Symmctry: Dynamical
Symmctry: Dynamical
symmcTry
Carl E Wulfman
World Scientific
This page is intentionally left blank
Carl E Wulfman
University of the Pacific, USA
World Scientific
NEW JERSEY • LONDON • SINGAPORE • BEIJING • SHANGHAI • HONG KONG • TA I P E I • CHENNAI
Published 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
DYNAMICAL SYMMETRY
Copyright © 2011 by World Scientific Publishing Co. Pte. Ltd.
All rights reserved. This book, or parts thereof, may not be reproduced in any form or by any means,
electronic or mechanical, including photocopying, recording or any information storage and retrieval
system now known or to be invented, without written permission from the Publisher.
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-13 978-981-4291-36-1
ISBN-10 981-4291-36-6
Printed in Singapore.
Dedication
v
This page is intentionally left blank
October 28, 2010 16:55 9in x 6in b943-fm
Preface
vii
October 28, 2010 16:55 9in x 6in b943-fm
References
[1] G. W. Bluman, S. Kumei, Symmetries and Differential Equations (Springer-
Verlag, N.Y., 1989).
[2] J. E. Campbell, Continuous Groups; reprint of the 1903 edition of Introduc-
tory Treatise on Lie’s Theory of Finite Continuous Transformation Groups
(Chelsea, Bronx, N.Y, 1966).
October 28, 2010 16:55 9in x 6in b943-fm
Preface ix
Acknowledgments
xi
This page is intentionally left blank
October 28, 2010 16:55 9in x 6in b943-fm
Contents
Dedication v
Preface vii
Acknowledgements xi
1. Introduction 1
1.1 On Geometric Symmetry and Invariance in the Sciences 1
1.2 Fock’s Discovery . . . . . . . . . . . . . . . . . . . . . . 6
1.3 Keplerian Symmetry . . . . . . . . . . . . . . . . . . . . 8
1.4 Dynamical Symmetry . . . . . . . . . . . . . . . . . . . 8
1.5 Dynamical Symmetries Responsible for Degeneracies
and Their Physical Consequences . . . . . . . . . . . . . 10
1.6 Dynamical Symmetries When Energies Can Vary . . . . 12
1.7 The Need for Critical Reexamination of Concepts of
Physical Symmetry. Lie’s Discoveries . . . . . . . . . . . 13
Appendix A: Historical Note . . . . . . . . . . . . . . . . . . . 18
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 19
xiii
October 28, 2010 16:55 9in x 6in b943-fm
Contents xv
Contents xvii
Contents xix
xx Dynamical Symmetry
Index 433
October 28, 2010 16:55 9in x 6in b943-ch01
CHAPTER 1
Introduction
1
October 28, 2010 16:55 9in x 6in b943-ch01
2 Dynamical Symmetry
Introduction 3
4 Dynamical Symmetry
Fig. 1.1 A unit cube viewed by an observer in its rest frame and by an oberserver
moving with two thirds the velocity of light in its rest frame.
Introduction 5
6 Dynamical Symmetry
accident as was the degeneracy of the 3s, 3p and 3d states (and so on). It is
natural to suspect that all these accidents afflicting every hydrogen atom
in the universe might have a cause.
Introduction 7
(a) (b)
(c) (d)
Fig. 1.2 (a) Stereographic projection of point on upper hemisphere. (b) Stereo-
graphic projection of point on lower hemisphere. (c) Stereographic projection of
circles of constant latitude onto an equatorial plane. (d) Stereographic projection
obtained when sphere in Fig. 1.2(c) is rotated 90◦ about the y axis.
interconvert the p and s type functions. However, they are indirectly inter-
converted by rotations in the three-space that carry a Px or Py function into
a Pz function. Such a rotation carries a p type function in the plane into an
sp type hybrid function, and then into a pure s type function. This illus-
trates a key feature underlying Fock’s analysis: stereographic projections
can interconvert functions not related by rotations in three-dimensional
Euclidean space, to functions related by rotations in a higher dimensional
Euclidean space.
Fock’s use of a stereographic projection to explain the degeneracy of
the hydrogen atom is subject to an evident objection: the projection could
be said to produce, rather than uncover, a symmetry. However, Bargmann
quickly gave a Lie-algebraic treatment which established the hyperspherical
symmetry without the use of any coordinate transformations.11 The work of
Fock and Bargmann provides an impressive demonstration that degenera-
cies can be due to symmetries that are definitely not symmetries in ordinary
space. For natural scientists it was an early hint that natural phenomena can
October 28, 2010 16:55 9in x 6in b943-ch01
8 Dynamical Symmetry
Introduction 9
The potential energy term, K/|r|, with K a constant, has ordinary rota-
tional symmetry; and the kinetic energy term, p2 /2m, with the mass, m,
constant, has ordinary rotational symmetry. This statement remains true in
Schrödinger quantum mechanics, where p is converted to −i∇. The hyper-
spherical symmetry is not a symmetry of the potential energy term in the
Hamiltonian nor is it a symmetry of its kinetic energy term, it is a symmetry
of the entire Hamiltonian. Operations of a group of transformations inter-
convert the kinetic and potential energy terms of the Hamiltonian H while
leaving E, the value of H, unchanged. These operations transform the posi-
tion vector r to r = f(r, p), and the momentum vector p to p = g(r, p), so
as to leave only the entire expression p2 /2m−K/|r|, and hence E, invariant.
For a two-dimensional atom, the operation that interconverts the kinetic
energy and potential energy can be put in one-to-one correspondence with
the Px ↔ Pz rotations on the sphere of Fig. 1.2, i.e. the rotations that
interconvert the s type and p type projections in the plane.
The hyperspherical symmetry of Kepler Hamiltonians is a truly dynam-
ical symmetry — a symmetry present when there is motion. It is a symme-
try that exists only when motion is allowed. It is extensively investigated
in Chapter 8.
An analogous situation occurs with the harmonic oscillator Hamiltonian
10 Dynamical Symmetry
Introduction 11
12 Dynamical Symmetry
Introduction 13
hidden symmetry, the united atom–nearly united atom energy level corre-
lation diagrams may appear to violate the noncrossing rule of Wigner and
von Neumann.25 As shown in Chapter 12, this hidden symmetry ensures
that nearly-united atoms obey the same rules that govern the geometrical
symmetry of actual triatomic molecules.26 These rules are due to Walsh,
who first pointed out that the ordinary spatial symmetry of triatomic and
tetraatomic molecules can be anticipated by simply counting the number
of valence electrons in the molecules.27
As the distance between atoms varies during the course of chemical reac-
tions, degeneracies, or near degeneracies, are often produced. Though often
considered accidental, many can be predicted from energy-level correlation
diagrams. Profound and rapid changes may take place at such crossings.
This happens, for example, when a neutral alkali atom approaches a neutral
halogen atom from afar. An electron jumps from the metal to the halogen
during a change of distance on the order of 10−2 angstroms.28
Finally, we would note that even when the entropy and other equilibrium
properties of highly degenerate, weakly interacting, systems are unaffected
by weak interactions, these interactions can govern energy flow between
subsystems. The resulting dynamical symmetry breaking then becomes
relevant in nonequilibrium statistical mechanics.
14 Dynamical Symmetry
This observation, though at first sight trivial, profoundly connects the study
of differential equations to studies of groups of transformations. In fact, Lie
used first-order ordinary differential equations, and sets of first-order partial
differential equations, to define continuous groups. He devoted much of his
life to developing the extensive mathematical implications of the connection
between differential equations and continuous groups. In doing so, he gave
us much of what is now called the theory of Lie groups.
The operations that convert solutions of a differential equation into
themselves and into other solutions, define the invariance group of the dif-
ferential equation. These groups are primarily Lie groups, and they can
be of great utility. In fact, most of the differential equations of the sci-
ences have Lie group structure which neatly organizes and simplifies their
solution. Unfortunately, for much of the past century the utility of the con-
nection between Lie groups and differential equations went unrecognized,
and was seldom exploited by mathematicians interested in solving differ-
ential equations. Several monographs that are outstanding exceptions to
this statement have been listed at the end of the Preface. In this book, the
group structure of the equations of dynamics is used to relate dynamics to
geometry by relating dynamical symmetries to geometrical symmetries.
The relevance of Lie’s work to the natural sciences is direct and fun-
damental. Whenever a continuous succession of causes and effects governs
natural phenomena, it becomes possible to formulate differential equations
October 28, 2010 16:55 9in x 6in b943-ch01
Introduction 15
16 Dynamical Symmetry
Introduction 17
18 Dynamical Symmetry
Introduction 19
References
[1] J. Kepler, Tertius Interveniens, Gesammelte Werke, edited by W. van Dycke,
M. Caspar; translated by A. Koestler, The Watershed (Doubleday, Garden
City, N.Y. 1960), pp. 65, 226.
[2] J. Kepler, Mysterium Cosmographicum (Tuebingen, 1596).
[3] E. Noether, Nachr. Koenig. Gesell. Wissen. Goettigen, Math.-Phys. Kl 235
(1918).
[4] J. C. Maxwell, Phil. Mag., I, 48 (1860); cf. The Scientific Papers of James
Clerk Maxwell, ed. W. D. Niven, Vol. 1, p. 379 (Librairie Scientific J. Her-
mann, Paris, 1890).
[5] a) J. C. Maxwell, Cambridge Phil. Soc. Trans., X, pt.1, (1855–6) Scientific
Papers, 155.
b) J. C. Maxwell, Phil. Trans. Roy. Soc. (London) CLV (1864) Scientific
Papers, 526.
[6] a) A. Einstein, Ann. der Physik 17, 891 (1905); translated by R. W. Lawson,
Relativity, The Special and General Theory, (Henry Holt, New York, 1921).
October 28, 2010 16:55 9in x 6in b943-ch01
20 Dynamical Symmetry
Introduction 21
22 Dynamical Symmetry
CHAPTER 2
23
October 28, 2010 16:55 9in x 6in b943-ch02
24 Dynamical Symmetry
w4 = 1. (2.1.1)
(0,1)
(− 1,0) (1,0)
(0,− 1)
Table 2.1. Operations that carry the square into itself and
interconvert solutions of w4 = 1, w = x + iy = 1, i, −1, −i.
A. Rotations:
R(0, 1) = (1, 0), R(1, 0) = (0, −1),
R(0, −1) = (−1, 0), R(−1, 0) = (0, 1).
B. Reflections:
S1 (x, y) = (x, −y), S2 (x, y) = (−x, y).
C. Inversions:
J1 (x, y) = (y, x), J2 (x, y) = (−y, −x).
and their action on a coordinate vector r = (x, y), are listed in Table 2.1.
Each operation corresponds to a substitution that leaves (2.1.1) invariant.
We wish to determine whether these operations give rise to a group
of operations. A set of operations comprise a group of operations iff the
following hold:
(a) The application of any operation Ta , followed by the application of any
operation Tb , is an operation in the set, say Tc . Symbolically,
Tb Ta = Tc . (2.1.3a)
(b) The set contains the identity operation, I, which satisfies
Ta I = I Ta = Ta . (2.1.3b)
(c) Every operation Ta has an inverse Ta in the set:
Ta Ta = I = Ta Ta . (2.1.3c)
(d) The operations obey the associative law
Ta (Tb Tc ) = (Ta Tb ) Tc . (2.1.3d)
Here, the inverse of a compound operation (Ta Tb ) is Tb Ta .
The operations of Table 2.1 give rise to the operations A’ in the first
column of Table 2.2. Their action on the operators A of the top row of the
table is indicated at the intersection of the corresponding rows and columns.
With the aid of the table the reader may verify that, taken together, the
operations are those of a group. Note that S1 S1 = I, S2 S2 = I, J1 J1 = I
and J2 J2 = I. It follows that each of these operations is its own inverse,
e.g. Sinv
1 = S1 . Consequently the operations S1 and I are themselves those
of a group. As this group is contained in the larger group, it is termed a
October 28, 2010 16:55 9in x 6in b943-ch02
26 Dynamical Symmetry
A A\ A I R R2 R3 S1 S2 J1 J2
I I R R2 R3 S1 S2 J1 J2
R R R2 R3 I J1 J2 S2 S1
R2 R2 R3 I R S2 S1 J2 J1
R3 R3 I R R2 J2 J1 S1 S2
S1 S1 J2 S2 J1 I R2 R R3
S2 S2 J1 S1 J1 I R2 R R3
J1 J1 S1 J2 S2 R3 R I R2
J2 J2 S2 J1 S1 R R3 R2 I
subgroup of the later. Each of the operations of reflection and inversion are
members of a subgroup of the larger group. Because R4 = I, R3 so R = I,
R3 is the inverse of R. Similarly, R2 is its own inverse. These rotations also
comprise a subgroup of the full group.
Because the operations carry solutions of Eq. (2.1.1) into solutions of
the same equation, it is left unchanged, i.e. it is invariant under the opera-
tion of the group. Given our geometric assumptions, the transformations in
Table 2.2 are also those of the symmetry group of the square of Fig. 2.1.
There are geometrical operations that do not leave Eq. (2.1.1) in-
variant, but do leave the symmetry of the square unchanged. These include
arbitrary rotations of the square about any point, translations of the square
in any direction, and the uniform dilatations, which enlarge or shrink the
square.
Rotations about the origin carry a point with coordinates (x, y) to the
point with coordinates (x , y ), where
and θ is the angle of rotation. Operations of the group will carry a circle
with center at the origin into itself. The group is termed a continuous group
because the operations of the group depend upon a parameter, θ, that can
vary continuously. The rotations that carry the square of Fig. 2.1 into itself
comprise a discrete subgroup of this continuous group of rotations.
Translations carry a point with coordinate (x, y) to the points at (x , y ),
where
x = x + α, y = y + β. (2.1.5)
Composing the operations of (2.1.4) and (2.1.5) one obtains E(2), the
Euclidean group of the plane. If we let T(α)x represent the operation that
converts x to x + α, the application of T(α1 ) followed by T(α2 ) is a trans-
formation
T(α2 )T(α1 )x = T(α2 )(x + α1 ) = x + α1 + α2 = T(α3 ), (2.1.6)
with
α3 = α1 + α2 .
The group E(2) is termed a three-parameter group because its operations
depend on the parameter θ, the angle of rotation, the parameter α, the
distance a point is translated parallel to the x axis, and the parameter β,
the distance it is translated parallel to the y axis.
Note that the operations of E(2) do not alter the distances Dij separating
points i and j.
Uniform dilatations in the plane carry out the transformations
x → eµ x, y → eµ y, −∞ < µ < ∞. (2.1.7)
Together with the operations of E(2), they comprise the operations of a
continuous group termed the similitude group of the plane. The operations
of this four-parameter group may be used to superpose objects of the same
symmetry — bring them into coincidence; they are unable to bring objects
of different symmetry into coincidence.
Operations of the similitude group that do not carry the square into
itself, do however, leave its symmetry unchanged. They might be said to
alter the perspective of a viewer of the square, changing the relation of the
square to this viewer, by actively rotating, translating, or dilatating it —
altering its apparent size. Operations of the similitude group of the plane
do not alter the relative distances Dij separating points i and j, and Di j
separating points i and j , because both are multiplied by eβ .
Now let us compare the effects of the discrete operations of inversion
and reflection applied to the vertices in Fig. 2.1.
The inversions through the origin transform r → −r. The same effect
can be obtained by rotating the vector r, with components x, y, through
180◦ . Inversions are operations contained in the similitude group.
However, the reflection x → −x, y → y, and the reflection y → −y,
x → x, may alter the symmetry of a “left-handed” or “right-handed” figure
in the plane, interconverting the one into the other. None of these reflec-
tions is an inversion. Also, reflections are not special cases of rotations.
October 28, 2010 16:55 9in x 6in b943-ch02
28 Dynamical Symmetry
Reflections, though they leave invariant the separation between points, can-
not be carried out by any operation of the continuous similitude group of
the plane because there is no continuous motion of the similitude group
that can move a point Px,y through all intermediate points P between Px,y
and P−x,y , or between Px,y and Px,−y . In short, it is impossible for the
similitude group acting in the plane to continuously deform a right-handed
object in the plane into a left-handed object.
A situation that is analogous to the Euclidean plane holds in three-
dimensional Euclidean space: operations of the corresponding continuous
similitude group may be used to rotate, translate, and enlarge or shrink
objects and thereby superpose geometrical objects of the same symmetry.
Extending the group by including the separate reflections, x → −x, y → −y,
z → −z, one obtains the group whose subgroups define symmetries in the
space. Again, the effect of the inversions r → −r can be obtained by opera-
tions of rotation and reflection. Right-handed and left-handed objects have
different symmetry, and reflections may interconvert them, i.e., alter their
symmetry. However, there is no continuous operation acting in the three-
dimensional space that will deform a right-handed object so as to convert
it to a left-handed object, while at the same time leaving the separation
between points in three-dimensional Euclidean space unchanged, uniformly
enlarged, or uniformly shrunk.
These observations on the symmetry of objects call attention to gen-
eral principles that are true in both two- and three-dimensional Euclidean
geometry:
1. The group defining the symmetry of any figure or lattice is a subgroup
of the group of all similitude and inversion transformations.
2. The symmetry of any figure or lattice is unaltered by any operation of
the continuous group of similitude operations acting on the entire figure
or lattice.
The human mind is able to recognize that two objects in space are
essentially identical if an operation of the similitude group of three-space
can convert the one into the other. Apparently our minds habitually and
rapidly rotate, translate, dilatate, and then compare memories of visual
images without us being conscious of this. When our minds impart such
structures in group theory, we may not notice that we have seen similar
objects from different perspectives.
October 28, 2010 16:55 9in x 6in b943-ch02
1. The real part of w was plotted on one axis, and the coefficient of its
imaginary part on the other. The coordinates of a point, (x, y), were
thus chosen to be real numbers.
2. The x and y axes were, by choice, made perpendicular to each other so as
to provide an orthogonal coordinate system. It is conceivable that there
might be relations between x and y which invalidate this supposition.
Gibbs phase diagrams provide an example of this.
3. Though four points do not necessarily lie in a Euclidean plane, the four
roots were plotted in a plane. One could have plotted them in any two-
dimensional space, such as that provided by the surface of a sphere
or a hyperboloid in three-dimensional Euclidean space. And on each
surface one could have chosen both x and y to be measured along the
axes of a variety of coordinate systems. Some examples are depicted in
Figs. 2.2. In each case one can imagine the points labeled a, b, c, d to
Fig. 2.2 Plots of roots of Eq. (2.1) on a hyperboloid (a) and on a sphere (b).
October 28, 2010 16:55 9in x 6in b943-ch02
30 Dynamical Symmetry
correspond to the solutions i, 1, −i, −1. Depending upon the physical (or
other) interpretations given to the roots, a plot on one of these surfaces
might be more appropriate than a plot in the plane.
4. One could have supposed the four points to be points in some space of
dimension greater than three.
5. We supposed a fixed coordinate system and moved points about it. The
relation between a point and a coordinate system is, however, a relative
one. We could just as well have considered the points to remain fixed and
carried out the inverse motion on the coordinate system. For example, we
rotated points by 90◦ about the point (0, 0) in the plane, but the resulting
relation of point to coordinate system is the same as that obtained by
rotating the coordinate system through −90◦ . We imagined uniformly
enlarging the square, but the resulting relation to the coordinate system
could have been obtained by uniformly shrinking units of measurement
on the axes. (Considering an operation to act on a geometrical figure
is the active interpretation of the operation; considering it to act on a
coordinate system is its passive interpretation.)
6. By plotting the roots of (2.1.1) in a Euclidean plane equipped with an
orthogonal coordinate system, we accepted a concept of measurement
in which the distance, ∆s, between a point with Cartesian coordinates
(x, y) and a point with Cartesian coordinates (x , y ) is predetermined;
if ∆x = x − x, and ∆y = y − y are two sides of a right triangle with ∆s
as the hypotenuse, then ∆s satisfies
This relation would not hold true on the curved surfaces indicated in
Figs. 2.2. The infinitesimal analog of (2.2.1a), obtained by allowing ∆x
and ∆y to become arbitrarily small but nonzero, is:
The transformations we will consider in this section have the property that
the transformed variables may be inhomogeneous linear functions of the
original variables. That is, in (2.3.1a,b)
x = a0 + a1 x + a2 y, y = b0 + b1 x + b2 y,
(2.3.1c,d)
x = a0 + a1 x + a2 y , y = b0 + b1 x + b2 y .
The point follows the indicated path as θ varies continuously from an inital
value of zero. For each value of θ the inverse transformation is obtained
by changing θ to −θ and applying the transformation to the final point.
That is,
32 Dynamical Symmetry
0.7 0.7
0.6 0.6
0.5 0.5
0.4 0.4
0.3 0.3
0.2 0.2
0.1 0.1
0.2 0.4 0.6 0.8 0.2 0.4 0.6 0.8 1.0 1.2
(a) (b)
1.2
1.0
0.8
0.6
0.4
0.2
(c)
P P
X
XXX
XX XX
(a) (b)
use the superscript c to denote points on the coordinate axes, the passive
interpretation of an active rotation through an angle θ is defined by
xxc = xc cos(−θ) − yc sin(−θ), yyc = xc sin(−θ) + yc cos(−θ).
(2.3.3a)
If xc , yc , are of unit length, then in vector notation, (2.3.3a) becomes
xxc = i cos(−θ) − j sin(−θ), yyc = i sin(−θ) + j cos(−θ). (2.3.3b)
October 28, 2010 16:55 9in x 6in b943-ch02
It follows that xxc xxc = 1, yyc yyc = 1, xxc yyc = 0, so xxc and yyc are
orthogonal unit vectors, and we write
34 Dynamical Symmetry
1.0 2
0.5 1
−0.5 −1
−1.0 −2
(a) (b)
−2 −1 1 2
−1
−2
(c)
theorem. Though, ∆r2 = ∆x2 + ∆y2 , ∆r2 = ∆x2 + ∆y2 . In fact, using
the relation cosh(γ)2 − sinh(γ)2 = 1 one finds it
unchanged by pseudo-rotations.
When studying the illustrations of the previous figures one naturally
supposes the paper to be held fixed, and may unconsciously use this per-
ception in interpreting the transformations being illustrated. A geometric
October 28, 2010 16:55 9in x 6in b943-ch02
Y
1.0
Y
2.0
0.5
1.5
1.0 X
− 1.0 − 0.5 0.5 1.0
0.5
− 0.5
0.0 X
0.0 0.5 1.0 1.5 2.0
− 1.0
(a) (b)
Action of Hyperbolic "Rotations"
1.0
0.5
− 0.5
− 1.0
(c)
36 Dynamical Symmetry
38 Dynamical Symmetry
with
H = RHR−1 . (2.5.2b)
Here R−1 , the inverse of R(θ), is R(−θ). Carrying out the matrix multipli-
cations in (2.5.2b) and using trigonometric identities to simplify the result,
one finds that the elements of H are
h11 (θ) = 1/2 h11 (1 + cos(2θ)) + 1/2 h22 (1 − cos(2θ))
− 1/2 (h12 + h21 )sin(2θ),
h12 (θ) = 1/2 h12 (1 + cos(2θ)) − 1/2 h21 (1 − cos(2θ))
+ 1/2 (h11 − h22 )sin(2θ),
(2.5.3)
h21 (θ) = 1/2 h21 (1 + cos(2θ)) − 1/2 h12 (1 − cos(2θ))
+ 1/2 (h11 − h22 )sin(2θ),
h22 (θ) = 1/2 h22 (1 + cos(2θ)) + 1/2 h11 (1 − cos(2θ))
+ 1/2 (h12 + h21 )sin(2θ).
Fig. 2.5.1 (a) Moebius band. (b) Surfaces swept through by the normal to the
band.
As both bands in Fig. 2.5.1b are Moebius strips, vectors in both behave
similarly.
40 Dynamical Symmetry
y
0. 0.5 1.
1.
ds
da
0.5z
db
0.
42 Dynamical Symmetry
or as
ds2 = ds2t = c2 dt2 − (dx2 + dy2 + dz2 ), (2.7.4b)
where c is the velocity of light in vacuo. The metric in (2.7.4a) is said to
be spacelike, and the metric in (2.7.4b) is said to be timelike. All observers
find for c a value of approximately 2.99776 × 108 meters/sec. For a point
at (x, y, z) moving to (x + dx, y + dy, z + dz) with the velocity of light,
dx2 + dy2 + dz2 = c2 dt2 , and so in both cases ds = 0. We use dssp
to denote the real-valued ds that is obtained when the spatial separation
dr = (dx2 + dy2 + dz2 )1/2 is greater than c dt, the time it takes for light to
travel between them; and dst to denote the real-valued ds that arises when
dr < c dt. Some authors use only one definition of the metric, allowing ds
to take on imaginary as well as real values. By preference we require ds
be real-valued, and shift between the spacelike dssp and the timelike dst to
keep ds real.
The Lorentz transformations of special relativity leave invariant the ds2
of Eqs. (2.7.4) and relate observations of the same pair of events perceived
by two observers moving at constant velocity relative to one another, but
using their own Cartesian spacetime coordinate systems, coordinate sys-
tems in which they are at rest. These observers will not agree on the spatial
distance dr separating events nor upon the time interval dt between them,
but they will agree upon the spacetime interval ds separating them.
In the simplest case, an observer A and another observer B at rest with
respect to one another arrange their clocks and meter sticks to agree as
closely as possible. They align their x, y, and z axes as closely as possible
with those of the other, and then set themselves, together with their mea-
suring systems, in relative motion with constant velocity vectors parallel
to their x axes. If vx is the velocity A measures for B’s system, then B
finds that A’s system has velocity −vx . A and B observe the same set of
events and each measures the spatial separation and time intervals between
the events, then transmits the results to the other observer. If one finds
that two events are separated by ∆x and ∆τ = c∆t, the other finds them
separated by ∆x and ∆τ = c∆t , with
∆x = ∆x cosh(α) − ∆τ sinh(α), ∆τ = ∆τ cosh(α) − ∆x sinh(α),
tanh(α) = v/c, α = ln((1 + v/c)/(1 − v/c))1/2 , v = ±vx .
(2.7.5a,b,c)
Each observer could conclude that the other’s clock has slowed, and meter
stick contracted, in the x direction. If, however, they plot their observed
October 28, 2010 16:55 9in x 6in b943-ch02
44 Dynamical Symmetry
Fig. 2.7.2 (a) Projection of B’s coordinate system onto the cartersian system of
A; (b) Projection of A’s coordinate system onto the cartersian system of B.
own (0, 0) point, then both may replace (∆x, c ∆t) by (x, t). Points on the
straight lines in the figures represent intervals measured for motions taking
place at the speed of light in vacuo, c. These lines separate regions where
∆s = ∆st , from the top and bottom regions where ∆s = ∆ssp .
October 28, 2010 16:55 9in x 6in b943-ch02
46 Dynamical Symmetry
Fig. 2.7.5 Curves of σ = constant; they are orthogonal to the curves of Fig. 2.7.4.
Fig. 2.7.7 (a) Triangle connecting three events, through observer A’s frame of
references; (b) triangle connecting same events, through observer B’s frame of
reference.
√ √
at time t = 3 in observer A’s system. In this
√ 2 case ∆x = √1, ∆t = 3, so
2 1/2
∆s is timelike, and ∆st has the value (( 3) − 1 ) = 2. Connecting
the observations by straight lines yields the equilateral triangle shown in
the figure.
October 28, 2010 16:55 9in x 6in b943-ch02
48 Dynamical Symmetry
50 Dynamical Symmetry
52 Dynamical Symmetry
The symmetry of the figure or lattice is invariant under the action of the
continuous group of transformations of the Poincare similitude group, which
leaves invariant the real-valued spacetime intervals defined by (2.10.2), or
multiplies these by a real number greater than zero.
0.6
0.4
0.4
0.2
0.2
− 0.4 − 0.2 0.2 0.4 − 0.6 − 0.4 − 0.2 0.2 0.4 0.6
− 0.2
− 0.2
− 0.4
− 0.4
(a) (b)
0.6
0.4
0.4
0.2
0.2
− 0.8 − 0.6 − 0.4 − 0.2 0.2 −1 − 0.8 − 0.6 − 0.4 − 0.2 0.2 0.4
− 0.2
− 0.2
− 0.4
− 0.4
− 0.6
(c) (d)
Fig. 2.10.2 Events from the perspectives of A, B, C and D in (a), (b), (c) and
(d) respectively.
54 Dynamical Symmetry
−0.5
0.0 −0.5
0.5 0.0 0.5
1 1 1
[ ,− , ]
2 2 2
1 1 1
1 1 1
[− , , ] [− ,− , ]
2 2 2
2 2 2 0.5 0.5
0.0
0.0
1 1 1
[ ,− ,− ] −0.5
2 2 2
−0.5
−0.5
0.0 −0.5
0.0
0.5
0.5
(a) (b)
−0.5 0.0
0.5 −0.5
0.0
0.5
1.0
0.5
0.5
0.0
0.0
−0.5
−0.5
−1.0
−1.0
−0.5
0.0
−0.5
0.5
0.0
1.0
0.5
(c) (d)
Fig. 2.10.3 Relation between eight events in spacetime from perspectives of four
observers in x, y, ct coordinates. (a) Events in A’s frame; (b) events in B’s frame,
α = 0, θ = π2 ; (c) events in C’s frame, α = 12 , θ = 0; (d) events in D’s frame,
θ = 12 , α = 12 .
with relative velocity 0.46c. All observers find s2 = 1/2. Observers A, with
(θ = 0, α = 0), and B, with (θ = π/4, α = 0), find the flashes occur at the
corners of a square. Observer C moving relative to A with (θ = 0, α = 1/2)
observes the events at corners of a rectangle; and observer D, with (θ = π/4,
α = 1/2), finds them to be at the vertices of a rectangular parallelepiped.
An analogous situation occurs in Euclidean geometry. If a square in an x–y
October 28, 2010 16:55 9in x 6in b943-ch02
plane through z = 1/2 is rotated about an axis in the plane, its projec-
tions on the x–y plane through z = 0 can be similar to the projections in
Figs. 2.10.2.
As the physical relationship defined by the four events depicted in
Fig. 2.10.2a has not itself changed, observers A,B,C,D are all recording
events defining a quadrilateral with all the spatial symmetry of the square
determined in the analysis of Fig. 2.1 and tabulated in Table 2.1. Stated in
another way, figures with the apparently different spatial symmetries pic-
tured in Figs. 2.10.2 have physical symmetries that are indistinguishable
in a spacetime with metric (2.10.2).
Figures 2.10.3 illustrate the spacetime perspectives of observers
A, B, C, D on two sets of four events, each defining quadrilaterals like those
in Figs. 2.10.2. For simplicity it is supposed that the clocks are so imagi-
natively set that for A the events occur at ct = −1/2 and ct = 1/2. All the
observers find s2 = 1/4. Again, all the figures are interconverted by actions
of the Poincare group, so their symmetries in spacetime are identical.
Figures 2.10.1 and 2.10.2 display a symmetry defined by an E(2) spa-
tial subgroup of the Poincare group: the E(2) operations act in x–y planes
with fixed t. Figures 2.10.3 display this same symmetry, and also rela-
tions defined in planes parallel to a t axis. To deal with these, we consider
briefly symmetries defined by operations of the Poincare group that are not
contained in its spatial subgroups.
Hyperbolic symmetry is defined by the invariance of curves of constant
∆s or s under the one-parameter subgroup of hyperbolic pseudo-rotations.
Hyperbolic symmetry can be thought of as a spacetime analog of rotational
symmetry: the Minkowski transformations z → ict, r2 → ±s2 convert circles
and spheres defined by
y 2 + z2 = r 2 , x2 + y2 + z2 = r2 , (2.10.3a,b)
56 Dynamical Symmetry
hyperbolae rather than circles. The same is true of the discrete opera-
tions which act on the top and bottom faces of Figs. 2.10.3a,b to yield
Figs. 2.10.3c,d.
For a conclusion, we note that arguments of the form we have used to
define the meaning of geometric symmetry both in Euclidean spaces and in
spacetime may be used to define geometric symmetry in any space with a
metric. We shall be using them later. In each case geometrical interpreta-
tions will depend upon physical interpretations.
58 Dynamical Symmetry
circle z1 = x/a, z2 = y/a. Note that both r and the scale parameter a must
have identical units, as the variables zj satisfy (2.11.5). Equations (2.11.4)
define the stereographic projection illustrated in Figs. 1.2, the second set
of figures in Chapter 1.
Figures 1.2a,b depict the manner in which points on the Northern
and Southern hemispheres are mapped into points in the plane. Points in
the Northern hemisphere map into points that lie inside the sphere, and
points in the Southern hemisphere map to points outside the sphere. This
property is independent of the value of the scale factor a. In the illustrated
case, the scale factor is unity, so the intersection of the sphere with the x–y
plane is a unit circle in both the x and y coordinates, and the zj coordinate
systems. When a is not unity, the circle of intersection is defined by
z21 + z22 = 1 in the one coordinate system, and by x2 + y2 = a2 in the other.
Using spherical polar coordinates (θ, φ), with 0 ≤ θ ≤ π and 0 ≤ φ ≤ 2π,
one has on a unit sphere
z1 = sin(θ) cos(φ), z2 = sin(θ) sin(φ), z3 = cos(θ). (2.11.7)
Using (2.11.4), (2.11.7) one finds that stereographic projection connects the
x, y coordinates to angular coordinates on the sphere via the relations:
a sin(θ) cos(φ)
x= = a cos(φ) tan(θ/2) ≡ X(θ, φ),
(1 + cos(θ))
(2.11.8)
a sin(θ) sin(φ)
y= = a sin(φ) tan(θ/2) ≡ Y(θ, φ).
