Sei sulla pagina 1di 19

Relativistic Angular Momentum

Nick Menicucci
(Dated: December 13, 2001)
1
I. INTRODUCTION
Conserved quantities are of the utmost importance in physics because they indicate a
symmetry that is built into nature. Translational isotropy begets the conservation of lin-
ear momentum. Time-translation symmetry gives rise to energy conservation. Rotational
isotropy gives birth to the conservation of angular momentum.
With the advent of Special Relativity near the turn of the last century, more symmetries
were discovered that reshaped our thinking of the universeliterally! Space and time, once
thought independent, have been merged inseparably into a single 4-dimensional spacetime.
Mass and energy have similarly been intertwined in Einsteins famous equation, E = mc
2
.
Force and power have suered a similar fate, becoming inexorably meshed in the 4-force.
Three-dimensional momentum has been augmented by a fourth term involving the mass
(energy) of the particle to beget the 4-momentum.
Angular momentum has not been forgotten in this universal conjugation of physical
quantities. Its partner is the mass moment (mass distance), which is proportional to the
center of mass, or centroid. Unlike in its classical formulation, the center of mass is not the
same in all frames, and this leads to interesting consequences for the size of nite systems
with angular momentum about the centroid, namely that the system has a minimum size
beyond which it cannot be compressed and still maintain the angular momentum present.
Additionally, the conservation of angular momentum can be seen to imply and be implied by
the uniformity of motion of the proper centroid (i.e., the centroid in the center-of-momentum
frame), thus linking these two physical principles into one law of the conservation of the
angular momentum 2-form.
Special Relativity arguably has as its crowning achievement the unication of many once-
believed separate concepts, including space and time, mass and energy, force and power,
angular momentum and center of mass. This paper focusses on the last of these unied
concepts by examination of the properties of the angular momentum 2-form. Analogy to
the classical angular momentum, relation to the classical center of mass, and some of the
consequences of this relativistic formulation will be discussed.
2
II. THE ANGULAR MOMENTUM 2-FORM
We begin with the classical expression for angular momentum in 3-space:

h r p . (2.1)
This denition of the angular momentum 3-vector

h is immediately useful in the relativistic
formulation if we dene p mv, the relativistic 3-momentum, as will be shown presently.
By its cross-product character, we can make an educated guess that it might be more useful
to write

h as a 1-form, h

. (In at 3-space, h

has the same components as



h.) The 1-form
equivalent of Eq. (2.1) is
h

=
3
(r

) , (2.2)
or perhaps more usefully,

3
h

= r

. (2.3)
The relativistic extension of this equation is generated simply by replacing the 3-dimension
1-forms, r

and p

, by their 4-dimensional counterparts, X

and P

, which are dened, as usual,


as
X

, t

and P

, m

, (2.4)
where, also as usual, we take c = 1 and use the metric
H =

+1 0 0 0
0 +1 0 0
0 0 +1 0
0 0 0 1

.
We then make the denition
L

. (2.5)
This is the relativistic version of Eq. (2.1), and it collects information about the relativistic
angular momentum into the 2-form L

. It is obviously Lorentz-invariant because it is created


from two Lorentz-invariant 1-forms. That the relativistic angular momentum is presented
as a 2-form is not really all that surprising, considering that any 1-form in 3-space can be
written as the dual of a 2-form, as in (2.2). Conversely, any 2-form in 3-space will generate
a unique vector based on the components of its dual. In 4-space, however, we have no such
3
luxury, since the 4-dimensional dual of a 2-form results in another 2-form. Nevertheless,
dispensing with the cross-product luxury of mapping two vectors onto a third, we can still
use the philosophy behind the creation of

h to dene a relativistically equivalent 2-form that


does the same job.
There is an obvious relationship between L

and h

, due to the similarity in their denitions:


