Download as pdf or txt
Download as pdf or txt
You are on page 1of 7

FYGB08 Project

Thomas precession

Author: SSN:
Daniel Kronberg 850810-5999

Abstract
This project is a derivation of Thomas precession, a relativistic correction to a vector traveling along
a curvilinear trajectory. We consider the precession as a consequence of consecutive restricted Lorentz
transformations between instantly inertial reference frames along the trajectory. These tranformations
˙
result in an observed retrograde precession of the vector, with the angular velocity ω~ T = −(γ − 1) ~vv×2~v .
1 Introduction
This project is an attempt at a straightforward explanation of Thomas precession by discussing Thomas-
Wigner rotation algebraically as a consequence of consecutive restricted Lorentz transformations. Fermi-
Walker transport is approached in a fairly conceptual manner as a motivation of how results based on a
set of inertial reference frames can be applied to accelerated systems. It is possible to derive the Thomas
angular velocity directly from the equations given in the Fermi-Walker section, but here we have chosen to
derive it from Thomas-Wigner rotation.
Illustrated reference frames are shown as a tetrad of vectors, where the {0}-vector is a time-like vector,
Minkowski-orthogonal to the three space-like vectors {1, 2, 3}.

2 Lorentz transformations
Lorentz transformations are used to connect observations between inertial reference frames. In particular
they give the correct transformation rules at relativistic velocities. A single Lorentz transformation along a
given direction is called a restricted Lorentz transformation, or boost, and affects the time component and
the spatial component parallel to the velocity. If the basis vectors are chosen so that x̂ = |~~vv| , then a boost
has the form
 
0 ~v x̂
t̂ = γ 1 − 2
c
x̂0 = γ(x̂ − ~v t)
ŷ 0 = ŷ
ẑ 0 = ẑ

1 ~v
γ=p , β= ,
1− β2 c

where ~v is the boost velocity, and c is the speed of light. A transformation from the primed frame into some
other frame with a collinear velocity has the same form, and is equivalent to a single boost from the first
frame with a total velocity given by

β + β0
β 00 = . (1)
1 + ββ 0
Any boost can be represented by a symmetric matrix, and the general boost form is
 
γ −γβx −γβy −γβz
β2
 β β

−γβ
x 1 + (γ − 1) βx2 (γ − 1) βx 2 y (γ − 1) ββx β2 z 
L= .
 
−γβy βy βx βy2 βy βz 
 (γ − 1) β 2 1 + (γ − 1) β 2 (γ − 1) β 2 
β β β2
−γβz (γ − 1) ββz β2 x (γ − 1) βz 2 y 1 + (γ − 1) βz2

However, the product of two non-collinear boosts is not equivalent to any single boost, but instead to some
boost and some rotation. For example, two perpendicular boosts
   
γ −γβ 0 0 γ0 0 −γ 0 β 0 0
   
−γβ γ 0 0  0 1 0 0
L=  , L0 = 
   

 0 0 1 0 −γ 0 β 0 0 γ0 0
   
0 0 0 1 0 0 0 1

1
have the products
 
γγ 0 −γβ −γγ 0 β 0 0
 
−γγ 0 β γ γγ 0 ββ 0 0
0
LL =  (2)
 

 −γ 0 β 0 0 γ0 0
 
0 0 0 1
 
γγ 0 −γγ 0 β −γ 0 β 0 0
 
 −γβ γ 0 0
L0 L =  , (3)
 
−γγ 0 β 0 γγ ββ 0 0
γ0 0
 
0 0 0 1

which are not symmetric; notice further that the product is non-commutative. As a result, the coordinate
axes of the final reference frame are not parallel with those of the first, but are instead rotated by some angle
θ about an axis orthogonal to the boosts. We can hence express the product of two boosts as a product of a
boost and a rotation. This rotation is called Thomas rotation, Wigner rotation, or Thomas-Wigner rotation,
after Llewellyn Thomas and Eugene Wigner. We will in the next section derive the explicit expression of
the rotation matrix represnting the Thomas-Wigner rotation.