(1 + cos(θ))
It follows that
r/a = tan(θ/2). (2.11.9)
In the x–y plane, r = a, θ = π/2, and (2.11.8) requires
x = r cos(φ), y = r sin(φ). (2.11.10)
Stereographic projections relate measurements of coordinates x, y and
displacements dx, dy in a plane to measurements of coordinates θ, φ and
displacements dθ, dφ in the curved two-dimensional space, termed S(2),
that may be considered to be the surface of the sphere in three-space. In
the Euclidean space of (z1 , z2 , z3 ), the metric is
ds2 = dz21 + dz22 + dz23 . (2.11.11)
On the unit sphere this becomes
ds2 = dθ2 + sin2 (θ)dφ2 . (2.11.12)
October 28, 2010 16:55 9in x 6in b943-ch02
(b)
Fig. 2.11.1 The projection shaded in each figure is onto the equatorial plane.
This metric is invariant under the group of rotations about the center
of the sphere, inversions z → −z, and reflections zj → −zj . The group,
denominated O(3), is termed an orthogonal group because it transforms
orthogonal vectors into orthogonal vectors. Both the metric and the sphere
are left invariant by the transformations of O(3).
Subgroups of the invariance group of the sphere will consequently define
the symmetry of a figure on the sphere. The symmetry is not altered by
rotations of the entire figure. Figures may be shrunk or enlarged without
October 28, 2010 16:55 9in x 6in b943-ch02
60 Dynamical Symmetry
in the plane:
60
40
20
0
1 2 3 4 5 6
θ
− 20
− 40
− 60
62 Dynamical Symmetry
All operations of the orthogonal group O(n) leave this quantity invariant.
Orthogonal groups leave orthogonal vectors orthogonal, hence their name.
Unit vectors vj = (0, . . . , 0, vi , 0, . . . , 0), with vj = 1, j = 1, . . . , n, may be
attached to a set of n orthogonal axes. The component cj of the vector
r = Σcj vi on the axis j is then defined by cj = r·vi /|vi |. While unit vectors
remain so when dilatations are considered to be active transformations, they
change length when dilatation transformations are interpreted passively.
In the most common realization of the group SO(n), the n(n − 1)/2
group parameters are the angles that define rotation operations. These act
in n(n−1)/2 mutually orthogonal two-planes in an n-dimensional Euclidean
space. Rotation of a vector z through an angle θjm about the origin in the
zj , zm two-plane only affects the coordinates (zj , zm ). It is defined by
64 Dynamical Symmetry
66 Dynamical Symmetry
a vector C, a normal to the plane from the origin. Figure 2.A.1 depicts the
relationship between C, and the plane normal to it when C has been rotated
about the z1 axis through an angle Θ, measured from the positive z3 axis.
Equation (2.A.1) applies when Θ = 90◦ . In this case, C = (0, (1 − s2 )1/2 , 0).
As τ ranges from 0 to 2π the terminus of the vector Z sweeps along the
circle at the intersection of the plane with the unit sphere. For all values of
s, the terminus of C is the center of the circles swept out by Z.
When the plane Ps is tilted by altering Θ, the normal from the origin
becomes
C(Θ) = (0, cos(Θ)(1 − s2 )1/2 , − sin(Θ)(1 − s2 )1/2 ). (2.A.2)
The vector Z of (2.A.1) then becomes a vector Z(Θ, s, τ ) with coordinates
z1 = s cos(τ ), z2 = cos(Θ)(1 − s2 )1/2 + sin(Θ)s sin(τ ),
(2.A.3)
z3 = cos(Θ)s sin(τ ) − sin(Θ)(1 − s2 )1/2 .
As τ ranges from 0 to 2π, the terminus of this vector describes circles at
the intersection of the tilted plane with the sphere, and center at C(Θ).
Consider, then, the projection of these circles onto the x–y plane. The
three-vector Z(Θ, s, τ ) is projected onto the x–y plane, i.e. z1 –z2 plane,
using the relations
x = az1 /(1 + z3 ), y = az2 /(1 + z3 ). (2.A.4)
This projection yields the two-vector (x, y) = r(Θ, s, τ ) whose compo-
nents are
x = a s cos(τ )/D,
y = a{cos(Θ)(1 − s2 )1/2 + sin(Θ)s sin(τ )}/D, (2.A.5)
October 28, 2010 16:55 9in x 6in b943-ch02
68 Dynamical Symmetry
where
D = 1 + cos(Θ)s sin(τ ) − sin(Θ)(1 − s2 )1/2 .
The centers of the circles on the sphere project to points defined by (2.A.4),
which have coordinates (x, y) given by
(x0 , y0 ) = (0, a cos(Θ)(1 − s2 )1/2 /{1 − sin(Θ)(1 − s2 )1/2 }). (2.A.6)
If the projections are themselves circles, (2.A.6) should fix their centers.
Expanding r(Θ, s, τ )·r(Θ, s, τ ), and using (2.A.6) and (2.A.5) to relate x, y
to τ , one finds that as τ ranges from 0 to 2π, the terminus of r(Θ, s, τ ) does
indeed sweep out a circle defined by
x2 + (y − y0 )2 = R2 ,
with radius
R = a s/{1 − sin(Θ)(1 − s2 )1/2 }, (2.A.7)
and
y0 = a cos(Θ)(1 − s2 )1/2 /{1 − sin(Θ)(1 − s2 )1/2 }.
Both y0 and R become infinite when {1 − sin(Θ)(1 − s2 )1/2 } becomes zero,
which happens for a great circle passing through the poles. This projection
of the circle is one that is indistinguishable from a straight line.
The circles defined by (2.A.3) lie in planes whose normal through the
origin lies in the y–z plane. Rotating this normal, C, about the z3 axis, and
varying Θ, allows one to obtain planes with arbitrary orientation. By vary-
ing s one can then obtain all possible circles produced by the intersection
of planes with the sphere. The rotation of C about the z3 axis through an
angle α produces a corresponding rotation of the vector (x0 , y0 ) and the
circles of (2.A.7). As all circles inscribed on the sphere can be produced by
varying s, Θ, and α, it follows from (2.A.7) that the stereographic projection
of any circle on the sphere is also a circle.
References
[1] B. A. Rosenfeld, A History of Non-Euclidean Geometry, translated by A.
Shenitzer, (Springer, New York, 1988).
[2] B. Riemann, Gesammelte mathematische Werke und wissenschaftlicher
Nachlass, Herausg. von H. Weber unter Mitwerkung von R. Dedekind
(Dover., N.Y., 1953).
[3] H. Helmholtz, Wissenschaftlich Abhandlungen, Bd. 2 (Barth, Leipzig, 1882).
[4] A. A. Michelson, E. W. Morley, Silliman’s. J., 34 (1887) 333.
October 28, 2010 16:55 9in x 6in b943-ch02
[5] A. Einstein, Annalen der Physik, 17 (1905) 891, cf. “On The Electrodynam-
ics of Moving Bodies”, in The Principle of Relativity, W. Perrett, translated
by G. B. Jeffery, (Dover, 1923).
[6] This involves a presupposition about the continuity of space and time which
is open to challenge. It also involves a presupposition that the infinitesi-
mal quantities can be integrated to give unique finite quantities. Cf. A. S.
Eddington, The Mathematical Theory of Relativity (Cambridge University
Press, 1924), pp. 198–9.
[7] S. Lie, Gesammelte Abhandlungen, Vol. 2, pp. 374–479 (Teubner, Leipzig,
1935) (Johnson Reprint, New York, 1973).
[8] F. Klein, Elementary Mathematics From An Advanced Standpoint, Vol. II,
Geometry (Dover, New York, 1939).
[9] E. S. Dana, A Textbook of Mineralogy, 4th edn. revised by W. E. Ford (Wiley,
New York, 1932) pp. 49–56.
[10] E. B. Wilson, Vector Analysis Founded upon the Lectures of J. Willard Gibbs
(1901).
This page is intentionally left blank
October 28, 2010 16:55 9in x 6in b943-ch03
CHAPTER 3
71
October 28, 2010 16:55 9in x 6in b943-ch03
72 Dynamical Symmetry
74 Dynamical Symmetry
in nonrelativistic mechanics. Here, for example, t is a cyclic variable (cf. Sec. 5.7).
October 28, 2010 16:55 9in x 6in b943-ch03
y y
R R'
x x
0 0
(a) (b)
Fig. 3.2.1 (a) A solution of Hamilton’s equations (3.2.6). (b) A solution of Eqs.
(3.2.6) with different initial conditions.
or, equivalently,
q → q = Q(τ ), p → p = P(τ ).
that is, they leave Hamilton’s equations invariant. They also leave invariant
the energy of the system:
E = (ω/2)(y2 + x2 ) = (ω/2)(y 2 + x 2 )
October 28, 2010 16:55 9in x 6in b943-ch03
76 Dynamical Symmetry
or, equivalently,
E = H(q, p) = H(q , p ). (3.2.9)
In Chapter 7 it is shown that these observations can be greatly generalized:
The family of solutions obtained by solving Hamilton’s equations for a
given mechanical system may be considered to define a continuous group
of transformations of its positions and momenta. This group is termed an
evolution group.
78 Dynamical Symmetry
p are also conjugate variables. When this happens, one may treat (q ,p )
in the same way one treats (q,p) in Hamilton’s equations.
The following one-parameter groups of transformations are all groups
of smooth (i.e., C∞ ) symplectic diffeomorphisms, and hence canonical dif-
feomorphisms:
q → q = q + a, p → p = p + b, (3.3.6a, b)
β −β
q → q = e q, p→p =e p, (3.3.6c)
q → q = q cos(α) − p sin(α), p → p = p cos(α) + q sin(α), (3.3.6d)
q → q = q cosh(γ) + p sinh(γ), p → p = p cosh(γ) + q sinh(γ).
(3.3.6e)
that convert H(q,p) to H(q ,p ) and convert Hamilton’s equations into
Hamilton’s equations, do not in general leave the Hamiltonian function,
H(q,p), unchanged. To determine whether H(q,p) is unchanged, one first
October 28, 2010 16:55 9in x 6in b943-ch03
80 Dynamical Symmetry
p
1
0.5
q
−1 − 0.5 0.5 1
− 0.5
−1
Lk = εijk qi pj , i, j, k = 1, . . . , 3. (3.5.1)
82 Dynamical Symmetry
p
1.0
0.8
0.6
0.4
0.2
θ=t
q
−0.2 0.2 0.4 0.6 0.8 1.0
p= j
1.0
j 1.0
0.8
j
0.8
p′ q′
0.6 0.6
0.4
0.4
0.2
0.2
i
−0.4 −0.2 0.2 0.4 0.6 0.8 1.0
i
q= i
0.2 0.4 0.6 0.8 1.0
(a) (b)
p p
j j
−C A A′ A
0.5 0.5
i i
q q
−0.5 0.5 −0.5 0.5
−B B B B
−0.5 −0.5
−A C C′ C
(c) (d)
84 Dynamical Symmetry
For any given function Q(q,p) this equation defines P(q,p) to within
an additive constant. Because one has available an uncountable infinity
of differentiable functions Q, the group of all symplectic transformations
contains an uncountable infinity of transformations. The inhomogeneous
linear symplectic group, ISp(2,R), is a five-parameter subgroup of this
group.
The linear symplectic transformation (3.3.6e), a hyperbolic rotation,
does not preserve the orthogonality of vectors. From this alone it follows
that the space defined by canonical transformations is not a Euclidean
space. However, as in spacetime, one may consistently erect an orthogonal
coordinate system in the space of active canonical transformations of q
and p. We have used a such coordinate system in Figs. 3.5.2. For it, the
Euclidean distance between nearby points with coordinates (q,p) and (q
+ dq, p + dp) is ds2 = dq2 + dp2 . Figures 3.5.3 depict the effect of
linear symplectic transformations (3.3.6) on four vectors A, B, C, D that
describe the square in Fig. 3.5.3a. The reader should imagine that in
each of the subsequent figures these vectors are transformed into vectors
A , B , C , D . Only translation and rotation transformations, which are
transformations of the Euclidean group, leave the lengths and orthogonality
of the vectors, and the Euclidean symmetry of the square unchanged. The
transformations (3.3.6 a,b,d,e) leave invariant a metric dq2 − dp2 , but the
symplectic dilatation (3.3.6c) does not. And, though (3.3.6c,d) leaves dq dp
invariant, (3.3.6e) does not.
In short, symplectic transformations do not in general preserve a
Euclidean metric, nor a metric analogous to that of special relativity, nor dq
dp. However, there is a quantity that is left invariant by transformations of
ISp(2,r). We shall shortly use it to define a metric in symplectic p,q space.
The cross product of Euclidean three-vectors may be used to define the
oriented area of each of Figs. 3.5.3 as (A × B)/k = Aq Bp − Bq Ap where
k = i × j. The oriented area of the square is (A×B)/k = Aq Bp = 1. In the
general case, if θa→b is the angle between A and B measured clockwise from
A to B, then (A×B)/k = |A||B| sin(θa→b ). The dimensionless area defined
here can be obtained by dividing the corresponding physical quantity, an
action, by the units of action, ml2 t−1 . These are units of energy, ml2 t−2 ,
multiplied by the units, t, of time.
October 28, 2010 16:55 9in x 6in b943-ch03
0.5
0.5
− 0.5
− 0.5
−1
−1
(a) (b)
1.5
1
0.5
0.5
−1 − 0.5 0.5 1
− 0.5
−1 − 0.5 0.5
−1
− 0.5
(c) (d)
0.5 0.6
0.4
0.2
−1
(e) (f)
Fig. 3.5.3 Action of transformation of ISp, (R). (a) Figure in phase space. (b)
Translation along q axis. (c) Translation along p axis. (d) Rotation. (e) Hyperbolic
rotation. (f) Sympletic dilatation.
October 28, 2010 16:55 9in x 6in b943-ch03
86 Dynamical Symmetry
The reader will find from (3.3.6) that the oriented areas enclosed in each
of Figs. 3.5.3, are related to one another by symplectic transformations, is
the same. The two triangles in Fig. 3.5.2c have the same oriented area. The
oriented areas in the two triangles in Fig. 3.5.2d have the same magnitude,
but opposite sign.
Adjacent oriented areas defined by the cross product have a property
illustrated in Figs. 3.5.3a. If the clockwise circuit A → B is continued
along the dotted line, and the clockwise circuit −A → −B along the dotted
line, the diagonal is traversed in opposite directions in the two different
circuits. It is evident that the total oriented area is the sum of the adja-
cent areas. If one of the oriented areas were defined by a clockwise circuit,
the other by a counterclockwise circuit, the dotted lines separating the
areas would be traversed in the same direction; the oriented areas would
have opposite signs, and the total oriented area would be the difference
of the two adjacent oriented areas. This is a general property of oriented
areas. It follows from the properties of the vector cross product. To sum-
marize, one has the following rule: If the boundary separating two adja-
cent oriented areas is traversed in opposite directions in circuits around
each sub-area, the oriented areas have the same sign, and the total ori-
ented area of the two is the sum of their individual oriented areas. This
is the area contained within the closed path that circumscribes the two
sub-areas.
In Hamiltonian mechanics one needs a generalization of the cross
product, one that defines oriented areas and actions in spaces of 2n dimen-
sions, rather than three dimensions. If A and B are vectors in a two-
dimensional p,q space we will use the (now standard) symbolism A ∧ B
to denote |A||B| sin(θa→b ). It is called the wedge product of A and B.
Because three points determine a plane, while four may not, it is best to
interpret A ∧ B as two times the oriented area contained within the trian-
gle formed by connecting the head of A to the tail of B, as has been done
in Figs. 3.5.2c and 3.5.2d.
So far we have only considered particular cases in which canonical trans-
formations leave areas invariant. To understand why all canonical transfor-
mations leave areas invariant, one need only understand how they leave
infinitesimal areas invariant — for any finite area can be obtained by inte-
grating infinitesimal areas. Two arbitrary infinitesimal vectors δA and δB
with a common origin determine a local two-plane. The infinitesimal area
defined by δA ∧ δB lies in this plane. One may install a coordinate system
with orthogonal i and j axes in the plane, then determine the components
October 28, 2010 16:55 9in x 6in b943-ch03
then
with
One may choose axes such that the untransformed dq, which we shall term
dqo , lies along i, and dpo , the untransformed dp, lies along j. The right-hand
side of (3.5.4c) can be considered to define a vector, dA, and the right-hand
side of (3.5.4d) can be considered to be dB. And neither of these need lie
along i or j. Equations (3.5.4c,d) can then be written
with
The oriented area defined by dqo ∧dpo in Fig. 3.5.4a is a special case;
the area defined by dq ∧ dp in Fig. 3.5.4b represents the more general
case, for which dq and dp are not orthogonal. As the point at (q,p) is
continuously shifted, this dq ∧ dp will change continuously to dq ∧ dp ,
the transformation of dq and dp being defined by (3.5.4). The oriented
areas dq ∧ dp and dq ∧ dp are related via (3.5.5) and (3.5.6), with the
October 28, 2010 16:55 9in x 6in b943-ch03
88 Dynamical Symmetry
dp
dp'
dq dq'
dq x dp dq'x dp'
general relation
dq ∧dp = (a dq + b dp) ∧ (c dq + d dp)
p 2
0.5
1
q
1 1 1 1
(a) (b)
p p
1 1
q q
1 1 1 1
1 1
(c) (d)
90 Dynamical Symmetry
Two such signed areas are indicated in Figs. 3.5.5c,d. Physically, they
represent the action developed in moving a mass point with initial position
qa and conjugate momentum pa around a closed trajectory in phase–space,
while imparting to it the action defined by the enclosed area. The trajec-
tories may be considered as those produced by a harmonic oscillator with
Hamiltonian p2 /2 + q2 /2, and energy E = 1/2. The clockwise trajectory in
Fig.
√ 3.5.5d will develop as time advances if both p0 and q0 have the value
1/ 2. The counterclockwise√ trajectory in Fig.
√ 3.5.5c can be developed as
time advances, if q0 = 1/ 2 and p0 = −1/ 2. Note that this change in the
sense of rotation has been brought about by changing an initial value, and
not by changing t to −t.
Now let us return to Eqs. (3.5.7), and consider a more general geo-
metric significance of the fact that symplectic transformations do not alter
the value of dq ∧ dp. The relationship between symplectic transforma-
tions and the invariant dq ∧ dp is analogous to the relationship between
Euclidean transformations in an x–y plane and the Euclidean metric ds ·
ds = dx2 + dy2 .
In fact, dq ∧ dp is the metric that characterizes geometry in the q,p
space termed a symplectic space. The geometry of this space is the geometry
of a space with an invariant metric that is a measure of the oriented area
that represents action.
This result establishes a direct connection between symplectic geometry,
and the classical mechanics of Hamilton. The concept of action took on
increased physical importance in 1913 when Paul Ehrenfest1 argued that
the magnitude of the action developed by any closed trajectory in phase–
space is quantized. It appeared that it must be an integer multiple of Plank’s
constant divided by 2π:
|dq∧dp| = n h/2π. (3.5.10)
R
October 28, 2010 16:55 9in x 6in b943-ch03
92 Dynamical Symmetry
We shall suppose that two objects must certainly have identical sym-
plectic symmetry if they can be superimposed by symplectic transforma-
tions. As translations and rotations of a rigid figure do not change the
area enclosed by it, they may be used to test for superimposibility of two
objects in symplectic space. However, this by no means exhausts the catalog
of operations that can be used to superpose figures in symplectic space. All
symplectic diffeomorphisms can be considered to change the relationship
of an object to the coordinate system of an observer, without altering the
symplectic symmetry of the object. Before going on to say anything about
the geometric roles of uniform dilatations of dq ∧ dp, and of transforma-
tions that change its sign, we will, in the next section, consider the physical
significance of these operations.
94 Dynamical Symmetry
96 Dynamical Symmetry
p
1
0.5
q
2 1 1 2
0.5
(a)
p p
2
1.5
1 1
0.5
q q
1.5 1 0.5 0.5 1 1.5 2 1 1 2
0.5
1 1
1.5
E = 1/2 E = -1/2
(b) (c)
Fig. 3.8.1 (a) Trajectories of a free particle in phase space. (b), (c) Trajectories
of republic oscillator in phase space.
transformation on the equation Hfp = const. has the same symplectic sym-
metry as Hfp = const. Any object obtained by action of symplectic trans-
formations on the equations HOSC = const. or Hrep = const., has the same
symmetry as HOSC = const. or Hrep = const., respectively.
The effect of several linear symplectic transformations that alter the
phase portrait of a free particle with Hfp = E = 1/2 is displayed in
Figs. 3.8.2. All have the same symplectic symmetry as Fig. 3.8.2a.
Figures 3.8.3b,c illustrate the effect of two linear symplectic trans-
formations that alter the phase portrait of an oscillator with HOSC = E
= 1/2. All these portraits in Figs. 3.8.3 are considered as having the cir-
cular symplectic symmetry of Fig. 3.8.3a.
Linear symplectic transformations that alter the phase portrait of a
repulsive oscillator with Hrep = E = 1/2 are depicted in Figs. 3.8.4b,c.
Again, the portraits have the same symplectic symmetry as the first por-
trait, Fig. 3.8.4a.
October 28, 2010 16:55 9in x 6in b943-ch03
1.0 2
0.5 1
−2 −1 1 2 −2 −1 1 2
− 0.5 −1
− 1.0 −2
(a) (b)
1.5
0.5
−2 −1 1 2
− 0.5
−1
− 1.5
(c)
98 Dynamical Symmetry
1.0
0.5
− 0.5
− 1.0
(a)
0.5
0.5
− 0.5
− 0.5
−1
−1
(b) (c)
Fig. 3.8.3 (a) Harmonic oscillator. (b) Hyperbolic rotation of 3a. (c) Sympletic
dilatation of 3a.
For them,
{p , q } = 1, (3.9.1b)
and
q = q exp( − αq p ), p = p exp(αq p ). (3.9.1c)
October 28, 2010 16:55 9in x 6in b943-ch03
2 2
1 1
−1 −1
−2 −2
(a) (b)
Note that
p q = p q. (3.9.1d)
0.5
−1 − 0.5 0.5 1
− 0.5
−1
1.5
1.0
0.5
− 0.5
− 1.0
− 1.5
−2 −1 1 2
−1
−2
−3 −2 −1 1 2 3
−1
−2
−3
In the light of this example, one might, at first sight, suspect that all
closed curves have the same symmetry in phase–space. In concluding this
section we wish to show that this is not the case. Consider, for example,
the equation
H(q, p) = E, H = p2 /2 + q4 /2, (3.9.3)
that defines closed curves. It may be obtained from the equation
p 2 /2 + q 2 /2 = E, (3.9.4)
by the transformation p → p, q → q2 . However, aside from the fact that
this is not a symplectic transformation, no diffeomorphism can convert q
to q2 because the inverse transformation must convert both +q and −q to
the same value of q .
In subsequent chapters we will consider transformations that are nonlin-
ear symplectic diffeomorphisms in 2N dimensional symplectic spaces. Some
of these will turn out to possess marvelous properties.
References
[1] P. Ehrenfest, Verh. Deutsch Physikal. Ges. 15 (1913) 451; Physik. Zeitschr.
15 (1914) 657; Phil. Mag. 33 (1917) 500.
[2] W. Heisenberg, Z. Physik 43 (1927) 172; cf., W. Heisenberg, Physical Prin-
ciples of the Quantum Theory (University Chicago Press, 1930).
[3] P. A. M. Dirac, Proc. Roy. Soc. A 109 (1926) 642.
October 28, 2010 16:55 9in x 6in b943-ch04
CHAPTER 4
4.1 Introduction
Previous chapters have considered examples of continuous groups of trans-
formations — groups such as those of translation, rotation, hyperbolic
rotation, and dilatation, which act on a set of real variables x with transfor-
mations that depend upon a set of continuous real variables, termed group
parameters. The parameters are not members of the set of variables x. This
chapter deals with fundamental properties possessed by groups of trans-
formations whose operations are labeled by one continuously variable real
parameter. In this and the next six sections we will sometimes distinguish
between initial values, final values, and running values lying between the
initial and final values of variables. Primes will be used to denote running
values and asterisks to denote final values. Variables with neither primes nor
asterisks will denote initial values. We begin with examples that illustrate
key points.
Figure 4.1.1 depicts an active interpretation of the dilatation trans-
formation
105
October 28, 2010 16:55 9in x 6in b943-ch04
χ
2
α
− 1.0 − 0.5 0.5 1.0
−1
−2
Fig. 4.1.1 Action of Dilatation Transformation x = xeα .
T(α0 )x = I x = x. (4.1.2a)
In this case, α0 = 0. Next, let the transformation T(α)x yield xα , and let
it be followed by a transformation of xα given by T(β) i.e.
Since (4.1.2a) and (4.1.2b) hold for a range of x, they state a more abstract
property of the group which may be expressed as
requires that
φ(γ, φ(β, α)) = φ(φ(γ, β), α). (4.1.2g)
For the transformation (4.1.1), with
x → T(α)x = x exp(α) = x∗ , (4.1.3)
the identity transformation is obtained by setting α = 0. For the composi-
tion of two transformations one has
T(γ)x = T(β)(T(α)x) = T(β)x exp(α) = x exp(β) exp(α) = x exp(α + β).
(4.1.4)
The composition law is γ = α + β. The identity transformation is produced
if β = −α; consequently the inverse of T(α) is T(−α). Note that this inverse
transformation, from x∗ to x in (4.1.3) may also be obtained by solving
equation (4.1.3) for x.
The reader can directly verify that all of the relations (4.1.2) hold true
for a continuous range of the parameters α, β, γ. Once the reader also
verifies that the associative law, Eq. (4.1.2f) holds, it will become evident
that Eq. (4.1.3) defines a continuous transformation group.
So far, in Eq. (4.1.2), we have not considered the range allowed for the
parameters α, β, and γ. A continuous transformation group acting on real
variables x is termed global if the group parameter may run through the full
range of the reals, −∞ < α < ∞, without carrying real variables x outside
of the range of the reals. In global groups, α and α mod some number, e.g.
2π, may give rise to the same transformation. The reader may verify that
the transformation group defined in this example is a global one-parameter
continuous group.
Figure 4.1.2 illustrates curves determined by the transformations x =
T(α )x with
x = x/(1 − α x). (4.1.5)
The identity transformation is obtained when α → 0. A transformation
with parameter α, followed by one with parameter β, gives
x = T(β)T(α )x = T(β)x , (4.1.6a)
so
x = x /(1 − βx ), with x = x/(1 − α x). (4.1.6b)
Thus
{x/(1 − αx)}
x = . (4.1.6c)
(1 − β{x/(1 − αx)})
October 28, 2010 16:55 9in x 6in b943-ch04
χ
30
20
10
α
− 0.5 0.5 1.0 1.5
− 10
− 20
− 30
− 40
The requirement that the Fj (x , α ) are continuous functions of their argu-
ments implies that as α → 0,
xj = Fj (x , α ) → xj = Fj (x , 0). (4.2.1b)
Let us also require that the derivatives
dxj /dα = dFj (x , α )/dα , j = 1, 2, . . . , n, (4.2.2a)
are continuous real functions of x , α . Define
fj (x ) = dFj (x , α )/dα )|α =0 , (4.2.2b)
then, as α → 0, Eqs. (4.2.2a) become
dxj /dα = fj (x ), j = 1, 2, . . . , n. (4.2.2c)
Given these requirements, the transformation group determines a set of
first-order ordinary differential equations in which −∞ < fj (x ) < + ∞.
One may suppose that (4.2.2c) relates infinitesimal changes, δα in the
parameter α and infinitesimal changes δx in the xj , via the n equations
δxj = fj (x )δα , j = 1, 2, . . . , n. (4.2.3a)
From a physical standpoint one may consider that, in some system of
measurement, these equations extrapolate an observed relationship
∆xj = fj (x )∆α (4.2.3b)
that continues to hold within experimental error as the observer makes
increasingly refined observations. As an example of this, consider the group
of mathematically defined functional relations
x∗ = x cos(α) − y sin(α), y∗ = y cos(α) + x sin(α). (4.2.4)
For running values of the variables one has
(dx /dα )|α =0 = (−x sin(α ) − y cos(α ))|α =0 = −y , (4.2.5)
and
(dy /dα )|α =0 = (−y sin(α ) + x cos(α ))|α =0 = x .
Also, when α → 0, the variables x → x , and y → y , so one may write
dx /dα = −y , dy /dα = x . (4.2.6)
An experimentalist studying rotations in the plane might make measure-
ments of increasingly minute values of ∆x , ∆y , and a change in angle, ∆α .
October 28, 2010 16:55 9in x 6in b943-ch04
1.0
1
0.5
0
− 1.0 − 0.5 0.5 1.0
1
− 0.5
−1
0 y
−1
0
− 1.0 −1
x
1
(a) (b)
1.0
0.5
− 0.5
− 1.0
(c) (d)
Fig. 4.2.1 (a) Vector field ∆r = [∆x, ∆y] defined by infinitesimal rotations;
(b) vector field [∆x, ∆y, ∆α] defined by infinitesimal rotations; (c) path curves
of successive finite rotations in the x–y plane; (d) path curve in the space of
(x, y, α).
The finite transformations that move points along the indicated lines can be
approximated by connecting the arrows. For the figure, this has been done
by numerically integrating Eqs. (4.2.10), beginning at an initial point on
each line. In this last example we have not produced a general equation that
defines the finite transformations. The reason for this will become apparent
in Sec. 4.8 below.
October 28, 2010 16:55 9in x 6in b943-ch04
y
y
0.0 x
3 0.5 1.0 1.5 2.0 2.5 3.0
2
− 0.5
− 1.0
x
− 0.5 0.5 1.0 1.5 2.0 2.5
− 1.5
−1
continuous real variables zi that may be independently varied over the full
open range of the reals, i.e. −∞ < zi < ∞. Variations of zi that are arbi-
trarily small but nonzero, that is, infinitesimal, will hereafter be denoted
by dzi or δzi . Except when otherwise noted, the variables zi will represent
Cartesian coordinates of points PZ in a Euclidean space with metric
This space is termed EN . Real variables that do not necessarily have all
the properties of the Cartesian variables zi, will be generically denoted as
xi ’s. Often the x’s will be geometrically interpreted as defining points on
a manifold in a Euclidean space. Thus the angular coordinates (θ, φ), of a
point on the unit sphere defined by z21 + z22 + z23 = 1, are x’s, as are the z’s,
but θ, φ, are not z’s. The time coordinate in spacetime will be considered
an ‘x’, except when otherwise stated.
We follow current mathematical usage and require that, by definition, a
function is single-valued. Thus, for example, the relation z3 = ±(z21 − z22 )1/2
does not define a function; it defines two functions which have misleadingly
been denoted by the same symbol, z3 .
As previously noted we will be dealing with transformations which, like
stereographic projections and canonical transformations, are more general
than the linear transformations of vector algebra. Letting x represent a set
of variables, x1 , . . . , xn , their transformation from x to x will be defined by
sets of functions:
x1 = Finv
1 (x ), x2 = Finv inv
2 (x ), . . . , xn = Fn (x ). (4.3.2b)
y
0.5 0 −0.5
0.5
0
z
−0.5
0.5
0
−0.5 x
δα = δy1 /ξ1 (y ) = δy2 /ξ2 (y ) · · · = δyn /ξn (y ), (4.5.1)
October 28, 2010 16:55 9in x 6in b943-ch04
Similarly,
c yc
δα = ∆cb = c − b = δyj /ξj (y ). (4.5.2b)
b yb
it follows that
b c c
δα + δα = δα , (4.5.4a)
a b a
just as
(b − a) + (c − b) = c − a. (4.5.4b)
Using this observation one easily establishes that integration can be con-
sidered an operation that defines a group in which the law of composition
is addition.
Because each of the terms δyj /ξj (y ) in (4.5.1) satisfy the relation
δyj /ξj (y ) = δα , one also has
yc yb yc
δyj /ξj (y ) = δyj /ξj (y ) + δyj /ξj (y ). (4.5.4c)
ya ya yb
y y
0.5 0 0.5 0
−0.5 −0.5
0.5 0.5
0 z 0 z
−0.5 −0.5
0.5 0.5
0 0
−0.5 −0.5
x x
(a) (b)
y
0.5 0 −0.5
0.5
0 z
−0.5
0.5
0
−0.5 x
(c)
The solution curves of the set of differential equations (4.5.1) are path curves
of a transformation group T(α). The operations of the group, integrations,
evolve solutions from initial data. Acting at any point on a given solution
curve, they carry the point along the solution curve. They may be said to
convert the solution curve into itself. Consequently:
ew = Σ∞ n
0 w /n!. (4.6.1)
exp(αU) = Σ∞ n
0 (αU) /n!. (4.6.2)
October 28, 2010 16:55 9in x 6in b943-ch04
If U = Σnj ξj (x )∂/∂xj , then when exp(αU) acts on functions of the xj ,
(4.6.3b) represents its action as the limit of a continuous succession of
infinitesimal transformations each with operator (1 + δαU), in which δα =
α/n. With this observation as a motivation, let us explore the idea that the
finite transformation
x → x∗ = T(α)x, x = (x1 , . . . , xn ), (4.6.4)
defined by
xj → x∗j = Xj (x, α), j = 1, . . . , n,
can be expressed as
xj → x∗j = exp(αU)xj , (4.6.5a)
with
Xj (x, α) = exp(αU)xj . (4.6.5b)
Because the transformation depends continuously upon x and α, it is nec-
essary that Eq. (4.6.5) hold true for all x ranging from x to x∗ . Because of
the differentiabilty requirement on (4.6.4), one must have
dxj /dα = dXj (x , α)/dα (4.6.6a)
so one must also have
dxj /dα = d(exp(αU)xj )/dα,
(4.6.6b)
= exp(αU)Uxj .
