Location via proxy:   [ UP ]  
[Report a bug]   [Manage cookies]                

Kleppner D. Kolenkow R.J. Introduction To Mechanics 2014

Download as pdf or txt
Download as pdf or txt
You are on page 1of 25
At a glance
Powered by AI
The document discusses how concepts of momentum and energy must be modified in relativity to ensure conservation in isolated systems. It also introduces the idea of massless particles that can carry energy and momentum.

The document defines relativistic momentum and energy in terms of a scalar quantity m(w) analogous to mass but which could depend on speed. Relativistic expressions for momentum and kinetic energy are introduced.

The document describes analyzing elastic collisions between particles in different frames to find expressions where total momentum is conserved. It shows that reversing y velocities while keeping x velocities unchanged results in conserved momentum.

13

RELATIVISTIC
DYNAMICS

13.1 Introduction 478


13.2 Relativistic Momentum 478
13.3 Relativistic Energy 481
13.4 How Relativistic Energy and Momentum are Related 487
13.5 The Photon: A Massless Particle 488
13.6 How Einstein Derived E = mc2 498
Problems 499
478 RELATIVISTIC DYNAMICS

13.1 Introduction
In Chapter 12 we saw how the postulates of special relativity lead to new
kinematical relations for space and time. These relations can naturally
be expected to have important implications for dynamics, particularly
for the meaning of momentum and energy. In this chapter we examine
the modifications to the Newtonian concepts of momentum and energy
required by special relativity. The underlying strategy is to ensure that
momentum and energy in an isolated system continue to be conserved.
This approach is often used in extending the frontiers of physics: by
reformulating conservation laws so that they are preserved in new situa-
tions, we are led to generalizations of familiar concepts. We can also be
led to the discovery of unfamiliar concepts, for instance the concept of
massless particles that can nevertheless carry energy and momentum.

13.2 Relativistic Momentum


y To investigate the nature of momentum in special relativity, consider a
A
glancing elastic collision between two identical particles A and B in an
isolated system. We want the total momentum of the system to be con-
B served, as it is in non-relativistic physics. We shall view the collision in
two frames: A’s frame, the frame moving along the x axis with A so that
x
A is at rest while B approaches along the x direction with speed V, and
then in B’s frame, which is moving with B in the opposite direction so
that B is at rest and A is approaching. (The term “frame” is used synony-
mously with “reference system.”) We take the collisions to be completely
symmetrical. Each particle has the same y speed u0 in its own frame be-
fore the collision, as shown in the sketches. The effect of the collision is
to reverse the y velocities but leave the x velocities unchanged.

u′
A
u0
A’s
frame u0 / γ
V
V B
u′/ γ

A V u′/ γ
u0 / γ V
B’s
frame
u0
B
u′

Before After

The relative x velocity of the frames is V. In A’s frame, the y velocity


of particle A is u0 , and by the transformation of velocities, Eqs. (12.8)
13.2 RELATIVISTIC MOMENTUM 479


and (12.9), the y velocity of particle B is u0 /γ where γ = 1/ 1 − V 2 /c2 .
The situation is symmetrical when viewed from B’s frame.
After the collision the y velocities have reversed their directions as
shown. The situation remains symmetric: if the y velocity of A or B in its
own frame is u , the y velocity of the other particle is u /γ.
Our task is to find a conserved quantity analogous to classical momen-
tum. We suppose that the momentum of a particle moving with velocity
w is
p = m(w)w,
where m(w) is a scalar quantity, yet to be determined, analogous to New-
tonian mass but which could depend on the speed w.
The x momentum in A’s frame  is due entirely to particle B. Before
the collision B’s speed is w = V 2 + u20 /γ2 and after the collision it

is w = V 2 + u2 /γ2 . Imposing conservation of momentum in the x
direction yields
m(w)V = m(w )V.
It follows that w = w , so that
u = u0 .
In other words, y motion is reversed in the A frame.
Next we write the statement of the conservation of momentum in the
y direction as evaluated in A’s frame. Equating the y momentum before
and after the collision gives
u0 u0
−m0 u0 + m(w) = m0 u0 − m(w)
γ γ
which gives
m(w) = γm0 .
In the limit u0 → 0, m(u0 ) → m(0), which we take to be the Newtonian
mass, or “rest mass” m0 , of the particle. In this limit, w = V. Hence
m0
m(V) = γm0 =  . (13.1)
1 − V 2 /c2
Consequently, momentum is preserved in the collision provided we de-
fine the momentum of a particle moving with velocity v to be
p = mv (13.2)
where
m0
m=  = γm0 .
1 − v2 /c2
The quantity m = γm0 is referred to as the “relativistic mass” or more
often simply as the mass of a particle. If the rest mass is intended, that
needs to be made specific.
480 RELATIVISTIC DYNAMICS

In relativity there is an upper limit to speed: the speed of light. How-


ever, there is no upper limit on momentum. Once a particle is moving
with speed close to c, an increase in momentum comes about primarily
through an increase in mass. High energy particle accelerators do not
make particles go substantially faster and faster. A particle is quickly ac-
celerated to speed close to c. After that, the accelerator principally makes
the particle more and more massive with only a very small increment in
speed.
The expression p = mv = γm0 v is sometimes taken as the starting
point for developing relativistic dynamics, but in the early days of rela-
tivity attention was focused not so much on momentum but on the appar-
ent dependence of mass on speed. Investigation of this problem provided
the first direct experimental evidence for Einstein’s theory.

Example 13.1 Speed Dependence of the Electron’s Mass


At the beginning of the twentieth century there were several specula-
tive theories based on various models of the structure of the electron
that predicted that the mass of an electron would vary with its speed.
One theory, from Max Abraham (1902), predicted m = m(u0 )[1 +
5 (v /c )] for v c and another from Hendrik A. Lorentz (1904) gave
2 2 2

m = m0 / 1 − v2 /c2 ≈ m(u0 )[1 + 12 (v2 /c2 )]. The Abraham theory,
which retained the idea of the ether drift and absolute motion,
predicted no time dilation effect. Lorentz’s result, while identical in
form to that published by Einstein in 1905, was derived using the
ad hoc Lorentz contraction and did not possess the generality of
Einstein’s theory. Experimental work on the effect of speed on the
electron’s mass was initiated by Kaufmann in Göttingen in 1902. His
data favored the theory of Abraham, and in a 1906 paper he rejected
the Lorentz–Einstein results. However, further work by Bestelmeyer
(1907) in Göttingen and Bucherer (1909) in Bonn revealed errors in
Kaufmann’s work and confirmed the Lorentz–Einstein formula.