3 Thomas-Wigner rotation
Consider two inertial reference frames, observed from the lab frame ˆl. When applying boosts first to the
ê frame and then from ê to the ê0 frame, we will find that even though the coordinate axes of ˆl and ê are
parallel to each other, and the coordinate axes of ê and ê0 are also parallel to each other, the axes of ˆl and
ê0 are not in general parallel.

ê00

ê01
ê03
ê02

β 00 β 0 L0
Rθ L00

l̂0 ê0

β
L
l̂1 ê1
l̂3 ê3
l̂2 ê2

Figure 1: Three inertial reference frames with individual relative velocities.

The Lorentz transformation from ˆl to ê0 is given by the matri L0 L. As previously discussed, this can be
written as the product of a rotation and a boost, i.e. L0 L = Rθ L00 . In order to find the rotation angle θ
between the two frames, we follow the method outlined in [1] by introducing an arbitrary rotation matrix
and using it to find an expression for the boost L00 . Since the coordinate axes can always be chosen parallel

2
to the boosts, it is sufficient to consider rotation in the x̂ ∧ ŷ plane.
 
1 0 0 0
 
0 cos(θ) sin(θ) 0
Rθ = 
 

0 − sin(θ) cos(θ) 0
 
0 0 0 1
 
γγ 0 −γγ 0 β −γ 0 β 0 0
 
 γ(γ 0 β 0 sin(θ) − β cos(θ)) −γ(γ 0 ββ 0 sin(θ) − cos(θ)) −γ 0 sin(θ) 0
L00 = Rθ−1 L0 L =  (4)
 

−γ(β sin(θ) + γ 0 β 0 cos(θ)) γ(sin(θ) + γ 0 ββ 0 cos(θ)) γ 0 cos(θ) 0
 
0 0 0 1

By requiring that the matrix L00 , as a boost, be symmetric, the elements of the rotational matrix can be
determined

0 0 0
−γ sin(θ) = γ(sin(θ) + γ ββ cos(θ))

0 0 0
−γγ β = γ(γ β sin(θ) − β cos(θ))
 0 0
−γ β = −γ(β sin(θ) + γ 0 β 0 cos(θ))

=⇒

γ + γ0 γγ 0 ββ 0
cos(θ) = , sin(θ) = − . (5)
γγ 0 + 1 γγ 0 + 1

Substituting the resulting expressions into L00 gives us


 
γγ 0 −γγ 0 β −γ 0 β 0 0
−γγ 0 β 1 + γ 2 γ 02 β 2 γγ 02 ββ 0
 
00 γγ 0 +1 γγ 0 +1 0
L = . (6)
 
 −γ 0 β 0 γγ 02 ββ 0 γ 0 (γ+γ 0 )
 γγ 0 +1 γγ 0 +1 0

0 0 0 1

This method can be applied to any two boosts by splitting them up into collinear and perpendicular
components and adding their velocities according to equation (1). For three or more arbitrary boosts, the
expressions are more complex, but in the situations that are interesting when discussing Thomas
precession, we need only deal with motion in a plane.

4 Fermi-Walker transport
The results from the previous section can be applied easily when dealing only with inertial frames,
however, many interesting systems are accelerated, and it is not immediately evident how Thomas-Wigner
rotation would apply to such systems. In order to describe accelerated systems, we will consider
Fermi-Walker transport, a form of parallel-transport used to move a vector along a curvilinear trajectory
with the minimum possible amount of rotation. In this model, we treat the trajectory of the accelerated
system as consisting of an infinite series of instantaneously inertial frames, each related to its neighbors by
an infinitesimal boost in the direction of acceleration. This is a commonly used approximation in relativity
theory [2].

3
ê00

ê03

ê01 ê02
ê0

ê1
ê3
ê2

Figure 2: Two instantaneously inertial frames on a curved world line, representing an accelerated system.