The group property requires that these equations must hold in the limit
α → 0. Now
(dXj (x , α)/dα)|α=0 = ξj (x ), (4.6.7a)
and
(exp(αU)Uxj )|α=0 = Uxj . (4.6.7b)
Thus if
U = Σn1 ξj (x )∂/∂xj , (4.6.7c)
October 28, 2010 16:55 9in x 6in b943-ch04
then exp(αU)xj satisfies the differential equations satisfied by the Xj (x, α).
Since both exp(αU)xj and Xj (x, α) yield the identity transformation when
α → 0, it follows that, as conjectured,
Xj (x, α) = exp(αU)xj . (4.6.8)
Thus, a transformation with generator U can be expressed as
T(α)x = exp(αU)x. (4.6.9)
Putting these last two equations together, and recalling that each
Xj (x, α) is defined by integrating Eqs. (4.4.1), one concludes that in (4.6.9)
the operation of integration has been converted into one of exponentiation.
It is important to note, however, that for the series and product expan-
sions of exp(αU)x to be valid, the functions ξj (x ) in (4.6.7c) must be
infinitely differentiable, and the expansion must converge.
When the group parameter α is so related to physical time, t, that
dα = dt, the group becomes the evolution group of the dynamical system
governed by the differential equations that define the group. As an example
of this, consider Hamiltonian’s differential equations (3.3.3), viz.,
dqj /dt = ∂H(q, p)/∂pj , dpj /dt = −∂H(q, p)/∂qj . (4.6.10a)
They can be rewritten in the form
dt = dqj /(∂H(q, p)/∂pj ) = dpj /(−∂H(q, p)/∂qj), (4.6.10b)
and define the group generator
U = −(∂H/∂qi ∂/∂pi − ∂H/∂pi ∂/∂qi ). (4.6.11)
This generator is often written as {−H·}, where, for any function A(q, p),
the operator
{A·} = Σi (∂A/∂qi ∂/∂pi − ∂A/∂pi ∂/∂qi ), (4.6.12)
is a “Poisson Bracket waiting to happen”. The corresponding evolution
operator exp(t{−H·}) acts on q, p to give q∗ = Q(q, p, t), p∗ = P(q, p, t),
the values of the position and momentum vectors at t, given their initial
values q, p at t = 0. If F(q, p) is an analytic function of q, p, then its value
evolves to F(Q(q, p, t), P(q, p, t)) = exp(t{−H·})F(q, p).
As the motions of a system of particles express the action of a one-
parameter group upon the positions and momenta of the particles, they
express a dynamical symmetry of the system. In subsequent chapters, oper-
ators exp(α{A·}) that convert solutions of the equations of motion into
October 28, 2010 16:55 9in x 6in b943-ch04
Here
One has
If x, y are the values of x , y at t = 0, the Lie series expression for the final
value of x at t = t, is
In most cases the series expansion one obtains is not that of easily rec-
ognized functions, and one simply retains a finite number of terms in the
expansion and so obtains approximate solutions of differential equations.
To determine the form U takes in the x coordinate system one need only
determine the functions ξk (x ). In the x system one has
ξk (x ) = Uxk , (4.7.3)
with xj = fj (x). Hence
Uxk = U fk (x) = Σj ξj (x)∂fk (x)/∂xj . (4.7.4)
Denoting ∂fk (x)/∂xj by fk,j (x), one uses the inverse transformation
x = f inv (x ), to convert the right-hand side of (4.7.4) to a function of x ,
obtaining
ξk (x ) = Uxk = Σj ξj (f inv (x ))fk,j (f inv (x )). (4.7.5)
As an example, let us express the generator of rotations in the z1 –z2 plane,
viz.,
U = z1 ∂/∂z2 − z2 ∂/∂z1 (4.7.6)
in the polar coordinates defined by
x1 = r = (z21 + z22 )1/2 , x2 = θ = arctan(z2 /z1 ). (4.7.7a)
The inverse transformation functions, f inv , are given by
z1 = r cos(θ), z2 = r sin(θ). (4.7.7b)
In the left-hand side of (4.7.2) one has ξ1 = −z2 , ξ2 = z1 , so the fundamental
Eqs. (4.4.1) take on the form
dzj /dα = ξ(z )j , (4.7.8)
and require
dz1 /dα = −z2 , dz2 /dα = z1 . (4.7.9)
Applying U to r and θ, one finds
ξr = Ur = 0, ξθ = Uθ = 1. (4.7.10)
Consequently, in the polar coordinate system, U = ∂/∂θ, and Eqs. (4.7.9)
are replaced by
dθ/dα = 1, dr/dα = 0. (4.7.11)
In this example, because the ξ turned out to be constants rather than
functions of the original variables, it is not necessary to use the inverse
transformation to express them as functions of the new variables. The exer-
cises contain examples that require the reader to use the inverse transfor-
mation to obtain the final results.
October 28, 2010 16:55 9in x 6in b943-ch04
In the example just given, the change of variables converted the differ-
ential equations (4.7.9) into the much simpler form (4.7.11). This simplifi-
cation is a special case of a very general result to be described in the next
section. (In it we will cease using primes to distinguish running values of
variables.)
1.0
0.5
x
− 1.0 − 0.5 0.5 1.0
− 0.5
− 1.0
(a)
r = y2
1.0
1.0
0.5
0.5
− 0.5
− 0.5
− 1.0
− 1.0
(b) (c)
before one arrives at the final Eqs. (4.8.2). However, some equations cannot
be integrated analytically. This is the case for Eqs. (4.2.10), which allow no
integrating factor. Nevertheless, the theorem guarantees that if the func-
tions are at least C1 , there exist diffeomorphisms which convert the x’s to
local or global canonical coordinates y.
In physical applications, the most common functions that are not at
least C1 are step functions, which are discontinuous and have discontinu-
ous first derivatives, and sawtooth functions, which have discontinuous first
derivatives. However in physically relevant equations, such as those of elec-
tric circuit theory, these functions approximate to smoother functions. They
October 28, 2010 16:55 9in x 6in b943-ch04
hares hares
1.0 1.0
0.5 0.5
foxes foxes
− 1.0 − 0.5 0.5 1.0 − 1.0 − 0.5 0.5 1.0
− 0.5 − 0.5
− 1.0 − 1.0
(a) (b)
hares
1.0
0.5
foxes
− 1.0 − 0.5 0.5 1.0
− 0.5
− 1.0
(c)
for the circular solution, θ and θ + 2π are equivalent, while on the spiral
solutions θ must range through all the reals to describe the system.
Let us determine local canonical coordinates of the system when r = 1.
Different transformations are required when r < 1 and when r > 1. Equa-
tions (4.8.6) imply that
dr/dθ = r − r3 . (4.8.8a)
October 28, 2010 16:55 9in x 6in b943-ch04
Hence
3 dln(r/(r2 − 1)1/2 ), r > 1,
dθ = dr/(r − r ) = (4.8.8b)
dln(r/(1 − r2 )1/2 ), 0 < r < 1.
Multiplying these equations through by two, one can replace them by
d(2θ) = dln(r2 /(r2 − 1)), r > 1, (4.8.8c)
2 2
d(2θ) = dln(r /(1 − r )), 0 < r < 1. (4.8.8d)
Integrating these, one obtains the solutions in a form valid over both ranges:
exp(−2θ)(r2 − 1)/r2 = k, (4.8.8e)
where
k = exp(−2θ0 )(r20 − 1)/r20 . (4.8.8f)
If the initial value of r is < 1, k is negative and r must remain less than 1.
If the initial value of r is > 1, k is positive and r must remain greater
than 1. If we define s = exp(−2θ)(r2 − 1)/r2 , then in each case s is constant
and ds = 0. If r is initially 1, k = 0, and s = 0, as is ds. In this case, in
(4.8.8g), exp(−2θ) is indeterminate.
Returning to (4.8.8a–d) we see that transforming to θ and s variables
converts these equations to the rectified form
dθ/dt = 1, ds/dt = 0. (4.8.9)
However we must no longer suppose that θ and θ + 2π are necessarily
equivalent, because, even though s remains constant, when s is nonzero, r
varies with s in such a manner that r(θ) and r(θ + 2π) do not have the same
value. Thus when r0 is neither one nor zero, θ must be considered to be a
noncyclic coordinate.
Equations (4.8.5) have limit cycle solutions that, in general, approach
but do not reach the cyclic solution which acts as a stable attractor. The
practical utilization of stable attractors has a long history. Since the days of
De Forest, electrical engineers have utilized negative feedback to stabilize
electrical oscillations. And, much before that, Watt designed a governor
which utilized centrifugal force to ensure that his steam engines would settle
down to a constant running speed.
Long before the arrival of Europeans, native American hunters were
aware that populations of predator and prey species oscillate in near syn-
chrony, the peaks of each population being separated from the other by
October 28, 2010 16:55 9in x 6in b943-ch04
a few years. Eighteenth and nineteenth century fur trade records of the
Hudson’s Bay Company appear to have led to the realization that these
populations obey rate equations with limit cycle solutions.5
A variety of chemical kinetic equations that produce concentration oscil-
lations can have the same property — a property which can also be respon-
sible for stable oscillatory behavior in biological systems.
If limit cycle oscillations become very small, the attractor approaches a
point, and the corresponding biological system exhibits homeostasis. Oscil-
latory chemical concentrations can develop when an autocatalytic reac-
tion and an autodepressive one are chemically coupled. They are of central
importance in Turing’s reaction–diffusion theory in which chemical concen-
tration waves produce cell differentiation and biological morphogenesis.6
Equations (4.8.5) were chosen to provide an instructive example of the
workings of the rectification theorem when the required change of variables
is different for different ranges of the variables x, y, or r, θ.
Differential equations with solution curves of differing topologies are of
common occurrence. For example, objects moving under gravitational or
electrical attractions typically have both closed orbits and open orbits. In
latter chapters it will be shown that this presence of both open and closed
orbits in classical Kepler systems affects the dynamical groups of quantum
mechanical Kepler systems as well as their classical counterparts.
together yield
ϕj (x, t) = φj (ψ(x, t), τ (t, x)). (4.9.10)
4.11 Conclusion
Any finite set of ordinary differential equations defined by sufficiently
smooth functions may be converted to a set of first-order autonomous equa-
tions
δyj /δα = ξj (y ), j = 1, . . . , n, (4.11.1a)
and expressed as
δy1 /ξ1 (y ) = δy2 /ξ2 (y ) = · · · = δyn /ξn (y ) = δα . (4.11.1b)
The rectification theorem guarantees that if the ξj (y ) are at least C1 func-
tions, then these equations at the very least define local one-parameter
groups acting within open intervals of the y’s, and the equations may define
a global transformation group. The generator of a group defined by (4.11.1)
is U = Σ ξj (y)∂/∂yj .
When Eqs. (4.11.1) are integrated over finite intervals, as the inte-
gration proceeds, the variables yj change from their initial values yj =
Yj (y, 0) to their final values yj∗ = Yj (y, γ). In each successive inter-
val, ∆α, ∆β, . . . , the integration process defines a finite transformation
October 28, 2010 16:55 9in x 6in b943-ch04
which it is defined, the diffeomorphism will map the compact region into
a compact region. A square and a circle in the Euclidean plane are not
diffeomorphic because components of the vector r = (z1 , z2 ) from the origin
to points on the boundary of the square do not change smoothly at the
corners of the square.
The following equations define the C∞ diffeomorphisms indicated:
y = exp(x), x = ln(y), −∞ < x < ∞ → 0 < y < ∞;
y = x/(1 − x), x = y/(1 + y), −∞ < x < 1 → −1 < y < ∞;
(x, y) → (x , y ) = (xea , yeb ), with a, b real.
The following do not define diffeomorphisms for the reasons indicated:
i. y = exp(|x|), |x| = ln(y), −∞ < x < ∞ → 0 < y < ∞;
dy/dx is discontinuous at x = 0.
ii. Any transformation from angular coordinates of points on a sphere
to angular coordinates on a torus; the topologies of the surfaces are
different.
A C∞ diffeomorphism becomes an analytic transformation iff the func-
tions fj (x), fj (x) can be expressed by convergent power series expansions
over the range allowed to x. The expansions need not be about a single
point, but may be about a set of points. (If the power series expansion of a
function is convergent about all points, the function is said to be an entire
function.)
Exercises
1. Determine the infinitesimal transformations and generators of the fol-
lowing finite transformations:
a. x = (x2 + 2α)1/2 , y = (y2 − α)1/2 ;
b. x = x/(1 − αx), y = y/(1 − αx);
c. x = eα (x cos(α) − y sin(α)), y = eα (x sin(α) + y cos(α)).
2. Let a be a continuously variable parameter. Can the transformation
x = 1/(x + a), be that of a Lie group with group parameter a, or
function of a?
3. Determine the finite transformations produced by the groups with gen-
erators
a. x(x∂/∂y − y∂/∂x);
b. x∂/∂y − y∂/∂x + k∂/∂z, where k is a constant.
4. Does the transformation defined by Eq. (4.2.10) have any critical points?
October 28, 2010 16:55 9in x 6in b943-ch04
5. What are the ordinary differential equations that define each of the fol-
lowing group generators?
U1 = x∂/∂y − y∂/∂x, U2 = x∂/∂(ct) + (ct)∂/∂x, U3 = xU1 ,
U4 = xU2 , U5 = ctU2 , U6 = x∂/∂x + y∂/∂y + t∂/∂t,
U7 = (x2 + y2 − c2 t2 )∂/∂x − 2xU6 ,
U8 = (x2 + y2 − c2 t2 )∂/∂(ct) + 2ctU6 .
6. Plot the two-dimensional vector fields associated with the Lie generators
U1 through U5 in problem (5).
7. Plot the three-dimensional vector fields associated with the generators
U6 , U7 , U8 , and plot the corresponding fields in the three two-planes in
which x, or y, or ct, is zero.
8. Consider the transformation defined by:
s1 = kx/(x2 +y2 −c2 t2 ), s2 = ky/(x2 +y2 −c2 t2 ), s3 = kct/(x2 +y2 −c2 t2 ),
where k is a positive constant and c is the velocity of light. Calculate
s21 +s22 −s23 , and determine the corresponding inverse transformation from
s = (s1 , s2 , s3 ) to (x, y, ct).
9. Transform the U’s of problem 5 to Lie generators of the form W =
Σξj (s)∂/∂sj .
Note: Problems 5 through 9 all relate to material that will be discussed in
subsequent chapters.
References
[1] H. Jerome Keisler, Foundations of Infinitesimal Calculus (Prindle, Weber &
Schmidt, Boston, 1976).
[2] J. C. Maxwell, Electricity and Magnetism (Cambridge, 1873).
[3] H. Whitney, Collected Papers, eds. J. Eels, D. Toledo, Vol. II (Birkhauser,
Boston, 1992).
[4] V. I. Arnold, Ordinary Differential Equations, translated by R. A. Silverman,
(MIT Press, Cambridge, 1973).
[5] A. J. Lotka, Elements of Physical Biology (Williams and Wilkins, 1925).
[6] A. M. Turing, Phil. Trans. Roy. Soc. (London) 237 (1952) 37.
October 28, 2010 16:55 9in x 6in b943-ch05
CHAPTER 5
Everywhere-Local Invariance
Because of the group property, every point on each path curve may be
considered an initial point of the group action as well as a point obtained
by an infinitesimal or finite transformation of the group, so one may let
α → 0 in (5.1.3) after taking the derivative. Then x → x, and Ux → Ux.
The right-hand side of (5.1.3) then becomes Uf(x), and
135
October 28, 2010 16:55 9in x 6in b943-ch05
Uf(x) = 0 (5.1.5)
f(x ) = 0 (5.1.7a)
when
f(x) = 0, (5.1.7b)
for all values of α, and for all values of x such that f(x) = 0. Let it be sup-
posed that f satisfies no other independently valid equations. In particular,
suppose that f(x) is not of form g(x)2 h(x), that contains repeated factors
g(x) which themselves vanish. With these restrictions in force, there are
only two ways in which exp(αU) can leave the equation invariant. Either:
i) the transformation leaves the function f(x) invariant for all values of x,
or,
ii) it leaves the equation f(x) = 0 invariant for values of x satisfying
f(x) = 0. (This can only happen if Uf contains f(x) as a factor.)
October 28, 2010 16:55 9in x 6in b943-ch05
0 = Uf(x)|f(x)=0 . (5.1.8)
and let
This expression vanishes on the unit circle defined by x21 + x22 = 1. Conse-
quently
0 = Uf(x)|f(x)=0 . (5.1.9e)
October 28, 2010 16:55 9in x 6in b943-ch05
x2
1.5
1.0
0.5
x1
− 1.5 − 1.0 − 0.5 0.5 1.0 1.5
− 0.5
− 1.0
− 1.5
f(x, y) = c (5.1.10a)
i.e. of
or
1.0
0.5
− 0.5
− 1.0
dx
Fig. 5.1.2 Transformation that interconvert curves of a flow defined by da = x.
October 28, 2010 16:55 9in x 6in b943-ch05
the family. This must be the case when the solution curves of (5.1.10b–d)
do not all have the same topology, for the diffeomorphism carried out by
exp(αU) cannot interconvert curves of different topology. An example of
this latter case is provided by the equations of motion of Kepler systems,
for they allow both open, hyperbolic and parabolic trajectories, and closed,
elliptical and circular trajectories.
δ(dx)
1.0
0.8 ∆r'
0.6
∆r
0.4
0.2
dx
0.0
0.0 0.2 0.4 0.6 0.8 1.0
(a) (b)
δv − Σi yi δxi , i = 1, . . . , n, (5.3.1a)
or
Σi yi δxi , i = 1, . . . , n, (5.3.1b)
that is linear in all its variables δv, δxi , yi . The variables are a priori inde-
pendent. A Pfaffian equation is an equation in which a Pfaffian form is set
equal to zero. An example of a Pfaffian equation is provided by the expres-
sion for the element of work, δw, done by a force F acting on a particle
that undergoes a displacement δr, viz.,
δw = F · δr. (5.3.2)
then, and only then, each yi is the partial derivative ∂v/∂xi of a differen-
tiable function v(x), and
For clarity of notation we will, in the next few lines, use the symbol δ
to signify variations carried out by the infinitesimal transformation with
generator U. The infinitesimal transformation of the yi is then
Along with the xi , the v and the yi are variables whose values may change
infinitesimally. Consequently
δ (δxi ) = δ α{Σk (∂ξi (x, y, v)/∂xk )δxk + Σk (∂ξi (x, y, v)/∂yk )δyk
+ (∂ξi (x, y, v)/∂v)δv}, (5.3.8)
δ (δv) = δ α{(∂ζ(x, y, v)/∂v)δv + Σk (∂ζ(x, y, v)/∂xk )δxk
The reader is invited to use (5.3.8) to show that this is the generator of a
transformation that leaves (5.3.9a) invariant.
We next turn to an investigation of the action of a continuous trans-
formation group on a metric. Not all metrics define separations between
points that are independent of the path taken between the points.1 Path-
dependent metrics are termed non-integrable. However the metrics of pri-
mary interest here are integrable and the separation between points may
be written as ds, with ds defined by
iff
i.e. iff
U = Σi ξi (x)∂/∂xi , (5.3.12a)
then
Inserting this into (5.3.11d) and requiring that the result vanish for all
values of δα, one determines that the metric is left invariant by the trans-
formation with generator U if
The case of two variables, when ds2 = dx21 + dx22 , is instructive. One
then has
Because dx1 and dx2 are independently variable, dx21 , dx1 dx2 , and dx22 are
linearly independent. Consequently, (5.3.14) can vanish for all choices of
October 28, 2010 16:55 9in x 6in b943-ch05
Since the first term in this is a function of x2 only, while the second is a
function of x1 only, and x1 and x2 are independently variable, one must
have
where a and b are also arbitrary constants. It follows that the generator of
the group that leaves the metric dx21 + dx22 invariant, is
such that
Σij ((1 + δαU)gij (x))(dxi + δαΣk ∂ξi /∂xk dxk )(dxj + δαΣm ∂ξj /∂xm dxm ).
(5.3.22c)
For this to vanish to first-order in δα, the following expression must vanish:
Σi Σj {(Ugij (x))dxi dxj +gij(x)(dxi Σm ∂ξj /∂xm dxm +dxj Σk ∂ξi /∂xk dxk )}.
(5.3.22d)
This implies
dx∗1 = dx1 + δα ((∂ξ1 /∂x1 ) dx1 + (∂ξ1 /∂x2 ) dx2 ),
(5.4.3a)
dx∗2 = dx2 + δα ((∂ξ2 /∂x1 ) dx1 + (∂ξ2 /∂x2 ) dx2 ).
As the group properties of the transformation are governed by terms first-
order in the group parameter, it is sufficient to expand this through first-
order in δα. To this order,
dx∗1 /dx∗2 = dx1 /dx2 + δα ξ1,2 ,
with
ξ1,2 = ∂ξ1 /∂x1 dx1 /dx2 + ∂ξ1 /∂x2
−(dx1 /dx2 )(∂ξ2 /∂x1 dx1 /dx2 + ∂ξ2 /∂x2 ). (5.4.3b)
When dx1 /dx2 is replaced by df/dx2 ,
ξ1,2 = ∂ξ1 /∂x1 df/dx2 + ∂ξ1 /∂x2 − (df/dx2 )(∂ξ2 /∂x1 df/dx2 + ∂ξ2 /∂x2 ).
(5.4.3c)
Note that ξ1,2 contains terms that depend upon the zero, first, and second
power of df/dx2 .
One may use (5.4.3) to determine the one-parameter Lie groups of point
transformations whose action will leave invariant a derivative dx1 /dx2 . It
will be invariant iff ξ1,2 vanishes for all values of dx1 /dx2 . As (dx1 /dx2 )2 ,
(dx1 /dx2 ), and (dx1 /dx2 )0 are linearly independent, this requires
∂ξ1 /∂x2 = 0,
∂ξ1 /∂x1 − ∂ξ2 /∂x2 = 0, (5.4.4)
∂ξ2 /∂x1 = 0.
The first of these equations implies that ξ1 must be a function of x1 only,
and the last implies that ξ2 must be a function of x2 only. The middle
equation therefore requires that
∂ξ1 (x1 )/∂x1 = ∂ξ2 (x2 )/∂x2 . (5.4.5)
Because x1 and x2 can be varied independently, this is only possible if each
of the two derivatives is equal to the same constant, say a. It follows that
ξ1 = a x1 + b, ξ2 = a x2 + c. (5.4.6)
The corresponding generator of the transformations of x1 and x2 is
U = (a x1 + b)∂/∂x1 + (a x2 + c)∂/∂x2 . (5.4.7)
October 28, 2010 16:55 9in x 6in b943-ch05
∂ξk /∂x1 dx1 /dx2 + ∂ξk /∂x2 → ∂ξk /∂f df/dx2 + ∂ξk /∂x2 = Dξk /Dx2 ,
(5.4.10)
the total derivative of ξk with respect to x2 . Consequently,
ξ1,2 → ∂ξ1 /∂f df/dx2 + ∂ξ1 /∂x2 − (df/dx2 )(∂ξ2 /∂f df/dx2 + ∂ξ2 /∂x2 ),
(5.4.11)
i.e.
δx − rδθ = 0 (5.5.1a)
Its solution is
F = −cotan(θ + g); g an arbitrary function of r. (5.6.13)
To express F(r, θ) as f(x, y) one may use the relations
cotan(θ + g) = cos(θ + g)/ sin(θ + g),
cos(θ + g) = cos(θ) cos(g) − sin(θ) sin(g), (5.6.14)
sin(θ + g) = sin(θ) cos(g) + cos(θ) sin(g).
On setting cos(θ) = x/r, sin(θ) = y/r, one finds
f = −(ax − by)/(ay + bx), (5.6.15a)
with a, b functions of r such that
a(r)2 + b(r)2 = 1. (5.6.15b)
With this restriction, the equation
dy/dx = −(ax − by)/(ay + bx) (5.6.16)
states the form of the most general ODE that is invariant under the group
of rotations carried out by exp(αU), U = x∂/∂y − y∂/∂x. U itself generates
the path curves defined locally by
dx/(−y) = dy/x = dα, (5.6.17a)
or equivalently, by
dθ = dα, r = const. (5.6.17b)
It follows that along any path curve generated by U,
dy/dx = −x/y, (5.6.18a)
and
xdx + ydy = 0 = d(r2 /2). (5.6.18b)
Equation (5.6.18a) coincides with Eq. (5.6.16) if b = 0. In all other cases
the circular path curves, evolved by the action of exp(αU) on r = (x, y), do
not coincide with the solution curves of the differential equation (5.6.16).
Thus, when b = 0, in Eq. (5.6.16), rotations interconvert its solution curves.
Lie1 showed that in such cases, the functions ξ and η in the generator
U = ξ(x, y)∂/∂x + η(x, y)∂/∂y determine an integrating factor for the dif-
ferential equation yx = f(x, y). This discovery is discussed in Appendix A
of this chapter.
October 28, 2010 16:55 9in x 6in b943-ch05
d2 y/dt2 + y = 0. (5.7.6)
dt = dx1 /f1 (x) = dx2 /f2 (x) · · · = dxn /fn (x). (5.9.1b)
Then, if
V = Σj fj (x)∂/∂xj , (5.9.2)
the evolution operator exp(tV) acts on the xj to convert them from their
values at t = 0, to their values at t = t .
If all the f’s are at least C1 functions, then, except in regions where
(5.9.1) has singular points, local diffeomorphisms
with solutions
y1 − t = c1 , yj = cj , j = 2, . . . , m. (5.9.5)
dτ = dt/1 = dx1 /f1 (x) = dx2 /f2 (x) · · · = dxn /fn (x). (5.9.6)
Thus Vt is converted to
Vt acts on the solution y1 − t = c1 to give zero, and it acts on all the
other yj to give zero. It follows that if φ(y, t) = c represents any solution
of (5.9.4), with or without explicit time dependence, then
and
On transforming back to the x system, φ(y, t) → Φ(x, t), and one obtains
the corresponding relations
Vt Φ(x, t) = 0, (5.9.11a)
and
Theorem 5.9.1. Let the solutions Φ(x, t) = c and Φ(x) = c, of the equa-
tions
Require that
The fact that for any solution, Φ = c, Vt Φ = 0, requires that the right-hand
side of the equation, as well as UVt Φ, vanish, whence
Vt (UΦ) = 0. (5.9.17)
One can distinguish the various possible cases if one knows the trans-
formation that puts the original differential equations into canonical form.
One can avoid this by solving the determining equations that arise from
the original differential equations themselves. The availability of computer
programs that set up and solve determining equations often make this the
most practical procedure. A basic mathematical discovery of Reid enables
some of these programs to produce series approximations to generators.4
This is particularly useful when one does not expect exact results in closed
form.5
Jacobi’s relation ensures that if U1 and U2 satisfy
then
U3 = [U1 , U2 ] (5.9.19a)
satisfies
[V, U3 ] = 0. (5.9.19b)
5.10 Conclusion
This chapter has developed basic methods for uncovering and using invari-
ance properties of ordinary functions and equations, and of expressions
involving infinitesimal changes — metrics, Pfaffians and derivatives. Par-
ticular attention has been paid to invariance properties of systems governed
by ordinary differential equations.
Trajectories or solutions defined by a given set of ODEs may be con-
verted into one another by one-parameter transformation groups if the tra-
jectories or solutions are diffeomorphic. The groups are invariance groups
of the differential equations; they define intrinsic symmetries possessed by
the equations.
October 28, 2010 16:55 9in x 6in b943-ch05
φ(x, y) = c, (5.A.2)
It follows that
for example,
e.g.
Tg1 (x)g2 (x) = Tg1 (x)Tinv Tg2 (x) = (Tg1 (x)Tinv )Tg2 (x),
(5.B.5)
= (Tg1 (x)Tinv )g2 (Tx).
takes when acting on the y variables of (5.B.1), we begin with the action of U
on xk , viz. Uxk , and use T to transform it to an action on the corresponding
y variable. Then
The first term in (5.B.7a) is evidently just Tξk (x) = ξk (Tx). The first term
in (5.B.7b) represents the action of the transformed generator on yk . In
Sec. 5.7, we investigated the effect of a change of variables on a group
generator. Setting yk = xk in Eqs. (4.7.3)–(4.7.5) gives
ξk (y) = Uyk |x=f inv (y) = (Σj ξj (x)fk,j (x))|x=f inv (y) , (5.B.8a)
a Some authors require (5.B.3) to be true, by definition, for any g(x). We have not done
where
fk,j (x) = ∂fk (x)/∂xj . (5.B.8b)
This establishes that the diffeomorphism x → y = Tx transforms
U = Σ ξj (x)∂/∂xj to
TUTinv = Σ ξk (y)∂/∂yk . (5.B.9)
We are now ready to prove that the commutator of two Lie generators
is invariant under diffeomorphisms.
Theorem 5.B.1. Let
W = [U, V], (5.B.10a)
with
U = Σ ξi (x)∂/∂xi , V = Σ υj (x)∂/∂xj , W = Σ ωk (x)∂/∂xk . (5.B.10b)
Let
U = TUTinv , V = TVTinv , W = TWTinv . (5.B.11)
Then
W = [U , V ]. (5.B.12)
Proof.
TWTinv = T[U, V]Tinv = T(UV − VU)Tinv
= TUTinv TVTinv − TVTinv TUTinv (5.B.13a–c)
in in
= [TUT , TVT ].
That is to say,
W = [U , V ]. (5.B.14)
Q.E.D.
address this subtlety. In Eqs. (5.9.4), fixing the value of c1 , and each of
the constants of motion, yk = Yk (x), k > 1, at a time t = t0 would com-
pletely fix the coordinates of the bodies in the space of the y’s, and in
the space of the x’s, at t0 . As t varies, the point representing the state
of the system traces out a line, which is the trajectory of the system in
the space of the x’s or y’s. Each successive point on this line determines
a state of the system at successive points of time. Because all the yk are
related to the x’s by diffeomorphisms, the Yk (x) are single-valued functions.
As a consequence trajectories evolving from different points cannot cross.
Furthermore, except in the region of singular points of (5.9.1), infinitesi-
mal changes δxj in initial values xj |t=0 , will produce infinitesimal changes
in the constants of motion Yk (x). However, small finite changes ∆xj |t=0
may produce changes ∆Yk |t=0 that grow exponentially with the ∆xj |t=0 ,
and small finite changes ∆yj |t=0 may produce changes in the ∆xk |t=0 that
grow exponentially with the ∆yj |t=0 . Under these circumstances, some tra-
jectories initially well separated may closely approach one another, then
recede — and in general wander about so much that they will pass through
any finite hypervolume of the space allowed for them by conservation of
energy, momentum, angular momentum, and any other isolating integrals.
An example is provided by the Henon and Heiles system,11 in which the
potential function is (x2 +y2 )/2+x2y−y3 /3. In these cases knowing a value
of the corresponding constant of motion can tell little about the evolution
of the system. The constants so affected are the ones that are said to be
non-isolating.
The study of the stability and reliability of predictions based upon con-
stants of motion of systems involving three or more gravitationally interact-
ing bodies has a long history. For a discussion of the contributions made by
Jacobi, Bruns, Painleve, and Poincare in the nineteenth and early twentieth
century, the reader is referred to Whittaker’s monograph.12 These workers
established fundamental mathematical results and showed that constants
of motion other than those of energy, momentum, and angular momentum
would generally have limited use in making long range predictions, and
retrodictions, of planetary and lunar motions. In the mid twentieth century
further fundamental advances were made by Kolmogorov,13 Arnold,14 and
Moser.15 Treating the sun and planets as point masses it was shown, for
example, that for two planetary orbits whose planes intersect at low angles
there are stable constants of motion which predict the constancy of the
inclination.
October 28, 2010 16:55 9in x 6in b943-ch05
Exercises
1. Is the equation dv − ydx = dc, with dc arbitrary, left invariant by a
non-denumerable, or a denumerable, infinity of one-parameter groups?
2. What equations of the form (5.6.16) have path curves that are themselves
circles?
3. Determine the generators of the transformations that leave invariant the
space metrics obtained from (2.7.4) by setting dz = 0.
4. What is the action of each of the generators (5.7.12a–d) on the general
solution of the equation of motion: y − (A sin(t) + B cos(t)) = 0? Which
U’s generate a transformation that leaves a particular solution invariant?
For further information see Ref. 2.
5. Complete the solution of Eqs. (5.7.9a–d) and compare your result with
(5.7.11).
6. Write down a set of first-order ODEs corresponding to the equation
d2 x/dt2 = 0. Obtain the determining equation for the point transfor-
mations that leave this equation invariant. Then use (5.9.16) to obtain
the same determining equation. What choice of the function w must
be made? Solve the determining equations. Show that the solution is a
linear combination of eight linearly independent U’s.
References
[1] S. Lie, Christ. Forh. Aar., 1874 (1875) 242–254; Collected Works, Vol. III,
No. XIII, p. 176.
[2] C. E. Wulfman and B. G. Wybourne, J. Phys. A 9 (1976) 507–518.
[3] S. Lie, Vorlesungen uber Continuierliche Gruppen (Chelsea, Bronx, N.Y.,
1971).
[4] G. Reid, J. Phys. A 23 (1990) L853; G. J. Reid, Eur. J. Appl. Math. 2
(1991) 293; loc. cit., 319.
[5] W. Hereman, Euromath Bull. 1 (1994) 45.
[6] G. W. Bluman and S. Kumei, Symmetries and Differential Equations
(Springer, N.Y., 1989).