the upper 3 3 submatrix, L
ij
, has the same components as r

=
3
h

. The remaining
3 independent components of this tensor are components of the mass moment (relativistic
mass-energy position vector) of the particle with respect to the origin at time t. When the
angular momentum is summed over a system of particles, this becomes an important and
useful quantity from which we may calculate the centroid, or center of mass, of the system.
We denote this 3-component object by
m(r vt) (2.6)
and note that it appears in the tensor L

as follows:
L

3
h

T
0

. (2.7)
Note that L

is dened for a particle with respect to an implicit origin O = (0, 0, 0) and


at a particular time t. For a single particle, this presents no problems, but when the total
angular momentum of a system is calculated in dierent frames, issues of simultaneity arise
(See Section III).
In nonrelativistic mechanics, angular momentum is conserved, so one might intuitively
believe that L

is constant in time, as well. However, it is useful to examine this point with


more scrutiny:
The relevant quantity in this case is the time-rate of change of L

, or
dL

dt
=
dX

dt
P

+X

dP

dt

dX

dt
P

dP

dt
. (2.8)
Assuming constant velocity, the P-derivatives drop out, leaving
dL

dt
=
dX

dt
P

dX

dt
P

= mU

mU

= 0 . (2.9)
Thus, L

is constant in time (i.e., conserved), just like its 3-dimensional counterpart. How-
ever, as we shall discover in Section IV, we get two conservation rules for the price of one
4
from this factthe uniform motion of the centroid is also implied by Eq. (2.9). We shall see
this later.
III. ANGULAR MOMENTUM FOR A SYSTEM OF PARTICLES
In the previous section, we generalized the concept of angular momentum to four dimen-
sions. The resulting object was a 2-form that included the 3-dimensional angular momentum
vector

h (in some suitable representation) and the 3-vector mass moment of the particle
(relativistic mass location vector). We also noted that we calculated the angular mo-
mentum 2-form with respect to a particular spatial origin and at a particular instant in
time.
We wish now to extend this notion of angular momentum to a system of particles.
1
We
will assume that these particles are subject to no forces other than collisions. Thus, total
4-momentum should always be conserved, and the individual (relativistic) angular momenta
of the particles should be conserved between collisions.
In dening the total angular momentum for this system of particles, however, one might
hastily assume that it is possible simply to add the angular momenta for all the particles
together to obtain the total, as long as we use the same origin and the same time coordinate
for all particles. While this is a valid mathematical operation, it is not immediately obvious
that the resulting object is a Lorentz-invariant 2-form. The problem lies in the phrase the
same time coordinate for all particles. In another frame, the set of particle locations used
in the previous frame are no longer simultaneous. A dierent set of locations will be used
based on this frames denition of simultaneous.
This point, while important to consider, turns out to be moot for the following reason.
Imagine two frames, S and S

, moving with some relative velocity. In frame S, the particles


locations will be determined by where their worldlines intersect a particular hypersurface
of simultaneity that we will call , as determined by an observer in S. An observer in S

will make an analogousbut dierentdetermination of the particles locations based on


a dierent surface of simultaneity, which we will call

. However, by our assertion that


all of the particles 4-momenta are constant between collisions, we can imagine the smooth
1
This derivation closely follows that of Rindler, sec. 34.
5
rotation of

into . As this surface sweeps across the worldlines of the particles, as


long as it crosses no collision events, the angular momenta of all particles intersecting the
surface will remain constant, by Eq. (2.9). In addition, if the sweeping operation does cross
a collision event, the total 4-momentum is still conserved across the collision. Furthermore,
at a collision event, the locations of the participating particles must be the same (since
they are colliding). This leads to the special form of the angular momentum summation for
collision events,


collision
X

, (3.1)
for the participating particles. Therefore, even though we cannot immediately conclude
that the angular momentum of each individual particle is conserved during a collision, by
Eq. (3.1) and by the conservation of

, the total angular momentum is still conserved


across a collision. Finally, since the individual angular momenta are unchanged by sweeping
over particles when they are freely moving, we may choose our sum using any surface of
simultaneityregardless of whether it is appropriate for that frame or notand still obtain
the same (frame-specic) value for