Let us describe Fermi-Walker transport in some more detail, following [3]. Consider an accelerated particle
and its world-line, observed from the Thomas conventional frame; an inertial frame related to the lab
frame by a single boost to some initial frame of the particle. Then consider the space-like hyperplane that
is orthogonal to the time-axis. Any purely space-like vector ~q in the accelerated system must be
continuously projected into the hyperplane of the next inertial frame. Each infinitesimal projection will be
in the direction of the time-axis, and so the evolution of the vector ~q will be
d~q
= kê0

where τ is the proper time and k is some proportonality factor.
Further,

~q · ê0 = 0

and
d d~q dê0
(~q · ê0 ) = · ê0 + ~q · = k + ~q · ~b = 0
dτ dτ dτ

where ~b = dê
dτ is the world acceleration and ê0 · ê0 = 1. Therefore the Fermi-Walker transport equation for
0

the vector ~q is
d~q
= −(~q · ~b)ê0

By differentiating two space-like vectors, ~q = −(~q · ~b)ê0 and p~ = −(~


p · ~b)ê0 , we find that the scalar product
is preserved by Fermi-Walker transport.
d
(~q · p~) = −(~q · ~b)(~ p · ~b) = 0
p · ê0 ) − (~q · ê0 )(~

Therefore the change in any space-like vector parallel-transported along this curved world-line is a rotation
only. The length of the vector remains constant.
With the basic concepts of Fermi-Walker transport in mind, we can now apply the results to the
description of Thomas-Wigner rotation.

5 Thomas precession
As a vector is Fermi-Walker transported over a curvilinear trajectory, it will experience countless
infinitesimal Thomas-Wigner rotations. Since the rotation angle is dependent on the relative velocities of
the boosts, the percieved rotation angle will depend on the reference system from which the
parallel-transported frame is observed. However, if the motion of a parallel-transported vector, such as a
spin-vector, is periodic, and obervations are made at integer multiples of the period, the rotations will add

4
up to produce a net retrograde precession that is independent of the observer’s frame of reference. The
angular velocity of this precession can be found by considering each individual infinitesimal set of boosts to
have the rotational angle given by (5). If we then assume that the second boost, representing the change in
velocity from one instant inertial frame to the next, is much smaller than the total relative velocity as seen
from the observer frame, β >> β 0 , and also small enough compared to the speed of light that γ 0 ≈ 1, we
can make the approximation

γββ 0
cos(θ) ≈ 1 , sin(θ) ≈ .
γ+1

As the rotational angle is infinitesimal, we can also make the approximation sin(θ) ≈ θ. We then subsitute
in the components of the velocity vectors as transformed by the associated boost L00 , to the first order, as
done in [2], i.e.

β0
βx00 ≈ β , βy00 = , β 002 ≈ β 2 , γ 00 ≈ γ .
γ
We can now find the expression for the Thomas angular velocity.

γ 2 ~v˙ × ~v ~v × ~v˙
ω
~T = = −(γ − 1)
γ + 1 c2 v2

Where ~v and ~v˙ are measured in the same inertial frame.


The Thomas angular velocity is observed, for example, when considering a spinning gyroscope traveling in
an orbit around a central force. In this case the spin vector of the gyroscope will precess as described.
Another common example is the spin-orbit interaction of electrons. In that case, however, Thomas
precession by itself can not account for the motion of the spin vector. Larmor precession must also be
taken into account.

6 Discussion
The greatest difficulty with this project, despite the rather paradoxical nature of Thomas-Wigner rotation
0
itself, was understanding the approximation βy00 ≈ βγ . It’s still not quite clear to me why this is a good
approximation, and I have been unable to find a satisfactory explanation. Because of this, obtaining the
correct result for the Thomas angular velocity was initially troublesome. Fortunately, accepting that the
acceleration in a curvilinear orbit has this value when seen from some inertial lab-frame allowed me to
calculate the angular velocity correctly based on both methods presented in [1] and [2], though only the
first is shown here. The more geometrically focused method demonstrated in [3] was more difficult to apply
to the end result, but I find the approach very interesting and I hope to be able to revisit it in the future,
in order to properly understand it.

5
References
[1] Fererro, R. & Thibeault, M. (2002) Generic composition of boosts: an elementary derivation of the
Wigner rotation arxiv.org/abs/physics/0211022v1 [physics.ed-ph]
[2] Goldstein, H. & Poole, C.P. & Safko, J.L. (2002) Classical Mechanics (3rd ed.). Pearson Education
Limited, Harlow.

[3] Notes 5, Thomas Precession. (2002). Retrieved January 15, 2015,


from http://bohr.physics.berkeley.edu/classes/221/0708/notes/thomprec.pdf

You might also like