[7] B. J. Cantwell, Introduction to Symmetry Analysis (Cambridge University
Press, Cambridge, 2002).
[8] A. Cohen, An Introduction to The Lie Theory of One-parameter Groups
(D. C. Heath, 1911), reprinted by (G. E. Stechert, N.Y., 1931).
[9] P. J. Olver, Application of Lie Groups to Differential Equations (Springer-
Verlag, N.Y., 1986).
[10] R. Sheshadri and T. Y. Na, Group Invariance in Engineering Boundary
Value Problems (Springer-Verlag, N.Y., 1985).
[11] M. Henon and C. Heiles, Astron. J. 69 (1964) 73–9.
October 28, 2010 16:55 9in x 6in b943-ch05
CHAPTER 6
In this chapter many-parameter groups are defined, and the concepts of Lie
group and Lie algebra are then developed along the lines initiated by Lie.1
The further topics selected are those most often needed in applied work.
169
October 28, 2010 16:55 9in x 6in b943-ch06
(the parameters are listed in reverse order because the operations must be
carried out in that order). Then
It requires that, for all k = 1, . . . , r, one must have αk = Ωk (ω , ω), with
z2 / e P'
T1
P
T2
P''
z1 / e
and
x → (1 + δα2 U2 )(1 + δα1 U1 )x,
= (1 + δα2 U2 + δα1 U1 + δα1 δα2 U2 U1 )x. (6.1.6a,b)
In Fig. 6.1.1a this transformation moves a point P with coordinates x =
(z1 , z2 ) to P and then to P with coordinates x = (z1 + δz1 , z2 + δz2 ). Note
that (6.1.6c) contains all terms in the expansion of (6.1.6a) that are linear in
both δα1 and δα2 . In the same approximation, the inverse transformation is
T(−δα1 , −δα2 ) = (1 − δα1 U1 )(1 − δα2 U2 )
= (1 − δα2 U2 − δα1 U1 + δα1 δα2 U1 U2 ). (6.1.7)
In Fig. 6.1.1a, the inverse transformation carries Px to Px and then to P.
Expanding T(−δα1 , −δα2 )T(δα2 , δα1 ) one obtains
1 + O(δα21 , δα22 ). (6.1.8)
If one carries out the general transformation T(β2 , βi )T(α2 , α1 )x, the
resulting operator must be able to be expressed as
T(γ2 , γ1 ) = exp(γ2 U2 ) exp(γ1 U1 ), (6.1.9a)
with
γ1 = Ω1 (α2 , α1 , β2 , βi ), γ2 = Ω2 (α2 , α1 , β2 , βi ). (6.1.9b)
A fundamental property of Lie groups is exposed if, instead of following
the transformation (6.1.6) by its inverse, it is followed by
T(−δα2 , −δα1 ) → (1 − δα2 U2 )(1 − δα1 U1 )
= (1 − δα2 U2 − δα1 U1 + δα1 δα2 U2 U1 ). (6.1.10)
October 28, 2010 16:55 9in x 6in b943-ch06
z2 / e
T1 P'
P''''
T2
T4 P''
T3
P'''
z1 / e
Expanding the RHS of this, one finds that through terms first order in
δα1 and/or δα2 ,
in which
in which the cµκλ are neither functions of the variables x nor of the group
parameters ω.
and
We have previously seen that each of the generators R annihilates the func-
tion x2 +y2 +z2 , and that the operations of the corresponding one-parameter
groups carry out rotations and leave invariant the Euclidean metric in three-
space.
October 28, 2010 16:55 9in x 6in b943-ch06
The generators of the group SO(2,1) which leaves invariant the value
x + y2 − z2 and the metric (dx2 + dy2 − dz2 ) are the Rz given above, and
2
group. Note also that the operators ∂/∂x, x∂/∂x, x2 ∂/∂x are generators of
a three-parameter group.
These results have an immediate consequence. Suppose that the mem-
bers of a set of generators are operators U(a) = ξa (x)∂/∂x whose ξa (x)
have power series expansions beginning with xa . The previous discussion
establishes that there is no r-parameter transformation group which con-
tains as generators both U(2) and U(3) , and so forth. Lie used extensions
from this observation to construct a wide variety of many-parameter groups
transforming many variables.3
In the previous examples, the commutator of two group generators is
itself a group generator, rather than the more general linear combination
of generators allowed by the theorem. However, it is easily shown that if a
set of r generators U is closed under commutation, then any set of r gener-
ators U obtained by taking independent linear combinations of U, is also
closed under commutation. As a consequence, a set of linearly independent
linear combinations of generators of a group may themselves be considered
generators of a group.
U = Σi ξi (x)∂/∂xi . (6.2.1b)
Theorem 6.2.1. If the uik (x) are linearly independent analytic functions,
the equations
∂xi /∂αj = Σk uik (x)λkj (ω), i = 1, . . . , n, j, k = 1, . . . , r, (6.2.3a)
and
Σk ∂(uik (x) λkj (ω))/∂αm = Σk ∂(uik (x) λkm (ω))/∂αj , (6.2.3b)
determine an r-parameter group whose generators are
Uk = Σj ujk (x) ∂/∂xj. (6.2.3c)
The transformations of the group xj → xj
are defined by = Φj (x; α) xj
in which the Φj (x; α) are analytic functions of their arguments in regions
surrounding (x; 0).
We have previously dealt with a case in which the region does not include
all of the space: the transformation x = x/(1 − αx) which is not defined
when αx = 1. Its generator is U = u(x) ∂/∂x, where u(x) is the analytic
function x2 .
Because the functions ξ(x) defining the generators of Lie groups are
assumed to be analytic functions they may always be approximated by
power series. The commutation relations of the generators may be deter-
mined from a knowledge of terms in the series that are of low order. This
property of Lie transformation groups is, as expected, a great help in applied
work.5
October 28, 2010 16:55 9in x 6in b943-ch06
not to mention a variety of other forms. The first form is particularly useful
when we deal with minuscule values of the parameters.
If the generators of a group do not commute, different orderings of the
operations in (6.2.4b) will have different actions. A number of formulae,
known as Baker–Campbell–Hausdorff relations, use commutation relations
among the group generators to interrelate these.6−8 Commutation relations
of the group generators may also, in some cases, be used to express the oper-
ators S(α1 , αr ), or T(α1 , αr ), as products of one-parameter group operators
exp(ακ Uκ ), in which not all the Uκ are different. Operators of the rotation
group are, for example, commonly put in the form
As before, the cµλκ are neither functions of the variables x, nor of the group
parameters ω. The r different U’s are required to be linearly independent,
October 28, 2010 16:55 9in x 6in b943-ch06
another, the rank is 1. A Lie algebra, all generators of which commute with
one another, is termed Abelian. All the structure constants of an Abelian
Lie algebra are zero.
Lie algebras with identical structure constants are said to be isomorphic.
The freedom represented by the transformation from (6.3.1a) to (6.3.4)
indicates that two Lie algebras of the same dimension, r, may be isomorphic
though their structure constants are different. A method for determining
whether Lie algebras are isomorphic, devised by Killing, is discussed in
Sec. 6.6. It is extensively developed by Cartan and Dynkin and used to
classify Lie algebras. The methodology is outlined in the Sec. 6.7, and the
classification of Lie algebras and groups is outlined in the appendix.
then
Uk → Uk = Σj ζj,k
(z ) ∂/∂zj . (6.4.3)
In this, as usual, the ζj,k (z ) are obtained by using the inverse transforma-
tion, (6.4.2b), to replace z by z in Uk zj = Uk fj (z). Now let us consider the
transformation from z to z passively, that is let us consider it to define
a change from one local coordinate system z to another z . Such a pas-
sive change in coordinates affects only the coordinate labels in Fig. 6.4.1.
Consequently, one expects that
y/ e
2.0
P'
1.5
P''''
1.0
P''
P'''
y' / e
0.5
x/ e
0.0
0.5 1.0 1.5 2.0
x' / e
whose operator carries points on each ellipsoid into points on the ellipsoid,
in exactly the same manner that the operator generated by the R carries
points on a sphere into points on a sphere.
Lie transformation groups generated by two different sets of gener-
ators with isomorphic commutation relations are said to be locally iso-
morphic because, for some (finite) range of the group parameters, the
actions T(α1 , αr )x and T (α1 , αr )y of the two groups can be put in one-
to-one correspondence. Locally isomorphic groups are globally isomorphic
if the one-to-one correspondence persists for all allowed values of the group
parameters. SO(3) and SU(2) are locally isomorphic, but not globally iso-
morphic. Different realizations of a group are produced when different group
generators produce globally isomorphic groups. The R and the R of the
previous paragraph are generators of different realizations of SO(3). Both
SO(3) and SU(2) groups appear in a variety of realizations in physics and
chemistry, some quite surprising.
All group parameters are allowed, in principle, to range through all the
reals, but, as is the case for one-parameter groups, it may happen that, for
any one-parameter subgroup, a group parameter and the same parameter
mod some number, e.g. 2π or 4π, will give rise to the same transformation.
One may, for convenience, then restrict the range of the group parameter so
that the effect of the group operation is not repeated. The group parameter,
say α, can then be contained within a compact region, e.g. −π ≤ α ≤ π.
Such a one-parameter group is termed a compact one-parameter group.
A one-parameter group that is not compact is termed noncompact. The
group parameter of a noncompact group ranges over an interval which may
or may not contain all the real numbers, examples being −∞ <α <∞ and
−π <α <π. The group SO(2,1) contains a compact one-parameter subgroup,
and two noncompact one-parameter subgroups. In the realization given
above, exp(αRz ) is the operator of a compact subgroup whose operation is
the same for α = π as it is for α = −π. On the other hand, the operators
exp(βSx ) and exp(γSz ) do not repeat their action as their group parameters
range through all the real numbers. The group SO(3) contains only compact
one-parameter subgroups.
If all one-parameter groups contained in a many-parameter group
are compact, the many-parameter group is termed compact. If a many-
parameter group contains one or more noncompact one-parameter groups,
the many-parameter group is noncompact. The groups SO(n) are, for all
values of n, compact. The groups SO(p, q) are noncompact for nonzero
p and q.
October 28, 2010 16:55 9in x 6in b943-ch06
The structure constants of the Lie algebra of SO(3) are different from
the structure constants of the Lie algebra of SO(2,1). This is an exam-
ple of a general truth: the commutation relations of a real Lie algebra
determine whether the corresponding Lie group is compact or noncom-
pact. Thus, rather surprisingly, the compactness or noncompactness of a
Lie group is determined by relationships among its infinitesimal trans-
formations. Subtleties may however be involved, as the following example
demonstrates.
The projective group in one variable, say t, has generators
∂/∂t, t∂/∂t, t2 ∂/∂t. (6.4.10)
The general transformation of the group can be given the form
t → t = (t + a1 )/(a2 t + a3 ). (6.4.11)
The identity transformation occurs when α1 = −a1 /a3 , α2 = −a2 /a3 , and
α3 = 1/a3 . The linear combinations of the generators (6.4.10),
U1 = (1/2)(1 − t2 ) ∂/∂t, U3 = (1/2)(1 + t2 ) ∂/∂t, (6.4.12)
together with U2 = t∂/∂t, obey the commutation relations
[U1 , U2 ] = U3 , [U2 , U3 ] = −U1 , [U3 , U1 ] = −U2 . (6.4.13)
The group is therefore locally isomorphic to SO(2,1), and the generator of
the compact subgroup is U3 . As the real line −∞ < t < ∞ is an open
interval it is rather surprising to find that a group that carries it into itself
contains a compact subgroup, which, as we shall see, can carry any point
on the line to any other point on the line.
The transformation (6.4.11), like that generated by t2 ∂/∂t, undergoes
a sudden change in sign from ±∞ to ±(−∞) when a2 t + a3 passes through
zero. In Chapter 2 we saw that a stereographic projection from the plane
onto the sphere projects onto the South pole of the sphere all points whose
(x, y) coordinates in the plane are ±∞. Let us consider the corresponding
projection of the line of t onto a unit circle. Set
x = 2t/(t2 + 1), y = (1 − t2 )/(t2 + 1). (6.4.14)
Then x2 + y2 = 1, and one may set x = sin(θ), y = cos(θ). Consequently
t = tan(θ/2), dt = (dθ/2)sec2 (θ/2). (6.4.15)
Figure 6.4.2 depicts the relation between t and θ. For −∞ < t < ∞ one
has a diffeomorphism that maps t, dt to θ, dθ respectively with θ lying in
October 28, 2010 16:55 9in x 6in b943-ch06
1.0
0.5
θ
t
−2 −1 1 2
− 0.5
− 1.0
Fig. 6.4.2 Stereographic projection of t from the real line onto a unit circle.
the open interval −π ≤ θ ≤ π. However θ also remains real over the closed
interval −π ≤ θ ≤ π with θ = −π representing the same point as θ = π.
The projection converts the Uj to the Vj defined below:
U1 = (q ∂/∂q − p ∂/∂p)/2,
U2 = (q ∂/∂p + p ∂/∂q)/2, (6.4.18)
U3 = (p ∂/∂q − q ∂/∂p)/2.
while
Rz = (x ∂/∂y − y ∂/∂x), (6.4.20b)
so
exp(αU3 ) = exp(α U3 ), α = α + 4n π, (6.4.20c)
while
exp(αRz ) = exp(α Rz ), α = α + 2n π, (6.4.20d)
where n is an integer. Thus the groups are not globally isomorphic.
It is usually assumed that physical time runs continuously from an indef-
inite past through the present to an indefinite future: −∞ < t < ∞, or
even −∞ ≤ t ≤ ∞. One expects that when exp(α ∂/∂t) acts on t to give
t = t + α, one must allow α the full range −∞ < α < ∞. However, even
in Newtonian mechanics, if one is not careful this view can lead to incon-
sistencies. In Chapter 5 we found eight generators of one-parameter groups
that leave invariant Newton’s equation of motion for a harmonic oscillator.
The commutation relations of three of these generators,
U1 = ∂/∂t, (6.4.21a)
U2 = (1 + y2 ) cos(t) ∂/∂y + y sin(t) ∂/∂t, (6.4.21b)
6.5 Transitivity
A transformation group acts transitively in a space or on a manifold if every
point of the space or manifold can be carried into every other point. A group
that fails this criterion, acts intransitively in the space or on the manifold.
Transitivity is a property that is not determined by commutation relations
alone, it depends upon the realization of the group, the functional form of
the ξj,k (z) in the group generators Uk = Σj ξj,k (z) ∂/∂zj . The translation
subgroup of the Euclidean group E(3) in three-dimensional Euclidean space
acts transitively in the space because it can move any point in the space
to any other point in the space. The SO(3) group that carries out rota-
tions about the origin acts transitively on the surface of a sphere centered
at the origin. However this group acts intransitively in three-dimensional
Euclidean space because it cannot move points from one sphere centered at
the origin to points on a concentric sphere of different radius. This exempli-
fies a general observation: a group that acts transitively in a space S cannot
simply move points about in a subspace of S. If S is a space whose points
October 28, 2010 16:55 9in x 6in b943-ch06
and Consequently,
Similarly, one finds that g33 = −2, g13 = g23 = 0. The Killing form is thus
−2 0 0
0 −2 0 . (6.7.5)
0 0 −2
Using these one finds that the only nonzero elements of the 6 × 6 dimen-
sional Killing form for ISO(3) are those of its SO(3) subgroup, the diagonal
elements in (6.7.5). The determinant of the form is thus zero.
The class of groups termed semi-simple all have Killing forms whose
determinant is nonzero. This class, defined in Appendix B, includes the
groups SO(p,q), SO(p + q), and groups locally isomorphic to them, but not
ISO(p, q), ISO(p + q).
October 28, 2010 16:55 9in x 6in b943-ch06
all are differential equations involving operators such as ∂/∂xj, and ∂/∂t.
However, in quantum theory the continuous evolution of causal chains in a
continuous spacetime is replaced by the conception of the continuous space-
time evolution of wave functions, ψ, ψ∗ , which determine the probabilities
that an observable will be found to have a given eigenvalue.
There are times in the evolution of any science when presuppositions
are challenged for good reason, and conjecture is required. After Gell-
Mann proposed SU(3) as a symmetry group interrelating the properties
of baryons, there was a period when theoretical physicists conjectured that
properties of the (infinitely) stable elementary particles, the short lived
elementary particles observed in high-energy collisions, as well as the prop-
erties assumed for quarks are the physical expressions of structure found
in a great variety of groups. It is however one thing to find group struc-
ture, and another to find the underlying space of variables upon which the
group acts, or to distinguish transformation groups from their parameter
groups, which as we have seen often have higher dimensionality. Current
attempts to provide unifying theories of physical phenomena have led to
proposals that the observed relationships are consequences of symmetries
expressed in multidimensional spaces with discontinuities such as strings
and wormholes.30
6.11 Conclusion
Lie established the result that every many-parameter transformation group
whose generators are first-order differential operators is defined by a
restricted class of first-order partial differential equations. He showed that
first-order differential operators are those of a many-parameter transforma-
tion group iff they are closed under commutation.
The commutator of two operators is the noncommutative multiplica-
tion operation used to define Lie algebras: operators that are closed under
commutation are, together with scalars, the elementary objects of a Lie
algebra.
Individual Lie groups are characterized by Lie-algebraic properties of
their generators, and the intervals over which the group parameters range —
the topology of their parameter space. If one is only concerned with the
interaction of the operators upon each other, their characterization as
abstract topological groups is complete.
Each differential equation of the sciences is invariant under the action of
a Lie transformation group; its solutions are interconverted by the group. It
October 28, 2010 16:55 9in x 6in b943-ch06
is now known that partial differential equations can be invariant under the
action of continuous transformation groups whose generators are differential
operators of order greater than one.a Each particular transformation group
can also be realized in a great number of ways. These truths have the con-
sequence that Lie groups have greater relevance to the natural sciences
than expected. When relations between observations are expressions of
Lie symmetries, this freedom of interpretation also allows one to consis-
tently conceive the Lie symmetries as geometric symmetries in a variety of
spaces — Euclidean three-space, Newtonian spacetime, Hamiltonian phase
spaces, Lorentzian spacetime, Einsteinian spacetimes with mass dependent
metrics, the spaces of string theories, etc.
If the first of two transformations is x = f(x; α), and the second is x =
f(x ; β), then
and
Define
and
The composition stated in (6.A.4b) requires that δx = dx . Given (6.A.3b),
this can only occur if
i.e.
δxi = Σλ ((∂fi (x ; α)/∂αλ )|α=0 )δαλ = Σλ uiλ (x )δαλ . (6.A.6b)
October 28, 2010 16:55 9in x 6in b943-ch06
This implies:
The differential equations defining a many-parameter group must be of
the form
This result states a necessary, but not sufficient, condition that must be
satisfied if the differential equations are to define a many-parameter group.
The problem is that it may or may not be possible to integrate them, and
so, for each value of κ, determine the transformations of a one-parameter
group with parameter ακ . For Eqs. (6.A.13b) to be integrable it is necessary
and sufficient that
with
Σρ ∂uiρ (x )/∂αµ ωρκ (α) − Σρ ∂uiρ (x )/∂ακ ωρµ (α)
= Σρ uiρ (x )(∂ωρµ (α)/∂ακ − ∂ωρκ (α)/∂αµ ). (6.A.17b)
Let us first deal with the derivatives in the first line of (6.A.17b). As the
α’s vary and produce variations in the x’s, the resulting local changes in
the uiρ (x ) are given by
As (6.A.16a) requires
one has
∂uiρ (x )/∂ατ = Σj Σσ ∂uiρ (x )/∂xj ujσ (x ) ωστ (α). (6.A.18b)
The appearance of both uiρ (x ) and ujσ (x ) suggests that (6.A.18b) may be
used to establish relations between several generators. Applying (6.A.18b)
to the first sum in (6.A.17b), by setting τ = µ, yields
Σρ ∂uiρ (x )/∂αµ ωρκ (α) = Σρ Σσ Σj ujσ (x ) ∂uiρ (x )/∂xj ωσµ (α) ωρκ (α).
(6.A.19a)
Applying (6.A.18b) to the second sum in (6.A.17b), by setting τ = κ, yields
Σρ ∂uiρ (x )/∂ακ ωρµ (α) = Σρ Σσ Σj ujσ (x ) ∂uiρ (x )/∂xj ωσκ (α) ωρµ (α).
(6.A.19b)
Much of the following analysis involves reorganizing terms in multiple sums.
This may be accomplished by altering summation indices. We begin this
process by interchanging the summation indices ρ and σ in (6.A.19b),
obtaining
Σσ Σρ {(Σj ujρ (x ) ∂uiσ (x )/∂xj ) ωρκ (α) ωσµ (α)}. (6.A.19c)
Subtracting it from (6.A.19a) yields the following expression for the first
line in (6.A.17b):
Σρ Σσ {Σj (ujσ (x ) ∂uiρ (x )/∂xj − ujρ (x ) ∂uiσ (x )/∂xj )}ωσµ (α) ωρκ (α).
(6.A.19d)
October 28, 2010 16:55 9in x 6in b943-ch06
Referring to Eq. (5.8.9) of Sec. 5.8, the reader will find that the expression
enclosed in curly brackets is the coefficient of xi defined by the commutator
of two generators Uσ = Σj ujσ (x )xj , and Uρ = Σi uiρ (x )xi . That is, one can
write
[Uσ , Uρ ] = {Σj (ujσ (x ) ∂uiρ (x )/∂xj − ujρ (x ) ∂uiσ (x )/∂xj )}xi . (6.A.20)
Σρ Σσ {Σj (ujσ (x ) ∂uiρ (x )/∂xj − ujρ (x ) ∂uiσ (x )/∂xj ) ωσ µ (α) ωρ κ (α)}
= Στ uiτ (x )(∂ωτ µ (α)/∂ακ − ∂ωτ κ (α)/∂αµ ). (6.A.21)
The terms ωσ µ (α)ωρ κ (α) on the left-hand side of this expression may be
removed with the aid of the general relation (6.A.11b): multiplying (6.A.21)
by νµσ and νκρ , and then summing over µ and κ converts (6.A.21) to
Σj (ujσ (x )xj uiρ (x ) − ujρ (x )xj uiσ (x )) = Στ uiτ (x ) Cτσρ (α), (6.A.22a)
where
Cτσρ (α) = Σµ Σκ (∂ωτ µ (α)/∂ακ − ∂ωτ κ(α)/∂αµ )νµσ (α)νκρ (α). (6.A.22b)
In (6.A.22a) we have a relation in which the x’s and the α’s can in
principle be varied independently. The reader is invited to use this freedom
and a power series expansion of the Cτσρ (α), to show that for (6.A.22a) to
hold true for any choice of the x’s and α’s, each Cτσρ (α) must be independent
of the α’s, and may be replaced by a constant, cτσρ .
Having established this, one may replace (6.A.22) by
Σj (ujσ (x )xj uiρ (x ) − ujρ (x )xj uiσ (x )) = Στ uiτ (x )cτσρ , (6.A.23a)
with each
cτσρ = Σµ Σκ (∂ωτ µ (α)/∂ακ − ∂ωτ κ (α)/∂αµ )νµσ (α) νκρ (α). (6.A.23b)
a constant.
As the right-hand side of (6.A.23a) is a sum of the coefficient func-
tions uiτ (x ) of generators Uτ = Σi uiτ (x )xi , and the left-hand side is
the corresponding coefficient of the xi defined by the commutator of
October 28, 2010 16:55 9in x 6in b943-ch06
A = B ⊕ C. (6.B.1)
and that all other c’s vanish. Show that in the Killing form,
g11 = 0, g12 = 0, g13 = −3, g22 = 5, g23 = 0, g33 = 0.
Evaluate the determinant of the form, and determine whether the group
is semi-simple.
8. Acting in two-dimensional Euclidean space, the Euclidean group E(2)
has generators
U1 = ∂/∂x, U2 = ∂/∂y, U3 = x∂/∂y − y∂/∂x.
Show that
c312 = 0, c123 = −1 = −c132 , c231 = −1 = −c213 ,
and that all other c’s vanish. Show that the Killing form is
0 0 0
0 0 0 .
0 0 −2
9. Let
dx/da = Ax, dx/db = Bx.
Let A, B respectively be linear and quadratic functions of a, b. In what
cases can the equations define a two-parameter group?
References
[1] (a) S. Lie, Math. Ann. 16 (1880) 451–528; Coll. Works, Vol. 6, pp. 1–94,
(Johnson Reprint, NY, 1973).
(b) Sophus Lie’s 1880 Transformation Group Paper, translated by M. Ack-
erman and R. Hermann (Math. Sci. Press, Broookline. Massachusetts).
[2] Ref. 1(b) above, pp. 161–169; S. Lie, Vorlesungen uber Continuirliche Grup-
pen, reprint (Chelsea, Bronx, N.Y. (1971)).
[3] S. Lie, Transformations Gruppen III, reprint (Chelsea, Bronx, N.Y, 1970).
[4] For a proof of the theorem, see S. Pontriagin, Topological Groups (Princeton,
1946).
[5] G. J. Reid, Eur. J. Appl. Math. 2 (1991) 293, 319.
[6] H. F. Baker, Proc. Lond. Math. Soc. 34(1) (1903) 347–360; 35(1) (1903)
333–374; 2(2) (1904) 293–296.
[7] J. E. Campbell, Proc. Lond. Math. Soc. (1) 29 (1898) 14–32.
[8] F. Hausdorff, Ber. Sach. Akad. Wiss. (Math. Phys. Klasse) Leipzig 58 (1906)
19–48.
[9] A. R. Edmonds, Angular Momentum in Quantum Mechanics (Princeton,
1957), pp. 5–7.
[10] C. E. Wulfman and B. G. Wybourne, J. Phys. A 9 (1976) 507–518.
October 28, 2010 16:55 9in x 6in b943-ch06
CHAPTER 7
Newton’s third law states that if particle i exerts a force Fij on particle j,
then particle j exerts a force Fji on particle i, with
207
October 28, 2010 16:55 9in x 6in b943-ch07
is zero. From this it follows that the total momentum of the system is
conserved:
Σj dpj /dt = d2 (Σj mj rj )/dt2 = 0. (7.1.5)
Newton proved that spherically symmetric masses of total mass mj behave
as point masses of mass mj when acted on by external inverse square law
forces, such as his force of gravity.
Equations (7.1.3) are second-order ordinary differential equations. A
knowledge of the Fj , and initial position and initial velocity vectors, is
therefore sufficient to predict the motions in a classical mechanical system.
This quite amazing feature of Newtonian mechanics withstood all experi-
mental tests and challenges until the advent of quantum theory.
Newton’s equations of motion for a system of interacting particles, are
left invariant by a group of transformations that relates measurements made
by two observers moving with constant relative velocity using clocks that
agree on time intervals. The equations are invariant under the passive trans-
formation t → t = t − (±) t0 . They are invariant under passive coordinate
transformations that convert one Cartesian system with axes (X, Y, Z) to
another with axes (X , Y , Z ), if the two origins are moving with relative
velocity −(±) (vx0 , vy0 , vz0 ), and at time t = 0, have origins whose posi-
tion vectors differ by −(±) x0 , y0 , z0 . The two systems of axes may differ
in orientation. Measurements of positions and velocities in the two systems
are related by the active transformations
t → t = t ± t 0 , (7.1.6a)
rj → rj = rj ± (x0 , y0 , z0 ), (7.1.6b)
T
rj = ±R(xj , yj , zj ) , (7.1.6c)
where R is a rotation matrix. The transformation
vj → vj = vj ± (vx0 , vy0 , vz0 ), (7.1.6d)
also converts rj to
rj = rj ± (tvx0 , vy0 , vz0 ). (7.1.6e)
The transformations (7.1.4a–e) are those of the Lie transformation
group termed the Galilei group.1 From a purely mathematical standpoint,
the transformations of the group may be active or passive ones. From a
physical standpoint, the transformations must relate measurements made
in one reference system to those in another by passive transformations of
October 28, 2010 16:55 9in x 6in b943-ch07
X → Y, Y → X, Z → −Z, (7.1.11)
At this point, the q’s and p’s are considered to be mathematical symbols
whose physical interpretations have yet to be supplied. To generalize the
discussion of Hamiltonian mechanics in Chapter 3, we consider transforma-
tions of the combined sets of variables (q, p) to sets (q , p ) in which
dqi → dqi = Σk (∂Qi /∂qk dqk + ∂Qi /∂pk dpk ) = Σk (aik dqk + bik dpk ),
dpj → dpj = Σk (∂Pj /∂qk dqk + ∂Pj /∂pk dpk ) = Σk (cjk dqk + ejk dpk ),
(7.2.3)
October 28, 2010 16:55 9in x 6in b943-ch07
with
∂Qi /∂qk = aik , ∂Qi /∂pk = bik , ∂Pj /∂qk = cjk , ∂Pj /∂pk = ejk .
s = (q1 , . . . , qn , p1 , . . . , pn ), (7.2.4a)
q = (q1 , . . . , qn , 0, . . . , 0) (7.2.4b)
and
p = (0, . . . , 0, p1 , . . . , pn ).
The pair of orthogonal unit vectors of Sec. 3.5 may generalized to a set of 2n
orthogonal unit vectors which we denote i1 ,. . . ,in and j1 ,. . . ,jn . Eq. (7.2.3)
then state
The coefficients aik and bik represent the components of the local displace-
ment vectors on the unit vectors of the reference system:
The vectors dq and dp have a common origin at the point Ss and together
define a two-plane. Figure 7.2.1a is an attempt to depict two such planes
that are produced as Ss moves in a space of dimension greater than two.
In Chapter 3, transformations which convert Hamilton’s equations of
motion into Hamilton’s equations of motion were characterized analyti-
cally with the aid of Poisson brackets, and geometrically by oriented areas
defined with the aid of the wedge product analog of the vector cross prod-
uct. To generalize these concepts to 2n dimensions, one first of all defines
October 28, 2010 16:55 9in x 6in b943-ch07
dp
dp
dq dq
the Poisson bracket {A(q, p), B(q, p)} of functions A(q,p) and B(q,p) by
One gives a geometric interpretation to the equations with the aid of the
wedge product defined in Chapter 3. At the points Ss and Ss the vectors
dq and dp, and the vectors dq and dp define local wedge products whose
magnitude is twice the area of the plane triangle obtained by connecting
October 28, 2010 16:55 9in x 6in b943-ch07
If dqk ∧ dpk is the wedge product in the local qk –pk subplane at s, and
dqk ∧ dpk is the wedge product in the local qk –pk subplane at s , then (cf.
the exercises), Eq. (7.2.9) imply that
dq ∧ dp = (dqk ∧ dpk ), (7.2.11a)
k
and that at s ,
dq ∧ dp = (dqk ∧ dpk ). (7.2.11b)
k
For the assumed form of L, Eq. (7.2.12b) determines the momenta pj cor-
responding to the qj . In Appendix A it is proved that the pj so defined are
canonical conjugates of the qj . The proof is prefaced by, and depends upon,
Thomson and Tait’s direct derivation of Lagrange’s equations of motion
from Newton’s.2
In the discussion of the Hamiltonian mechanics in Sec. 3.7, the phase
space of Hamiltonian mechanics differs from symplectic space, because in
October 28, 2010 16:55 9in x 6in b943-ch07
physics one must consider scale changes q → aq, p → bp, that need not
be the scale changes q → aq, p → a−1p, which leave the wedge product
dq ∧ dp invariant. The same arguments that were applied to shifts from
clockwise directed, to counterclockwise directed, two-dimensional wedge
products, also apply to (7.2.11a). Thus, we adopt the following.
On defining
H = − L − pj qj , (7.3.5)
j
The fact that dH is a function of the dp’s and dq’s suggests that it might
be useful to convert H in (7.3.5) to a function of q’s and p’s. This is done
by defining Hamilton’s function, H, by
H(q, p) ≡ pj vj − L |v=g(q,p) . (7.3.7)
It has
dH = (∂H(q, p)/∂pj dpj + ∂H(q, p)/∂qj dqj ). (7.3.8)
j
Comparing this with (7.3.6) one observes that dH and dH become identi-
cal if
and so
dH/dt = (dqj /dt dpj /dt − dqj /dt dpj /dt) = 0. (7.3.10b)
This vanishes for all values of E, simply because H–E is a factor of f. This
is of little import. The physically interesting cases arise when f does not
directly contain H–E, but contains a factor that substitutions convert to
H–E. The situation is then similar to that illustrated in Fig. 5.1.1.
In the general case, to convert a Lie generator, ξk (q)∂/∂qk of point
transformations in position space, Q, to a PB operator that acts in the
phase space PQ, one first associates the position space operator with a
function of q,p by letting
ξk (q)∂/∂qk → ξk (q)(−pk ) = − ξk (q)pk . (7.4.3a)
One then associates this function with a Lie operator in phase space by
letting
− ξk (q) pk → {− ξk (q)pk , ·}, (7.4.3b)
{A, {B, X}} + {X, {A, B}} + {B, {X, A}} = 0. (7.4.4)
{qi , {A, pj }} − {pj , {A, qi }} = {qi , {A, pj }} + {pj , {qi , A}}. (7.4.5c)
October 28, 2010 16:55 9in x 6in b943-ch07
From Jacobi’s relation one sees that adding {A, {pj , qi }} to this gives zero.
As {pj , qi } is a number, {A, {pj , qi }} has the value zero. Consequently
{qi , {A, pj }} + {pj , {qi , A}} = 0, (7.4.5d)
and
{qi , pj } = {qi , pj } = δij . (7.4.5e)
Using the properties of Lie transformation groups established in previ-
ous chapters, one sees that (7.4.5e) holds to all orders in α. Applying the
same argument to {qi , qj } and {pi , pj }, one finds that they vanish. Thus, if
A(q, p) is a C2 function, {A, ·} generates canonical transformations.
