You are on page 1of 87

Lecture notes for the spring semester of 2005

MODERN PHYSICS

Fam Le Kien
Department of Applied Physics and Chemistry,
University of Electro-Communications, Chofu, Tokyo 182-8585, Japan

Textbook: CONCEPTS OF MODERN PHYSICS (sixth edition, 2003)


By Arthur Beiser
McGraw-Hill

i
Contents

I. The special theory of relativity 1


A. The Michelson-Morley experiment 1
B. The special theory of relativity 4
C. The Galilean transformation 5
D. The Lorentz transformation 7
E. Length contraction 9
F. Time dilation 11
G. Doppler effect 13
1. Doppler effect in sound 13
2. Doppler effect in light 14
H. The relativity of mass 15
I. Mass and energy 18
J. Velocity addition 21

II. Wave properties of particles 23


A. De Broglie waves 23
B. Wave function: de Broglie waves are waves of probability amplitude 24
C. Describing a wave 25
D. Phase and group velocities of de Broglie waves 27

III. Particle diffraction 32

IV. Uncertainty principle 37


Average value and standard deviation 40
Compatible observables 40
Proof of the uncertainty principle 40
Uncertainty principle from the particle approach 41
Uncertainty principle for energy and time 42

V. Atomic spectra 46
A. Spectral series 46
B. The Bohr atom 47

ii
C. Energy levels and spectra 50
D. Origin of line spectra 51

VI. Correspondence principle 55

VII. The lasers 57

VIII. Quantum mechanics 61

IX. Schrödinger equation 64

X. Particle in a box 69

XI. Finite potential well 75

XII. Tunnel effect 79

iii
I. THE SPECIAL THEORY OF RELATIVITY

A. The Michelson-Morley experiment

In the past, it was assumed that light is a wave propagating in an all-pervading elastic
medium called the ether. In other words, it was assumed that there exists a universal frame
of reference. Let us see what this idea means by considering a simple analogy.
Consider a river of width D which flows with the speed v, see Fig. 1. Two boats start
out from one bank of the river with the same speed V (with respect to the water). Boat
A crosses the river to a point on the other bank directly opposite the starting point and
then returns to the starting point. Boat B heads downstream for the distance D and then
returns to the starting point. Let’s calculate the time required for each round trip.
We first consider boat A. In order to compensate the water current, boat A must head
somewhat upstream, see Fig. 2. The upstream component of the velocity of boat A should
be exactly −v in order to cancel out the river current v. The perpendicular component V 0
is the actual speed across the river. We have the relation

V 2 = V 02 + v 2 . (1)

Hence, the actual speed across the river is


√ p
V0 = V 2 − v2 = V 1 − v 2 /V 2 . (2)

The time for the initial crossing is D/V 0 . The total round-trip time tA is twice D/V 0 , that
is,
2D/V
tA = p . (3)
1 − v 2 /V 2
The case of boat B is somewhat different. As boat B heads downstream, its speed relative
to the shore is V + v, and it travels the distance D in the time

D
. (4)
V +v

On the return trip, the speed of boat B relative to the shore is V − v, and boat B travels
the distance D in the time
D
. (5)
V −v

1
FIG. 1: Boat A goes directly across the river and returns to its starting point, while boat B heads
downstream for an identical distance and then returns.

FIG. 2: Boat A must head upstream in order to compensate for the river current.

The total round-trip time tB is the sum of these times, namely,

D D 2D/V
tB = + = . (6)
V +v V −v 1 − v 2 /V 2

The ratio between the times tA and tB is

tA p
= 1 − v 2 /V 2 . (7)
tB

If we know the common speed V of the two boats and measure the ratio tA /tB , we can
determine the speed v of the river current.

2
ether current

FIG. 3: The Michelson-Morley experiment.