P Physicists were in agreement that the force on a moving electron in an


S
applied electric field E and magnetic field B is q(E + v × B) (the units
C
V
are SI), where q is the electron’s charge and v its velocity. Bucherer
A d
C
employed this force law in the apparatus shown in the sketch. The ap-
paratus is evacuated and immersed in an external magnetic field B per-
pendicular to the plane of the sketch. The source of the electrons A is
a button of radioactive material, generally radium salts. The emitted
electrons (“beta-rays”) have a broad energy spectrum extending to 1
MeV or so. To select a single speed, the electrons are passed through a
“velocity filter” composed of a transverse electric field E (produced be-
tween two parallel metal plates C by the battery V) and the perpendic-
ular magnetic field. E, B, and v are mutually perpendicular. The trans-
verse force is zero when qE = qvB, so that electrons with v = E/B are
undeflected and are able to pass through the slit S.
13.3 RELATIVISTIC ENERGY 481

Beyond S only the magnetic field acts. The electrons move with con-
stant speed v and are bent into a circular path by the magnetic force
qvB. The radius of curvature R is given by mv2 /R = qvB, or R = mv/
qB = (m/q)(E/B2 ).

The electrons eventually strike the photographic plate P, leaving a


trace. By reversing E and B, the sense of deflection is reversed. R is
found from a measurement of the total deflection d and the known ge-
ometry of the apparatus. E and B are measured by standard techniques.
Finding R for different velocities allowed the velocity dependence of
m/q to be measured. Physicists believe that charge does not vary with
velocity (otherwise an atom would not stay strictly neutral in spite of
how the energy of its electrons varied), so that the variation of m/q can
be attributed to variation in m alone.

m/m0 The graph shows Bucherer’s data togetherwith a dashed line corre-
1.50 sponding to the Einstein prediction m = m0 / 1 − v2 /c2 . The agreement
1.40 is striking.
1.30
1.20 Today, the relativistic equations of motion are used routinely to design
1.10 high energy particle accelerators. For protons, accelerators have been
1.00 operated with m/m0 up to 104 , while for electrons the ratio m/m0 = 105
0.30 0.40 0.50 0.60 0.70 has been reached. The successful operation of these machines leaves no
υ/c doubt of the validity of relativistic dynamics.

13.3 Relativistic Energy


By generalizing the Newtonian concept of energy, we can find a corre-
sponding relativistic quantity that is also conserved. Recall the argument
from Chapter 5: the change in kinetic energy K of a particle as it moves
from ra to rb under the influence of force F is
 b
Kb − Ka = F · dr
a
 b
dp
= · dr.
a dt
For a Newtonian particle moving with velocity u the momentum is given
by p = mu, where m is constant. Then
 b
d
Kb − Ka = (mu) · dr
a dt
 b
du
= m · u dt
a dt
 b
= mu · du.
a
482 RELATIVISTIC DYNAMICS

Using the identity u · du = 12 d(u · u) = 12 d(u2 ) = u du, we obtain


1 1
Kb − Ka = mub 2 − mua 2 .
2 2
It is natural to try the same procedure
 starting with the relativistic ex-
pression for momentum p = m0 u/ 1 − u2 /c2 :
 b
dp
Kb − Ka = · dr
a dt
 b
dp dr
= · dt
a dt dt
 b ⎡⎢ ⎤
⎥⎥⎥
d ⎢⎢⎢ m0 u ⎥⎥⎦ · u dt
= ⎢⎣ 
a dt 1 − u2 /c2
 b ⎡ ⎤
⎢⎢ ⎥⎥⎥
u · d ⎢⎢⎢⎣ 
m0 u ⎥⎥⎦ .
=
a 1 − u2 /c2
The integrand has the form u · dp. Using the relation u · dp = d(u · p) −
p · du gives
  b
Kb − Ka = (u · p) ba − p · du
a
b 
m0 u2  b
m0 u du
=   −  ,
1 − u2 /c2 a a 1 − u2 /c2
where we have again used the identity u · du = u du. The integral is
elementary, and we find
b  b
m0 u2  2
 + m0 c2 1 − 2  .
u
Kb − Ka = 
1 − u /c a
2 2 c 
a

Let point b be arbitrary and take the particle to be at rest at point a so


ua = 0:

m0 u2 u2
K =  + m0 c2
1 − 2 − m0 c2
1 − u2 /c2 c
m0 [u2 + c2 (1 − u2 /c2 )]
=  − m0 c2
1 − u /c
2 2

m0 c2
=  − m0 c2
1 − u2 /c2
or
K = (γ − 1)m0 c2 . (13.3)
This expression for kinetic energy bears little resemblance
 to its classical
counterpart. However, in the limit u c, γ = 1/ 1 − u /c2 ≈ 12 u2 /c2 .
2
13.3 RELATIVISTIC ENERGY 483


Using the expansion 1/ 1 − x = 1 + 12 x + · · · we obtain

1 u2
K ≈ m0 c2 1 + − 1
2 c2
1
= m0 u2 .
2
The kinetic energy arises from the work done on the particle to bring it
from rest to speed u. Using the relation mc2 = γm0 c2 , we can rearrange
Eq. (13.3) to give
mc2 = K + m0 c2
= work done on particle + m0 c2 . (13.4)
Einstein proposed the following bold interpretation of this result: mc2
is the total energy E of the particle. The first term arises from external
work; the second term, m0 c2 , represents the “rest” energy the particle
possesses by virtue of its mass. In summary,
E = mc2 . (13.5)
It is important to realize that Einstein’s generalization goes far beyond
the classical conservation law for mechanical energy. Thus, if energy ΔE
is added to a body, its mass will change by Δm = ΔE/c2 , irrespective of
the form of energy. ΔE could be mechanical work, heat energy, the ab-
sorption of light, or any other form of energy. In relativity the classical
distinction between mechanical energy and other forms of energy disap-
pears. Relativity treats all forms of energy on an equal footing, in con-
trast to Newtonian physics where each form of energy must be treated as
a special case.
The conservation of total energy E = mc2 is a consequence of the
structure of relativity. In Chapter 14 we shall show that the conservation
laws for energy and momentum are actually different aspects of a single,
more general, conservation law.
The following example illustrates the relativistic concept of energy
and the application of the conservation laws in different inertial frames.