We next prove the following relation between the PB’s of C2 functions
and PB operators:
If
{A, B} = C, (7.4.6a)
then
[{A, ·}, {B, ·}]X(q, p) = {C, ·}X(q, p). (7.4.6b)
This is a direct consequence of Jacobi’s relation amongst A,B and X, since
the relation can be rewritten as
{A, {B, X}} − {B, {A, X}} = −{X, {A, B}} = {{A, B}, X}. (7.4.7)
As the lhs is [{A, ·}, {B, ·}]X, and the rhs is {{A, B}, ·}X, the relationship
of (7.4.6) follows.
The reader may prove that if A, B and C are functions that satisfy the
Jacobi relation
{A, {B, C}} + {C, {A, B}} + {B, {C, A}} = 0, (7.4.8a)
then {A, ·}, {B, ·} and {C, ·} satisfy the Jacobi relation
[[{A, ·}, {B, ·}], {C, ·}] + [[{C, ·}, {A, ·}], {B, ·}]
+ [[{B, ·}, {C, ·}], {A, ·}] = 0. (7.4.8b)
If a finite set of C2 functions fi (q,p) satisfies (7.4.8a) and
{fi , fj } = ckij fk , (7.4.8c)
then
[{fi , ·}, {fj , ·}] = ckij {fk , ·}. (7.4.8d)
October 28, 2010 16:55 9in x 6in b943-ch07
1. If r linearly independent operators, {fi , ·}, satisfy (7.4.8d), and the fi are
C2 functions, then the {fi , ·} are elements of a Lie algebra with composi-
tion operation [ , ], and they generate an r-parameter Lie transformation
group.
2. The fi are elements of a Poisson bracket Lie algebra whose composition
operation is { , }, each giving rise to a function exp(αi fi ). The group
which the fi generate is termed a function group.
3. The commutator and PB Lie algebras are formally isomorphic.
H(p, q) − E = 0, (7.5.1)
defines a manifold in a 2n-dimensional phase space, PQ space. For each
value of E it requires 2n − 1 numbers to fix a point in the manifold.
Hamilton’s equations of motion define evolution trajectories on these
(2n − 1)-dimensional hypersurfaces. The manifold defined by (7.5.1) pos-
sesses Euclidean geometric symmetries. Furthermore it possesses symplec-
tic geometric symmetries expressed by a dynamical group. The operations
of the latter leave Hamilton’s equations invariant, and define dynamical
symmetries.
We have seen that each constant of motion u(q, p) determines a Poisson
bracket operator U = {u, ·} that satisfies
0 = UH(q, p)|H=E , (7.5.2a)
and that, if UH can not contain H – E as a factor, then U also satisfies
UH(q, p) = 0. (7.5.2b)
Each U generates a one-parameter Lie group that transforms evolution
trajectories into one another or into themselves. All such transformations
October 28, 2010 16:55 9in x 6in b943-ch07
carry points on the hypersurface of fixed energy into points on the same
hypersurface.
If one knows two constants of the motion, u1 and u2 , then U1 = {u1 , ·},
and U2 = {u2 , ·} satisfy
U1 H(q, p)|H=E = 0, U2 H(q, p)|H=E = 0. (7.5.3)
Then (cf. the exercises) [U1 ,U2 ] satisfies
[U1 , U2 ]H(q, p)|H=E = 0. (7.5.4)
If [U1 , U2 ] is not identically zero, then u3 = {u1 , u2 } is a further constant
of the motion. If a set of r linearly independent constants of motion, uj , is
closed under the PB operation, they generate an r-parameter Lie function
group.
Each u, defines a PB operator U. If a set of r constants of motion,
uj , provides a set of PB operators, Uj , that are linearly independent and
closed under commutation, they generate an r-parameter Lie transforma-
tion group. For the given value of E, the operations of this dynamical group
convert solutions of Hamilton’s equations into solutions. The group is the
classical mechanical analog of a quantum mechanical degeneracy group,
whose operations interconvert wave functions of the same energy.
As illustrated below, knowing the constants of motion, u, of a system
enables one to solve problems involving systems with Hamiltonians H =
H+f(u). Constants of motion and the groups they determine can also imply
a variety of revealing connections between the dynamics of a system and
its geometrical properties in phase space.
Let
s = (s1 , . . . , sn , sn+1, . . . , s2n ) = (q1 , . . . , qn , p1 , . . . , pn ), (7.5.5)
be a vector whose components are conjugate q,p variables in this phase–
space. The components may be, but are not necessarily, Cartesian. Suppose
that s is a vector from the origin to a point ρs in the Hamiltonian hyper-
surface defined by H(p, q) = E. If uk = µk (s) is a constant of motion, then
at ρs each group generator, Uk = {uk , ·}, and infinitesimal parameter, δαk ,
determines an infinitesimal displacement, δsk = δαk Uk s, defined by
s → s + δαk Uk s + O(δα2k ) (7.5.6)
This may be imagined to lie in the hypersurface defined by H(q,p) =E. If
one lets ∆sk = Uk s, the components of ∆sk are just the functions ξkm (s)
in the generator Uk = m ξkm (s) ∂/∂sm :
∆sk = (ξk1 , . . . , ξkn , ξkn+1 , . . . , ξk2n ). (7.5.7)
The finite vector ∆sk is tangent to the hypersurface at ρs .
October 28, 2010 16:55 9in x 6in b943-ch07
Because the generators are linearly independent, the ∆sk are linearly
independent. They may be used to establish a local coordinate system at
ρs . If, for a fixed value of s, the tangent vectors ∆sk associated with each
generator, including s itself, all transform as Cartesian vectors, then the
∆sk may be transferred to the origin of s and used to set up a Cartesian
coordinate system there. Evaluating these vectors at a particular time, t0 ,
provides a set of fixed vectors to which motions of ρs may be referred.
When the constants of motion uj transform as components of a Carte-
sian vector, they can also provide a fixed Cartesian coordinate system.
When the coordinate system provided by these fixed vectors, or by the
tangent vectors, is Cartesian, one is best able to use one’s geometric intu-
ition to guide investigations of the dynamics governed by the equations of
motion.
Harmonic oscillators with Hamiltonian
in the space. If s is a radius vector that stretches from the origin to a point
defined by H = E, then, for the Hamiltonian (7.5.8), all such points satisfy
the relation
The solutions of (7.5.8b,c) depend upon the four initial values of the q’s
and p’s at t0 = 0; q1 (0), p1 (0), q2 (0), and p2 (0). They are
q2
q1
−4 −2 2 4
−1
−2
−3
One lets
u= cm 1 n 1 m 2 n 2 qm 1 n 1 m2 n 2
1 p1 q2 p2 , (7.5.11a)
and requires
cm1 n1 m2 n2 {H, qm n1 m2 n2
1 p1 q2 p2 } = 0.
1
(7.5.11b)
All of the {U, ·}’s necessarily commute with the evolution generator
{−H, ·} = {2u4 , ·} = p1 ∂/∂q1 + p2 ∂/∂q2 − q1 ∂/∂p1 − q2 ∂/∂p2 . (7.5.19)
The group generated by {−H, ·} could, in the abstract, be any one-
parameter Lie group, but as we know, it generates symplectic transfor-
mations, the group is always a realization of Sp(1). In analogy to U1 , U2
and U3 , we defined U4 to be {−H, ·}/2. The group generated by the four
U’s is then Sp(2) × Sp(1). The quadratic Casimir invariant of the Sp(2)
groups is equal to the square of the generator of the Sp(1) group. It is the
symplectic group Sp(2) × Sp(1) that defines dynamical symmetries of the
oscillator.
As noted earlier, the Hamiltonian (7.5.8a) has evident four-dimensional
rotational symmetry. Let us investigate the relationship between its SO(4)
symmetry group and the Sp(2) × Sp(1) symplectic group uncovered to this
point. To this end, we let
Jij = si ∂/∂sj − sj ∂/∂si , (7.5.20)
be the generator of rotations in the i–j plane of the four-space in which the
Hamiltonian is defined. Six of the Jij are linearly independent. Together they
generate rotations of the SO(4) group that leaves invariant the Hamiltonian
(7.5.8a). Inspecting (7.5.16) and (7.5.18) one finds that
U1 = −(J14 + J23 )/2, U2 = −(J24 + J31 )/2, U3 = (J34 + J12 )/2,
(7.5.21a)
and
U4 = (J31 − J24 )/2. (7.5.21b)
Note that each U is composed of two rotation generators that commute.
Equations (7.5.21) allow one to connect the operations of the symplectic
group Sp(2) × Sp(1) generated by the U’s, to operations of the group that
leaves H invariant.
In addition to these U’s, there are two linear combinations of the rota-
tion generators (7.5.20), which annihilate H but do not generate canonical
transformations. These are
V2 = (J14 − J23 )/2, V3 = (J34 − J12 )/2. (7.5.22)
Let us re-label U4 , which equals {−H, ·}/2, and call it V1 . Then
[V1 , V2 ] = −V3 , [V3 , V1 ] = −V2 , [V2 , V3 ] = −V1 . (7.5.23)
Though V2 H = 0 and V3 H = 0, the operators V2 , V3 generate neither
symplectic transformations nor transformations that convert solutions of
October 28, 2010 16:55 9in x 6in b943-ch07
and
Fig. 7.5.2 (a) Radius vector Sa and unit tangent vectors. (b) Tangent vectors
translated to origin of Sa .
contains all the solutions of the equations of motion which have a common
energy.
In the foregoing we have used the Lie generators {uj , ·} to define coor-
dinate systems and motions in the four-dimensional phase space. An alter-
native approach is suggested by the fact that {u3 , ·} is a rotation generator
divided by two. Relations (7.5.8) imply that
the manifold defined by this equation is not a sphere. To sum up: the
constants of motion u1 , u2 , and u3 do not provide a coordinate system
as intuitively useful as that provided by the tangent vectors of the PB
operators they determine. In Chapter 8, we will find that just the reverse
is true in Kepler systems.
t = τ + const (7.6.4e)
and
of particle i by
These, together with p, enable one to express the effects of active Galilei
transformations on a many-particle system.
{H − E, E} = 0 if ∂H/∂t = 0, (7.6.9a)
{H, p} = 0 if ∇q H = 0, (7.6.9b)
{H, L} = 0 if RH = 0, (7.6.9c)
Kx = t∂/∂qx + ∂/∂vx ,
one finds
Ki = −tpi + mi qi , (7.6.10b)
and
u=− Ki . (7.6.10c)
Now, let m = mi , then ρ ≡ mi qi /m is the vector from the origin to
the center of mass of the system. Also let tpi =tP, where P is the total
momentum vector of the system. Then
and
Thus exp(v0 ·U) has the same effect as exp(v0 ·K) on q and v. One finds
that
exp(v0 · U)E = E + v0 · pi + (1/2) mi v0 · v0 , (7.6.11c)
and
exp(v0 · U) pi ·pi /2mi = v0 · pi + (1/2) mi v0 ·v0 . (7.6.11d)
October 28, 2010 16:55 9in x 6in b943-ch07
Consequently, if
H= pi · pi /2mi + v(|qi − qj |), (7.6.12a)
then
U(H − E) = 0. (7.6.12b)
Thus, though observers in systems moving at constant velocity with respect
to each other will assign different values to E and H, they will agree that
H = E.
[U+ , U− ] = 0. (7.7.12)
Define
α = α+ + α− , β = i(α+ − α− ), (7.7.17a)
and
Then
T(α, β) carries out real transformations that may be obtained from the
composition of the two transformations (7.7.13). Evaluating the action of
T(α, β) at t = 0, one finds that it carries out the conversions
p0 → p0 + α, q0 → q0 + β, E0 → E0 + αp0 + βq0 + (α2 β 2 )/2. (7.7.18)
Using T(α, 0) and T(0, β), one is able to independently shift the two
initial values q0 and p0 , so the operations of the group can convert any
solution of the equations of motion into any other solution. Using the defi-
nition of transitivity in Chapter 5, the group is seen to act transitively on
the space of solutions.
The generators Uα and Uβ satisfy the commutation relations
[{H, ·}, Uα ] = Uβ , [{H, ·}, Uβ ] = Uα , [Uα , Uβ ] = 0. (7.7.19)
The corresponding group, denoted N(3), is called the Heisenberg group.
To apply the results of this section to the oscillator of Sec. 7.6, with
H = H1 + H2 , as defined in (7.5.8a), one uses operators T(α1 , β1 ) and
T(α2 ,β2 ) to carry out the operations of (7.7.19) in the subspaces of p1 ,
q1 , E1 and p2 , q2 , E2 . This enables one to independently change each of
the four initial values p1 (0), q1 (0), p2 (0), q2 (0), and thereby change E1 ,
E2 and E. The operations of the two Heisenberg groups carried out by
T(α1 ,β1 ),T(α2 ,β2 ), combined with the operations of the degeneracy group,
produce a subgroup of Sp(3) that acts transitively on the space of solutions
of the oscillator equations of motion (7.5.8b,c).
For classical harmonic oscillators, knowing a dynamical group that acts
transitively is of little practical value. For more complex systems, knowing
analogous groups can be very revealing, and of great utility — even if the
required group generators are only approximately known. For this reason,
as well as conceptual reasons previously set forth, we will often be working
with Hamiltonian dynamics and symmetries in PQET space, or with their
quantum mechanical analogs.
Subsequent chapters will also make much use of the PB operators
defined in this chapter, the commutator Lie algebras they define, and the
corresponding Poisson bracket Lie algebras. Chapter 9 establishes quantum
mechanical versions of a number of results contained in this chapter.
the group Sp(2n,r). The following Lie generators provide a basis for the Lie
algebra of Sp(2n,r) when i and j run from 1 to n:
{q2i /2, ·} = qi ∂/∂pi , {p2i /2, ·} = −pi ∂/∂qi ,
{qi pi , ·} = pi ∂/∂pi − qi ∂/∂qi , {qi pj , ·} = pj ∂/∂pi − qi ∂/∂qj ,
(7.8.1)
{qj pi , ·} = pi ∂/∂pj − qj ∂/∂qi , {qi qj , ·} = qi ∂/∂pj + qj ∂/∂pi ,
{pi pj , ·} = −(pi ∂/∂qj + pj ∂/∂qi ).
The commutator and PB Lie algebras obtained from the functions above
are isomorphic. Table 7.1 below gives Poisson bracket relations that are
sufficient to determine the Lie algebras of both types of symplectic groups.
Sp(2n,r) has n(2n + 1) generators and possesses the following subgroup
chains:2
Sp(2n, r) ⊃ U(n) ⊃ SU(n) ⊃ SO(n),
(7.8.2)
Sp(2n, r) ⊃ SL(n, r) × R ⊃ SL(n, r) ⊃ SO(n).
Writing 2n = 2a + 2b, with a and b integers, one also has
Sp(2a + 2b, r) ⊃ U(a, b) ⊃ SU(a, b) ⊃ SO(a, b),
(7.8.3)
Sp(2a + 2b, r) ⊃ Sp(2a, r) × Sp(2b, r).
Numerous group–subgroup chains of SO(n) and SO(a,b) are tabulated in
Ref. 2.
The group N(3) of the previous section is a symplectic group though the
functions determining that its PB generators are not quadratic in p,q,E,t.
Nonlinear transformations can convert the transformations of an
Sp(2n,r) group that acts linearly in phase space into those of Sp(2n,r)
groups that act nonlinearly. This enables Sp(2n,r) to take on a particu-
larly important role in Hamiltonian dynamics. An example of a nonlinear
realization of Sp(2n,r) will be found in Chapter 8.
the transformations are said to belong to an infinite group. If the U(α) are
Poisson bracket operators, the resulting group is a symplectic one. As an
example of this generalization, let u(α) = p exp(αq). Then
and
To investigate the evolution of the kinetic energy under the action of forces,
one considers dT/dt. As
dT/dt = mi vi dvi /dt, (7.A.3b)
where dW is the element of work done by all the forces during the displace-
ments dx.
Lagrange’s equations are obtained by determining the way in which
relations (7.A.3b–d) are altered when one replaces Cartesian coordinates
and velocities by generalized coordinates and their time derivatives.
October 28, 2010 16:55 9in x 6in b943-ch07
where
Fj = fi ∂Xi /∂qj = fi dxi /dqj . (7.A.4c)
i i
Here, and hereafter, each of the generalized coordinates qj are considered
independently variable, so each Fj becomes the local component of the force
that acts in the direction of dqj .
The connection between the Fj and the fi is put to use in converting
the left-hand side of (7.A.2) to an expression analogous to (7.A.3c), one
in which the qj and their time derivatives replace the xi and their time
derivatives. To begin, for each value of i in Newton’s equations (7.A.2),
multiply both sides by dxi /dqj , to obtain
mi (dvi /dt)dxi /dqj = fi dxi /dqj . (7.A.5)
The term multiplying mi on the left-hand side of this may be re-expressed
as
d/dt(vi dxi /dqj ) − vi (d/dt)dxi /dqj . (7.A.6a)
Now (cf. the exercises) it can be shown that
dxi /dqj = (dxi /dt)/(dqj /dt). (7.A.6b)
On writing qj for dqj /dt, (7A.6a) can thus be re-expressed as
(d/dt)(vi dvi /dqj ) − vi dvi /dqj , (7.A.6c)
and as
(d/dt)(d(vi2 /2)/dqj ) − (d(vi2 /2)/dqj ). (7.A.6d)
Using (7.A.6d) to replace the coefficient of mi in the left-hand side of
(7.A.5), Eq. (7.A.5) becomes
d(vi2 /2) 2
mi (d/dt) − d(vi /2)/dqj = fi dxi /dqj . (7.A.7)
dqj
October 28, 2010 16:55 9in x 6in b943-ch07
Thus,
(d/dt)∂T/∂qj − ∂T/∂qj = Fj . (7.A.8b)
These equations, Lagrange’s equations of the first kind, are valid for all
coordinate systems that can be defined by Eqs. (7.A.1). If the forces are
derivable from a potential, V(q), then
Fj = −∂V(q)/∂qj , (7.A.9a)
and, on defining
L(q, q ) = T(q, q ) − V(q), (7.A.9b)
(7.A.8b) becomes
d(∂L/∂qj )/dt − ∂L/∂qj = 0. (7.A.9c)
These are Lagrange’s equations of the second kind.
To utilize Lagrange’s equations it is necessary to express T as a function
of the q’s and their time derivatives. Using (7.A.1) one finds
T = (1/2) gjk (q)qj qk , (7.A.10)
j k
with
gjk (q) = mi ∂Xi /∂qj ∂Xi /∂qk .
i
The variables,
pj ≡ ∂T/∂qj = ∂L/∂qj , (7.A.11)
are the canonical momenta of the system.
We next show that the pj and qj are, in fact, members of canonically
conjugate sets of variables. Let the Cartesian positions and momenta be
denoted by χk and πk , with πk = mk vk . Using Poisson brackets defined
using derivatives with respect to χk and πk , one has
{qi , pj } = {qi , ∂T/∂qj}, (7.A.12a)
= {qi , ∂ (mk vk2 /2)/∂qj } = {qi , mk vk dvk /dqj }. (7.A.12b)
October 28, 2010 16:55 9in x 6in b943-ch07
Hence
{qi , pj } = {qi , πk dvk /dqj }. (7.A.12c)
Consequently,
{qi , pj } = {qi , πk dχk /dqj }. (7.A.12e)
Because dχk /dqj can be expressed as a function of the χ’s only, for the
Poisson bracket on the right-hand side of (7.A.12e) one has
{qi , πk dχk /dqj } = {qi , πk }dχk /dqj
k k
= ( ∂qi /∂χm ∂πk /∂πm ) dχk /dqj . (7.A.12f)
k m
(7.B.1) becomes
τ2
S ≡ L (mj , qj , qj /t )dτ. (7.B.8b)
τ1
Exercises
1. In matrix notation, (7.2.3) may be expressed as ds = M dsT , with
M composed from n x n submatrices A, B, C and E, with the upper
left sub-matrix, A, having elements aij , the upper right submatrix, B,
having elements bij , and the lower left and right sub-matrices, C, E
having elements cij , eij respectively. What relations between the matrix
elements are required by Eqs. (7.2.8)?
2. a) Produce plots, analogous to Fig. 7.5.1, of trajectories in a common
q–q plane and a common p–p plane. b) Plot motions implied by (7.5.10)
in each of the six two-planes.
3. Show that u1 ,u2 ,u3 satisfy Eq. (7.5.11b).
4. Expand (7.5.11b), collect terms that multiply the same powers of each
q and p, and determine the relations that the coefficients c must satisfy
when H is given by (7.5.8a).
5. Derive (7.2.11) from (7.2.9).
6. Derive (7.7.5a) from (7.7.4).
References
[1] J.-M. Levy-Leblond, Galilei Group and Galilei Invariance, in Group Theory
and it Applications, II, ed. E. M. Loebl (Academic Press, N.Y., 1971).
[2] R. Gilmore, Lie Groups, Lie Algebras, and Some of Their Applications
(Wiley-Interscience, N.Y., 1974).
[3] Sir W. Thomson and P. G. Tait, Treatise on Natural Philosophy (Cambridge
University Press, Cambridge, 1879), p. 302.
[4] C. Lanczos, The Variational Principles of Mechanics (University Toronto
Press, Toronto, 1949), p. 133.
October 28, 2010 16:55 9in x 6in b943-ch08
CHAPTER 8
247
October 28, 2010 16:55 9in x 6in b943-ch08
to the mass of the sun, κ is nearly the same for all, so one may in good
approximation set c = 1. For such systems, a2 /b3 = 1, and the time and dis-
tance scalings must produce the approximate relation (T /T)2 = (R /R)3 .
As the planetary orbits are all ellipses of small eccentricity, they may be
approximated by circles of radii R and R . Letting T and T be the period
of time it takes for two planets with orbital radii R and R , to traverse such
a circle, one obtains Kepler’s law of periodic times:
(T /T)2 = (R /R)3 . (8.1.2)
A discussion of the accuracy of this relation will be found in Ref. 1 at the
end of the chapter. The exact law will be derived below.
We next consider time-independent properties of planetary orbits that
are consequences of two different vector constants of motion of Keplerian
systems. Subsequent sections of the chapter will investigate the way in
which these constants of motion are related to dynamical symmetries.
Taking the cross product of (8.1.1) with r, one obtains
r × dv/dt = −f(r) r × r = 0. (8.1.3)
Now
d(r × v)/dt = (v × v) + (r × dv/dt) = (r × dv/dt). (8.1.4a)
It then follows from (8.1.3) that
d(r × v)/dt = 0. (8.1.4b)
Hence,
h ≡ r×v (8.1.4c)
is a vector that remains constant as r and v evolve. It has a length, h,
given by
h2 = (h · h) = r2 v2 − (r · v)2 . (8.1.4d)
From (8.1.4c) it follows that
r × dr = h dt, (8.1.5a)
so r and dr lie in a fixed plane perpendicular to h. This plane contains
the center
of mass of the planet and that of the sun. In every time interval
∆t = dt, the area swept out by the radius vector r has the same value,
(r(t) × dr(t))/2 = h∆t/2. (8.1.5b)
This implies Kepler’s law of equal areas: a line from the sun to the planet
sweeps out equal areas in equal times. This law may also be considered to
October 28, 2010 16:55 9in x 6in b943-ch08
arise from the constancy of the angular momentum vector of the planet,
L = r × p, p = µv. (8.1.6)
(a × b) × c = (c · a)b − a(c · b)
to expand
h × r = (r × dr/dt) × r,
one obtains
Consequently
b ≡ (h × v) + κr/r, (8.1.9a)
0.4
0.2
b
0.5 1.0 1.5 2.0
− 0.2
− 0.4
(a)
2 3
2
1
1
b b
1 2 3 4 1 2 3
−1
−1
−2
−2 −3
(b) (c)
Fig. 8.1.1 Runge–Lenz vectors for elliptical, parabolic, and hyperbolic orbits,
(a) |b| = 0.8, |b| = 0.71, (b) |b| = 1, 0, |b| = 0.71, (c) |b| = 1.2, |b| = 0.71.
All the orbits available to bodies moving under an inverse square law
of force can be determined by considering the scalar product of r with b.6
One has
Letting θ be the angle between b and r, one may express this equation as
(h2 /κ)
r= (8.1.11c)
(1 − (b/κ) cos(θ)).
The dynamical quantities h = |h| and b = |b| are then related to the
geometrical variables Λa and Λb by
Using (8.1.12a) to eliminate h from (8.1.7), one finds that the period T
satisfies
E = −µ/2Λa.. (8.1.16a)
For ε > 1,
E = µ/2Λa. . (8.1.16b)
As these energies do not depend upon the length of the minor axis, for given
values of µ and a, varying Λb produces a family of degenerate orbits —
orbits of the same energy. The dynamical symmetries responsible for this
degeneracy are investigated in the upcoming sections of the chapter.
A development analogous to that of (8.1.10) may be used to obtain the
relation among velocity components that defines the hodograph of Keplerian
motion.6 The component of v on the j axis is vy , so (8.1.10) implies
vc
0.5
va
− 1.0 − 0.5 0.5 1.0
− 0.5
− 1.0
− 1.5
− 2.0
− 2.5
(a)
vc vc
va va
− 1.0 − 0.5 0.5 1.0 − 1.5 − 1.0 − 0.5 0.5 1.0 1.5
− 0.5
− 0.5
− 1.0
− 1.0
− 1.5
− 1.5
− 2.0
− 2.0
− 2.5
− 2.5 − 3.0
(b) (c)
Fig. 8.1.2 Hodographs corresponding to orbits of Figs. 8.1.1, for (a) |b| =
0.8, |h| = 0.71, (b) |b| = 1.0, |h| = 0.7, and (c) |b| = 1.2, |h| = 0.71.
October 28, 2010 16:55 9in x 6in b943-ch08
L = µh, (8.2.4a)
A ≡ −(L × p + q/q) = −b/κ ≡ a/κ. (8.2.4b)
The negative sign has been introduced in the definition of A and a, because
doing so substantially simplifies the discussion in subsequent sections. In
position space, the vectors A and a point from the center of mass toward
perhelion. After expanding the L × p term in A to (q · p)p − qp2 , one can
use (8.2.1d,e) to make the conversions
and
H = −(1/2)/(A2< + L2 ). (8.2.9)
The Lie algebra is of rank 2, and consequently possesses a quartic Casimir
operator, which is (L · A< )2 . Because L is perpendicular to A, for all values
of E, (L · A> )2 vanishes.
The operators exp(αi A<i ) and exp(βi Li ) may be shown to be those of the
group SO(4). They alter the components of A<j , and Lj , j = i, while leaving
(A2< +L2 ) unaltered. Because of this, when E is negative, the equation H = E
October 28, 2010 16:55 9in x 6in b943-ch08
r
− a(− E)
a(E)
Fig. 8.3.1 Relation of position vector to Runge–Lenz vectors a(E), a(−E) when
E = 12 .
and
and define
Then, letting
(4)
p(4) → P± = 2p(4) /(p(4) · p(4) )± , (8.3.10c)
and setting
Here,
thus orthogonal q(4) and p(4) produce orthogonal Q(4) and P(4) .
When the positive signs in (8.3.10) are chosen,
and
(4)
P± = 2p(4) /(p(4) · p(4) )± = (P0± , P± ); (8.3.12b)
with
with
z
z
1.0
1.0
P0 π
0.5
0.5
ρ x'
y q
− 1.0 − 0.5 0.5 1.0 − 1.0 − 0.5 0.5 1.0
ρ
− 0.5
− 0.5
− 1.0
− 1.0
(a) (b)
P0 π
y'
x'
ρ
(c) (d)
Fig. 8.4.1 (a) Projection connecting P0 π and ρ; (b) projection connecting ρ and
q; (c) relation between the ρ and p0 π vectors; (d) S< , the sum of the two vectors
in the previous figure, terminates on the surface of the sphere.
It requires that
The potential function, 1/|q1 |, becomes infinite at the origin, so for the
kinetic energy to remain finite, q1 cannot go to zero, and the sign on the
right-hand side of (8.4.15b) cannot change during the motion. We will
assume that the positive sign remains. Then, on substituting (8.4.15b)
October 28, 2010 16:55 9in x 6in b943-ch08
into (8.4.14b),
ρ< → ((p20< − p21 )/(p21 + p20< ), 2p1 p0< /(p21 + p20< )). (8.4.16)
ρ< (q1 , p1 ) = ((p20< − p21 )(q1 /2), 2p1 p0< /(p21 + p20< )). (8.4.17a)
Then
and
Consequently,
d2 ρ< /dτ 2 = −p20< ρ< , (8.4.18c)
d2 π < /dτ 2 = −p20< π < . (8.4.18d)
If the initial values of ρi< and πi< are ρi< (0) and πi< (0) respectively, then
integration of each component of these vector equation yields
ρi< (τ ) = ρi< (0) cos(p0< τ ) + p0< πi< (0) sin(p0< τ ),
(8.4.19)
πi< (τ ) = πi< (0) cos(p0< τ ) − (ρi< (0)/p0< ) sin(p0< τ ).
These describe uniform circular motions of the vectors with period 2π/p0< .
The relation between τ and t is considered in the exercises.
Because a(0) = a(E) + a(−E), a Lie algebra that contains a(E) and a(−E)
also contains a(0). The same can be said of the corresponding PB operators.
The evolution of the vectors ρ0 and π 0 is determined by the equations
Consequently,
d2 ρ0 /dτ 2 = 0, (8.5.10c)
2 2
d π 0 , /dτ = 0. (8.5.10d)
and when E = 0,
For each range of E, the points that represent the system evolve in phase
space on a three-dimensional manifold determined by these equations.
These manifolds have everywhere-local geometric symmetry determined by
the Lie generators of the corresponding degeneracy group, all of which are
first-order differential operators that are of degree one or zero in the vari-
ables ρj and πj .
October 28, 2010 16:55 9in x 6in b943-ch08
= p−1 2 2 2
0 (q(q p /2 − q · p p) − p0 q/2) = A< (−p0 /2).
(8.6.3a)
Similarly
S(−α)a(1/2) = A> (p20 /2). (8.6.3b)
(Because a(0) = (a(−1/2)+a(1/2))/2, it is not necessary to treat the E = 0
case separately.) The general effect of S(α) on a(−1/2) and a(1/2) is
S(α){a(1/2), ·} = cosh(α){a(1/2), ·} + sinh(α){a(−1/2), ·}, (8.6.4a)
October 28, 2010 16:55 9in x 6in b943-ch08
and
S(α){a(−1/2), ·} = cosh(α){a(−1/2), ·} + sinh(α){a(1/2), ·}. (8.6.4b)
These relations enable one to use D and the components of a(−1/2) and
a(1/2) to provide PB operators that constitute a basis for the Lie algebras
and groups that act in the E < 0, 0− ≤ E ≤ 0+ , and E > 0, subspaces of
PQET space.11 One may set
D = K45 ≡ K54 , (8.6.5a)
and for i < j, j = 1, 2, 3, define the operators
Jij = {Lij , ·} = −Jji , (8.6.5b)
Jj4 = {aj(−1/2), ·} ≡ −J4j , (8.6.5c)
Kj5 = {aj(1/2), ·} ≡ K5j . (8.6.5d)
These Jij , Jj4 satisfy the commutation relations of the SO(4) bound-state
degeneracy group, and the Jij , Kj5 satisfy the commutation relations of the
SO(3,1), E > 0, degeneracy group. For E = 0, the generators of the E(3)
degeneracy group may be taken to be Jij and (Jj4 + Kj5 )/2. Together, the
Jij , Jj4 , Kj5 , and K45 generate an SO(4,1) invariance group of W(E) and
the equations of motion.
Exercises:
1. Is q · p invariant under the rotations generated by qj ∂/∂pj − qj ∂/∂pj , or
transformations generated by qj ∂/∂pj + qj ∂/∂pj ?
2. Let the elliptical orbit of a planet lying in the q1 –q2 plane have its
Hamilton–Runge–Lenz vector b on the q1 axis. Determine the effect on
q of infinitesimal transformations with the PB generators determined by
Lij , Ai . Average the effect over a circular orbit. What do these infinitesi-
mal transformations do to p? Average their effect over the corresponding
hodograph.
3. Let s = p0 τ , and use the equations of motion (8.3.3) to determine the
relation between physical time, t, and s. Set both t and τ equal to zero
when a planet or comet is closest to the sun. For the planetary case verify
that your periodic time agrees with the exact form of Kepler’s Law.
4. Determine the stereographic projection of Keplerian hodographs onto a
unit sphere (cf. Am. J. Phys. 33 (1965) 570). Generalize your results by
considering rotations of the plane of the hodograph, and compare with
the work of Fock in Ref. 10.
October 28, 2010 16:55 9in x 6in b943-ch08
References
[1] B. J. Cantwell, Introduction to Symmetry Analysis (Cambridge University
Press, Cambridge, 2002).
[2] E. A. Milne, Vectorial Mechanics (Methuen, London, Interscience, N.Y.,
1948), pp. 236–7.
[3] H. Goldstein, Am. J. Phys. 43 (1975) 737.
[4] E. B. Wilson, Vector Analysis Founded upon the Lectures of J. Willard Gibbs
(Scribners, N.Y., 1901).
[5] W. Pauli, Z. Phys. 36 336.
[6] E. A. Milne, ibid. p. 238.
[7] A. L. Fetter, J. D. Walecka, Theoretical Mechanics of Particles and Continua
(McGraw-Hill, N.Y, 1980); H. Goldstein, Classical Mechanics (Addison-
Wesley, Reading, Mass, 1980).