. Therefore, we may conclude that

is both
Lorentz-invariant and constant in time.
Now that we have determined that we may create a Lorentz-invariant total angular mo-
mentum 2-form by summing individual angular momenta of particles, lets do so and examine
its properties. We dene the total (relativistic) angular momentum to be
L

. (3.2)
We may, however, at some point wish to calculate the angular momentum of a continuous
medium. It is perhaps useful to note, then, that this 2-form can also be written in terms of
the stress-energy tensor T

, which contains the momentum density of the system in its T


i4
components. Thus, Eq. (3.2) becomes
L

dx dy dz (x

T
4
x

T
4
) , (3.3)
where the integration is over a surface of simultaneity (and hence, we use dx dy dz). This is
the continuous limit of Eq. (3.2) above.
Since we noted above that any surface of simultaneity will do, with our continuous limit
equation (3.3) above, we see that we can calculate L

over any spacelike hypersurface o.


6
Thus, we may write the total angular momentum as
L

S
d
3

, (3.4)
where

, (3.5)
and

is the dual of a volume 3-form. This integral assumes that o is a spacelike surface,
and by the plane-tilting argument above, the calculated value at each point is independent
of the particular surface of simultaneity on which it is calculated. Thus, we may approximate
the surface o by a set of innitesimal hypersegments of constant-t surfaces, and the validity
of Eq. (3.4) holds. Furthermore, the law of conservation of angular momentum can be
written as the null-result of integrating about a closed hypersurface:

V
d
3

= 0 , (3.6)
where 1 is a 4-dimensional volume enclosed by the surface ( 1).
Now that we have many ways of dening the total relativistic angular momentum, an
interesting question to ask is, can we write it in a form similar to that of a single particle?
The answer is no, but going through a simple example of why this is the case illustrates
an interesting connection inherent to the form of L

. We will attempt to write the total


angular momentum 2-form in the following form:
L

= X

,
where X

is any generalized coordinate 1-form, and similarly for P

. Now imagine a 2-
particle system in the low-velocity limit (v c), for simplicity.
2
These two particles both
have mass m, and in a frame S at time t, they are located at r
1,2
= r e
x
, travelling with
velocities v
1,2
= v e
y
(particle 1 gets the + signs). We now calculate the components of
L

in terms of

h = 2mvr e
z
, and

= 0 .
2
Taking the low-velocity limit loses no generality in our conclusion: our result is false for this special case
and therefore cannot be true in general.
7
We want to write these expressions in the form

h = r

, and
= m

t ,
which are the low-velocity versions of Eqs. (2.1) and (2.6), respectively. That this is impos-
sible can be seen by substituting
r

=

m

+
p

t ,
from the second of the above equations, into the rst, yielding

h =
p

= 0 ,
a contradiction.
In this example, the centroid (center of mass) of the system is at rest, yet there is still
angular momentum present. Such angular momentum in this situation is therefore called
spin angular momentum. More importantly, even though we used the low-velocity limit for
simplicity in our example, it is the relativistic linkage of mass moment (or center of mass)
and (classical) angular momentum that tells us immediately that we cannot treat the spin
angular momentum as if it were due to a single particle. Without this linkage, we might
have postulated any number of combinations of r

, p

, and m

to satisfy the requirement


that

h = 2mvr e
z
. The requirement for a consistent eliminates all such possibilities. In
the next section, we examine this connection more closely.
IV. THE CENTROID
In the previous section, it was discovered that the coupling of mass moment (related
to center of mass) with the classical angular momentum

h leads to an intrinsic dierence
between one-particle angular momentum and the spin angular momentum of a composite
system of particles. Let us look more closely at just why this is.
We start by writing
= mr
C
pt , (4.1)
8
where
m

m , (4.2)
p

p , and (4.3)
r
C

mr/m. (4.4)
The 3-vector r
C
is location of the centroid, or center of mass, of the system in a given frame.
Note that this is the relativistic center of mass-energy and is therefore frame-dependent.
Returning to the two-body example in the previous section, and taking v c, we see that in
the original frame, the centroid of the system is still at O. However, if we were to transform
to the rest frame of particle 1, moving with velocity +v e
y
, synchronous with the original
frame at t = t