The reasoning used in this problem may be transferred to the analogous problem of the
passage of light waves through the ether. If there is an ether pervading space, we move
through it with at least the speed of the earth’s orbital motion about the sun. From the
point of view of an observer on the earth, the ether is moving past the earth. To detect this
motion, we can use a pair of light beams formed by a beamsplitter instead of a pair of boats,
see Fig. 3. One of these light beams is directed to a mirror along a path perpendicular to the
ether current. The other beam goes to a mirror along a path parallel to the ether current.
The optical arrangement is such that both beams return to the same viewing screen. The
path lengths of the two beams are chosen to be exactly the same.
If there is no ether current, the two beams will arrive at the screen in phase and will
interfere constructively to yield a bright field of view. The presence of an ether current,
however, would cause the beams to have different transit times, so that they would no
longer arrive at the screen in phase but would interfere destructively. This is the essence of
the famous experiment performed by American physicists Michelson and Morley in 1887.
In the Michelson-Morley experiment, although the sensitivity was enough to detect the
expected ether current, no ether current was detected.
The negative result of the Michelson-Morley experiment had two consequences. First, it

3
said that the hypothesis of the ether is wrong. Second, it suggested that the speed of light
in free space is the same everywhere, regardless of any motion of the source or the observer.

B. The special theory of relativity

When we speak of motion, we mean motion relative to a frame of reference. Without a


frame of reference the concept of motion has no meaning. The frame of reference may be
a road, the earth’s surface, the sun, the center of our galaxy; but in every case we must
specify it. The absence of an ether means that there is no universal frame of reference.
Therefore, all motion exists solely relative to the person or instrument observing it. Should
we be isolated in the universe, there would be no way in which we could determine whether
we are in motion or not.
Inertial frame of reference: An inertial frame of reference is a frame in which Newton’s
first law of motion holds true. In such a frame, an object at rest remains at rest and an
object in motion continues to move at constant velocity if no force acts on it. Any frame
of reference that moves at constant velocity with respect to an inertial frame is itself an
inertial frame.
The special theory of relativity, developed by Einstein in 1905, treats problems involving
the motion of inertial frames of references at constant velocity with respect to one another.
It has a profound influence on all of physics.
The special theory of relativity is based upon two postulates:
1) The first postulate: The laws of physics may be expressed in equations having the
same form in all inertial frames of reference moving at constant velocity with respect to one
another. This postulate expresses the absence of a universal frame of reference.
2) The second postulate: The speed of light in free space has the same value in all inertial
frames of reference. This postulate follows directly from the result of the Michelson-Morley
experiment.
The above postulates subvert almost all of the intuitive concepts of time and space we
form on the basis of our daily experience. A simple example will illustrate this statement.
Example: We have two boats on a lake, with boat A stationary in the water and boat B
drifts at the constant velocity v. At the instant that B is abreast of A, a flare is fired by
somebody on one of the boats. The light from the flare travels uniformly in all directions,

4
according to the second postulate of special relativity. An observer in either boat must see a
sphere of light expanding with himself at its center, even though one of the boat is changing
its position with respect to the point where the flare went off.
The above situation is unusual. Why? Let us consider a more familiar analog. Instead
of firing a flare, one of the observers drops a stone into water when the boats are abreast of
each other. A circular pattern of ripples spreads out. The center of the circular pattern is
the point where the stone was dropped. Therefore, the pattern appears different to observers
on each boat.
It is important to recognize that motion and waves in water are entirely different from
motion and waves of light in space; water is in itself a frame of reference while space is not,
and the wave speed in water varies with the observer’s motion while the wave speed of light
does not.

C. The Galilean transformation

Suppose that we are in a frame of reference S and find that an event occurs at the time
t and has the coordinates x, y, z. Consider a different frame of reference S 0 moving with
respect to S at the constant velocity v, see Fig. 4. An observer located in S 0 will find
that the same event occurs at the time t0 and has the coordinates x0 , y 0 , z 0 . How are the
measurements x, y, z, t related to x0 , y 0 , z 0 , t0 .
For simplicity, we assume that v is in the +x direction and that time in both systems is
measured from the instant when the origins of S and S 0 coincide. We intuitively expect that
the measurement x will exceed the measurement x0 by the amount vt while the measurements
y, z, t are the same as the measurements y 0 , z 0 , t0 :

x0 = x − vt,
y 0 = y, (8)

z 0 = z.

We also intuitively expect that times are the same in both frames of reference:

t0 = t. (9)

The set of equations (8) and (9) is known as the Galilean transformation.