[8] H. Poincare, New Methods of Celestial Mechanics I, NASA Technical Trans-
lation TTF- 450 (Washington, D.C., 1967) pp. 14, 24.
[9] C. Caratheodory, Calculus of Variations and Partial Differential Equations
of the First Order, translated by R. B. Dean, J. J. Brandstatter (Holden-Day,
San Francisco, 1965).
[10] V. Fock, Z. Phys. 98 (1935) 145.
[11] The Lie generators in equations (8.6.5) are classical mechanical versions of
operators introduced by M. Bednar in Ann. Phys. (N.Y.) 75 (1973) 305;
see also A. O. Barut and H. Kleinert, Phys. Rev. 156 (1967) 1541, and 157
(1967) 1180.
[12] V. Bargmann, Z. Physik 99 (1936) 576.
[13] A. O. Barut, Dynamical Groups and Generalized Symmetries in Quantum
Theory (University of Canterbury Press, Christchurch, N. Z. 1972).
[14] M. Bander, C. Itzykson, Rev. Mod. Phys. 38 (1966) 330, 346.
[15] H. V. Macintosh, Symmetry and Degeneracy, Group Theory and its Appli-
cations, Vol. 2, ed. E. M. Loebl (Academic Press, N.Y., 1971).
[16] C. E. Wulfman, J. Phys. A 42 (2009) 185301.
This page is intentionally left blank
October 28, 2010 16:55 9in x 6in b943-ch09
CHAPTER 9
275
October 28, 2010 16:55 9in x 6in b943-ch09
Here dv is the volume element in position space, and ψk∗ is the complex
conjugate of the Schrödinger wave function ψk , which is in general a func-
tion of t as well as the position variables x. Dirac proposed the bra and
ket notation that is commonly used to represent these integrals, and matrix
elements in quantum theory. Because Dirac’s notation can lead to confusion
when one is dealing with operators that are not self-adjoint, we prefer to
use the mathematician’s notation (f, g), or f, g, to denote a general scalar
product of f and g. Using it, one would express (9.2.1) as
Mab = (ψa∗ , µ ψb ), or Mab = ψa∗ , µ ψb . (9.2.2)
October 28, 2010 16:55 9in x 6in b943-ch09
The reader may wish to establish that if µ is self-adjoint and ψa and ψb are
eigenfunctions of it with eigenvalues, ma , mb , then the eigenvalues must be
real numbers.
Because adjoints of Schrödinger operators are defined by the scalar
products, (9.2.1), in which there is no integration over t, in Schrödinger
mechanics the adjoint of i∂/∂t and of t, is not defined. However, in
time dependent-Schrödinger mechanics, energies are eigenvalues of the self-
adjoint operator H(q, pop ) and the operator i∂/∂t. If
H(q, pop )φk (q) = Ek φk (q), (9.2.3a)
then
ψk (q, t) = exp(−iEk t/)φk (q) (9.2.3b)
is the time-dependent Schrödinger function that is a solution of (9.2.3a)
and
H(q, pop )ψk (q, t) = i ∂ψk (q, t)/∂t. (9.2.4a)
Though the adjoint of i ∂/∂t is not defined, it is evident from (9.2.3) and
(9.2.4a) that ψk (q, t)∗ satisfies
H(q, pop )ψk (q, t)∗ = (i ∂/∂t)∗ ψk (q, t)∗ = −i ∂ψk (q, t)/∂t∗ . (9.2.4b)
The adjoint of the product of two operators, AB, is (AB)∗T = B† A† ; so
the adjoint of the commutator of two operators A and B is
[A, B]† = [B† , A† ]. (9.2.5a)
If A and B are self-adjoint, their commutator is skew-adjoint, that is
[A, B]† = [B, A] = −[A, B]. (9.2.5b)
If A and B are self-adjoint, and c is a real constant, either of the equations
C = ic[A, B] or C = −ic[A, B] (9.2.6)
must produce from A and B a self-adjoint operator C.
Born and Jordan observed that in Cartesian coordinates the commuta-
tion relations
[qi , −i ∂/∂qj ] = i δij , [−i ∂/∂qi, −i ∂/∂qj] = 0,
[qi , qj ] = 0,
(9.2.7a)
are in one-to-one correspondence with the Poisson bracket relations
{qi , pj } = δij , {qi , qj } = 0, {pi , pj } = 0. (9.2.7b)
October 28, 2010 16:55 9in x 6in b943-ch09
are not the spaces one needs to relate classical and quantum mechanical
symmetries in which the group parameters a,b can vary.
There is another way of viewing group-theoretic consequences of the
Born–Jordan correspondence. Let g(q) be an analytic function and consider
the expression
(−i/ )[exp(iapop ), g(q)] = (−i/ )[exp(a ∂/∂q), g(q)]
= (−i/ )(g(q + a ) − g(q)) exp(a ∂/∂q).
(9.2.13a)
Its Poisson bracket analog is
{exp(ap), g(q)} = −(∂ exp(ap)/∂p) ∂g(q)/∂q
(9.2.13b)
= −a ∂g(q)/∂q exp(ap).
A Maclaurin series expansion of the term g(q+ a)− g(q) in (9.2.12a) yields
−i(a∂g(q)/∂q + (a2 )(∂ 2 g(q)/∂q2 )/2 + · · · ) exp(a∂/∂q). (9.2.13c)
Comparing this with the corresponding term, −a ∂g(q)/∂q, in (9.2.12b) one
sees that relations (9.2.12a) and (9.2.12b) are in one-to-one correspondence
only to first order in .
Both these observations indicate that a change in viewpoint is required if
one is to understand why it is that some, but not all, classical and quantum
mechanical systems have identical local symmetry groups.
Intrinsic local symmetries of Schrödinger equations are determined
by Lie algebras whose generators are functions of q’s and pop ’s, t and
−Eop = −i ∂/∂t. On the other hand, as we have seen, the local dynam-
ical symmetries of classical Hamiltonian systems are determined by Lie
algebras whose generators are Poisson bracket operators. In Schrödinger
mechanics
[exp(iapkop ), g(q)] = [exp(a ∂/∂qk), g(q)]. (9.2.13)
In classical mechanics
[exp(−a{pk , · }), g(q)] = [exp(a∂/∂qk ), g(q)]. (9.2.14)
The group operators in these expressions are in one-to-one correspon-
dence to all orders in As a consequence, correspondences between the
group actions of Poisson bracket operators in phase space and the group
actions of quantum mechanical operators in position or momentum sub-
spaces of phase space are valid to all orders in .
It will be shown in Secs. 9.3 and 9.4, below, that these observations lead
to a change in viewpoint that does much to clarify the relation between
October 28, 2010 16:55 9in x 6in b943-ch09
or
and
{a, · } = 0. (9.4.3)
Thus
[{Gm (q, p), · }, {Gn (q, p), · }] = Σk ckmn {Gk (q, p), · }. (9.4.6)
It was also noted that the apparent isomorphism between (9.4.5) and (9.4.6)
was misleading if any Gk (q,p) could be a constant. If this occurs, a {Gk , · }
will vanish, and the Lie algebras determined by (9.4.5) and (9.4.6) will not
be isomorphic, the vanishing of a {Gk , ·} being equivalent to the vanishing
of every ckmn with the same value of k.
This is illustrated by the Lie algebra that is put in correspondence with
the Heisenberg algebras of (9.2.10) by using (9.4.1). In this algebra
and
[qj , pj ] = i . (9.4.8)
When these conditions are satisfied one may say that Eqs. (9.4.10) define
a faithful Lie-algebraic extension of the correspondence principle. In this
case, the Lie algebra of quantum mechanical operators and the Lie alge-
bra of PB operators define locally isomorphic dynamical groups. Heisen-
berg’s uncertainty principle, a consequence of the commutation relation
October 28, 2010 16:55 9in x 6in b943-ch09
[qop , pop ] = i , does not affect this correspondence. The SO(4,2) local Lie
groups of quantum-mechanical and classical Kepler systems are an exam-
ple of this isomorphism which may be said to exist despite the uncertainty
principle.
Harmonic oscillators provide a contrasting example. The Lie algebra of
their Dirac energy shift operators is a Heisenberg algebra. Because of this,
there is no one-to-one correspondence between the consequent quantum
mechanical and classical dynamical groups of harmonic oscillators.
When a Poisson bracket Lie algebra contains a pair of operators −∂/∂qj ,
and ∂/∂pj , they generate translations in the q and p subspaces of classical
phase space. Thus translation invariance in phase space produces systems
with quantum mechanical and classical symmetries that need not be iso-
morphic. In Schrödinger mechanics, −∂/∂qj innocuously becomes −i∂/∂qj .
On the other hand, ∂/∂pj becomes −iqj and generates gauge transforma-
tions which multiply ψ(q) by exp(iaqj ).
To determine whether corresponding quantal and classical systems with
locally isomorphic symmetry groups have globally isomorphic groups, one
must determine the interval over which their group parameters can range
without producing an identity operation. In the classical mechanical cases
the group moves a point in a phase space, so one determines the range which
the group parameters can span before restoring points in the phase space to
their initial positions. In Schrödinger quantum mechanics the wave function
Ψ (q, t) defines points with coordinates (ψ, q, t), with ψ having the value
Ψ(q, t). One determines the range which group parameters can span before
restoring q,t to their original values, and ψ to its original value. When
the correspondence between classical and quantum mechanical global Lie
groups exists, intrinsic quantum mechanical global symmetries correspond
to global geometrical symmetries in phase space.
An example is provided by the correspondence between quantum angu-
lar momentum operations and rotational symmetry. For L12 , the component
of the angular momentum operator on the q3 axis, one has
and
Note that
Allowing the operators exp(iαjk Ljk ) to act upon the spherical harmonics
Ylm (θ, φ) in order to reproduce the identity operation, the group parameters
must have the same ranges that are required of exp(−αjk Rjk ) when it acts in
phase space — the group in both cases is SO(3).
when
(H − i∂/∂t)ψ(x, t) = 0.
when
(H − E)ψ(q) = 0. (9.5.3b)
October 28, 2010 16:55 9in x 6in b943-ch09
Also, if it is, E may, or may not, equal E. We shall, for example, see in
Sec. 9.6 that if
H = −(1/2)∂ 2/∂x2 , (9.5.4a)
then,
Gop = −i(x∂/∂x + 2t∂/∂t) (9.5.4b)
satisfies (9.5.2). It converts solutions of the time-dependent equation into
solutions, but does not convert an eigenfunction of H into an eigenfunction.
Similarly, if
H = −(1/2)∂ 2/∂x2 + (1/2)kx2 , (9.5.5a)
the shift operator
exp(−ikt)(∂/∂x − ikx) (9.5.5b)
satisfies (9.5.2), and converts eigenfunctions of energy E into eigenfunctions
with E not equal to E.
If G(q; p) = G(x, t; px , −E) is an analytic function of the components
of p in a region about p = 0, we may expand it in a Maclaurin series in
the components of p, E. We suppose this is the case, and let
G(q, p) = Σgijk··· (p1 )i (p2 )j (p3 )k · · · , (9.5.6)
where the g’s are functions of the q’s or constants. We use the form (9.5.6) in
determining Lie generators, then subsequently convert them to self-adjoint
operators A(G + G† )/2 or their symmetrized counterparts.
Before proceeding further let us make a change in notation to conform
to usage in much literature on invariance transformations of Schrödinger
equations. Let us express the Schrödinger operator corresponding to the
power series expansion (9.5.6) as
Gop = Q(q, ∂/∂q) = Σgijk···(∂/∂q1 )i (∂/∂q2 )j (∂/∂q3 )k · · · , (9.5.7)
ijk···
the g being functions of the position variables, q = x, and/or t. The
operators may have a term with all indices equal to zero, which contains
no derivative operator, as well as terms containing derivatives of arbitrary
order. If the coefficient of a term linear in a ∂/∂x is real, one may wish
to multiply Q by i. Call the result Qc . Then, if Q generates invariance
transformations of a Schrödinger equation, (Qc ± Q†c )/2 can be made self-
adjoint or skew-adjoint. Doing this for each of the generators Qk of interest,
one can construct a set of operators with definite adjointness and definite
commutation relations.
October 28, 2010 16:55 9in x 6in b943-ch09
Among the transformations which leave this equation invariant, some leave
invariant the corresponding time-independent equation
We will seek invariance generators of (9.6.1) that have the simple form
for all y satisfying (9.6.1a) and all of its relevant differential consequences.
Given the assumed form of Q, (9.6.3) contains a third-order derivative,
October 28, 2010 16:55 9in x 6in b943-ch09
∂Q/∂t = Qt = q00 10 01
t (x, t) + qt (x, t)∂/∂x + qt (x, t)∂/∂t, (9.6.4)
∂ 2 Q/∂x2 = Qxx = q00 10 01
xx (x, t) + qxx (x, t)∂/∂x + qxx (x, t)∂/∂t.
+ 2i(q00 10 01
t y + qt y x + q t y t ) = 0 (9.6.7)
We choose y, yx , yt , ytx to be independently variables, and use (9.6.6b) to
eliminate yxx from (9.6.7). This converts (9.6.7) to
(q00 00 10 00 10
xx + 2i qt )y + (qxx + 2 qx + 2i qt )yx
+ q01 01 10 01
xx yt + 2i(qt − 2 qx )yt + 2qx yxt = 0. (9.6.8)
As y, yx , yt , ytx can be assigned arbitrary values, for (9.6.8) to hold, the
coefficients of these functions in the equation must themselves vanish. We
thus obtain the following determining equations for Q:
q00 00
xx + 2iqt = 0, (9.6.9a)
q10 00 10
xx + 2qx + 2iqt = 0, (9.6.9b)
q01 01 10
xx + 2i(qt − 2qx ) = 0, (9.6.9c)
q01
x = 0. (9.6.9d)
October 28, 2010 16:55 9in x 6in b943-ch09
The last equation is the simplest, so we start out with it and work our way
upward. For q01
x to vanish, q
01
can not be a function of x, so one may set
This causes the first term in (9.6.9c) to vanish — with the consequence that
q10 01
x = (1/2)qt = (1/2)∂A(t)/∂t. (9.6.11)
Hence
q10 2 2
t = (x/2)∂ A(t)/∂t + ∂B(t)/∂t. (9.6.13)
q00 10 2 2
x = −iqt = −i(x/2)∂ A(t)/∂t − i∂B(t)/∂t, (9.6.14)
so
∂ 3 A(t)/∂t3 = 0, (9.6.17a)
2 2
∂ B(t)/∂t = 0, (9.6.17b)
and
so
Inserting these results in (9.6.2), that is in Q = q00 + q10 ∂/∂x + q01 ∂/∂t,
yields
Collecting terms that are multiplied by the same cj , one may express this as
Q = c1 Q 1 + c 2 Q 2 + c 3 Q 3 + c 4 Q 4 + c 5 Q 5 + c 6 Q 6 . (9.6.21)
Q6 = −i → 1. (9.6.23f)
The commutation relations of the Q are given in Table 9.1 below.
Q1 , Q2 , Q3 , Q4 ,
and Q5 are the quantal analogs of classical operators which
generate a five-parameter subgroup of the projective group in the two vari-
ables x, t. On inspecting the table one observes that Q4 and Q5 are members
of a Heisenberg subalgebra, so the classical and quantal subgroups are not
isomorphic. In this subgroup, Q6 generates transformations which multiply
a wave function by the phase factor eiα . The generators Q1 , Q2 , Q3 , are
members of a Lie algebra that has no Abelian or Heisenberg subalgebra.
[QA , QB ]
One has
and
Using the Schrödinger scalar product, these yk are orthonormal in the gen-
eralized sense of Dirac’s delta function:
∞
yk , yk = dx yk∗ (x, t)yk (x, t) = δ(k − k). (9.7.3)
−∞
of the Galilei group. The previous statement allows exceptions. For exam-
ple, the time and space translation invariant eigenfunction y0 (x, t) = (1)
satisfies Q4 y0 = 0.
The eigenfunctions of Q4 with eigenvalue λ are of the general form
Φλ (x, t) = N(t) exp((i/t)(x2 /2 + λx)), (9.7.8)
where N(t)√is an arbitrary normalizable function of t. Thus one may choose
N(t) = (1/ 2π) exp(if(t)), with f being an arbitrary function of t. Observers
whose clocks and meter sticks agree at t = 0, will agree on the value of λ if
they are moving with uniform relative velocity.
Now, because Q1 commutes with Q4 , one can replace N(t) with a func-
tion of t which makes Φλ (x,t) an eigenfuntion of Q1 . Letting C be a con-
stant, one finds the resulting functions
Φnl (x, t) = C t−1/2 exp(−in/t) exp((i/t)(x2 /2 + λx))
(9.7.9)
= C t−1/2 exp((i/t)(x2 /2 + λx − n)).
One finds that these functions satisfy Schrödinger’s equation if n = −λ2 /2.
The resulting functions,
yλ (x, t) = C t−1/2 exp((i/t)(x + λ)2 /2), (9.7.9)
also satisfy the relations
Q1 yλ (x, t) = −(λ2 /2)yλ(x, t), Q4 yλ (x, t) = λyλ (x, t). (9.7.10)
The probability function is
yλ (x, t)∗ yλ (x, t) = C∗ C/t, (9.7.11)
so
yλ , yλ = t−1 dxC∗ C = t−1 C∗ Cδ(λ − λ).
Consequently, yλ is not normalizable for all times. The t−1/2 factor ensures
that the resulting wave functions can only be normalized for fixed nonzero
values of t. These observations lead to the conclusions that:
a. Time-dependent Schrödinger equations may be left invariant by trans-
formations that do not leave the Schrödinger scalar product invariant.
b. Time-dependent Schrödinger equations may have solutions, yn (x, t), with
constents of motion whose eigenvalues, n, do not change with time, but
the probability of finding a system in the eigenstate yn (x,t) can vary with
time, even though the eigenstate remains the same. This can be true
even when the eigenvalues correspond to values of classical constants of
motion.
October 28, 2010 16:55 9in x 6in b943-ch09
√
Setting ω = k, define
Q3 → bt = (2ω)−1/2 exp(iωt)(ωx + ∂/∂x),
(9.8.4)
Q4 → b+
t = (ω)
−1/2
exp(−iωt)(ωx − ∂/∂x).
Here b+t is the adjoint of bt , and the operators are those introduced by
Dirac. They obey the following commutation relations:
[i∂/∂t, bt ] = −ωbt , [i∂/∂t, b+ +
t ] = ω bt ,
(9.8.5a–c)
[bt , b+
t ] = 1.
algebraic properties of the system establish that its energy spectrum has a
bottom level.
An integration gives the result that the normalized wave function of this
bottom level is
(b+ n
t ) y0 = yn (9.8.14)
yn = (1/(n!)1/2 )(b+ n
t ) y0 . (9.8.16)
The raising and lowering operators act upon the normalized functions as
follows:
b+
t yn = (n + 1)
1/2
yn+1 ,
(9.8.18)
bt yn = n 1/2 yn−1 .
corresponding to bt and b+
t . One finds
[Bt , B+
t ] = 0. (9.8.20)
The quantal group and the group generated by the PB operators are not
isomorphic. The group generated by bt , b+
t and 1 has no isomorphic ana-
log in the phase space PQET. In contrast, the Lie algebra of the quan-
tum mechanical double-jump operators Qm = (bt )2 , Qp = (b+ 2
t ) , and the
+ +
labeling operator Qpm = bt bt + bt bt is isomorphic to the corresponding
classical Lie algebra of PB operators.
Hence
Exercises
(1,2)
1. Compute the commutator of QPj with qj and with p2jop .
2. Establish the condition under which the Q1 and Q2 of Sec. 9.6 can be
expressed in terms of products of Q3 , Q4 , Q5 .
October 28, 2010 16:55 9in x 6in b943-ch09
3. Does using the symmetrized self-adjoint form of it2 ∂/∂t in Q1 produce
an operator that leaves invariant the surface defined by Sop y(x, t) = 0
in Eq. (9.6.1a)?
4. Solve the determining Eqs. (9.8.1) when the potential v = kx or k/x2 ,
or simply a constant, k.
5. Using the Bt , B+ t given in Eqs. (9.8.19), determine the finite transfor-
mations of p, q, E, t carried out by the groups generated by (Bt + B+ t )/2
and i(Bt − B+ t )/2.
6. Obtain real linear combinations of the oscillator shift operators and com-
pute their commutation relations. Do the same with the corresponding
Poisson bracket operators. Determine the finite transformations of p, q,
E, t generated by the Poisson bracket operators.
7. Use the scaling properties of the harmonic oscillator Hamiltonian to
prove that
Y(x), Top Y(x) = Y(x), VY(x).
References
[1] W. Heisenberg, Z. Physik, 33 (1925) 879.
[2] M. Born and P. Jordan, Z. Physik 34 (1925) 858. For translations, of this and
the previous reference, see Sources of Quantum Mechanics, ed. B. L. Van Der
Warden (North-Holland, Amsterdam, 1967).
[3] a) H. J. Groenewald, Physica 12 (1946) 405–460.
b) L. Van Hove, Acad Roy. Belgique Bull. Cl. Sci. (5) 37 (1951) 610–620.
c) For a recent mathematical review see, Mark J. Gotay, in Mechanics: From
theory to computation, J. Nonlinear Science (2000) 271–316.
[4] C. Wulfman, J. Phys. Chem. A 102 (1998) 9542-8; J. Phys. A 42 (2009)
185301.
[5] a) E. Noether, Nachr. Konig. Gesell. Wissen. Gottingen, Math.-Phys. p. 23;
English translation: Transport Theory Stat. Phys. I, Vol. 186 (1971).
b) H. H. Johnson, Proc. Am. Math. Soc. 15 (1964) 432, 675 .
c) E. A. Muller and K. Matschat, Miszellaneen der Angewandten Mechanik
(1962) 190.
d) Z. Khukhunashvili, Izvestiy Fizika Vyss. Ucheb. Zavadenii 3 1971 95–103.
e) S. Kumei, M. Sc. Thesis, University of the Pacific (1972); R. L. Anderson,
S. Kumei, C. Wulfman, Phys. Rev. Lett. 28 (1972) 988.
f) S. Kumei, J. Math. Phys. 16 (1975) 2461; 18 (1977) 256.
g) N. Ibragimov and R. L. Anderson, Dokl. Akad. Nauk. SSSR 227 (1976)
539.
h) G. W. Bluman and S. Kumei, Symmetries and Differential Equations
(Springer, N.Y., 1989).
This page is intentionally left blank
October 28, 2010 16:55 9in x 6in b943-ch10
CHAPTER 10
Introduction
The generators of invariance transformations of time-dependent Schrödin-
ger equations provide operators that can change the eigenstates and
eigenenergies determined by the corresponding time-independent equa-
tions. The previous chapter considered two examples of time-dependent
Schrödinger equations whose invariance groups provide operators that com-
muted with H, along with those that did not commute with H. The eigen-
values of the group generators that commuted with H were used to label the
energy eigenstates, and the generators that did not commute with H were
used to change eigenstates with a given energy into states with a different
energy. Both group generators and group operators were used for the lat-
ter purpose. Groups whose generators and/or operators convert an energy
eigenstate into such sets of eigenstates are said to be spectrum generating.
Infeld and Hull developed a method which uncovered spectrum gen-
erating groups,1 long before the extension of Lie’s theory of differential
equations to Schrödinger equations was available.2 Their method, the fac-
torization method, depends upon factoring a Hamiltonian into a product of
raising and lowering operators, and determines the generators of a three-
parameter group. A generalization produces Hamiltonians termed super-
symmetric, along with the Lie algebra of their three-parameter groups.3
In this chapter we develop a characterization of spectrum-generating Lie
algebras and use the generalized Lie theory of differential equations to
determine generators of the dynamical groups of a number of quantum
mechanical systems. Appendix 10.A lists references in which the dynamical
groups of other Schrödinger equations are displayed.
305
October 28, 2010 16:55 9in x 6in b943-ch10
H = ΣN
1 Hj . (10.3.2)
The oscillator is isotropic if all the ωj have the same value, ω. The ground
state of such an oscillator has energy
E0 = (N/2)ω. (10.3.3)
with energy
En = E0 + nω, (10.3.4)
has degeneracy
(N + n − 1)!
D= . (10.3.5)
n!(N − 1)!
Fowler had previously determined D by working out the way in which n
quanta can be assigned N degrees of freedom.6 The operator that shifts a
quantum of energy from degree of freedom j to degree of freedom k is
Its adjoint, b†j bk , carries out the inverse transfer. These N(N − 1) opera-
tors together with the N operators Hj comprise a set of N2 operators that
generate the degeneracy group U(N).
Several different types of dynamical groups are known for N-dimensional
isotropic harmonic oscillators. One of them, denoted NU(N) ,7 is generated
by these operators, together with the 2N Dirac shift operators and the
identity operator of the Heisenberg group “N”.
A dynamical group of three-dimensional isotropic harmonic oscillators
that is particularly useful is the group Sp(6), whose Lie algebra contains
the double-jump operators mentioned above. It may be used to interconvert
states containing odd numbers of excitation quanta, or interconvert states
containing even numbers of these quanta. For a more extensive discussion,
the reader is directed to Chapter 20 of Wybourne’s book,13 and references
therein.
The dynamical group of non-isotropic oscillators is the same as that
of the isotropic ones, but the generators b†k bj involving degrees of freedom
with different ω’s become time-dependent operators which alter energies,
and cease being generators of the degeneracy group. However, if the ω’s in
two different degrees of freedom are so related that there are two integers
Ia and Ib such that ωa /ωb = (Ia /Ib ), then degeneracies will appear when
na /nb = (Ib /Ia ).
The dynamical symmetries of harmonic oscillators have come to play
a large role in nuclear shell theory. This begins with the assumption that
nucleons move in approximate harmonic oscillator potential wells. It then
develops the consequences of j–j coupling between spins of the nucleons.
Appendix 10.B lists a number of references which will introduce the reader
to this application of dynamical symmetries which is now very extensive.
October 28, 2010 16:55 9in x 6in b943-ch10
to
D(H − i∂/∂t) Σcn Φn (q) exp(−iE(n)t)
= (H − D i∂/∂tD−1 )DΣcn Φn (q) exp(−iE(n)t)
= (H − (λ(N)/N)i∂/∂t)Σcn Φn (q) exp(−int). (10.4.5)
As
(H − λ(i∂/∂t))Ψ = 0. (10.4.6d)
As both N and i∂/∂t have linear spectra, both may be considered
to be Cartan labeling operators. Consequently, on seeking operators
Q(q, t, pop , i∂/∂t) that generate transformations which leave equation
(10.4.6d) invariant, one expects to find Cartan shift operators Q± for N
and i∂/∂t.
The determination of the spectrum-generating group of a rigid rotor
provides an example of the foregoing general discussion. Because, of the
rotor’s assumed rigidity, its Schrödinger wave function does not have a
radial dependence, and the Schrödinger equations for a rotor with moment
of inertia I, are
(L2 /2I − i∂/∂t)Ψ(t, θ, φ) = 0, (L2 /2I − El )Ψl (t, θ, φ) = 0,
(10.4.7a,b)
with
El = l (l + 1)2 /2 I. (10.4.7c)
Here, L2 is the total angular momentum operator. The operator Lz =
−i ∂/∂φ commutes with L2 , and has eigenvalues m. The two operators,
together have eigenstates Φlm (t, θ, φ), with m an integer constrained by
the condition |m| ≤ l . The Ψ(t, θ, φ) can be a linear combination of such
October 28, 2010 16:55 9in x 6in b943-ch10
Seeking Q that contain derivatives no higher than the second, Kumei set
f, ft , fφ , fx , fφφ , fxφ , ftφ , ftt , fxt , fφφφ , fxφφ , fttφ , fxtφ , ftφφ (10.4.16)
Inserting (10.4.13b) and (10.4.15) into (10.4.13a) and collecting the terms
that multiply each of the functions in (10.4.16) he used the linear indepen-
dence of the functions to obtain the following 14 equations:
Qφφ
t = 0, Qxφ
t = 0, Qφφ
x −y
−4 xφ
Qφ = 0, Qxφ
x + xy
−2 xφ
Q = 0,
Qφφ
φ − xy
−2 xφ
Q = 0, Qt
x −y
−2 x
Qt = 0, Qt x
t − Qx − xy
−2 x
Q = 0,
Qφ
t −y
−2 t
Qφ = 0, AQxφ − 2Qxφ − 2y−2 Qx 2 φ
φ − 2y Qx = 0,
AQφφ + 2y−2Qx
x − 4x y Q − 2y2 Qφ
2 −4 x
φ = 0, (10.4.17)
x x
AQ − 2Q − 4xQx
x − 4x y 2 −2 x
Q − 2y 2
Q0
x = 0,
t
AQ + 2Q0
t + 2iQx
x + 2ixy −2 x
Q = 0,
φ −2
AQ − 2y Q0
φ = 0, AQ = 0.0
Qj = D−1
t Qj Dt . (10.4.19)
with
1/2
c+ (l , m) = {(2 l + 3)/(2 l + 1)} ((l + 1 + m)(l + 1 − m))1/2 ,
and
[L0 , L+ ] = L+ , [L0 , L− ] = −L− , [L+ , L− ] = 2L0 . (10.5.12)
The operators
and
corrects this, and removes the factors in curly brackets from the expressions
for the c± (l , m) in (10.5.11), so that
and
The transformation does not affect the other group generators L’. All may
be subjected to the inverse of the spectrum-linearizing transformation,
which does not change the relations (10.5.6) and (10.5.16).
When written in spherical polar coordinates, Laplace’s equation,
∆Ψ = 0, (10.5.17a)
becomes
In this case it has been the radial variable r, rather than time, t, which has
been used in the linearization of the spectrum of an operator.
In Cartesian coordinates, the invariance generators of the Laplace equa-
tion are:
the generators of the Euclidean group E(3):
with
defined by
Uf(x) = (x−2 )f(x/x2 ). (10.5.23)
The reflections and the inversion are their own inverse, and
UTj U−1 = −S, URjk U−1 = Rjk , UDU−1 = −D. (10.5.24)
On the space of solutions, Ψ, of the Laplace equation, because ∆Ψ = 0,
one has
T · T Ψ = 0, S · S Ψ = 0, T · R Ψ = 0, (10.5.25a–c)
and
(R · R+D2 )Ψ = 1/4Ψ. (10.5.25d)
In Chapter 11 the reader will find that the invariance group of Laplace’s
equation and its generalization to Euclidean four-space becomes relevant to
the theory of the hydrogen atom. In Chapter 14, the generalization of the
group to Minkowski space provides invariance transformations of Maxwell’s
equations.
0.5
0
1 2 3 4 5
−0.5
−1
The operator
When Kumei’s operators are expressed in terms of z,t one obtains opera-
tors Q = DND−1 , in which the N are the n-labeling and n-shifting operators
N0 = i∂/∂t, (10.7.5a)
N+ = exp(−it)(i∂/∂z − N0 /z)(2N0 + 1) + d, (10.7.5b)
N− = exp(it)(i∂/∂z + N0 /z)(2N0 − 1) − d. (10.7.5c)
The operators
and, as in (10.4.14),
Defining
are then
Q = D−1 −1
t Dr Q Dr Dt . (10.8.10)
Allowing Q to contain derivative operators of zero, first, and second order in
all the variables, Kumei obtained and solved a set of 22 coupled determining
October 28, 2010 16:55 9in x 6in b943-ch10
equations. The 16 generators Q that are not products of others, are Q16 = 1,
and Q1 , . . . , Q15 , which satisfy SO(4,2) commutation relations.
Of these generators, Q, three are the standard angular momentum m
labeling operator M0 = L12 , and m shift operators M± = L23 ± iL31 .
Another three represent the components of Az = L34 , Ax ±iAy = L14 ±iL24 ,
that act on the functions ψnlm (r/2, θ, φ). These change the value of the
angular momentum quantum number, l . One of the operators, −i∂/∂t, is
the n labeling operator for the functions exp(−int)ψnlm (r/2, θ, φ). The
remaining eight operators all change the value of n, and hence, E, and are
consequently explicit functions of t. In the expressions for the generators
given below, x = cos(θ), y = sin(θ). Generators of compact subgroups are
labeled Lab (= −Lba ). Generators of non-compact subgroups are labeled
Kab (= +Kba). The generators satisfy the commutation relations
[Lab , Lbc ] = i Lac , [Lab , Kbc ] = i Kac , [Kab , Kbc ] = i Lac . (10.8.11)
Using these commutation relations, one is able to obtain the explicit form
of all the generators from the following:
L12 = −i∂/∂φ, (10.8.12)
−1
M± = exp(±iφ)(ixy ∂/∂φ − ±y∂/∂x) = (L23 ± iL31 ),
L34 = 2 −y2 ∂ 2 /∂r∂x + r−1 xy2 ∂ 2 /∂x2 + r−1 xy−2 ∂ 2 /∂φ2 + x∂/∂r
−2r−1x2 ∂/∂x − (i/2)x∂/∂t , (10.8.12a–e)
K45 = cos(t)(−ir∂/∂r − i) − sin(t)(i∂/∂t + r/2), (10.8.13)
L56 = i∂/∂t.
The generators not listed above are L14 , L24 , K15 , K25 , K35 , K16 , K26 , K36 ,
and K46 . The generator K46 may, for example, be determined from the
relation
[L56 , K45 ] = −iK46 , (10.8.12f)
which implies that
K46 = − sin(t)(−ir∂/∂r − i) − cos(t)(i∂/∂t + r/2). (10.8.12g)
When acting on the functions in (10.8.7), i∂/∂t multiplies them by n,
so in the generators (10.8.12c–e,g) it may be replaced by Nop = Z(P0op )−1 .