= 0, we would nd at t

= 0 that the centroid has shifted to the negative side


of x-axisi.e., toward the other, now more energetic, particleby a factor of ( 1)/( +1)
times the original distance r from the origin, where is that of particle 2 as measured by
particle 1. That the centroid is frame-dependent is an important property that leads to the
inability to shrink a rotating system of particles to an arbitrary size. This will be discussed
more in Section VI.
For now, it is useful to look at how the constancy of L

aects the centroid. After all, its


location is derived from , so the fact that
d
dt
L

= 0 should have a calculable eect, and it


does:
d
dt
L

= 0
d
dt
= 0 , (4.5)
which implies that
dr
C
dt
=
1
m
d
dt
( + pt) =
p
m
. (4.6)
This shows that the centroid moves uniformly with the same velocity as that of a particle of
mass m and momentum p. It is interesting to point out that the constancy of L

has two
consequences: not only does it imply the constancy of

h, the classical angular momentum,
it also implies uniformity of the centroids motion. This is another beautiful example of the
kind of 2-for-1 deals that relativistic unication brings to the laws of physics.
There is another way to determine that the centroid must move uniformly if
d
dt
L

= 0,
and it is arguably a more powerful reason. It is based on the zero-component lemma,
described below. Briey, this lemma states that if any one o-diagonal component of a
9
Lorentz-invariant 2-form is zero in all frames, then the entire tensor must be zero. The
proof for the zero-component lemma is as follows:
Given a Lorentz-invariant 2-form T

, choose an o-diagonal component to be 0 in all


frames. (The choice of component is arbitrary because a rotation can make another of the
components take its place in the matrix.) We choose T
12
= T
21
= 0. Next, apply each of the
following Lorentz transformations to the original tensor, and after the transformation, make
a note of which component moves into the new (1, 2) position or into any other position
designated to be 0 in all frames by a previous operation like thisthat component of the
original tensor then must also be 0:
R(

2
, e
x
) T
13
= T
31
= 0 ,
R(

2
, e
y
) T
23
= T
32
= 0 ,
B(v e
x
) T
24
= T
42
= T
34
= T
43
= 0 ,
B(v e
y
) T
14
= T
41
= 0 ,
and therefore, the entire tensor T is 0, and the lemma is proved.
By this lemma, and by the linkage of the mass moment and angular momentum vector

h in a Lorentz-invariant 2-form, conservation of angular momentum is one and the same


with uniform motion of the centroid:
d
dt

h = 0 leaves a 33 submatrix of zeroes in the upper-


left corner of
d
dt
L

in all frames; by the zero-component lemma, therefore, the remaining 3


independent elements must be zero, as well. Summarily, by the relativistic construction of
L

,
d

h
dt
= 0
dr
C
dt
=
p
m
= const. (4.7)
We mentioned above that the centroid is frame-dependent. However, there is one unique
frame, designated as the center-of momentum frame, or CM frame, that bears some scrutiny.
This frame is uniquely determined by the requirement that the total momentum of all
particles in the system sum to 0. By Eq. (4.7), however, we see that this is the same as
saying that the centroid is at rest:
p = 0
dr
C
dt
= 0 . (4.8)
The fact that the centroid is at rest in the CM frame is not surprising. Nevertheless, with
the variety of possible centroids (one for each frame), it is useful to designate one centroid
10
as the proper centroid. This honor is bestowed upon the centroid in the CM frame, which
is designated by the position 3-vector, r
PC
.
V. THE SPIN VECTOR AND THE PAULI-LUBANSKI VECTOR
While there is not uniformity of denition, both the so-called spin vector and the
Pauli-Lubanski vector are useful for representing the part of the angular momentum that is
intrinsic to the system (i.e., the spin angular momentum). While some texts treat these two
vectors as one, they are presented separately here with trivial (but important) dierences
in denition. Before we dive into dening these quantities, it would be good to discuss the
spin angular momentum a bit more. We begin with a discussion of how and where it is
calculatedin the CM frame.
An important aspect of the CM frame is that the angular momentum calculated therein
is independent of the pivot, which we will show presently. In Section II, we mentioned briey
that L