Example 13.2 Relativistic Energy and Momentum in an


Inelastic Collision
Suppose that two identical particles each of mass M collide with equal
and opposite velocities and stick together. In Newtonian physics,
the initial kinetic energy is 2( 12 MV 2 ) = MV 2 . By conservation of
momentum the mass 2M is at rest and has zero kinetic energy. In the
language of Chapter 4 we say that mechanical energy MV 2 was lost as
heat. As we shall see, this distinction between these different classical
forms of energy does not occur in relativity.

Now consider the same collision relativistically, as seen in the original


frame x, y, and in the frame x , y moving with one of the particles. By
484 RELATIVISTIC DYNAMICS

the relativistic transformation of velocities, Eqs. (12.8) and (12.9), the


relative velocity in the x , y frame is
2V
U= (1)
1 + V 2 /c2
in the direction shown.
y y
V V

x x

y′ y′
U V

x′ x′
Before After

Let the rest mass of each particle be M0i before the collision and M0 f
after the collision. In the x, y frame, momentum is obviously
 con-
served. The total energy before the collision is 2M0i c2 / 1 − V 2 /c2 ,
and after the collision the energy is 2M0 f c2 . No external work was
done on the particles, and the total energy is unchanged. Therefore
2M0i c2
 = 2M0 f c2
1 − V 2 /c2
or
M0i
M0 f =  . (2)
1 − V 2 /c2
Physically, the final rest mass is greater than the initial rest mass be-
cause the particles are warmer. To see this, we take the low-velocity
approximation

1 V2
M0 f ≈ M0i 1 + .
2 c2
The increase in rest energy for the two particles is 2(M0 f − M0i )c2 ≈
2( 12 M0i V 2 ), which corresponds to the loss of Newtonian kinetic energy.
Now, however, the kinetic energy is not “lost”—it is present as a mass
increase.

By the postulate that all inertial frames are equivalent, the conservation
laws must hold in the x , y frame as well. Checking to see if our as-
sumed conservation laws possess this necessary property, we have in
the x , y frame
M0i U 2M0 f V
 =  (3)
1 − U 2 /c2 1 − V 2 /c2
13.3 RELATIVISTIC ENERGY 485

by conservation of momentum and


M0i c2 2M0 f c2
M0i c2 +  =  (4)
1 − U 2 /c2 1 − V 2 /c2
by conservation of energy.

The question now is whether Eqs. (3) and (4) are consistent with our
earlier results, Eqs. (1) and (2). To check Eq. (3), we use Eq. (1) to
write
U2 4V 2 /c2
1− =1−
c2 (1 + V 2 /c2 )2
(1 − V 2 /c2 )2
= . (5)
(1 + V 2 /c2 )2
From Eqs. (1) and (5),
U 2V (1 + V 2 /c2 )
 =
1 − U 2 /c2 (1 + V /c ) (1 − V /c )
2 2 2 2

2V
=
1 − V 2 /c2
and the left-hand side of Eq. (3) becomes
M0i U 2M0i V
 = . (6)
1 − U 2 /c2 1 − V /c
2 2

From Eq. (2), M0i = M0 f 1 − V 2 /c2 , and Eq. (6) reduces to
M0i U 2M0 f V
 =  ,
1 − U 2 /c2 1 − V 2 /c2
which is identical to Eq. (3). Similarly, it is not hard to show that Eq.
(4) is also consistent.

We see from Eq. (6) that if we had assumed that rest mass was un-
changed in the collision, M0i = M0 f , the conservation law for momen-
tum (or for energy) would not be correct in the second inertial frame.
The relativistic description of energy is essential for maintaining the
validity of the conservation laws in all inertial frames.

Example 13.3 The Equivalence of Mass and Energy


In 1932 J.D. Cockcroft and E.T.S. Walton, two young British physi-
cists, successfully operated the first high energy proton accelerator and
succeeded in causing a nuclear disintegration. Their experiment pro-
vided one of the earliest confirmations of the relativistic mass–energy
relation.

Briefly, their accelerator consisted of a power supply that could reach


600 kV and a source of protons (hydrogen nuclei). The power sup-
ply used an ingenious arrangement of capacitors and rectifiers to
486 RELATIVISTIC DYNAMICS

quadruple the voltage of a 150 kV supply. The protons were supplied


by an electrical discharge in hydrogen and were accelerated in vacuum
by the applied high voltage.

Proton beam Cockcroft and Walton studied the effect of the protons on a target of
7
Li (lithium atomic mass 7). A zinc sulfide fluorescent screen, located
nearby, emitted occasional flashes, or scintillations. By various tests
they determined that the scintillations were due to alpha particles, the
nuclei of helium, 4 He. Their interpretation was that 7 Li captures a pro-
Lithium target ton and that the resulting nucleus of mass 8 immediately disintegrates
into two alpha particles. We can write the reaction as
Screen 7
Li + 1 H → 4 He + 4 He.
The mass–energy equation for the reaction can be written
Kinitial + Minitial c2 = Kfinal + Mfinal c2
where the masses are the particle rest masses. Applied to the lithium
bombardment experiment, this gives
K(1 H) + [M (1 H) + M (7 Li)]c2 = 2K(4 He) + 2M(4 He)c2
where K(1 H) is the kinetic energy of the incident proton, K(4 He) is
the kinetic energy of each emitted alpha particle, M(1 H) is the proton
rest mass, etc. (The initial momentum of the proton is negligible, and
the two alpha particles are emitted back-to-back with equal energy by
conservation of momentum.)