5
FIG. 4: Frame S 0 moves in the +x direction with the speed v relative to frame S.

To convert velocity components measured in the frame S to their equivalents in the frame
S 0 , we simply differentiate Eqs. (8) with respect to time. The results are

Vx0 = Vx − v,

Vy0 = Vy , (10)

Vz0 = Vz .

The Galilean transformation and the corresponding velocity transformation are in accord
with our intuitive expectations. However, they violate both of the postulates of special
relativity. The first postulate calls for identical equations of physics in both the S and S 0
frames of reference, but the fundamental equations of electricity and magnetism assume
very different forms when Eqs. (8) and (9) are used. The second postulate calls for the
same value of the speed of light whether determined in S or S 0 . If the speed of light in
the direction x in the S system is c, then in the system S 0 we have c0 = c − v, not correct.
Clearly, a different transformation is required if the postulates of special relativity are to be
satisfied.

6
D. The Lorentz transformation

We now develop a set of transformation equations directly from the postulates of special
relativity. A reasonable guess is
x0 = k(x − vt), (11)

where k is a factor of proportionality that does not depend on x and t but may be a function
of v. The choice of Eq. (11) follows from several considerations: it is linear in x and x0 , so
that a single event in the frame S corresponds to a single event in the frame S 0 ; it is simple;
and it has the possibility of reducing to the equation x0 = x − vt, which is valid in ordinary
mechanics. Because the equations of physics must have the same form in both S and S 0 , we
need only change the sign of v to write the corresponding equation for x in terms of x0 and
t0 :
x = k(x0 + vt0 ). (12)

As in the case of the Galilean transformation, there is nothing to indicate that there
might be differences between y and y 0 and between z and z 0 . Hence we again take

y0 = y
(13)
z 0 = z.

To get the time transformation, we substitute Eq. (11) into Eq. (94). The result is

x = k 2 (x − vt) + kvt0 , (14)

from which we find that


1 − k2
t0 = kt + x. (15)
kv
To find k, we use the second postulate. At the instant t = t0 = 0, the origins of the two
frames of reference S and S 0 are in the same place. Suppose that a flare is set off at the
common origin at t = t0 = 0, and the observers in each system proceed to measure the speed
with which the light spreads out. Both observers must find the same speed c, which means
that the light propagation in the frames S and S 0 is governed by the equations

x = ct (16)

and
x0 = ct0 , (17)

7
respectively. Substituting Eqs. (11) and (15) into Eq. (17) yields
1 − k2
k(x − vt) = ckt + cx. (18)
kv
Solving for x, we find
ckt + vkt 1 + v/c
x= 1−k2
= ct . (19)
k − kv c 1 − (1/k 2 − 1)(c/v)

The use of Eq. (17) gives


1 + v/c
= 1. (20)
1 − (1/k 2 − 1)(c/v)
Hence, we obtain
1
k=p . (21)
1 − v 2 /c2
Thus, the transformation equations are
x − vt
x0 = p
1 − v 2 /c2
y0 = y
(22)
z0 = z
t − (vx/c2 )
t0 = p .
1 − v 2 /c2
The above transformation is called the Lorentz transformation.
The inverse Lorentz transformation is
x0 + vt0
x= p
1 − v 2 /c2
y = y0
(23)
z = z0
t0 + (vx0 /c2 )
t= p .
1 − v 2 /c2
The Lorentz equation reduce to the ordinary Galilean equation when the relative velocity
v of S and S 0 is small compared to the velocity of light c. Therefore, the relativistic effects
to be explored in the remainder of this section are usually small except for the case where
enormous velocities are encountered.
According to the Lorentz transformation, measurements of time and position depend on
the frame of reference of the observer, so that two events occuring simultaneously in one
frame at different places need not be simultaneous in another.

8
Simultaneity
The relative character of time as well as space has many implications. Notable, events
that seem to take place simultaneously to one observer may not be simultaneous to another
observer in relative motion, and vice versa.
Consider two events–the setting off of a pair of flares–that occur at the same time t0 to
somebody at two different locations x1 and x2 . What does the pilot of a spacecraft in flight
see? To him, the flare at x1 and t0 appears at the time

t0 − vx1 /c2
t01 = p , (24)
1 − v 2 /c2

while the flare at x2 and t0 appears at the time

t0 − vx2 /c2
t02 = p . (25)
1 − v 2 /c2

Since x1 6= x2 , we have t01 6= t02 . Hence two events that occur simultaneously to one observer
are separated by a time interval

v(x1 − x2 )/c2
t02 − t01 = p (26)
1 − v 2 /c2

to another observer moving at the speed v relative to the first observer. Thus, simultaneity
is a relative concept.