(and hence, L

) was dened with respect to a particular pivot, taken implicitly to be


the origin O = (0, 0, 0). This is not a necessary condition, and it is useful to know how to
move the pivot and calculate L

about a new location.


Let us designate this new location with a 1-form Y

. We will calculate the angular


momentum about the new pivot Y

as follows:
L

(Y

) =

(X

) P

= L

(O) Y

, (5.1)
where P

. This has the fortunate consequence that in the CM frame, where p = 0,


L
ij
(

Y ) = L
ij
(O)

h(

Y ) =

h(O) s , (5.2)
indicating that the spin angular momentum, s, dened in Section III as the angular mo-
mentum of the system in the frame in which the centroid is at rest (i.e., the CM frame), is
independent of the pivot about which it is measured.
With the pivot-independence of the spin angular moment established, we can proceed to
discuss its representation as a 1-form. In Section II, we discussed the requirement of using
11
forms in four dimensions, instead of cross-products of vectors, as in three dimensions. With
the spin angular momentum set apart as unique to the CM frame and its pivot-independence
establish, we continue now to dene a spin vector (dened as a 1-form)
SSS

(L

) , (5.3)
where U

/m. This 1-form contains all relevant information about the intrinsic, or spin,
angular momentum of a system. To see this, consider specializing to the CM frame, where
U

= (

0, 1), as we have already discussed. In this frame,


SSS

=
CM

s, 0

(5.4)
If we had just taken s above and tried to make a 1-form from it simply by giving it a fourth
component, we might not have ended up with a Lorentz-invariant quantity. SSS

, however, is
Lorentz-invariant, due to its construction as the dual of a (Lorentz-invariant) 3-form.
We can also reconstruct the spin angular momentum 2-form from the 4-velocity of the
proper centroid and the spin vector:
S

= i (SSS

) S

= SSS

, (5.5)
(Note the font dierence used to dierentiate between the 2-form S

and the spin vector 1-


formSSS

). Being calculated about the particles location, this 2-form has the same components
as that of L

calculated about the centroid of a system, namely,


S
ij
=
ijk
s
k
, and (5.6)
S
i4
= 0 . (5.7)
While we will see in Section VI that a nite system of particles cannot be reduced to a point,
we have dened S

based on the assumption that it represents the spin angular momentum


of a nite system. However, this progression, L

CM
S

is not the only way to arrive at an


intrinsic spin angular momentum 2-form. We may instead introduce directly the concept
of a point-particle with spin and require that its intrinsic angular momentum 2-form S

satisfy Eqs. (5.6) and (5.7). In addition, this particle is assumed to have a straight worldline
representing its history and a momentum

P parallel to this worldline, and it is assumed that
the momentum vector and spin 2-form satisfy P

= 0. This last requirement is derived


from the more general equation for a system of particles,
P

(r
PC
) = 0 , (5.8)
12
which is immediately derivable from the requirement that the proper centroid be at rest
analogous to Eq. (5.7) above, but using L
i4
= 0, instead.
With this notion of a point-particle with spin, we now introduce the Pauli-Lubanski
vector, which has a similar denition to that of the spin vector, namely,
W