We can rewrite the mass–energy equation as


K = ΔMc2 ,
where K = 2K(4 He) − K(1 H), and where ΔM is the initial rest mass
minus the final rest mass.

The energy of the alpha particles was determined by measuring their


range in matter. Cockcroft and Walton obtained the value K =
17.2 MeV (1 MeV = 106 eV = 1.6 × 10−13 J.

The relative masses of the nuclei were known from mass spectrome-
ter measurements. In atomic mass units, amu, the values available to
Cockcroft and Walton were
M(1 H) = 1.0072
M(7 Li) = 7.0104 ± 0.0030
M(4 He) = 4.0011.
Using these values,
ΔM = (1.0072 + 7.0104) − 2(4.0011)
= (0.0154 ± 0.0030) amu.
13.4 HOW RELATIVISTIC ENERGY AND MOMENTUM ARE RELATED 487

The rest energy of 1 amu is ≈ 931 MeV and therefore


ΔMc2 = (14.3 ± 2.7)MeV.
The difference between K and ΔMc2 is (17.2 − 14.3) MeV = 2.9 MeV,
slightly larger than the experimental uncertainty of 2.7 MeV. However,
the experimental uncertainty always represents an estimate, not a pre-
cise limit, and the result from these early experiments can be taken as
consistent with the relation K = ΔMc2 . It is clear that the masses must
be known to high accuracy for studying the energy balance in nuclear
reactions. Modern techniques of mass spectrometry have achieved an
accuracy of better than 10−10 amu, and the mass–energy equivalence
has been amply confirmed to within experimental accuracy. According
to a modern table of masses, the decrease in rest mass in the reaction
studied by Cockcroft and Walton is ΔMc2 = (17.3468 ± 0.0012) MeV.

13.4 How Relativistic Energy and Momentum


are Related
Often it is useful to express the total energy of a free particle in terms of
its momentum. In Newtonian physics the relation is
1 2 p2
E= mv = .
2 2m
To find the equivalent relativistic expression we can combine the rela-
tivistic momentum
m0 u
p = mu =  = γm0 u (13.6)
1 − u2 /c2

with the energy


E = mc2 = γm0 c2 . (13.7)

Squaring Eq. (13.6) gives

m20 u2
p2 = .
1 − u2 /c2
We can solve for γ as follows:

u2 p2
= 2
c2
p + m20 c2
1
γ= 
1 − u2 /c2

p2
= 1 + 2 2.
m0 c
488 RELATIVISTIC DYNAMICS

Inserting this in Eq. (13.7), we have



p2
E = m0 c2 1+ .
m0 2 c2
The square of this equation is algebraically simpler and is the form usu-
ally employed:
E 2 = (pc)2 + (m0 c2 )2 . (13.8)
For convenience, here is a summary of the important dynamical formulas
we have developed so far.
p = mu = m0 uγ (13.9)
K = mc − m0 c = m0 c (γ − 1)
2 2 2
(13.10)
E = mc = m0 c γ
2 2
(13.11)
E = (pc) + (m0 c ) .
2 2 2 2
(13.12)

13.5 The Photon: A Massless Particle


In 1905, in the annus mirabilis when Albert Einstein published four pa-
pers each worthy of a Nobel Prize, the first paper, and the only one to
actually receive the Prize, had the unlikely title On a Heuristic Viewpoint
Concerning the Production and Transformation of Light. A heuristic the-
ory is a theory based partly on guesswork, intended to stimulate think-
ing. The paper ostensibly provided an explanation for the photoelectric
effect, the process by which electrons are ejected from a surface when it
is irradiated with light. It is now recognized that the paper provided the
foundation for the quantum theory of light, contributed significantly to
the development of quantum mechanics, and made applications such as
the laser possible.
Because light is inherently relativistic, Einstein’s paper actually
opened a chapter on relativity even before relativity had been announced.
At the heart of his argument is a concept that makes little sense in New-
tonian physics but perfect sense in relativistic physics: massless particles
that carry momentum.
A little background is needed: In December, 1900, quantum physics
was born when Max Planck proposed that the energy of a harmonic os-
cillator cannot be varied at will but only by discrete steps. If the fre-
quency of the oscillator is ν, then the energy steps had size hν where h
is a constant, now called Planck’s constant, h ≈ 6.6 × 10−34 m2 kg2 /s =
6.6 × 10−34 joule · second. Planck proposed this idea to solve the mys-
tery of thermal radiation, often called blackbody radiation. The shape
of the spectrum of radiation emitted by a warm body could not be
accounted for by the known laws of physics, based on Newtonian
mechanics and Maxwell’s electromagnetic theory. Planck put forward
his hypothesis more in the spirit of a mathematical conjecture than
13.5 THE PHOTON: A MASSLESS PARTICLE 489

a physical theory, but in 1905 Einstein came to a similar conclusion,


though by totally different reasoning, and his theory had some startling
implications.
Einstein was thoroughly aware of the wave nature of light. He knew all
about Maxwell’s equations and how they predicted the existence of elec-
tromagnetic waves—light waves—that can travel through empty space
with speed c. The wavelength of an electromagnetic wave λ and its fre-
quency ν are related by λν = c. There was a considerable body of ex-
perimental evidence that confirmed the wave nature of light, for instance
the colors in soap bubble films that are a signature of light waves in-
terfering, not to mention the fringes in Michelson’s interferometer. Ein-
stein, however, pointed out that these phenomena involve observations at
the macroscopic (large scale) level. Macroscopic behavior results from
the effect of many microscopic events. He pointed out that little was
known about how light interacted with matter at the atomic or indi-
vidual particle level. He went on to argue that light could also be un-
derstood from a particle point of view. He suggested that a light wave
could behave as if it were a gas of particles, each possessing energy
 = hν = hc/λ, where ν is the frequency of the wave. This particle
hypothesis seemed to be in direct contradiction to the wave theory of
light.
We now understand that light displays either wave-like or particle-like
behavior depending on the situation. To understand light from the wave
point of view, one starts by writing Maxwell’s wave equations. Their
solution reveals that time-varying electric and magnetic fields in space
support each other to create an electromagnetic wave that travels at the
speed of light. Furthermore, no matter which inertial coordinate system
one chooses for describing the radiation process, the wave always prop-
agates at speed c. In other words, Maxwell’s equations are intrinsically
relativistic. Einstein showed that we can also understand the relativistic
behavior of light starting from a particle point of view, and that is the
approach we now follow.
A startling consequence of the relativistic energy–momentum relation
is the possibility of “massless” particles, particles that possess momen-
tum and energy but have zero rest mass. The essential point is that a
particle can possess momentum without possessing mass. This follows
from the definition of relativistic momentum
⎛ ⎞
⎜⎜⎜ 1 ⎟⎟⎟

