f (r), α·p] + 2mKγ5 βf (r) = iKγ5 α· r̂f (r) + 2mKγ5 βf (r)
= iK(Σ· r̂)f (r) + 2mKγ5 βf (r).
Solution:
K 2 (H − βm)2 = K 2 [H 2 + m2 − m(Hβ + βH)].
We have
Zα
Hβ = m + γ5 βΣ·p − β ,
r
Zα
βH = m + βγ5 Σ·p − β ,
r
Zα
Hβ + βH = 2m − 2β .
r
Adding these together yields the desired relationship.
Exercise 14.5 Prove
−2iβK 2
(Σ·r̂)Kγ5 (H − βm) − (H − βm)γ5 K(Σ·r̂) = .
r
Solution:
(Σ·r̂)Kγ5 (H − βm) − (H − βm)γ5 K(Σ·r̂)
= Kγ5 Σ·r̂(H − βm) − γ5 K(H − βm)Σ·r̂
K −2iβK 2
= Kγ5 [Σ·r̂, H] = Kγ5 (−2iγ5 )β = .
r r
August 27, 2011 9:22 9in x 6in The Supersymmetric Dirac Equation b1217-solution
Bibliography
Books
[Ada94] B. G. Adams. Algebraic Approach to Simple Quantum Systems. Berlin: Springer-
Verlag, 1994.
[BD64] J. D. Bjorken and S. D. Drell. Relativistic Quantum Mechanics. New York:
McGraw-Hill, 1964.
[BL81] L. C. Biedenham and J. D. Louck. Angular Momentum in Quantum Physics:
Theory and Applications. Reading, MA: Encyclopedia of Mathematics and its
Applications 8, Addison-Wesley, 1981.
[Bre91] D. Bressoud. Second Year Calculus. New York: Springer, 1991.
[Bro24] L. de Broglie. Licht und Materie. Hamburg: H. Goverts Verlag, 1930.
[BS57] H. A. Bethe and E. E. Salpeter. Quantum Mechanics of One- and Two-Electron
Atoms. Berlin: Springer-Verlag, 1957.
[Dir53] Paul A. M. Dirac. The Principles of Quantum Mechanics. 3rd ed. Oxford:
Clarendon Press, 1953.
[DL03] C. Doran and A. Lasenby. Geometric Algebra for Physicists. Cambridge:
Cambridge University Press, 2003.
[ER74] R. Eisberg and R. Resnick. Quantum Physics of Atoms, Molecules, Solids, Nuclei,
and Particles. New York: J. Wiley, 1974.
[GR65] I. S. Gradshteyn and I. M. Ryzhik. Tables of Integrals, Series, and Products. New
York: Academic Press, 1965.
[GS90] V. Guillemin and S. Sternberg. Variations on a Theme by Kepler. Providence, R.I.:
The American Mathematical Society, 1990.
[Jac98] J. D. Jackson. Classical Electrodynamics. 3rd ed. New York: J. Wiley, 1998.
[Kel83] A. Keller. The Infancy of Atomic Physics: Hercules in his Cradle. Oxford:
Clarendon Press, 1983.
[LL77] L. D. Landau and E. M. Lipshitz. Quantum Mechanics. 3rd ed. Oxford: Pergamon
Press, 1977.
[Mer61] E. Mertzbacher. Quantum Mechanics. New York: J. Wiley, 1961.
[MR99] J. E. Marsden and T. S. Ratiu. Introduction to Mechanics and Symmetry. 2nd ed.
New York: Springer-Verlag, 1999.
[PGS01] C. Poole, H. Goldstein and J. Safko. Classical Mechanics. New York: J. Wiley,
1998.
[Rig03] J. S. Rigden. Hydrogen: The Essential Element. Cambridge, MA: Harvard
University Press, 2003.
[Ros61] E. M. Rose. Relativistic Electron Theory. New York: J. Wiley, 1961.
[Sak67] J. J. Sakurai. Advanced Quantum Mechanics. Reading, MA: Addison-Wesley,
1967.
195
August 27, 2011 9:22 9in x 6in The Supersymmetric Dirac Equation b1217-Bib
Journals
[AB78] P. R. Auvil and L. M. Brown. “The Relativistic Hydrogen Atom: A Simple
Solution”. In: Am. J. Phys. 46 (1978), p. 679.
[BB80] L. Basano and A. Bianchi. “Rutherford’s scattering formula via the Runge–Lenz
vector”. In: Am. J. Phys. 48 (1980), p. 400.
[BHJ26] M. Born, W. Heisenberg and P. Jordan. “Zur Quantenmechanik II”. In: Z. Phys.
35 (1926), p. 557.
[Bie62] L. C. Biedenharn. “Remarks on the Relativistic Kepler Problem”. In: Phys. Rev.
126 (1962), p. 845.
[Bie83] L. C. Biedenharn. “The ‘Sommerfeld Puzzle’ Revisited and Resolved”. In:
Foundations of Physics. 13 (1983), p. 13.
[BJ25] M. Born and P. Jordan. “Zur Quantenmechanik”. In: Z. Phys. 34 (1925), p. 858.