(L

) = mS

, (5.9)
where S

is the spin vector. It nds its most important use in quantum eld theory. Because
of this, the 4-momentum (operator) is favored over the 4-velocity in its denition since the
former becomes the generator of 4-space translations in quantum eld theory. In such an
application, the spin angular momentum 3-vector s is taken as an intrinsic particle-spin
operator.
In quantum eld theory, everything becomes an operator: energy (the Hamiltonian)
and 3-momentum (combined into the 4-momentum operator), angular momentum, Lorentz
transformations, and spin. The square of the Pauli-Lubanski vector, W

= m
2
s
2
, ac-
counts for the last of these, where m is the mass of the eld particle, and s is the spin
operator, which, when squared, has eigenvalue s(s +1), where s is the spin of the eld par-
ticle. This quantity is conserved through time and independent of the pivot about which it
is calculated, so it commutes with the generator of 4-translations, the 4-momentum. Since
it is invariant, it commutes with all Lorentz transformation operators.
Because W

commutes with all the generators of the Poincare group, it is called a


Casimir operator, and its eigenvalues may be used to label states of the quantum eld.
The other Casimir operator for the Poincare group is the square of the momentum operator,
P

= m
2
. Together, these two operators provide a complete set of commuting operators
for the Poincare group. Thus, they provide a unique way of labelling the quantum eld states
of this group. Using the eigenvalues of these operators to label states is analogous to using
the eigenvalues of J
2
and J
z
to label states in O(3): these operators commute, and so their
eigenvalues may be used to label the states used in O(3).
VI. INABILITY TO SHRINK A ROTATING SYSTEM TO A POINT
In Section IV, we discussed the centroidthat it is frame-dependent and that it is at rest
in and only in the CM frame. Two arguments will be made here to outline the consequence
13
of this relativistic law that a nite system with a given angular momentum cannot be shrunk
to an arbitrarily small size.
The rst argument given is based on our intuitive understanding of the conservation of
angular momentum. In a freshman physics course, students learn that a gure-skater can pull
her arms in while spinning to increase her rotational velocity because of the conservation of
angular momentum. Applying this principle with relativistic constraints, we nd that there
must be a minimum radius for at least one of the particles, namely [

h
CM
[/m
CM
, where

h
CM
and m
CM
are measured in the CM frame. If this were not the case, then in order to attain
the required angular momentum, one or more of the particles would have to be travelling
faster than the speed of light.
The second argument for this principle of the minimum size of a composite system is
simply a mathematical consequence of what was discussed in Section IV, namely, that the
centroid is dierent in all frames. It can be shown, based on Eqs. (4.1) and (5.1), that
(r
C
) = 0 (6.1)
(Cf. Eq. (5.7)) for the calculation of angular momentum about the centroid of a system in
any frame. However, if the centroid were the same in all frames, then the zero-component
lemma discussed in Section IV would apply, meaning that L

= 0. Since L

(and, hence, )
is being calculated about dierent pivots (i.e., the dierent centroids) in dierent frames,
this lemma does not apply, and therefore, we are allowed to have