p = m0 u ⎜⎝  ⎟⎟⎠ .
1 − u2 /c2

If we consider the limit m0 → 0 while u → c, then p can remain finite.


Evidently a particle without mass can carry momentum, provided that it
travels at the speed of light. From Eq. (13.12),

E 2 = (pc)2 + (m0 c2 )2 ,
490 RELATIVISTIC DYNAMICS

and if we take m0 = 0, then we have, denoting photon energy by the


symbol ,

 2 = (pc)2 ,
 = pc. (13.13)

We have taken the positive square root because the negative solution
would predict that in an isolated system the momentum of a photon could
increase without limit as its energy dropped. Combining Eq. (13.13) with
Einstein’s relation  = hν, we find that a photon possesses momentum p
of magnitude

p= . (13.14)
c
The direction of the momentum vector is along the direction of travel of
the light wave.
Einstein’s quantum hypothesis was designed to solve a theoretical
dilemma—the spectrum of blackbody radiation—but its first application
was to a totally different problem—the photoelectric effect.

Example 13.4 The Photoelectric Effect


In 1887 Heinrich Hertz discovered that metals can give off electrons
when illuminated by ultraviolet light. This process, the photoelectric
effect, represents the direct conversion of light into mechanical energy
(here, the kinetic energy of the electron). Einstein predicted that the
energy a single electron absorbs from a beam of light at frequency v
is exactly the energy of a single photon, hν. For the electron to escape
from the surface it must overcome the energy barrier that confines it to
the surface. The electron must expend energy W = eΦ to escape from
the surface, where e is the charge of the electron and Φ is an electric
potential known as the work function of the material, typically a few
volts. The maximum kinetic energy of the emitted electron is therefore
K = hν − eΦ.
The work function depends on the poorly known chemical state of the
surface, making the photoelectric effect difficult to investigate. Never-
theless, Robert A. Millikan overcame this problem in 1914 by working
with metal surfaces prepared in a high vacuum system. He plotted the
reverse voltage V needed to prevent the photoelectrons from reaching
a detector as a function of the frequency of light. The voltage is given
by
eV = K = hν − eΦ.
The slope of the plot of V versus ν is
dV h
= . (13.15)
dν e
13.5 THE PHOTON: A MASSLESS PARTICLE 491

The photoelectric effect:

+Volts
experimental results on the energy
of photoelectrons and the
0
frequency of light. The graph is V = 439 × 105
from R.A. Millikan. From R.A.

−Volts
Millikan, Physical Review 7, 355
(1916).
1
½mυ2 = hv − P = PDe
d PD d Volts 10 8 h
= ⋅ =
dν dν 3 × 10 10 e
dV 3
= = 4.124 × 10−15
dν (121.00 − 48.25) × 10°
2 e dV 4.774 × 10−25
h= = × 4.124 = 6.56 × 10−25
300 dν 300

40 × 1013 50 60 70 80 90 100 110 120


Frequency

The graph of Millikan’s results shows the linear relation between en-
ergy and frequency predicted by Einstein, and the slope of the line
provides an accurate value for the ratio of two fundamental constants,
Planck’s constant and the charge of the electron.

The fact that light can interfere with itself, as in the Michelson
interferometer, is compelling evidence that light has wave proper-
ties. Nevertheless, the photoelectric effect illustrates that light also
has particle properties. Einstein’s energy relation, E = hv, pro-
vides the link between these apparently conflicting descriptions of
light by relating the energy of the photon to the frequency of the
wave.

Example 13.5 The Pressure of Light


The photon picture of light provided an immediate explanation for a
phenomenon that was also predicted by Maxwell’s electromagnetic
theory: the pressure of light. If a beam of light is absorbed or reflected
by a body, it exerts a force on the body. The force per unit area, the
radiation pressure, is too small to feel when we are in sunlight but
it can have visible effects. Radiation pressure causes comets’ tails
to always point away from the Sun. On the astronomical scale, it
helps prevent stars from collapsing under their gravitational attraction.
In ultra-high intensity laser beams radiation pressure can be large
enough to compress matter to the high density needed to initiate fusion
reactions.

Energy flow in a light beam is often characterized by the beam’s inten-


sity I, which is the power per unit area of the light beam. If the number
of photons crossing a unit area per second is Ṅ and each photon carries
energy , then I = Ṅ.
492 RELATIVISTIC DYNAMICS

Consider a stream of photons in a monochromatic light beam striking


a perfectly reflecting mirror at normal incidence. The initial momen-
tum of each photon is p = /c directed toward the mirror, and the
total change in momentum after the reflection is 2p = 2/c. The to-
tal momentum change per unit area per second due to the reflection is
2Ṅ p = 2Ṅ/c. This is the force on the light beam due to the mirror. The
reaction force is the pressure P on the mirror due to the light. Hence
2Ṅ 2I
P= = .
c c
The average intensity of sunlight falling on the Earth’s surface at nor-
mal incidence, known as the solar constant, is ≈ 1000 W/m2 . The
radiation pressure of sunlight on a mirror is therefore
P = 2I/c
= 7 × 10−6 N/m2
which is very small compared, for example, to atmospheric pressure
105 N/m2 .