E. Length contraction

A rod is lying at rest along the x0 axis of a frame of reference S 0 . The coordinates of its
ends are x01 and x02 . The length L0 of the rod is

L0 = x02 − x01 . (27)

Suppose that we measure the length of the rod from a frame of reference S, parallel to which
the rod is moving with the velocity v. Will the length L measured in S be the same as the
length L0 measured in S 0 ?
The length L of the rod in the frame S is determined as

L = x2 − x1 , (28)

9
where x1 and x2 are the coordinates of the rod ends measured at the same time t. According
to the inverse Lorentz transformation, we have
x1 − vt
x01 = p
1 − v 2 /c2
(29)
x2 − vt
x02 = p .
1 − v 2 /c2
Hence, we obtain
x2 − x1
x02 − x01 = p . (30)
1 − v 2 /c2
The use of the definitions (27) and (28) yields
L
L0 = p (31)
1 − v 2 /c2
or, equivalently,
p
L = L0 1 − v 2 /c2 . (32)

According to the above equation, the length of an object in motion with respect to an
observer appears to be shorter than when it is at rest with respect to him. This phenomenon
is called the Lorentz-FitzGerald contraction or the length contraction.
The length of an object is a maximum when measured in a reference frame in which the
object is at rest.
The relativistic length contraction is negligible for ordinary speeds, but it is an important
effect at speeds close to the speed of light. A speed of 3000 km/s seems enormous to us, but
it results in a shortening in the direction of motion by a factor of only
s µ ¶2
L p 3000
= 1 − v 2 /c2 = 1 − = 0.99995 = 99.995%. (33)
L0 3 × 105

On the other hand, a body traveling at 0.8 the speed of light is shortened by a factor of
s µ ¶2
L p 0.8c
2 2
= 1 − v /c = 1 − = 0.6 = 60%. (34)
L0 c

The Lorentz-FitzGerald contraction occurs only in the direction of the relative motion:
if v is parallel to x, the y and z dimensions of a moving object are the same in both S and
S 0.
The Lorentz-FitzGerald contraction is a real physical phenomenon and is different from
the visual effects.

10
F. Time dilation

Time intervals, too, are affected by relative motion. Clocks moving with respect to an
observer appear to tick less rapidly than they do when at rest with respect to him. This
effect is called time dilation.
Let’s consider an event happening at a point x0 in the frame S 0 . An observer in S 0
measures the time. He finds that the event happens from the time t01 to the time t02 . The
duration of the event is
T0 = t02 − t01 . (35)

Meanwhile, an observer in S also measures the time. He finds that the above event happens
from t1 to t2 , where
t01 + (v/c2 )x0
t1 = p (36)
1 − v 2 /c2
and
t02 + (v/c2 )x0
t2 = p . (37)
1 − v 2 /c2
To the observer in S, the duration of the event is

t02 − t01
T = t2 − t1 = p . (38)
1 − v 2 /c2

Hence, we obtain
T0
T =p . (39)
1 − v 2 /c2
Clearly, T > T0 . Thus, a clock moving with respect to an observer appears to tick less
rapidly than it does when at rest with respect to him. In other words, a moving clock runs
more slowly than a stationary clock.
Example 1: µ mesons
We show an interesting manifestation of both the time dilation and the length contraction
in the decay of unstable particles called µ mesons. A µ meson decays into an electron an
average of T0 = 2 × 10−6 s after it comes into being. µ mesons are created high in the
atmosphere by fast cosmic-ray particles arriving at the earth from space. µ mesons reach
sea level in profusion. The typical speed of µ mesons is v = 2.994 × 108 m/s, which is 0.998
of the velocity of light c. In the mean lifetime T0 , µ mesons can travel a distance of only

L = vT0 = (2.994 × 108 m/s) × (2 × 10−6 s) = 600 m. (40)