h(r
C
) ,= 0 even when
(r
C
) = 0.
Now consider for a moment what would happen if we were to shrink a system of particles
to an arbitrarily small volume. The only possible location for the centroid in this case
would be at the resulting particle itself, leaving us with but a single centroid for all frames,
and therefore zero angular momentum by the above argument. That particles have been
determined experimentally to have spin forces us back to the drawing board in our particle
models. The laws of relativistic angular momentum forbid a point-particle to have spin if
we regard them as the innitesimal size-limit of a nite system (a tiny spinning sphere,
for instance). Obviously, we cannot regard them as such. Formally, however, this is of no
concern because we can still declare the existence of a point-particle with spin, if we take
into account the requirements listed before Eq. (5.8).
14
VII. THOMAS PRECESSION
An interesting and important consequence of Special Relativity and its relation to angular
momentum is that orbital motion creates a precession in the spin of the orbiting particle.
This phenomenon is called the Thomas precession and is based on the fact that boosts
in dierent directions do not commute. The product of two boosts is not a third boost
but a boost and a rotation. This rotation, due to revolution about an axis, is the Thomas
precession.
Arguably the most straightforward conceptual derivation of the Thomas precession is
given here.
3
Imagine a particle in uniform circular motion with radius r and velocity v
about a xed rotation point. The circular path of the particle may be approximated by an
N-sided regular polygon with sides of length l = 2r/N. As the object moves along this
polygonal path, it moves from one vertex to the next, turns, and repeats this process ad
innitum (or until it is stopped). In the (inertial) rest frame of the laboratory, it is clear
that at each vertex, the object must rotate by 2/N to align itself with the next segment.
However, this is not the case in the frame of the object. (While the circular motion implies
a non-inertial frame, our approximation allows us to assume uniform motion along each of
the sidesit is only at the vertices that the acceleration becomes evident.) As the object
moves to its next destination vertex, it sees the projection of the next segment in its current
direction of motion foreshortened by a factor of 1/, but the lateral component of the next
segment remains unchanged. Because of this eect, from the point of view of the object it
must rotate by a (larger) angle, 2/N, in order to align with the following segment.
After N successive rotations of this sort, the object returns to its starting vertex, having
rotated by a total angle of 2, even though the researchers in the laboratory measure its
total rotation to be only 2. The orientation of the object after one revolution is not in
question. It points in the direction of the rst segment, ready to go around again. However,
its internal sense of coordinate axis has been altered by = 2( 1) in the opposite sense
of its orbital revolution. (This internal sense of coordinate axis may be realized physically
as an inertial object, such as an arrow or a bar, that does not undergo the same rotation
as the rotating object but is co-moving with it.) In the lab, it appears that the object has
3
Taken from the appendix of the paper by Muller.
15
only revolved about a point, but in the frame of the object, it believes that it has rotated
by more than 2 radians. This discrepancy is resolved if we allow the objects sense of
direction (i.e., the arrow or bar) to rotate in the opposite direction in the lab frame. This
eect is called the Thomas precession, and it links orbital motion with spinning motion.
Relating the precession frequency,
P
/T, to the angular velocity 2/T, where
T is the period of one revolution, we discover that

=
/T
2/T
= 1 . (7.1)
Because our result for was independent of the number of segments N, we may let N
and obtain the same result, which is the exact result for the Thomas precession. In the low-
velocity limit, Eq. (7.1) becomes


1
2
v
2
, (7.2)
which accounts for the miscalculation (overshoot by a factor of 2) of the spin-orbit coupling
energy in the hydrogen atom by considering the orbital motion of the electron without the
Thomas precession.
4
VIII. PARADOXES RELATED TO ANGULAR MOMENTUM
Perhaps an important question to ask in this situation is, where is the torque that is
causing the Thomas precession? For a point-particle with spin, this is dicult to answer
and will not be attempted here. However, we can answer the question if it is asked of a
multi-particle system,
5
and the answer lies in our old friend, the centroid. Because the
centroid is frame-dependent, a (good) rotating gyroscope will have its proper centroid on
the axis of rotation. This is the centroid in the gyroscopes own frame. Since the centripetal
force on the rotating gyroscope acts at the centroid, it creates no torque in this frame,
and the 2 total rotation is simply the result of length contraction. However, in the lab
frame, some parts of the gyroscope will be seen to be rotating faster than others, resulting
in an asymmetric distribution of mass and shifting the centroid o-axis. Nevertheless, the
centripetal force still acts at the centroid, wherever it is, and it is this torque (centripetal
force displacement of r
C
from r
PC
) that rotates the gyroscope in the lab frame.
4
More information on this calculation can be found in Cohen-Tannoudji, volume 2, page 1215.
5
Taken from the paper by Muller.
16
The relativistic version of the torque-angular momentum equation is of interest, as well.
It will briey be discussed here. The right-angle lever paradox illustrates the problem
well. It begins with a rigid L-shaped rod with sides of equal length with 4 forces acting on
it. Let A and C be two points on the rod near each of the legs of the L, and let B be a
point on the rod near the corner. Now let two forces act at B, one horizontally and one
vertically, both in the plane of the L-shaped rod, and let one force each act transversely on
the arms at points A and C, these too in the plane of the rod. These forces are all of equal
magnitude in the CM frame of the rod, and so the system is said to be in equilibrium.
Now imagine another frame moving with velocity v with respect to the rst, its velocity
oriented along one of the legs of the rod. Relativistically, the forces acting parallel to the
velocity remain unchanged, but those that act perpendicular to it are decreased, resulting
in a net external torque on the system. Why does the system not rotate? This paradox
is resolved if we realize that while, in nonrelativistic mechanics,
(classically)
d