Newtonian particles can be neither created nor destroyed. If they are


combined, their total mass is constant. In contrast, massless particles can
be created and annihilated. The emission of light occurs by the creation
of photons, while the absorption of light occurs by the destruction of
photons. The familiar laws of conservation of momentum and energy,
as expressed in the theory of relativity, let us draw conclusions about
processes involving photons without a detailed knowledge of the inter-
actions, as the following examples illustrate.

Example 13.6 The Compton Effect


The photon description of light seemed so strange that it was not widely
accepted until an experiment by Arthur Compton in 1922 made the
photon picture inescapable: by scattering x-rays from electrons in
matter, and showing that the x-rays scattered like particles undergoing
elastic collisions, and that the dynamics were correctly described by
special relativity.

A photon of visible light has energy in the range of 1 to 2 eV, but pho-
tons of much higher energy can be obtained from x-ray tubes, particle
accelerators, or cosmic rays. X-ray photons have energies typically in
the range 10 to 100 keV. Their wavelengths can be measured with high
accuracy by the technique of crystal diffraction.

When a photon scatters from a free electron, the conservation laws re-
quire that the photon loses a portion of its energy due to the recoil of
the electron. The outgoing photon therefore has a longer wavelength
13.5 THE PHOTON: A MASSLESS PARTICLE 493

than the incoming photon. The shift in wavelength, first observed by


Compton, is known as the Compton effect.

f

θ

i

me φ
u
Ef

Suppose that a photon having initial energy i and momentum i /c is


scattered at angle θ and has final energy  f . The electron has rest mass
me and relativistic mass m = γme . The electron is assumed to be ini-
tially at rest with energy Ei = me c2 . The scattered electron leaves at
angle φ with momentum p and energy E f = mc2 . Here m = me γ =

m0 / 1 − u2 /c2 , where u is the speed of the recoiling electron.

The initial photon energy i is known and the final photon energy  f
and the scattering angle θ are measured. The problem is to calculate
how  f varies with θ.

Conservation of total energy requires


i + me c2 =  f + E f (1)
and conservation of momentum requires
i  f
= cos θ + p cos φ (2)
c c
f
0= sin θ − p sin φ. (3)
c
Because Compton detected only the outgoing photon our object is
to eliminate reference to the electron and find  f as a function of θ.
Equations (2) and (3) can be written
(i −  f cos θ)2 = (pc)2 cos2 φ
( f sin θ)2 = (pc)2 sin2 φ.
Adding,
i 2 − 2i  f cos θ +  2f = (pc)2 . (4)
To solve for  f , we introduce the energy–momentum relation in
Eq. (13.12), which can be written (pc)2 = (mc2 )2 − (me c2 ). Combining
this with Eq. (4) gives
 f 2 − 2i  f cos θ +  2f = (i + me c2 −  f )2 − (me c2 )2 ,
494 RELATIVISTIC DYNAMICS

which reduces to
i
f = . (5)
1 + (i /me 2
c )(1 − cos θ)
Note that the photon’s final energy  f is always greater than zero, which
means that a free electron cannot absorb a photon, but can scatter
it.

Compton measured wavelengths rather than energies in his experiment.


From the Einstein frequency condition, i = hνi = hc/λi and  f = hc/λ f ,
where λi and λ f are the wavelengths of the incoming and outgoing
photons, respectively. In terms of wavelength, Eq. (5) takes the simple
form
h
λ f = λi + (1 − cos θ).
me c
The quantity h/me c is known as the Compton wavelength λC of the
electron and has the value
h
λC =
me c
= 2.426 × 10−12 m
= 0.02426 Å,
where 1 Å = 10−10 m. (Å, called the angstrom, is a non-SI unit
formerly used for wavelength measurements.)

The shift in wavelength at a given angle is independent of the initial


photon energy:
λ − λ0 = λC (1 − cos θ).
The figure shows one of Compton’s results for λ0 = 0.711 Å and
θ = 90◦ . The peak P is due to primary photons while the peak T is for
photons scattered from a block of graphite. The measured wavelength
shift is approximately 0.0246 Å and the calculated value is 0.02426
Intensity

Å. The difference is less than the estimated uncertainty due to the
experimental limitations.

P T We have assumed that the electron was free and at rest. For sufficiently
high photon energies, this is a good approximation for electrons
λ, Å
0.7110 0.7356 in the outer shells of light atoms. If the motion of the electrons is
taken into account, the Compton peak is broadened or can have
structure.

If the binding energy of the electron is comparable to the photon


energy, momentum and energy can be transferred to the atom as a
whole, and the photon can be completely absorbed.
13.5 THE PHOTON: A MASSLESS PARTICLE 495

Example 13.7 Pair Production


We have seen two ways by which a photon can lose energy in matter:
photoelectric absorption and Compton scattering. If a photon’s energy
is sufficiently high, it can also lose energy in matter by the mechanism
of pair production. The rest mass of an electron is m0 c2 = 0.511 MeV.
Can a photon of this energy create an electron? The answer is no, since
this would require the creation of a single electric charge. As far as we
know, electric charge is conserved in all physical processes. However,
if equal amounts of positive and negative charge are created, the total
charge remains zero and charge is conserved. It is therefore possible to
create an electron–positron pair (e− , e+ ), two particles having the same
mass but opposite charge.

v+ A single photon of energy 2m0 c2 or greater has enough energy to form


an e− , e+ pair, but the process cannot occur in free space because it
would not conserve momentum. To show why, imagine that the process

occurs. Conservation of energy gives
v−
hν = m+ c2 + m− c2 = (γ+ + γ− )m0 c2 ,
or

= (γ+ + γ− )m0 c,
c
while conservation of momentum gives

= |γ+ v+ + γ− v− | m0 .
c
These equations cannot be satisfied simultaneously because
(γ+ + γ− )c > |γ+ v+ + γ− v− | .
Pair production is possible if a third particle is available for carrying off
e+ the excess momentum. For instance, suppose that the photon collides
M with a nucleus of rest mass M0 and creates an e− , e+ pair at rest. We

M V have
e−
hν + M0 c2 = 2m0 c2 + M0 c2 γ.
Since nuclei are much more massive than electrons, let us assume that
hν M0 c2 . (For hydrogen, the lightest atom, this means that hν 940
MeV.) In this case the atom will not attain relativistic speeds and we can
make the classical approximation
hν = 2m0 c2 + M0 c2 (γ − 1)
1
≈ 2m0 c2 + MV 2 .
2
To the same approximation, conservation of momentum yields

= MV.
c
496 RELATIVISTIC DYNAMICS

Substituting this in the energy expression gives


1 (hν)2
hν = 2m0 c2 + ≈ 2m0 c2 ,
2 Mc2
since we have already assumed hν Mc2 . The threshold for pair pro-
duction in matter is therefore 2m0 c2 = 1.02 MeV. The nucleus plays an
essentially passive role, but by providing for momentum conservation
it allows a process to occur that would otherwise be forbidden by the
conservation laws.