11
However, µ mesons are actually created at attitudes more than 10 times greater than the
distance L. How to explain this paradox?
a) Let’s examine the problem from the frame of reference of an observer on the ground.
The lifetime of the meson in our reference frame has been extended, due to the relative
motion, to the value
T0 2 × 10−6 2 × 10−6
T =p = √ s= m = 32 × 10−6 s. (41)
2
1 − v /c 2 1 − 0.998 2 0.063

This value is almost 16 times greater than when it is at rest with respect to us. In 32 × 10−6
s, a meson can travel a distance

L0 = vT = (2.994 × 108 m/s) × (32 × 10−6 s) = 9600 m. (42)

This distance is larger than the attitude at which the meson is created.
b) Let’s examine the problem from the frame of reference of the meson. In this frame,
the meson is at rest, its lifetime is T0 = 2 × 10−6 s, the earth ground is moving toward the
meson. Compared to the distance L0 = 9600 m in the frame of the ground, the distance L
p
appears to be shortened by the factor 1 − v 2 /c2 = 0.063, that is,
L p
= 1 − v 2 /c2 . (43)
L0
Hence, we have
p √
L = L0 1 − v 2 /c2 = 9600 1 − 0.9982 m = 9600 × 0.063 m = 600 m. (44)

The travel time in the frame of reference of the meson is L/v = (600 m)/(3 × 108 m/s)
= 2 × 10−6 s. This time is the same as the lifetime of the meson. Thus, the two points of
view give identical results.
Example 2: Twin paradox
Consider the famous relativistic effect known as the twin paradox. This paradox involves
two identical clocks, one remains on earth and the other one is taken on a trip into space at
the speed v and eventually is brought back. It is customary to replace the clocks with the
pair of twins Dick and Jane. Dick is 20 years old when he takes off a space trip at a speed
of 0.8c to a star 20 light-years away. His trip takes 50 years. To Jane, who stays behind,
the pace of Dick’s life is slower than her pace by a factor of
p √
1 − v 2 /c2 = 1 − 0.82 = 0.6 = 60%. (45)

12
To Jane, Dick’s heart beats only 3 times for every 5 beats of her heart; Dick thinks only
3 thoughts for every 5 thoughts of hers. Finally, Dick returns after 50 years according to
Jane’s calendar, but to Dick the trip has taken only 30 years. Dick is therefore 50 years old
whereas Jane is 70 years old.
Where is the paradox? If we consider the situation from the point of view of Dick in the
spacecraft, Jane on the earth is in motion relative to him at a speed of 0.8c. Should not
Jane then be 50 years old when the spacecraft returns, while Dick is then be 70–the precise
opposite of what was conducted above?
But the two situations are not equivalent. Dick changed from one inertial frame to a
different one when he started out, when he reversed direction to head home, and when he
landed on the earth. Jane, however, remained in the same inertial frame during Dick’s trip.
Therefore, the time dilation formula applies to Jane’s observations of Dick, but not to Dick’s
observations of Jane.
To look at Dick’s trip from his perspective, we must take into account that the distance
L he covers is shortened to
p
L = L0 1 − v 2 /c2 = 20 light-years × 0.6 = 12 light-years. (46)

Hence, to Dick, his trip took

2L/v = 2 × 12 light-years/0.8c = 30 light-years. (47)

Thus, the aging of the twins is nonsymmetric. This effect has been verified experimentally.

G. Doppler effect

1. Doppler effect in sound

We are familiar with the increase in frequency of a sound when its source approaches us
(or we approach the source) and the decrease in frequency when the source recedes from us
(or we recede from the source). These changes in frequency constitute the Doppler effect.
The origin of this effect is straightforward. Indeed, successive waves emitted by a source
moving toward an observer are closer together than normal because of the advance of the
source. Consequently, the separation between the waves, i.e. the wavelength, is shorter, and
hence the corresponding frequency is higher.

13
The relation between the source frequency ν0 and the observed frequency ν is
1 + v/c
ν = ν0 , (48)
1 − V /c
where c is the speed of sound, v is the speed of the observer, and V is the speed of the source.
The signs of v and V are plus for approaching and minus receding. Transverse motion does
not cause a Doppler shift in sound.