h
dt
=

N
ext
, (8.1)
where

N
ext
is the net external torque, the relativistic version of this equation contains another
term: the net internal torque,

N
int
:
(relativistically)
d

h
dt
=

N
ext
+

N
int
. (8.2)
This internal torque is 0 in the objects rest frame if the object is in equilibrium, but it can
be nonzero in other frames. Without the aid of diagrams, it is dicult to explain succinctly
where this torque comes from.
6
However, it is sucient to summarize by saying that the
same transformation laws that created the net external torque in the second frame for our
L-shaped rod also create a perfectly balancing net internal torque, equal and opposite to the
external. This torque is specically due to the relativistic transformation of intermolecular
forces within the object itself.
6
For more details, including helpful gures that could not be reproduced in this L
A
T
E
X document, see the
paper by Nickerson.
17
IX. CONCLUSION
Relativity has been known to link things that one would normally not put together. No
one imagined that mass itself might be exchangeable with energy before Einstein explained
why this must be so (and experiment conrmed it). Similarly, who would have thought that
the conservation of angular momentum would imply the uniform motion of the center of
mass of the system? Yet, one cannot exist without the other.
This paper has served as an introduction to the study of relativistic angular momentum,
including its denition for nite systems of particles, both discrete and continuous. Its
relation to its 3-vector cousin was discussed, as was its co-dependent co-existence with
the mass moment and resulting consequence of uniform motion of the centroid. Further
consequences of the frame-dependent centroid were explored, as well, the most striking being
the inability to compress a system of particles to innitesimal size, requiring new thoughts
on just what a point-particle with spin really is. The spin vector and Pauli-Lubanski
vector were discussed, as well as an overview of the latters use in quantum eld theory. The
Thomas precession was explained and calculated, and two paradoxes involving torque and
angular momentum were explored and dissolved into understanding, as well.
Symmetry in the universe is a remarkable mathematical tool, and it is aesthetically
pleasing. Symmetries are realized in physical laws through conserved quantities. The free-
ow of one quantity into another, like space into time or angular momentum into mass
moment, is perhaps the most beautiful principle that Relativity has brought to physics.
18
[1] C. Cohen-Tannoudji, B. Diu, and F. Laloe, Quantum Mechanics (John Wiley and Sons, New
York, NY, 1977), Volume II.
[2] M. Kaku, Quantum Field Theory: A Modern Introduction (Oxford University Press, Oxford,
England, 1993).
[3] A. P. Lightman, W. H. Press, R. H. Price, and S. A. Teukolsky, Problem Book in Relativity
and Gravitation (Princeton University Press, Princeton, NJ, 1975).
[4] C. W. Misner, K. S. Thorne, and J. A. Wheeler, Gravitation (Freeman and Co., San Francisco,
CA, 1970).
[5] R. A. Muller, Am. J. Phys. 60, 313 (1992).
[6] J. C. Nickerson and R. T. McAdory, Am. J. Phys. 43, 615 (1975).
[7] W. Rindler, Introduction to Special Relativity, Second Edition (Oxford University Press, Ox-
ford, England, 1991).
[8] H. M. Schwartz, Introduction to Special Relativity (McGraw-Hill, New York, NY, 1968).
19

Potrebbero piacerti anche