If one is primarily interested in the action of the generators on the position-
space wave functions ψnlm (r/2, θ, φ), one may set t = 0 in the operators that
result from this replacement of i∂/∂t. This process carries out conversions
such as
K45 → K45 = −ir∂/∂r − i, K46 → K46 = Nop + r/2. (10.8.13a,b)
October 28, 2010 16:55 9in x 6in b943-ch10
If one considers the ψm to be unit vectors, then (10.9.1) and (10.9.3) may
be represented, respectively, by the vectors
C = (c−1 , c0 , c+1 ), C = ( − c−1 , 0, c+1 ). (10.9.4)
L0 is then represented by a 3 × 3 matrix, L0 , whose off-diagonal elements
are zero and diagonal elements are −1, 0, +1. The functions ψm are said
to provide the basis for the matrix representation of the operator.
The analogous matrix representations of L+ = Lx+iLy, and L− = Lx−i
Ly, are defined by their action on ψm , for which one finds
L+ ψm = ψm+1 ,
L− ψm = ψm−1 . (10.9.5)
The reader may verify that the resulting matrix representations, L+ and
L− , satisfy the commutation relations
L0 L+ − L+ L0 = L+ ,
L0 L− − L− L0 = −L, (10.9.6)
L+ L− − L− L+ = 2L0 ,
which are isomorphic to the commutation relations obeyed by the corre-
sponding differential operators. As there is a one-to-one correspondence
between the matrices and the differential operators and a one to one cor-
respondence between their Lie algebras, it is said that the basis provides a
faithful representation of the Lie algebra. When such one-to-one correspon-
dence does not exist, the commutation relations of the matrices and that
of the differential operators are merely homomorphic.
The reader will find that all the matrices M in (10.9.6) are unitary,
that is
M−1 = M∗T . (10.9.7)
It can also be proven that these matrices cannot be reduced to the form
A B
(10.9.8)
B D
in which the elements of the submatrices B and/or B are all zero. Matrices
that cannot be put into this form are said to be irreducible. (When both
B and B are zero matrices, the matrix (10.9.8) is the product of the two
matrices A and D). All these conditions being fulfilled, the basis ψ−1 , ψ0 ,
ψ+1 , of the vector in (10.9.1) is said to be that of a unitary irreducible
representation, abbreviated UIR.
October 28, 2010 16:55 9in x 6in b943-ch10
at hand illustrates some of these. If one defines L2ab to be the sum of the
squares of the generators in (10.12.4), then the quadratic Casimir operator
of the group is
a. Adding a constant times L2ab to the Hamiltonian (10.12.2) leaves L2a and
L2b , and hence (La + Lb )2 as constants of motion, so it alters the energy
of the unpertubed states by an amount determined by the eigenvalues
of C2 .
b. As SO(6) is of rank 3, it also has two further Casimir operators. Adding
these to the Hamiltonian (10.17.2) also leaves la , lb and L as labeling
eigenvalues of constants of motion.
Exercises
1a. Show that the quantum mechanical version, -iD, of the dilation operator
D defined in (10.5.18), is not self-adjoint under the Schrödinger scalar
product, but becomes self-adjoint if one replaces r2 dr by rdr.
1b. Construct symmetrized quantum-mechanical versions of all the other
generators in (10.5.18) that are self-adjoint under this scalar product.
2a. Determine the commutation relations of the generators defined in
(10.5.18).
2b. Prove that they are isomorphic to commutation relations of the genera-
tors of SO(4,1), and establish the relation of the generators in (10.5.18)
October 28, 2010 16:55 9in x 6in b943-ch10
to the generators Lab = −Lba of the SO(4) subgroup of SO(4, 1), and
those, Kab = Kba , of its noncompact subgroup.
3a. Completion of the direct-product group SO(2)⊗SO(2) produces SO(4).
Consider the circles defined by the one-parameter groups generated by
L12 and L34 , and suppose that these are contained in orthogonal planes,
so that (x1 , x2 )·(x3 , x4 ) = 0. How are these circles affected by rotations
generated by L23 , L14 ?
3b. Determine the small rotations in the SO(4) group that leave invariant
the scalar product (x1 , x2 ) · (x3 , x4 ).
References
[1] L. Infeld and T. E. Hull, Rev. Mod. Phys. 23 (1951) 21.
[2] Chap. IX Ref. 5.
[3] F. Cooper, A. Khare and U. Sukhatme, Supersymmetry in Quantum
Mechanics (World Scientific Publishers, Sinagapore, 2001).
[4] R. L. Anderson, S. Kumei and C. E. Wulfman, Rev. Mex. de Fisica 21 (1972)
1–33.
[5] G. A. Baker Jr., Phys. Rev. 103 (1956) 1119.
[6] R. H. Fowler, Statistical Mechanics (Cambridge University Press, N. Y.,
1955), Sec. 2.21.
[7] A. O. Barut and A. Bohm, Phys. Rev. B 139 (1965) 1107.
[8] S. Kumei, Group Theoretic Properties of Schrödinger Equations– Systematic
Derivation, M.Sc. Thesis, Department of Physics, University of the Pacific
(Stockton, CA, 1972).
[9] A. R. Edmonds, Angular Momentum in Quantum Mechanics (Princeton
University Press, Princeton, 1957).
[10] P. M. Morse, Phys. Rev. 34 (1929) 57.
[11] P. Cordero and S. Hojman, Lett. Nuovo Cimento 4 (1970) 1123.
[12] A. Messiah, Quantum Mechanics (John Wiley, N.Y., 1962), Vol. I, Chap. XI.
[13] B. G. Wybourne, Classical Groups for Physicists (Wiley-Interscience, N.Y.,
1974).
[14] A. O. Barut and R. Raczka, Theory of Group Representations and Applica-
tions, 2nd edn., (World Scientific, Singapore, 1986).
[15] I. M. Gelfand and M. L. Tsetlein, Dokl. Akad. Nauk. SSSR 71 (1950) 25;
ibid. 1017.
[16] T. Shibuya, Fock’s Representation of Molecular Orbitals, M.Sc. Thesis,
Department of Physics, University of the Pacific (Stockton, CA, 1965);
cf., T. Shibuya and C. E. Wulfman, Proc. Roy. Soc. A 286 (1965) 376.
[17] C. E. Wulfman, Phys. Rev. A 54 (1996) R987.
[18] T. Kato, Perturbation Theory for Linear Operators (Springer, N. Y. 1966),
pp. 410–413.
This page is intentionally left blank
October 28, 2010 16:55 9in x 6in b943-ch11
CHAPTER 11
Introduction
This chapter relates the dynamical symmetries of one-electron atoms to
the dynamical symmetries of regularized classical Keplerian systems. It
first describes the quantum mechanical position-space formulation of the
regularized system and the generators of its SO(4,2) group, then sets forth
the corresponding momentum-space formulation and Fock’s projection of
its momentum space onto the hypersphere S(3). The chapter concludes
with a description of the SU(2) ⊗ SU(2) structure of the SO(4) degeneracy
group. To simplify the presentation in this and the subsequent chapter, we
will use units in which the charge of the electron, e, has unit value, and as
in the two previous chapters, will use units in which both the reduced mass
of the electron and = h/(2π) also have unit value.
337
October 28, 2010 16:55 9in x 6in b943-ch11
one obtains
ψ ∗ r(Hop − E)ψ dv. (11.1.3b)
This implies
ψ, r(Hop − E)ψ = ψ ∗ (Hop − E)ψ dv, (11.1.4b)
and
ψ, rψ = ψ ∗ ψ dv. (11.1.4c)
Though r and −r∇2 are self-adjoint under the scalar product (11.1.4a),
the operator −i∂/∂x is not. However, operators, such as ordinary angular
momentum generators, which commute with r and are self-adjoint under
the Schrödinger scalar product, remain self-adjoint under (11.1.4a).
The dynamical symmetry of the classical Keplerian systems investigated
in Chapter 8 can be enlarged to an SO(4,2) symmetry.1−3 The simplest way
to do this is to add to the SO(4,1) generators of Sec. 8.6, the five operators:
The relations
[Jab , Jbc ] = iJac , [Jab , Kbc ] = iKac , [Kab , Kbc] = iJac , (11.1.6)
L = (J12 , J23 , J31 ) ∼ {(r × p), ·}, A− = (J14 , J24, J34 ) ∼ {A(−1/2), ·},
A+ = (K15 , K25 , K35 ) ∼ {A(1/2), ·}, Γ = (K16 , K26 , K36 ) ∼ {(rp), ·}.
(11.1.7a)
K45 ∼ {(r · p), ·}, K46 ∼ {W(1/2), ·}, J56 ∼ {W(−1/2), ·}, (11.1.7b)
Table 11.1.
SO(3) L L2 → l (l + 1)
SO(2,1) K45 , K46 , J56 (J56 )2 − (K45 )2 − (K46 )2 → l (l + 1)
SO(4) L, A− L2 + A2− → n 2 − 1; (A− · L)2 ≡ 0
SO(3,1) L, A+ L2 − A2+ → ν 2 − 1; (A+ · L)2 ≡ 0
SO(4,1) L, A− , A+ , K45 A2+ + (K45 )2 − A2− − L2 ≡ 2
SO(4,1) L, A− , Γ, K46 Γ2 + (K46 )2 − A2− − L2 ≡ 2
October 28, 2010 16:55 9in x 6in b943-ch11
has unit value. Both higher order SO(4,2) Casimir operators vanish. The
SO(4,2) basis that is defined by the values of these Casimir operators is
countably-infinite discrete. Because A is perpendicular to L, the quar-
tic Casimir operator of SO(4) vanishes identically. As a result the SO(4)
representations contained in the SO(4,2) have dimension n 2 . The n 2 -fold
degeneracy of hydrogen atom energy levels is a physical expression of this
group-theoretic property. The basis states of the SO(4) representations,
with a given value of n, are most commonly chosen to be those of n different
SO(3) representations, labeled by l, which takes on the values 0, 1, . . . , n −1,
and by the integer m which takes on the 2l + 1 values for which |m| ≤ l.
The resulting basis for the representation of SO(4,2) is provided by the
functions which satisfy the equations
[2]
CSO(4,2) χnlm = χnlm , J56 χnlm = nχnlm ,
J12 χnlm = mχnlm , L2 χnlm = l (l + 1)χnlm , (11.1.9)
with
has been assigned the phase factor, (−i)l .5 The functions χnlm , now com-
monly called Sturmians, differ from those introduced in Sec. 10.8, and are
not the same as the usual normalized bound-state electronic wave functions,
October 28, 2010 16:55 9in x 6in b943-ch11
When E is negative, the coefficient of J56 is greater than that of K46 , and
when E is positive the reverse is true. The unitary dilatation operators
exp(±iαJ45 ) can also be used to relate a system with arbitrary E to the
systems with the E values ±1/2. For E < 0 one uses the relation
exp(−iαJ45 )J56 exp(iαJ45 ) = cosh(α)J56 + sinh(α)K46 . (11.1.18)
Letting k to be a constant, one may in (11.1.17) set
(1/2 − E) = k cosh(α), (1/2 + E) = k sinh(α). (11.1.19a)
This requires
(1/2 − E)2 /k2 − (1/2 + E)2 /k2 = 1, (11.1.19b)
whence
k = (−2E)1/2 , (11.1.19c)
and
cosh(α) − sinh(α) = e−α = (−2E)1/2. (11.1.19d)
It follows that, when E < 0, the equation
(Wop − 1)ψ = 0 (11.1.20a)
can be expressed as
(exp(−iαJ45 )J56 exp(iαJ45 ) − 1)/(−2E)1/2)ψ = 0. (11.1.20b)
Multiplying this through by exp(iαJ45 ) gives
(J56 − 1/(−2E)1/2 )ψ = 0, ψ = exp(iαJ45 )ψ. (11.1.20c)
Thus ψ is an eigenfunction of J56 , which has eigenvalues n, and the eigen-
functions χnlm defined Lie-algebraically in (11.1.9). Consequently,
n − 1/(−2E)1/2 = 0, e−α = 1/n, E = −1/(2n 2 ). (11.1.21)
One can thus write
ψnlm = exp(−iαJ45 )χnlm (r, θ, φ) = e−α χnlm (e−α r, θ, φ), (11.1.22a)
and
ψnlm = (1/n)χnlm (r/n, θ, φ). (11.1.22b)
The Schrödinger eigenfunctions Ψnlm (r, θ, φ) defined in (11.1.11) differ from
these ψnlm only because of the difference in normalization that arises from
October 28, 2010 16:55 9in x 6in b943-ch11
Thus
and the functions exp(−(π/2)J45 )χnlm are eigenfunctions of K46 with eigen-
values in. Let us set in equal to ν, and denote these eigenfunctions by χνlm .
Using (11.1.22a) establishes then that
and
∞
= (2π)−1 exp(ipx)dpδ(x). (11.2.2b)
−∞
and
f(r)δ(r − r )d3 r = f(r ). (11.2.3b)
Using (11.2.3) one may derive that the inverse of transformation (11.2.1) is
ψ(r) = (2π)−3/2 exp(ip · r)φ(p)d3 p, d3 p = dpx dpy dpz . (11.2.4)
(11.2.5a)
(11.2.5b)
To prove this relation one may apply (11.2.1), to its left-hand side, and
obtain
∗
(2π)−3 exp exp(−ip · r)Ψ(r)d3 r exp(−ip · r )Ψ (r )d3 r d3 p
−3/2 −3/2
= (2π) (2π) exp(ip · (r − r )) d3 p Ψ(r)∗ d3 r ψ (r ) d3 r
= δ(r − r )ψ(r)∗ d3 rψ (r )d3 r = ψ ∗ (r)ψ (r)d3 r,
(11.2.7b)
and thus Eq. (11.2.7a) is established.
The momentum-space analog of the Schrödinger equation
(p2op /2 + v(r) − E)φ = 0, (11.2.8)
is
(p2 /2 + v(rop ) − E)φ = 0, rop = −∇2 . (11.2.9)
Because working with the function of rop is often very difficult, one pro-
ceeds differently. Taking the Fourier transform of the original Schrödinger
October 28, 2010 16:55 9in x 6in b943-ch11
which becomes
(p2 /2 − E)φ(p) = (2π)−3/2 exp(−ip · r)v(r)ψ(r)d3 r
= φ(p )V(|p − p|)d3 p , (11.2.11)
these relations into the integrand, and multiplying both sides of (11.2.12)
by (p2 + p20 ), converts the equation to
(p2 + p20 )2 φ(p) = (Z/2p0 )(1/π)2 (p20 + p2 )2 φ(p )/|x − x |2 dΩ .
(11.2.16a)
On setting
(p2 + p20 )2 φ(p) = Φ(x), (11.2.16b)
and multiplying (11.2.16a) through by p0 , it takes on the simpler form
p0 Φ(x) = (Z/π ) Φ(x )/|x − x |2 dΩ .
2
(11.2.17)
the orthonormality of the Y’s requires that it becomes unity if Φ is Ykjm (x).
Equation (11.2.20) then states that
If both Φ(x) and Yk j m (x) are normalized, this relation requires that
where
= (p2 + p20 )/2p20 = ((p/p0 )2 + 1)/2,
and N is a normalization factor that may also contain a phase factor.
Conversely,
3/2
Yklm (α, θ, φ) = N−1 p0
2 φnlm (p/p0 , θ, φ), (11.3.2b)
The momentum-space functions φnlm , are of the form
ρnl (p/p0 )Ylm (θ, φ). (11.3.2c)
The hyperspherical harmonics are directly defined in terms of Gegen-
bauer polynomials Cλν with ν = n − (l + 1) = k − l , and λ = l + 1. One
has11
Yklm (α, θ, φ) = Nl Cλν (cos(α)) sinl (α)Ylm (θ, φ), (11.3.3a)
with
Nl = (−i)l (2π 2 )−1/2 . (11.3.3b)
The phase factor, (−i)l , which makes the normalization factor l dependent,
has been chosen so that the hyperspherical harmonics satisfy a standard
phase convention5
∗
Yklm = (i)l−m Ykl−m . (11.3.3c)
The term sinl (α)Ylm (θ, φ) in (11.3.3), is a function of x1 , x2 , x3 that
is obtained from the usual Ylm (θ, φ) when one replaces x, y, z defined in
momentum-space by the x1 , x2 , x3 defined above. This leads us to define a
unit vector y in p-space, analogous to the unit vector x in the projective
space. Then one of the effects of Fock’s projection is to convert Ylm (y) to
Ylm (x).
−3/2
Its other effect is to convert the radial function ρnl (p/p0 ) to Nl p0
times
((p/p0 )2 + 1)−2 Cλν (cos(α)) = (1 + cos(α))2 Cλν (cos(α)) = (1 + x4 )2 Cλν (x4 )
= (1 + x4 )2 Cλν (x4 ).
(11.3.4)
Variations in the length of the momentum vector are mapped into variations
of x4 , and the consequent variations in the radial function ρnl (p/p0 ) are
expressed through variations in the Gegenbauer functions and the factor
(1 + x4 )2 . The color figures in Chapter 1 depict the stereographic projection
October 28, 2010 16:55 9in x 6in b943-ch11
The Gegenbauer polynomials Cλν (x) play the same role in the theory of
the rotation groups SO(N), N > 3, that Legendre polynomials play in the
theory of SO(3). The Cλν (x) have a generating function, G, defined by
dx1 dx2 dx3 dx4 = ρ3 dρ dΩ, ρ2 = x21 + x22 + x23 + x24 , (11.3.7)
Harmonic (2π 2 )1/2 Yklm (α, θ, φ) (2π 2 )1/2 Yklm (x) Radial Sturmian
Y000 1 1 2e−r
√
Y100 −2 cos(α) −2x4 2 2(1 − r)e−r
√ √
Y110 i 2 sin(α) cos(θ) i 2x3 (2 2/ 3)re−r
√ √
Y111 i 2 sin(α) sin(θ)eiφ −i 2(x1 + ix2 )/21/2 (2 2/ 3)re−r
√ √
Y11−1 2 sin(α) sin(θ)e−iφ i2(x1 − ix2 )/21/2 (2 2/ 3)re−r
√
Y200 4 cos2 (α) −1 4x24 − 1 2 3(1 − 2r + 2r2 /3)e−r
√ √
Y210 −i61/2 sin2 (α) cos(θ) −i61/2 x4 x3 (4 2/ 3)r(1 − r/2)e−r
√ √
Y21±1 ±(3)1/2 sin(2α) sin(θ)e±iφ ±i(3)1/2 2x4 (x1 ± ix2 ) (4 2/ 3)r(1 − r/2)e−r
√
Y220 −21/2 sin2 (α)(3 cos2 (θ) − 1) −21/2 (3x23 + x24 − 1) (4/( 30))r2 e−r
√
Y22±1 ±(3)1/2 sin2 (α) sin(2θ)e±iφ ±i(3)1/2 2x3 (x1 ± ix2 ) (4/( 30))r2 e−r
√
Y22±2 −(3)1/2 sin2 (α) sin2 (θ)e±2iφ −(3)1/2 (x1 ± ix2 )2 (4/( 30))r2 e−r
October 28, 2010 16:55 9in x 6in b943-ch11
The area element bears the following relation to the volume element d3 p in
momentum space:
dΩ =
−3 p−3 3
0 d p. (11.3.8)
also have the same values. The phase factors have been so chosen that
d3 p p−3
0
−2 ∗
Yklm
Xp
−2 Yk l m
= dΩ
−3
−2 Yklm ∗
Xp
−2 Yk l m
∗
= dΩ Yklm
2 Xp
−2 Yk l m . (11.3.24)
Thus if
Xp → Ξ =
2 Xp
−2 , (11.3.25a)
then
(Yklm , Ξ Yk l m ) = φnlm , Xp φn l m . (11.3.25b)
The inverse of the similarity transformation defined by (11.3.25a), that is
Ξ → Xp =
−2 Ξ
2 , (11.3.25c)
connects the SO(4) rotation generators Ji4 , to momentum-space versions of
Runge–Lenz operators.
In Fock’s treatment, the dynamics of the atom is governed by the equa-
tion,
p0 Φ(x) = (Z/π ) Φ(x )/|x − x |2 dΩ ,
2
(11.3.26a)
has two commuting generators, which may be choosen to be J12 and J34 .
The quantum mechanical position-space operators corresponding to the Jab
satisfy commutation relations in which the negative sign on the right-hand
side is replaced by its square root, and the same may be said of the corre-
sponding momentum-space operators. In the case of the SO(4) degeneracy
group of hydrogen-like atoms, one may label degenerate eigenstates | by
the quantum number n, and by the eigenvalue m12 of J12 and the eigenvalue
m34 of J34 . Then one has the n2 states defined by
and
one finds that all the M operators commute with all the N operators. It
follows that SO(4) is the direct product of two three-parameter groups.
The quantum mechanical version of the operators satisfy
The labeling schemes (11.4.2) and (11.4.5) are connected by the relations
One has
Using the relations r = (J56 − K46 ), and z = (K35 − J34 ), one obtains:
and
by converting (11.B.1a) to
{(1 − x24 )(∂/∂x4 )2 + (2λ + 1)x4 ∂/∂x4 + i∂/∂τ (i∂/∂τ + 2λ)} (11.4.1)
× exp(−iντ )Cλν (x4 ) = 0. (11.B.1b)
The determining equations for the generators of the invariance group of this
equation yield the required labeling and v-shift operators for the Cλν when
one allows Q’s that correspond to first-order differential operators, U. The
solutions of the determining equations are defined to within a multiplicative
constant. He chose them to be
Q1 = i∂/∂τ, Q2 = 1, (11.B.2a)
Q− = exp(iτ ){(1 − x24 )∂/∂x4 + i x4 ∂/∂τ }, (11.B.2b)
Q+ = exp(−iτ ){(1 − x24 )∂/∂x4 − i x4 ∂/∂τ − 2λ x4 }. (11.B.2c)
Q0 = Q1 + λQ2 . (11.B.2d)
Consequently
They shift ν by ±1 without shifting l , and hence shift n by ±1. The Casimir
operator is
Q exp(−iντ )Cλν (x4 ) → N exp(−iντ )(1 − x24 )l/2 Cλν (x4 ), (11.B.4a)
October 28, 2010 16:55 9in x 6in b943-ch11
with
N = (1 − x24 )l/2 Q(1 − x24 )−l/2 . (11.B.4b)
Q0 = i∂/∂τ + λ ≡ N0 is not altered by the transformation. If we let
|nlm = exp(−iντ )Yklm (11.B.5a)
then
N0 |nlm = n|nlm. (11.B.5b)
Also let
N± |nlm = N± exp(−iντ )Yklm = {(1 − x24 )l/2 Q± (1 − x24 )−l/2 }
× exp(−iντ |Nl (1 − x24 )l/2 Cλν (x4 )
= (1 − x24 )l/2 Q± exp(−iντ ) Nl Cλν (x4 ) (11.B.5c)
= (1 − x24 )l/2 c± (n, l ){exp(−i(ν ± 1)τ ) Nl Cλν±1 (x4 )}
= c± (n, l )|n ± 1, l , m
One finds
N− = exp(+iτ ){(1 − x24 )∂/∂x4 + x4 (N0 − 1)}, (11.B.5d)
N+ = exp(+iτ ){(1 − x24 )∂/∂x4 − x4 (N0 + 1)}. (11.B.5e)
Because they have been obtained by similarity transformation, these Q’
operators satisfy the same commutation relations as the original Q opera-
tors. The corresponding Casimir operator has eigenvalues l (l + 1).
If N+ and N− are matrices with elements n lm, N± |nlm, then one
finds that N− = −N†+ : the operator N− is the negative of the adjoint of
N+ . In the general case,
N± |nlm = c± (n, l )|n ± 1, l , m (11.B.6)
with
c± (n, l ) = ±((l + 1 ± n)(±n − l ))1/2 .
The operators Q also yield operators Q = exp(iντ )Q exp(−iντ ) that can
be used to produce operators which act directly on the Yklm . After removing
the time-dependent exponents from the Q , the operator Q0 → N0 , and
one obtains operators Θ± that convert Yklm to c± (k + 1 , l ) Yk ±1lm . These
operators are
Θ− = (1 − x24 )∂/∂x4 + (k − l )x4 ,
(11.B.7)
Θ+ = (1 − x24 )∂/∂x4 + (k − l )x4 − 2(l + 1)x4 .
October 28, 2010 16:55 9in x 6in b943-ch11
References
[1] A. O. Barut and H. Kleinert, Phys. Rev. 156 (1967) 1541.
[2] C. Fronsdal, Phys. Rev. 156 (1967) 1665.
[3] Y. Nambu, Phys. Rev. 160 (1967) 1171.
[4] M. Bednar, Ann. Phys. (N.Y.) 75 (1973) 305.
[5] B. R. Judd, Angular Momentum Theory for Diatomic Molecules (Academic
Press, N. Y., 1975), pp. 32–34.
[6] P. A. M. Dirac, Quantum Mechanics, 3rd edn. (Oxford, 1947), p. 53.
[7] V. Fock, Z. Phys. 98 (1935) 145.
[8] J. Avery, Hyperspherical Harmonics and Generalized Sturmians (Kluwer,
Dordrecht, 2000), p. 63.
[9] W. Pauli, Z. Phys. 36 (1926) 336.
[10] V. Bargmann Z. Phys. 99 (1936) 576.
[11] J. Avery, ibid. (Ref. 8) p. 25.
[12] T. Shibuya and C. E. Wulfman, Amer. J. Phys. 33 (1965) 570.
[13] J. Avery, Hyperspherical Harmonics (Kluwer, Dordrecht, 1989).
[14] B. R. Judd, ibid. (Ref. 5) p. 222.
[15] T. Shibuya, C. E. Wulfman, Proc. Roy. Soc. (London) A 286 (1965) 376.
[16] L. C. Biedenharn, J. Math. Phys. 2 (1961) 433.
[17] S. Kumei, M. Sc. Thesis (University of the Pacific, 1972).
This page is intentionally left blank
October 28, 2010 16:55 9in x 6in b943-ch12
CHAPTER 12
12.1 Introduction
In 1885, Annalen der Physik published a paper on the physical relevance of
certain small integers.1 In the paper, the Basel secondary-school teacher,
J. J. Balmer, pointed out that the observed wavelengths, λ, of the Hα , Hβ ,
Hγ , and Hδ lines in the sun’s Fraunhofer spectrum2 obeyed, with amazing
accuracy, a relation
361
October 28, 2010 16:55 9in x 6in b943-ch12
with eigenvalues m12 and m34 , respectively. These then become the zero-
order eigenstates of the Stark effect. The |nlm can be considered as being
obtained from them by the coupling defined by
|nlm = (j1 m1 j2 m2 |j1 j2 lm)|j1 m1, j2 m2 , (12.2.4a)
m1 m2
E(a), is
H = p2 /2 − (Za /|q − ak| + Zb /|q + ak|), (12.2.10)
and the constant of motion can be expressed as
F = L2 + a2 p2z + 2az(Za /|q−ak| − Zb /|q + ak|). (12.2.11)
The corresponding Poisson bracket operator is the generator of a one-
parameter symmetry group in classical phase space. As F commutes with
Lz , the degeneracy group is Abelian.
The non-crossing rule applies to sets of energy level curves such as E(a).
Wigner and von Neumann proved that varying a single parameter cannot
produce crossings of levels whose wave functions have the same symmetry;
if they appear to do so, a symmetry has been overlooked.15 One-electron
diatomics provide one of the simplest such examples. Some levels of the
same spatial symmetry cross because they have different dynamical symme-
try. In diatomics with two or more electrons this symmetry is removed, and
levels that previously crossed approach, but avoid, each other. The reader
is referred to Judd’s monograph11 for discussions of a number of interest-
ing aspects of the dynamical symmetry of one-electron diatomics. Included
within it is a discussion of the simple solutions that arise for special sets
of nuclear charges. In some later calculations on systems with Za = 1, and
Zb = 3, . . . , 8, Ponomarev and Puzynia observed pseudo-intersections.16
This raises a suspicion that as the bond between the proton and the other
nucleus weakens, such systems may have a useful approximate dynamical
symmetry larger than that of homonuclear systems.
B1 A2
B1 B2
A1
A2
A1
B2
Fig. 12.3c Relation between the p-orbitals after a twist through 45◦ .
Fig. 12.3d Relation between the p-orbitals after a twist through 180◦ .
as the excited state, which may be produced when the molecule in (c)
absorbs a photon. Such photo-excited molecules often undergo cis-trans
isomerism — that is twist through 180◦ to produce the ground state
of the molecule, (b). If A1 and B1 are identical to A2 and B2 respec-
tively, there must be a level crossing that produces a degeneracy at the
90◦ twisted configuration.17 In systems with less geometrical symmetry,
degeneracies develop when the p orbitals are approximately perpendic-
ular to each other. The resulting degenerate states are very sensitive to
the presence of polarizing fields with components in the C1 –C2 direction,
and the molecules begin to develop large dipole moments when the twist-
ing becomes a few degrees less than 90◦ . In the case of ethylene, a pro-
ton situated along this axis and a C–C bond length away produces the
dipole moment obtained by moving a charge of 0.9e from one C atom
to the other.17 In molecules with a conjugated system of double bonds,
asymmetries due to substituents may lead to sudden polarizations in the
twisted states that produce very large dipole moments. Bruckner and Salem
proposed that just such a sudden photo-polarization of retinal is respon-
sible for initiating nerve impulses in the retina.18 For further informa-
tion the reader is referred to the proceedings of a conference on sudden
polarization.19
When variation of a parameter changes the spatial point-group sym-
metry of a nonlinear molecular system, the theorem of Jahn and Teller
can come into play.20 It dictates that such a variation creates a degener-
acy of energy levels, when in the region of the crossing the energies are
linear functions of the parameter. It follows that if the degeneracy were to
occur naturally, the molecule would undergo a distortion that removes the
symmetry.
October 28, 2010 16:55 9in x 6in b943-ch12
Fig. 12.3e Correlation between the united-atom and separated-atom energy lev-
els of one-electron linear homonuclear triatomic molecules with identical bond
lengths.
or 19 valence electrons are nonlinear. As the systems are bent, the sym-
metry becomes C2v . However, levels of the same C2v symmetry, but with
different l, cross, and the πu orbitals correlate with orbitals of much lower
energy. Systems with 15 or 16 valence electrons remain linear, but those
with 17 to 19 valence electrons are bent. In 20 electron systems, bend-
ing increases the electron density at the center of nuclear charge, and
this stabilizes the bent system. Extensive analyses have fully established
the connection between all of Walsh’s rules and the extended Jahn–Teller
Theorem.25
The examples in this section have been chosen to illustrate a general
strategy that is available whenever one is dealing with a system governed
by differential equations. If one begins to attack the problem by seeking the
dynamical symmetry that, when broken, can produce the phenomenon of
interest, several advantages result:
|PQLM (n1 , n2 )
= B(n1 l1 m1 , n2 l2 , m2 |n1 n2 , PQLM )|n1 l1 m1 |n2 l2 m2 . (12.4.1)
l1 l2 m1 m2
1
S(2s2 ) = |00(0, 0, 2, 2)|1σ, 1
S(2p2 ) = |00(1, 1, 2, 2)|1σ. (12.4.5)
The first state is simply a product of two 2s functions and a spin func-
tion, |1 σ, that is antisymmetric under exchange. In the second state, prod-
ucts of two 2p functions are coupled to produce a state of zero angular
momentum that is symmetric under particle exchange. Multiplying this
by |1 σ produces the 1 S(2p2 ) wave function. Together, these produce the
|PQLM (n1 , n2 ) states29
For them
ja1 = (L1 + A1 )/2, jb1 = (L1 − A1 )/2,
(12.5.4)
ja2 = (L2 + A2 )/2, jb2 = (L2 − A2 )/2,
and
The rule states that in the neutral atoms in a given period, orbitals of
lowest n + l are filled first, and that orbitals having the same n + l value
are filled in order of increasing n.39 Unfortunately, as Ostrovsky notes, the
SO(3,1) subgroups do not contain the orbitals required to deal with period
doubling.42
Advances that follow Barut’s are described in Ref. 37 in contributions
from Novaro,40 Kibler,41 and Ostrovsky, who provides a very extensive list
of references to the earlier literature.42 These authors explore the relevance
of the groups SU(2) ⊗ SU(2) ⊗ SU(2), SO(4, 2) ⊗ SU(2), and SO(4, 2) ⊗
SU(2) ⊗ Z2 . (In each of these direct-product groups the SU(2) subgroup is
not due to electron spin).
Most periodic charts portray one period of length two, Period 1. Then,
Periods 2 and 3 both have the same length, eight; Periods 4 and 5, both
have length eighteen; Periods 6 and 7 both have length thirty-two.
Though SO(4) has unitary irreducible representations of the required
dimensions, 1, 4, 9, and 16, each of these representations occurs only once
in SO(4,2). Thus a definitive feature of the periodic chart can have no
explanation within SO(4,2). A dynamical explanation of the observed order
of filling of orbitals with nodal properties of hydrogen-like n, l, m orbitals
must deal with this fundamental problem. Each of the authors mentioned
above has approached it in a different way, but much remains to be done
to develop a definitive deduction of period doubling.
The discussions in previous chapters make it evident that a dynamical
explanation of period doubling will also provide a knowledge of the relevant
subgroup(s) of the dynamical group of n-electron atoms.
In this connection, a key group-theoretic aspect of period doubling
emerges from an investigation carried through by Kitigawara.43 He estab-
lished that the smallest special orthogonal groups with UIR of the required
dimensions are the rank three group SO(7) and its noncompact general-
izations SO(5,2), etc. No known single-particle system appears to have the
required dynamical symmetry.