2. Doppler effect in light

We consider a light source as a clock that ticks ν0 times per second and emits a wave of
light with each tick.
a) Transverse Doppler effect in light
Assume that the observer is moving perpendicular to a line between him and the light
source. In the reference frame of the source, the proper time between two adjacent ticks is
T0 = 1/ν0 . In the reference frame of the observer, the time between two adjacent ticks is
p
T = T0 / 1 − v 2 /c2 . The frequence measured in the reference frame of the observer is
1 1p
ν= = 1 − v 2 /c2 . (49)
T T0
Hence, we have
p
ν = ν0 1 − v 2 /c2 . (50)

The observed frequency ν is always lower than the source frequency ν0 . The same formula
is true when the source is moving perpendicular to the line between the source and the
observer. Thus, transverse motion does cause a Doppler shift in light, unlike in the case of
sound.
b) Longitudinal Doppler effect in light
Assume that the observer is receding from the source. We call S the reference frame
of the source, and call S 0 the reference frame of the observer. Assume that light source is
positioned at x = 0, the first tick occurs at t=0, and the second ticks occurs at t = T0 = 1/ν0 .
Assume that the observer is positioned at x0 = 0. In the frame S 0 , the first tick is received
by the observer at t0 = 0. The second tick occurs at
x − vt −vT0
x0 = p =p
1 − v 2 /c2 1 − v 2 /c2
(51)
t − (vx/c2 ) T0
t0 = p =p .
1 − v 2 /c2 1 − v 2 /c2

14
Since x0 < 0, the second light wave takes a time |x0 |/c to reach the observer. Therefore, the
observer receives the second wave at the time t0 + |x0 |/c, that is, at the time
s
T0 (v/c)T0 1 + v/c 1 + v/c
T =p +p = T0 p = T0 . (52)
1 − v 2 /c2 1 − v 2 /c2 1 − v 2 /c2 1 − v/c
Hence, the observed frequency ν = 1/T is
s
1 − v/c
ν = ν0 . (53)
1 + v/c
The observed frequency ν is lower than the source frequency. The same formula is true for
the motion of the source away from the observer.
In the case where the observer is approaching the source, we have
s
1 + v/c
ν = ν0 . (54)
1 − v/c
In this case, the observed frequency ν is higher than the source frequency. The same formula
is true for the motion of the source toward the observer.
The expanding universe
The Doppler effect in light is an important tool in astronomy. Stars emit light of certain
characteristic frequencies called spectral lines. Motion of a star toward or away from the
earth results in a Doppler shift in these frequencies. The spectral lines of distant galaxies
of stars are all shifted toward the lower-frequency end and hence are called red shifts. Such
shifts indicate that galaxies are receding from us and from each other, that is, the universe
is expanding.

H. The relativity of mass

We have seen that fundamental physical quantities such as length and time have meaning
only when the reference frame in which they are measured is specified. We now show the
mass of a body also depends on the reference frame.
For this purpose, we consider an elastic collision between two particles A and B, see Fig.
5. In this collision, kinetic energy is conserved. The properties of A and B are identical
when determined in the reference frames in which they are at rest. We observe the collision
in two different reference frames S and S 0 which are in uniform relative motion. The frame
S 0 is moving in the +x direction with respect to S at the velocity v.

15
=2L

FIG. 5: An elastic collision as observed in two different frames of reference. The balls are initially
2L apart, which is the same distance in both frames since S 0 moves only in the x direction.

Before the process, particle A had been at rest in frame S, at the point (x = 0, y = −L)
and particle B in frame S 0 , at the point (x0 = 0, y 0 = L). Then, at the same instant
tA = t0B = −L/V , A was thrown in the y direction at the speed VA while B is thrown in the

16
−y 0 direction at the speed VB0 , where

VA = VB0 = V. (55)