[BJ53] W. L. Bade and H. Jehle. “An Introduction to Spinors”. In: Rev. Mod. Phys. 25
(1953), p. 714.
[Boh13] N. Bohr. “On the Quantum Theory of Line Spectra”. In: Phil. Mag. 26 (1913),
p. 476.
[Dah95] J. P. Dahl. “On the Origin of the Runge–Lenz Vector”. In: Int. J. of Quant. Chem.
53 (1995), p. 161.
[Dar28] C. G. Darwin. “The Wave Equations of the Electron”. In: Proc. R. Soc. London.
Ser. A118 (1928), p. 654.
[Dir26] P. A. M. Dirac. “Quantum Mechanics and a Preliminary Investigation of the
Hydrogen Atom”. In: Proc. Roy. Soc. A110 (1926), p. 561.
[Dir28] P. A. M. Dirac. “The Quantum Theory of the Electron”. In: Proc. Roy. Soc. A118
(1928), p. 610.
[DJ95] J. P. Dahl and T. Jørgensen. “On the Dirac–Kepler Problem: The
Johnson–Lippmann Operator, Supersymmetry, and Normal-Mode Repre-
sentations”. In: Int. J. of Quant. Chem. 53 (1995), p. 161.
[FK05] M. R. Francis and A. Kosowsky. “The Construction of Spinors in Geometric
Algebra”. In: Ann. Phys. 317 (2005), p. 383.
August 27, 2011 9:22 9in x 6in The Supersymmetric Dirac Equation b1217-Bib
Bibliography 197
Index
199
August 27, 2011 9:22 9in x 6in The Supersymmetric Dirac Equation b1217-Index
Kramer, 4, 127, 137, 149, 154, 155, 157, quantization condition, 35, 36, 38, 47, 55,
159, 168 99, 102
Kramer’s equation, 4, 127, 137, 149, 154, quantum electrodynamics, 1, 80
155, 157, 159, 168 quantum Laplace vector, 60
quantum rules, 2, 47
Laguerre polynomials, 73, 74, 171, 174, quantum Runge–Lenz vector, 13
175
Lamb-shift, 4, 169 radial Dirac functions, 129, 130, 142
Laplace, 2, 3, 6, 8, 9, 11, 13, 15, 29, radial functions, 74, 111, 123, 129, 135,
31–34, 59, 60, 65, 66, 82, 88, 115, 116, 136, 140, 143, 146, 151, 161
119, 126, 167, 168 Rutherford, 2, 31, 34, 36
Laplace vector, 3, 6, 59, 115 Rydberg constant, 35, 43
Laplace–Lenz vector, 65
Lévy-Leblond, 77 Schrödinger, 2, 3, 69, 75, 87, 168
Lie Algebra, 3, 15, 18–26, 28, 29, Sommerfeld, 2, 3, 47–49, 52, 53, 55, 57,
65, 167 58, 99, 101, 103, 104, 167, 169
Sommerfeld formula, 99, 103
matrix formalism, 59 Sommerfeld puzzle, 99
minimal substitution, 79 spherical harmonics, 71, 72, 83, 111
spherical polar coordinates, 70, 86, 105,
Nieto, 126 111, 156
normalization, 4, 72–74, 132, 136, 137, spin, 2–4, 58, 67, 68, 77, 78, 80, 88, 90,
147, 169, 175, 179, 182, 183, 185, 186 94, 97, 99, 105, 107, 111, 129, 140,
nucleus, 6, 31, 32, 34, 35, 37, 39, 40, 157, 159, 167, 168
42–44, 51, 52, 69, 167 spinors, 77
stationary orbit, 52
Old Quantum Theory, 2, 99, 104 Stern–Gerlach experiment, 77
SU(2) supersymmetry, 122
parity, 91, 92, 96, 109–111, 119, 120, 122, superalgebra, 92, 93, 97
128, 139, 141 superalgebra s(2), 92
parity operator, 92, 96, 119, 120, 122, 128, supercharges, 92, 95, 96
139, 141 supersymmetric charge, 93, 121
Pauli, 3, 4, 13, 59, 62, 65, 68, 77–80, 82, supersymmetric eigenvalue problem, 146
83, 87, 94, 95, 105, 106, 108, 111, 114, supersymmetric generators, 122
161, 164, 167–169 supersymmetric Hamiltonian, 92, 93, 122,
Pauli formalism, 3, 77, 167 125, 133, 144
Pauli hydrogenic realization, 62, 65 supersymmetric quantum, 93, 94, 154
Pauli matrices, 79, 80, 106, 108, 168 supersymmetric quantum mechanics, 154
Pauli Spinors, 82, 105, 111 supersymmetric structure, 128, 133, 137,
Planck, 4, 35–37, 47–50 139, 144
Planck’s constant, 4, 36, 37, 50 supersymmetry, 1, 3, 4, 77, 84, 91–95, 97,
Pochhammer symbols, 74, 171, 172 98, 115, 116, 121, 122, 125–128, 137,
probability interpretation, 69 139–141, 147, 161, 168
August 27, 2011 9:22 9in x 6in The Supersymmetric Dirac Equation b1217-Index
Index 201