Fortunately, Kitigawara’s analysis provides further insight into what is
required to derive period doubling from the dynamics — and hence dynam-
ical symmetries — encoded in many-electron Schrödinger equations. The
length of each period is a reflection of the fact that the energy gaps between
two successive periods is greater than the step-by-step energy changes that
develop as each periods fills. One might conjecture that in the simplest
of approximations the dynamical symmetry responsible for this could be a
degeneracy group which is a subgroup of SO(8,4). It could be that of systems
October 28, 2010 16:55 9in x 6in b943-ch12
corresponds to
and
To determine the effect of the translation one may expand the product
exp(ip · R)Yplm (Ω) as a sum of hyperspherical harmonics, writing,44
exp(ip · R)Yplm (Ω) = Snlm
n l m (p0 R) Yp l m (Ω). (12.9.2a)
Then
Snlm
n l m (p0 R) = Yp l m (Ω)∗ exp(ip · R)Ynlm (Ω)dΩ. (12.9.2b)
(P − p0 I)C = 0, (12.9.4b)
The matrix
P= Zj W(p0 Rj ) = Zj S(p0 Rj )−1 ΠS(p0 Rj ). (12.9.4d)
October 28, 2010 16:55 9in x 6in b943-ch12
and the total potential V(x) is the sum of the attractive potentials, V0 (x),
the total electrostatic repulsion potentials, and any other relevant poten-
tials.
Let us for the moment ignore the Pauli principle and spin, and expand
ψ as a linear combination of many-particle Sturmian functions
ψ= Bν Φν (x), (12.10.3)
in which the Φν are required to satisfy the equations
(−∆/2 + βν V0 (x) − E)Φν (x) = 0. (12.10.4)
Here βν is chosen in such a way that the energry E in these equation does
not depend on ν. The members of the set of solutions to Eq. (12.10.4)
all correspond to the same energy. This energy is chosen according to the
October 28, 2010 16:55 9in x 6in b943-ch12
state that is being represented. It can be shown that the set of isoenergetic
solutions to (12.10.4) obeys a potential-weighted orthogonality relation.
It is convenient to normalize the configurations in such a way that these
orthogonality relations take the form
dx Φ∗ν (x) V0 (x) Φν (x) = (2E/βν )δν,ν , (12.10.5)
Next, suppose that the functions Φν (x) are products of single particle
Sturmians φη (x), η = nlm, and spin functions σ± , and then write
χµ (x1 ) = φη (x)σ± , µ = η ± 1/2. (12.10.8a)
Then let
φν (x) = χµ (x1 ) χµ (x1 ) χµ (x1 ) . . . (12.10.8b)
and require:
(A) The functions χµ (xj ) are solutions of the equations
(−∆j + ku2 − 2nk u /rj ) χµ (xj ) = 0 (12.10.9a)
and
(B)
d3 xj χµ (xj )∗ (1/rj ) χµ (xj ) = (kµ /n)δµ,µ , (12.10.9b)
(C)
d3 xj |χµ (xj )|2 = 1. (12.10.9c)
(D)
Equations (A) and (D) taken together assign to the argument of each χµ (xj )
a scale factor kµ , and hence eigenvalues
Thus Zβν is an effective nuclear charge that becomes assigned to the Stur-
mian with single-particle energy
Let
Next define
dx Φ∗ν (x) V0 (x) Φν (x) = −p0 T0ν ν , (12.10.11c)
dx Φ∗ν (x) V1 (x) Φν (x) = −p0 T1ν ν . (12.10.11d)
= (p0 /βν )δν ν = (1/βν )( ku2 )1/2 δν ν . (12.10.12)
If there are Nj electrons with principle quantum number nj , it becomes
convenient to define
1/2
Rν = Nj /nj2 , (12.10.13)
a function whose value only changes when orbital occupancies change. Then
(12.10.9e) implies
T0ν ν = ZRν δν ν . (12.10.14)
Inserting (12.10.11b) into (12.10.7), and dividing through by p0 produces a
set of secular equations. If one lets T0 = T0 I be the diagonal matrix with
elements ZRν , and T1 be the matrix with elements T1ν ν , then the secular
equations take on the matrix-vector form:
(T1 − (p0 − T0 )I)B = 0. (12.10.15)
Some of the key consequences of Eqs. (12.10.9) are revealed when the
methodology is applied to neutral several-electron atoms while configura-
tion interactions are neglected. If the many-electron states obtained by
solving (12.10.15) are labeled by an index κ, and their momenta p0 are
labeled p0κ , then
p0κ = ZRν − T1κκ , (12.10.16a)
and
Eκ = −(1/2)(ZRν − T1κκ )2 . (12.10.16b)
For the ground states of neutral atoms with 2 to 9 electrons, as Z runs from
2 through 9, T1κκ changes from 0.441942 to 2.41491, an average increase
October 28, 2010 16:55 9in x 6in b943-ch12
References
[1] J. J. Balmer, Ann. Phys. 25 (1885) 80.
[2] J. V. Fraunhofer, Encylopedia Britannica, 11th edn. (Cambridge, 1910).
[3] N. Bohr, Phil. Mag. 26 (1913) 1.
[4] Cf. A. Sommerfeld, Atomic Structure and Spectral Lines, 3rd German edn.,
translated by H. L. Brose (Methuen, London, 1923).
[5] E. U. Condon and G. H. Shortley, The Theory of Atomic Spectra (Cambridge
University Press, 1935, 1951–1967).
[6] (a) C. Eckart, Rev. Mod. Phys. 2 (1930) 305;
(b) E. Wigner, Gruppentheorie (Viewig, 1931);
(c) E. Wigner, Group Theory and its Application to the Quantum Mechanics
of Atomic Spectra, (Academic Press, N.Y., 1959).
[7] G. Racah, Phys. Rev. 61 (1942) 186; Phys. Rev. 62 (1942) 438; Phys. Rev.
63 (1943) 367; Phys. Rev. 76 (1949) 1352.
[8] A. R. Edmonds, Angular Momentum in Quantum Mechanics (Princeton,
1957).
[9] (a) L. C. Biedenharn and J. D. Louck, Angular Momentum in Quantum
Physics (Addison Wesley, Reading, Mass, 1981);
(b) L. C. Biedenharn and J. D. Louck, The Racah-Wigner Algebra in Quan-
tum Theory (Addison Wesley, Reading, Mass, 1981).
[10] B. G. Wybourne, Symmetry Principles and Atomic Spectroscopy (Wiley-
Interscience, N.Y., 1970).
[11] B. R. Judd, Atomic Shell Theory Recast, Phys. Rev. 162 (1967) 28.
[12] J. Avery, Generalized Sturmians and Atomic Spectra, (World Scientific, Sin-
gapore, 2006).
[13] P. J. Redmond, Phys. Rev. 133 (1964) B1352.
[14] H. A. Erikson and E. L. Hill, Phys. Rev. 75 (1949) 29.
[15] J. von Neumann and E. Wigner, Z. Phys. 30 (1929) 467.
[16] L. I. Ponomarev and T. P. Puzynina, Soviet Physics JETP 25 (1967) 846.
[17] C. E. Wulfman and S. Kumei, Science 172 (1971) 1061.
[18] L. Salem, Isr. J. Chem. 12 (1979) 8.
October 28, 2010 16:55 9in x 6in b943-ch12
CHAPTER 13
Rovibronic Systems
13.1 Introduction
Lie-algebraic methods were introduced into the study of the spectra of
rotating and vibrating molecules by physicists who had contibuted to
the development of the liquid drop model of nuclei and its connections
to models composed of discrete nucleons. Their early success in nuclear
physics led to several thoroughly developed mathematical models of rovi-
brational motion: the harmonic oscillator model, the symplectic model and
the vibron model. For information about the first two, the reader is referred
to reviews by Moshinsky and Rowe.1,2 A U(6) vibron model was intro-
duced in nuclear physics by Arima and Iachello,3 and Iachello introduced
a U(4) vibron model into molecular physics.4 They have been applied to
diatomics,5 triatomics,6 tetra-atomics,7 and to molecules at least as com-
plex as benzene.8 Several monographs have been devoted to applications of
the model in nuclear physics and molecular physics.9–11
The vibron models are semi-empirical ones, and a major criticism leveled
against their application to molecular physics has been that there does not
appear to be a definitive connection between their Lie-algebraic structure
and that associated with Schrödinger equations with definite potentials.
Nevertheless, they represent an outstanding example of the inductive use
of the mathematics of Lie algebras in the sciences. We will concentrate
attention on the connection they establish between molecular spectroscopy
and Lie algebras.
389
October 28, 2010 16:55 9in x 6in b943-ch13
(a)
(b)
Fig. 13.2.1 (a) Morse potential function; (b) Morse energy surface.
seen in Sec. 10.7, for the Morse oscillator, the m = 0 case is excluded, so in
the second case there are j energy levels.
To make contact with the vibron model which originated in nuclear
shell theory, one expresses the group generators as binary products of
boson creation and annihilation operators b†1 b1 , b†2 b2 , b†1 b2 , and b†2 b1 .
October 28, 2010 16:55 9in x 6in b943-ch13
Nop ≡ Nopa + Nopb = b†1a b1a + b†2a b2a + b†1b b1b + b†2b b2b , (13.3.2a)
Bopp and Haag have proposed that the D functions with half-integer j
values could provide spatial realizations of the wave functions of particles
with half-integer spin.21
To determine the dynamical group of (13.4.1), Kumei22a began by using
Dop = exp{t∂/∂t ln((Jop + 1/2)/H)} (13.4.4a)
to linearize the spectrum of j. In (13.4.4a),
Jop + 1/2 ≡ (1/2){1 + 8I1 [H + ((I1 − I3 )/2I1 I3 )∂ 2 /∂γ 2 ]}1/2 . (13.4.4b)
The dilatation operator Dop converts (13.4.1) to
{(1 − x2 )∂ 2 /∂x2 − 2x∂/∂x + (1 − x2 )−1 (∂ 2 /∂α2 + ∂ 2 /∂γ 2 )
−2x(1 − x2 )−1 )∂ 2 /∂α∂γ + (i∂/∂t)2 − 1/4}Φ = 0.
October 28, 2010 16:55 9in x 6in b943-ch13
with
Φ= Cjmn φjmn , φjmn = Ψjmn (α, β, γ) exp(−i(j + 1/2)t).
(13.4.5a,b)
Setting up and solving the 14 determining equations required to obtain
first- and second-order differential operators yields 16 generators Q of the
invariance group of (13.4.5). One of these is the generator iΦ∂/∂Φ of the
group that can change the phase of Φ. The other 15 were denoted Qjk . They
obey the SL(4,R) commutation relations
[Q ij , Q k
m]
i
= δm Q k
j − δjk Q i
m. (13.4.6)
Here we summarize some of the general properties of this dynamical group.
For a fuller discussion the reader is referred to the original paper.22b
Converting the Q kj to skew-adjoint or self-adjoint generators of invari-
ance transformations of the original Schrödinger equation produces the gen-
erators of its dynamical group. The action of the resulting labeling operators
−i∂/∂α, −i∂/∂γ and i∂/∂t, on the eigenfunctions
φjmn = Ψjmn (α, β, γ) exp(−i(j + 1/2)t)
(13.4.7)
= ((2j + 1)1/2 /4π)Djmn (α, β, γ) exp(−i(j + 1/2)
is given by
−i∂/∂α φjmn= mφjmn , −i∂/∂γ φjmn= n φjmn ,
i∂/∂t φjmn= (j + 1/2)φjmn .
Kumei found that all of the generators which shift j shift it by 1/2 unit.
They must consequently shift both m and n by 1/2 unit. The Lie algebra is
consequently that of the group SU(2,2). The group contains operators that
mix states with integral and half-integral values of j. The physical interpre-
tation of these mixed states present problems that have been addressed by
Bohm and Teese.23 Double-jump operators for j may be constructed from
products of the single-jump operators. However they do not close under
commutation with the other SL(4,R) generators.
We next consider the degeneracy groups of the systems. When all three
moments of inertia are equal, the resulting spherical top has SU(2) ⊗ SU(2)
as its degeneracy group. In the notation of Ref. 22 the shift and labeling
operators of the group are
Q9 = Q†10 = exp(iα){(1 − x2 )1/2 ∂/∂x − ix(1 − x2 )−1/2 ∂/∂α
+ i(1 − x2 )−1/2 ∂/∂γ}, (13.4.8a)
Q11 = Q†12 2 1/2 2 −1/2
= exp(iγ){(1 − x ) ∂/∂x − ix(1 − x ) ∂/∂γ
+ i(1 − x2 )−1/2 ∂/∂α}. (13.4.8b)
October 28, 2010 16:55 9in x 6in b943-ch13
It acts on the term exp(±iγ) in Q11 and Q12 , converting both generators
to functions of time.
The transformation from the generators Qa to Qab is a similarity trans-
formation. Application to all the generators, Qa , of the Lie algebra asso-
ciated with Hamiltonian Ha does not change their commutation relations:
the corresponding set of generators Qab are those of an isomorphic Lie
algebra. Its generators are the quantum mechanical representatives of time-
dependent constants of motion of system b. An analogous change occurs
when the symmetry of the top is reduced to that of the asymmetric top
whose degeneracy group is simply SO(3).
The inverse of the transformation Sab is the transformation S−1
ab , which
can convert time-dependent generators of the dynamical group of the sym-
metric top, to the generators Q11 and Q12 of the degeneracy group of
the spherical top. Usually one expects perturbations to shrink degeneracy
groups. Here one has a perturbation that enlarges a degeneracy group, and
does so in a most obvious manner. As pointed out in Chapter 1, if one does
not know the dynamical group of a system, the perturbations that produce
increases or changes in observed symmetries are very difficult to detect.
In 1984 Harter and Patterson introduced the concept of rotational
energy surface. Such surfaces depict the relation between the energy of
a rotating object and the orientation of its angular momentum vector in
a body fixed system. The angular momentum vector is assumed to have a
given direction and magnitude in a space-fixed system. For a classical top
with angular momentum of magnitude L2 and principal moments of inertia
Ix = 1/2A, Iy = 1/2B, Iz = 1/2C,
If the Euler coordinates of L in the body-fixed frame are −α, −β, −γ, then
the energy, E, of the top is given by the expression25
Figures 13.4 illustrate the resulting constant energy surfaces for prolate,
oblate and spherical tops. They are obtained by setting β and γ equal to the
polar coordinates −θ and −φ, respectively, then considering E = r(θ, φ).
Because the classical and quantum mechanical dynamical groups of these
October 28, 2010 16:55 9in x 6in b943-ch13
If w is the SO(4) quantum number with values 1, 2,. . . , then the rovibra-
tional state |N wlm will have
Ea = E0 + α1 n + α2 n(n + 3),
(13.5.5)
E0 = −(1/4)(β 2 /2µ), α1 = (5/2)(β 2 /2µ), α2 = −(β 2 /2µ).
If one wishes to generalize to the l = 0 case, one must also take into account
the fact that n is an integer which jumps by unity from 0 up to N , while |m|
is half integer and jumps from 1/2 to j. This can only be arranged by setting
|m| = n + 1/2, and N = j + 1/2. This, however, requires l to be shifted by 2
units, so the change in l is that of a two-photon transition. This indicates
that model (a) should be interpreted as applying to systems other than a
Morse oscillator. For further discussion of this algebraic Hamiltonian the
reader is referred to the review by Oss.26
To identify Hb with a Morse Hamilton when l = 0, one must take into
account that w is an integer which jumps by two units from 0 to N . This
requires that |m| = w/2 + k, where k = 0 if N is odd and k = 1/2 if N is
even. One then finds that A = −(1/4)(β 2 /2µ) = −(k/2)(β 2 /2µ), so k must
be chosen as 1/2. It follows that N must be an even number. Also, E0 must
have the same value it has under case (a). Consequently
As in the atomic case considered in previous chapters, the SO(4) Lie algebra
contains operators that shift l by one unit.
October 28, 2010 16:55 9in x 6in b943-ch13
In both cases (a) and (b), the energy expression is the sum of a kinetic
energy of rotation and an internal vibrational energy. As the internal energy
of the vibration motion increases, the number of rotational states allowed to
the bound system must decrease. If ν is the vibrational quantum number,
as it increases n and w must decrease. This is accommodated in model (a)
by setting n = j − ν − 1/2. In model (b) one sets (w + 1)/2 = j − ν.
The generators of U(4) and SO(4) that couple U(2) ⊗ U(2) relate the
vibrational and rotational motions, while converting product wave functions
or the form R(r)Ylm (θ, φ) to functions Z(r, θ, φ) that could be expressed as
a sum of functions of the general form Rl (r)Ylm (θ, φ). The existing work
on the U(4) model is examined in the monograph of Iachello and Levine,
who conclude that chain (b) “provides a reasonably good description of the
spectra of rigid diatomics”, but “quite often the acual situation deviates
somewhat from that of a simple Morse oscillator.” They find that some
general expansions which are polynomial functions of the Casimir operators
of model (b) converge more rapidly than standard Dunham expansions.
They then add to their Hamiltonian the coupling term in the quadratic
Casimir operator of U(4)12 . This removes the degeneracies of the local-
modes that it requires only minor degenerate perturbation calculations to
accomodate. This is because the term only couples states with identical
SO(4)12 quantum numbers, i.e., with τ1 = τ2 . They find that with this
sophistication, the algebraic Hamiltonian of chain IIIa determines overtone
frequencies of the nonlinear molecules H2 O16 , H2 O18 , D2 O16 , H2 S32 and
S32 O16
2 , with root mean square deviations of 1 to 5 cm
−1 16c
. A similar, but
slightly more elaborate treatment of the spectra of N2 O, C12 O2 , C13 O2 ,
OCS and HCN, produced results of the same accuracy.16d
Harter and his coworkers developed extensions to semi-rigid rotors of
the energy surfaces of rigid rotors noted above. This produced amazing
insights that led to a number of useful simplifications in treatments of the
rotation of systems at least as complex as SF6 .3
Exercises
1. Hold a claw hammer by its handle, with its head–claw axis in a horizon-
tal plane, and note the direction in which the head points. Throw the
hammer up in the air, giving it a rotation about its intermediate axis of
inertia. Catch it by its handle as it descends after one revolution about
this axis. In what direction does the head point? The figures on page
327 of Harter’s article, Ref. 25, can help in developing an explanation of
your observations.
References
[1] M. Moshinsky, Group Theory and the Many-Body Problem (Gordon &
Breach, New York, 1967).
October 28, 2010 16:55 9in x 6in b943-ch13
CHAPTER 14
In his 1905 paper introducing the theory of relativity, Einstein prefaced his
arguments with the observation, “. . . the unsuccesful attempts to discover
any motion of the earth relative to the (aether), suggest that the phenomena
of electrodynamics as well as mechanics possess no property corresponding
to the idea of absolute rest.”1 He then made four key assumptions in devel-
oping his theory of special relativity. These are:
i. Sets of rigid rods and clocks may be used to define sets of local coordi-
nate systems that assign identical space and time intervals to successive
events.
ii. The rods and clocks are not affected by motions at constant velocity
with respect to one another.
iii. Measurements made using these comoving coordinate systems are com-
pared via electromagnetic radiation.
iv. Electromagnetic radiation moves with a constant velocity, c, with
respect to all observers.
409
October 28, 2010 16:55 9in x 6in b943-ch14
question arises whether the theory of relativity also holds for the types of
motion of an electromagnetic system derived from one another by such a
transformation.”
The following pages briefly describe the transformations of Maxwell’s
equations carried out by their Poincare invariance group, and their physical
interpretation. The action of the Bateman–Cunningham inversion transfor-
mation is then described, and questions associated with its physical inter-
pretation are considered. The primary concern throughout this chapter
is the physical interpretation of invariance transformations of Maxwell’s
equations — none of which can change the velocity of light in vacuo.
equation
Also,
H · E = 0. (14.1.4c)
i.e. by (14.1.8)
Cunningham and Bateman showed that Eqs. (14.1.4) and (14.1.5) are
invariant under Lorentz transformations, and that the same is true of
Eq. (14.1.2). Figures. 14.1.1a,b depict, in the frame of a, the electric
field E = Ey k, between the plates of a capacitor, and the magnetic field
H generated by a current running through a wire on the x axis a. Fig-
ures. 14.1.1c,d depict the fields E = and H that are observed by b,
moving at velocity v = 0.7ci with respect to a.
1.0 1.0
0.9 0.9
0.8 0.8
0.7 0.7
0.6 0.6
0.5 0.5
0.4 0.4
0.4 0.5 0.6 0.7 0.8 0.9 1.0 0.4 0.5 0.6 0.7 0.8 0.9 1.0
(a) (b)
1.0 1.0
0.9 0.9
0.8 0.8
0.7 0.7
0.6 0.6
0.5 0.5
0.4 0.4
0.4 0.5 0.6 0.7 0.8 0.9 1.0 0.4 0.5 0.6 0.7 0.8 0.9 1.0
(c) (d)
Fig. 14.1.1 (a) Electric field in the y–z plane between plates of a capacitor;
(b) magnetic field surrounding a wire along the y-axis; (c) the electric field of
Fig. 14.1.1a observed in the reference frame of a moving observer; (d) the magnetic
field of Fig. 14.1.1.b observed in the reference frame of a moving observer.
converts solutions of Maxwell’s equations into solutions. For all real values
of k, one may define the dimensionless four-vector
s = s /s2 (14.2.2c)
Thus, IST is its own inverse.
In discussing the effect of the inversion (14.2.1) on electromagnetic fields,
it is useful to define
σ = (s1 , s2 , s3 ) = (x, y, z)/k = r/k, (14.2.3a)
τ = s4 , (14.2.3b)
2
u = s21 + s22 + s23 + s24 2
=σ +τ . 2
(14.2.3c)
Cunningham3a showed that the inversion converts solutions E and H of
Maxwell’s equations, to solutions E and H , which may be defined by
E = (−s2 ){u2 E − 2(σ · E)σ + 2τ (σ × H)}, (14.2.3d)
and
H = (s2 ){u2 H − 2(σ · H)σ − 2τ (σ × E)}. (14.2.3e)
This result holds true for all choices of the origin of the spacetime coordi-
nates for which s does not vanish.
Figures 14.2.1 illustrate the effect of inversions on the fields E and H
of Figs. 14.1.1a,b. Figures 14.2.2 depict the corresponding solutions, E
and H , of Maxwell’s equations produced by the relations (14.2.3).
In dealing with the effect of the Cunningham–Bateman inversion on the
Lorentz metric
dx2 + dy2 + dz2 − d(ct)2 = dx21 + dx22 + dx23 − dx24 , (14.2.4a)
we will usually use the dimension free metric
ds2 = (dx2 + dy2 + dz2 − d(ct)2 )/k2 . (14.2.4b)
October 28, 2010 16:55 9in x 6in b943-ch14
1.0 1.0
0.9 0.9
0.8 0.8
0.7 0.7
0.6 0.6
0.5 0.5
0.4 0.4
0.4 0.5 0.6 0.7 0.8 0.9 1.0 0.4 0.5 0.6 0.7 0.8 0.9 1.0
(a) (b)
1.0 1.0
0.9 0.9
0.8 0.8
0.7 0.7
0.6 0.6
0.5 0.5
0.4 0.4
0.4 0.5 0.6 0.7 0.8 0.9 1.0 0.4 0.5 0.6 0.7 0.8 0.9 1.0
(c) (d)
Fig. 14.2.1 (a) Electric field in the y–z plane between plates of a capacitor;
(b) effect of the inversion on the electric field at t = 0; (c) effect of the inversion
on the electric field at t = 0.5c; (d) effect of the inverson on the electric field at
t = 0.9c.
1.0 1.0
0.9
0.8
0.8
0.6
0.7
y
y
0.4
0.6
0.2
0.5
0.4 0.0
0.4 0.5 0.6 0.7 0.8 0.9 1.0 0.0 0.2 0.4 0.6 0.8 1.0
x x
(a) (b)
1.0 1.0
0.9 0.9
0.8 0.8
0.7 0.7
y
0.6 0.6
0.5 0.5
0.4 0.4
0.4 0.5 0.6 0.7 0.8 0.9 1.0 0.4 0.5 0.6 0.7 0.8 0.9 1.0
x x
(c) (d)
Fig. 14.2.2 (a) Effect of spacetime inversion on the electric field between plates
of a capacitor, ct = 0; (b) effect of spacetime inversion on the electric field between
plates of capacitor, ct = 0; (c) effect of spacetime inversion on the magnetic field
surrounding a current-carrying wire, ct = 0; (d) effect of spacetime inversion on
the magnetic field surrounding a current-carrying wire, ct = 1.
They satisfy the commutation relations of the Poincare group, cf. (14.1.7),
and the following commutation relations:
one finds
With this result in hand, one can easily obtain all the finite transformations
of the full 15 parameter conformal group. The operations of this continuous
group do not, in general, leave the Lorentz metric invariant. The group is
(see the exercises) locally isomorphic to a SO(4,2) group.
Cunningham argued that, just as in the case of the Poincare trans-
formations of special relativity, the inversion and special conformal trans-
formations of Maxwell’s equations could not make it possible “for an
observer to discriminate between the sequence of electromagnetic phenom-
ena as he knows them and the sequence obtained by this transforma-
tion.” Cunningham arrived at this conclusion before the advent of quantum
mechanics. A fundamental contribution by Barut and Haugen12 implies
that it can only continue to be correct if one replaces the usual concept of
(Poincare invariant) mass by a concept of mass that is invariant under the
subgroup of special conformal transformations. Barut and Raczka point
out that the operations of this subgroup of the Bateman–Cunningham
invariance group of Maxwell’s equations also leave invariant relativistic
wave equations for massless particles of spin 0 and 1/2.12 However, wave
October 28, 2010 16:55 9in x 6in b943-ch14
equations for massive particles are only invariant under its actions if the
mass m is also transformed by a dilation operator that also carries out the
transformation
and
with
Thus
Thus
Note that the conformal metric smoothly reduces to the Minkowski metric
as the group parameter α approaches zero. This metric is specific to C4 . As
it is the only special conformal metric we will deal with in the remainder
of the chapter, we will sometimes, simply call it the conformal metric.
The function γ introduces a relation between coordinates that destroys
the Poincare invariance of spacetime with metric ds2 . In this respect the
conformal metric is qualitatively similar to the Robertson–Walker metric.
However, the conformal metric produces a local multiplicative scaling of
both dr and dx4 , while the Robertson–Walker metric only applies such a
rescaling to dr. Most importantly, in the conformal metric the choice of
spacetime origin is arbitrary. However because the origin r = 0, t = 0, is
an invariant point of the conformal transformation, there one also has the
special property r = 0, t = 0.
October 28, 2010 16:55 9in x 6in b943-ch14
Consequently,
This Doppler shift will not be observed if the propagation of the wave
has taken place in a space with Minkowski metric: it may be considered
to have arisen because the conformal wavelength, λ , changes as the wave
propagates.
October 28, 2010 16:55 9in x 6in b943-ch14
When the observer has velocity V = dR/dT relative to the source, then
the wavelength λ in the (R,T) system undergoes the additional Doppler
shift given by the usual formula
1/2
1 + V/c
∆λ/λ = − 1 = V/c + O(V2 /c). (14.3.19)
1 − V/c
If the sign of V/c is positive when the source is receding from the observer,
a reshift is produced. On neglecting terms of O((αR)2 /c) and O(V2 /c),
(14.3.18) then yields Hill’s relation
V = αR + V. (14.3.20)
if the metric in which the waves move is the special conformal metric of
(14.3.12). To understand the consequences it is helpful to reorganize Hill’s
approximate equation, which for V c, can be written as
VC − VM = αR. (14.3.23a)
As the distance R traversed by the radar waves was twice the distance r
in (14.3.22), in Eq. (14.3.23c) one has −2αr = −(2.8 ± 0.42) × 10−18 sec−1
r, so α = (1.4 ± 0.42) × 10−18 sec−1 . This converts (14.3.23a) to
VM − VC = αR , i.e. VC
− VM = −αR . (14.3.23b,c)
Conclusion
Few physical implications of Lie’s discoveries were noticed prior to the dis-
coveries of Bateman and Cunningham. Hill seems to have been one of the
first to investigate possible physical consequences of their conformal group.
The increasing interest in the role of group theory in quantum mechan-
ics that began in the 1960’s led to an increasing recognition of possible
consequences of the conformal symmetry of Maxwell’s equations at the
submicroscopic level. Some applications of the conformal group in quan-
tum theory abandon the requirement of Lorentz invariance on the submi-
croscopic scale. Hill’s work and the Pioneer results suggest that Poincare
invariance of electromagnetic phenomena on the supermacroscopic scale
should not be assumed. Though no conclusions as startling as that of Hill
have emerged, a number of workers have investigated relativistic and non-
relativistic consequences of conformal invariance.25 If further determina-
tions of the conformal group parameter α show that it is nonzero, much
physics will have to be revised. If it is found that α has a value of the same
order of magnitude as Hubble’s constant, the consequences for physics and
cosmology will be even more extensive.
Over a century has passed since Bateman and Cunningham established
the conformal invariance group of Maxwell’s equations. However, as has
become evident in this chapter, many fundamental physical implications of
this invariance have yet to be established.
Exercises
1. Use the extended generators of the Poincare group to show that the
group leaves the Lorentz metric (14.2.4a) invariant.
October 28, 2010 16:55 9in x 6in b943-ch14
References
[1] A. Einstein, Ann. Phys., 17 (1905) 891.
[2] H. Lorentz, Proc. Amstrdam Acad. 6 (1904) 809.
[3] (a) E. Cunningham, Proc. London Math Soc. 8(2) (1909) 77;
(b) H. Bateman, ibid. 223;
(c) H. Bateman, Proc. London Math Soc. 7(2) (1909) 70.
[4] J. C. Maxwell, Electricity and Magnetism (Cambridge University Press,
Cambridge, 1873).
[5] a) W. F. Magie, Source Book in Physics (McGraw-Hill, N.Y., 1935), p. 408;
(b) E. Whittaker, A History of the Theories of Aether and Electricity
(Thomas Nelson and Sons, Edinburgh, 1951), p. 83.
[6] W. D. Niven, ed., The Scientific Papers of James Clerk Maxwell (Cambridge
University Press, Cambridge, 1880); photographic reprint (J. Hermann,
Paris), (Steichert, N.Y., 1930).
[7] L. P. Wheeler, Josiah Willard Gibbs (Yale University Press, New Haven,
1952).
[8] J. A. Kong, Electromagnetic Wave Theory, 2nd edn. (Wiley Interscience,
New York, 1990).
[9] J. C. Maxwell, Trans. Royal. Soc. (London) CLV, (1864), Collected works,
XXV, A dynamical theory of the electromagnetic field, Part I, Sec. (20).
[10] S. Lie, Transformations Grupen, III (Leipzig, 1893), p. 351; reprinted
(Chelsea, New York, N.Y., 1970).
[11] B. G. Wybourne, Classical Groups for Physicists, (Wiley, New York, 1974),
pp. 345–7.
[12] (a) A. O. Barut and R. Raczka, Theory of Group Representations and Appli-
cations, 2nd revised edn. (World Scientific, Singapore, 1986);
(b) A. O. Barut and R. B. Haugen, Ann. Phys. 71 (1972) 519.
[13] T. Fulton, F. Rohrlich and L. Witten, Rev. Mod. Phys. 34 (1962) 442.
[14] C. Wulfman, J. Phys. A: Math. Theor. 42 (2009) 185301.
[15] H. A. Kastrup, Ann. Phys. 17 (2008) 631.
[16] M. Born, Einstein’s Theory of Relativity, translated by H. Brose (Methuen,
London, 1924), p. 240.
[17] E. Hubble, The Realm of the Nebulae (Yale University Press, New Haven,
1936).
[18] (a) P. J. E. Peebles, Principles of Physical Cosmology (Princeton, 1993),
pp. 24, 71, 82;
(b) ibid., pp. 20, 106–7;
(c) ibid., pp. 73, 4.
[19] E. L. Hill, Phys. Rev. 68 (1945), 235.
[20] C. Wulfman, [arXiv: 1003.0507]; [arXiv: 1010.2139] (2010).
October 28, 2010 16:55 9in x 6in b943-ch14
Index
433
October 28, 2010 16:55 9in x 6in b943-Index
Index 435
Index 437
Sturmians, 340, 349, 350, 380–384 tangent vector, 111, 112, 115, 119
sudden polarization, 366 topology, 132
superconductivity, 15 triatomic molecules, 367
superposition invariance, 275
symmetry breaking and creation U(4) ⊗ U(4) subgroup chains, 404
Bernard convection, 16 U(4) subgroup chains, 401
cell differentiation, 15 unit circle
crystallography, 15 invariance transformation, 137
crystals, 15 united atom and separated atom
fluid flow, 16 correlation diagram, 367
geology, 16
Hartree–Fock calculations, 11 vector, 62, 63
reaction–diffusion equation, 16 vector fields, 110–112
superconductivity, 11 vibron models, 389
symmetry in phase space, 92–95
symplectic diffeomorphism, 76 Walsh’s rules, 368
symplectic space, 82, 91 wavelength
symplectic transformations, 80–84, conformal, 425
86–90, 95–97, 210–214 wedge product, 86
Whenever systems are governed by continuous chains of causes
and effects, their behavior exhibits the consequences of dynamical
symmetries, many of them far from obvious. Dynamical Symmetry
introduces the reader to Sophus Lie’s discoveries of the connections
between differential equations and continuous groups that underlie
this observation. It develops and applies the mathematical relations
between dynamics and geometry that result. Systematic methods
for uncovering dynamical symmetries are described, and put to
use. Much material in the book is new and some has only recently
appeared in research journals.
World Scientific
ISBN-13 978-981-4291-36-1
ISBN-10 981-4291-36-6