Example 13.8 The Photon Picture of the Doppler Effect


E In Chapter 12 we analyzed the relativistic Doppler effect from the
p standpoint of waves but we can also understand it from the photon
picture. Consider first an atom with rest mass M0 , held stationary. If
M0
the atom emits a photon of energy hν0 , the atom’s new mass is given
by M0 c2 = M0 c2 − hν0 .

Next, we suppose that before emitting the photon the atom moves
freelywith velocity u. The atom’s energy is E = Mc2 = γM0 c2 , where
γ = 1 − u2 /c2 and the atom’s momentum is p = Mu = M0 γu. After
emitting a photon of energy hν the atom has velocity u , rest mass M0 ,
energy E  , and momentum p . For simplicity, we consider the photon
to be emitted along the line of motion.

E′ By conservation of energy and momentum we have


p′
E = E  + hν (1)

M ′0

p = p + . (2)
c
Rearranging Eqs. (1) and (2) gives
(E − hν)2 = E 2
(pc − hν)2 = (p c)2 .
Subtracting, and using Eq. (13.12), E 2 − (pc)2 = (m0 c2 )2 , we have
(E − hν)2 − (pc − hν)2 = E 2 − (p c)2 = (M0 c2 )2 (3)
by the energy–momentum relation. Expanding the left-hand side and
using E 2 − (pc)2 = (M0 c2 )2 , with M0 c2 = M0 c2 − hν0 , we obtain
(M0 c2 )2 − 2Ehν + 2(pc)(hν) = (M0 c2 )2
= (M0 c2 − hν0 )2 .
Simplifying, we find
(2M0 c2 − hν0 )
v = v0 .
2(E − pc)
13.5 THE PHOTON: A MASSLESS PARTICLE 497

However,
u

E − pc = M0 c2 γ 1 −
c

1 − u/c
= M0 c2 .
1 + u/c
Hence

hν0 1 + u/c
ν = ν0 1 − .
2M0 c2 1 − u/c

The term hν0 /2M0 c2 represents a decrease in the photon energy due to
the recoil energy of the atom. Usually the recoil energy is so small that
it can be neglected, leaving

1 + u/c
ν = ν0 ,
1 − u/c

in agreement with the wave analysis that led to Eq. (12.12). However,
the wave picture does not readily take into account the recoil of the
atom. In modern experiments using high precision lasers and ultra-cold
atoms, the recoil cannot be overlooked. On the contrary, it plays a cru-
cial role in many studies.

Example 13.9 The Photon Picture of the Gravitational


Red Shift
In Chapter 9 we derived an expression for the effect of gravity on
time—the gravitational red shift—by invoking the equivalence prin-
ciple. However, the effect of gravity on time can also be understood
using the photon description of light and the conservation of energy.

Atoms can absorb or emit photons at certain characteristic frequen-


cies. For a frequency ν0 , the atom loses energy hν0 when it emits a
photon, going from an upper energy state E1 to a lower energy state,
E0 , and it can gain energy hν0 when it absorbs a photon, reversing the
process.

Consider an atom with rest mass M0 in its ground state with energy
E0 = M0 c2 , in a gravitational field g. It absorbs a photon that increases
its energy to E1 = E0 + hν0 . The mass of the atom is M1 = E1 /c2 =
(E0 + hν0 )/c2 . If we lift the atom to height H in a gravitational field g
the work that we do is M1 gH, so the final energy Wa of the atom is
Wa = E1 + M1 gH
= (E0 + hν0 )(1 + gH/c2 )
= E0 + hν0 + hν0 gH/c2 + E0 gH/c2 .
498 RELATIVISTIC DYNAMICS

Consider an alternative scenario: the atom is first lifted to height H


while it is in state E0 , and then a photon of energy hν is radiated up-
ward to put the atom in state E1 . The energy Wb of the atom with this
procedure is
Wb = E1 + M0 gH = E0 + hν + E0 gH/c2 .
The final state of the system is the same in both scenarios. Conse-
quently Wa = Wb and it follows that
hν = hν0 (1 + gH/c2 ).
In fractional form, the gravitational red shift is
ν − ν0 gH
= 2.
ν0 c
A word of explanation about the adjective “red.” Our result reveals that
if radiation travels outward from the Earth to a region of higher (less
negative) gravitational potential, its energy decreases. Consequently,
radiation emitted by a massive body such as the Sun is observed to
shift to lower energy, equivalently to longer wavelengths, toward the
red end of the spectrum. In contrast, radiation that comes down to the
Earth from a satellite, for instance the signal from an atomic clock, is
shifted to higher energy, which might be called a blue shift.

13.6 How Einstein Derived E = mc2


Einstein’s famous equation E = mc2 is not to be found in his historic pa-
per on relativity but only appeared a few months later in a short note
titled Does the Inertia of a Body Depend Upon Its Energy Content?
(translated from German). His argument was elegant in its simplicity,
based entirely on elementary considerations of energy, momentum, and
the Doppler shift.
Consider a body in system S at rest at the origin. The body has en-
ergy E0 initially, and then sends out a pulse of light with energy /2 in
the +x direction, and simultaneously a pulse with energy /2 in the −x
direction. The body remains at rest after the emission by conservation of
momentum, and its energy is then E1
1 1
E0 = E1 +  + .
2 2

In system S moving with velocity v with respect to S , the initial energy
of the body is H0 and its energy after the emission is H1 . Taking the
Doppler shift into account,
⎛ ⎞ ⎛ ⎞
1 ⎜⎜⎜⎜ 1 − v/c ⎟⎟⎟⎟ 1 ⎜⎜⎜⎜ 1 + v/c ⎟⎟⎟⎟
H0 = H1 +  ⎜⎝  ⎟⎠ +  ⎜⎝  ⎟⎠
2 1 − v2 /c2 2 1 − v2 /c2

= H1 +  .
1 − v2 /c2
PROBLEMS 499

The energy differences in the two systems are


⎛ ⎞
⎜⎜⎜ ⎟⎟
− 1⎟⎟⎟⎠ .
⎜ 1
(H0 − E0 ) − (H1 − E1 ) =  ⎜⎝ 
1 − v /c
2 2

Einstein argued that the difference H − E must equal the kinetic energy
K of the body, to within an additive constant C that is independent of the
relative velocity
H0 − E0 = K0 + C
H1 − E1 = K1 + C.
Thus
⎛ ⎞
⎜⎜ ⎟⎟
K0 − K1 =  ⎜⎜⎜⎝  − 1⎟⎟⎟⎠
1
1 − v2 /c2
2
1 v
≈  .
2 c2
Classically,
1
K0 − K1 = Δmv2 .
2
Einstein then obtained his famous equation by comparing the two results
for K0 − K1 :

Δm = 2 .
c
Einstein concluded his brief paper by asserting that the equivalence of
mass and energy must be a general law, holding for any form of energy,
not just radiation.

Problems
For problems marked *, refer to page 525 for a hint, clue, or answer.
13.1 Energetic proton
Cosmic ray primary protons with energy up to 1020 eV (almost
10 J) have been detected. Our galaxy has a diameter of about 105
light years.
(a) How long does it take the proton to traverse the galaxy, in
its own rest frame (proper time)? (1 eV = 1.6 × 10−19 J, M p =
1.67 × 10−27 kg.) What is the proper time for a photon to traverse
our galaxy?
(b) Compare the proton’s energy to the kinetic energy of a base-
ball, mass = 145 g, traveling at 100 miles/hour.
13.2 Onset of relativistic effects
When working with particles it is important to know when rela-
tivistic effects have to be considered.
500 RELATIVISTIC DYNAMICS

A particle of rest mass m0 is moving with speed v. Its classical


kinetic energy is Kcl = m0 v2 /2. Let Krel be the relativistic expres-
sion for its kinetic energy.
(a) By expanding Krel /Kcl in powers of v2 /c2 , estimate the
value of v2 /c2 for which Krel differs from Kcl by 10 percent.
(b) For this value of v2 /c2 , what is the kinetic energy in MeV
of
(1) an electron (m0 c2 = 0.51 MeV)?
(2) a proton (m0 c2 = 930 MeV)?
13.3 Momentum and energy
In Newtonian mechanics, the kinetic energy of a mass m moving
with velocity v is K = mv2 /2 = p2 /(2m) where p = mv. The change
in kinetic energy due to a small change in momentum is dK =
p · dp/m = v · dp.
Show that the relation dK = v · dp also holds in relativistic
mechanics.
13.4 Particles approaching head-on*
Two particles of rest mass m0 approach each other with equal and
opposite velocity v in the laboratory frame. What is the total en-
ergy of one particle as measured in the rest frame of the other?
13.5 Speed of a composite particle after an inelastic collision*
A particle of rest mass m0 and speed v collides with a stationary
particle of mass M and sticks to it. What is the final speed of the
composite particle?
13.6 Rest mass of a composite particle*
A particle of rest mass m0 and kinetic energy xm0 c2 , where x is
some number, strikes an identical particle at rest and sticks to it.
What is the rest mass of the resultant particle?
13.7 Zero momentum frame*
In the laboratory frame a particle of rest mass m0 and speed v is
moving toward a particle of mass m0 at rest.
What is the speed of the inertial frame in which the total mo-
mentum of the system is zero?
13.8 Photon–particle scattering*
A photon of energy i collides with a free particle of mass m0 at
rest. If the scattered photon flies off at angle θ, what is the scatter-
ing angle φ of the particle?

f


i θ
m0 φ
u
Ef
PROBLEMS 501

13.9 Photon–electron collision*


E0 A photon of energy E0 and wavelength λ0 collides head-on with a
V
m0 free electron of rest mass m0 and speed V, as shown. The photon
is scattered at 90◦ .
(a) Find the energy E of the scattered photon.
(b) The outer electrons in a carbon atom move with speed v/c ≈
6 × 10−3 . Using the result of part (a), estimate the broadening
in wavelength of the Compton scattered peak from graphite for
λ0 = 0.711 × 10−10 m and 90◦ scattering. The rest mass of an
electron is 0.51 MeV and h/(m0 c) = 2.426 × 10−12 m. Neglect the
binding of the electrons. Compare your result with Compton’s
data shown in Example 13.6.
13.10 The force of sunlight
The solar constant, the average energy per unit area from the Sun
falling on the Earth, is 1.4 × 103 W/m2 .
(a) How does the total force of sunlight compare with the Sun’s
gravitational force on the Earth?
(b) Sufficiently small particles can be ejected from the solar
system by the radiation pressure of sunlight. Assuming a specific
gravity of 5, what is the radius of the largest particle that can be
ejected?
13.11 Levitation by laser light
A 1-kW light beam from a laser is used to levitate a solid alu-
minum sphere by focusing it on the sphere from below. What is
the diameter of the sphere, assuming that it floats freely in the
light beam? The density of aluminum is 2.7 g/cm3 .
13.12 Final velocity of a scattered particle
A photon of energy i = hν scatters from a free particle at rest of
mass m0 . The photon is scattered at angle θ with energy  f = hν ,
Light beam
and the particle flies off at angle φ.
Find an expression for the final velocity u of the particle.

You might also like