Hence the behavior of A as seen from S is exactly the same as the behavior of B as seen
from S 0 . At the time t = tA + L/V = 0 and t0 = t0B + L/V = 0, the particles A and B reach
the origins O and O0 , respectively. Since O = O0 at the time t = t0 = 0, the two particles
collide with each other at this time. When tho two particles collide, A rebounds in the −y
direction at the speed VA as seen from S, while B rebounds in the +y direction at the speed
VB0 as seen from S 0 . The round-trip time T0 for A as measured in the frame S and for B as
measured in S 0 is therefore
2L 2L
T0 = = 0. (56)
VA VB
Due to the time dilation effect, the round-trip time T for B as measured in S is longer than
p
T0 by the factor 1/ 1 − v 2 /c2 , that is,
T0
T =p . (57)
1 − v 2 /c2
Hence, the speed VB of B as measured in S is
p
2L 2L 1 − v 2 /c2 p
VB = = = VA 1 − v 2 /c2 . (58)
T T0
Since the collision is elastic, the total momentum in conserved, that is,

mA VA − mB VB = −mA VA + mB VB , (59)

where mA and mB are the masses of A and B as measured in S. This leads to

mA VA = mB VB . (60)

Inserting Eq. (58) into Eq. (60), we find


p
mA = mB 1 − v 2 /c2 . (61)

In the above example, both A and B are moving in S. In order to obtain a formula for the
mass m of a moving body in terms of its mass m0 when measured at rest, we consider the
limit where VA and VB0 are very small. Then, we have mA = m0 and mB = m and so
m0
m= p . (62)
1 − v 2 /c2

17
We call m0 the rest mass or the proper mass. Thus, the mass m of a moving body is larger
p
than its rest mass m0 by the factor 1/ 1 − v 2 /c2 . The increase of the mass due to motion
is a relativistic effect. We call m the relativistic mass of the body.
The relativistic momentum is defined as

m0 v
p = mv = p . (63)
1 − v 2 /c2

In the theory of special relativity, the Newton’s second law of motion is expressed by the
formula " #
dp d d v
F = = (mv) = m0 p . (64)
dt dt dt 1 − v 2 /c2
According to the above formula, we have

d d d d
F = (mv) = m v + v m = ma + v m. (65)
dt dt dt dt

Here a = dv/dt is the acceleration. If v varies with time, m also varies with time. In this
case, we have F 6= ma.

I. Mass and energy

The most famous formula Einstein obtained from the postulates of special relativity is
the relationship between mass and energy.
We recall from elementary physics that the work done on an object by a constant force F
that acts through a distance s is F s. Here we have assumed that F is in the same direction
as s. If no other forces act on the object and the object starts from rest, all work done on it
becomes kinetic energy K, so K = F s. In the general case where F need not be constant,
the formula for kinetic energy is the integral
Z s
K= F ds. (66)
0

Using the relativistic form of the second law of motion

d(mv)
F = , (67)
dt

18
Eq. (66) becomes
Z Z Z Ã !
s v v
d(mv) m0 v
K= ds = v d(mv) = vd p
0 dt 0 0 1 − v 2 /c2
Z v
m0 v 2 v
=p − m0 p dv
1 − v 2 /c2 1 − v 2 /c2
0
¯v (68)
m0 v 2 p ¯
=p + m0 c2 1 − v 2 /c2 ¯
1 − v 2 /c2 0

m0 c2
=p − m0 c2 .
1 − v 2 /c2

The above equation may be rewritten as

mc2 = K + m0 c2 . (69)

We interpret mc2 as the total energy of the body. It follows from Eq. (69) that, when the
body is at rest, that is, when K = 0, the body nevertheless possesses the energy m0 c2 . The
energy
E0 = m0 c2 (70)

is called the rest energy. The total energy can be written as

m0 c2
E = mc2 = p . (71)
1 − v 2 /c2

In terms of E and E0 , Eq. (69) becomes

E = E0 + K. (72)

Equation (71) says that mass and energy are not independent. Mass can be created or
destroyed, but when this happens, an equivalent amount of energy simultaneously vanishes
or comes into being, and vice versa. Mass and energy are different aspects of the same thing.
The conversion factor between the unit of mass (kg) and the unit of energy (J) is c2 , so
1 kg of matter has an energy of mc2 = 1 kg × (3 × 108 m/s)2 = 9 × 1016 J. This energy is
enough to send a payload of a million tons to the moon.
Kinetic energy at low speeds
The relativistic formula for the kinetic energy is

m0 c2
K = mc2 − m0 c2 = p − m0 c2 . (73)
1 − v 2 /c2

19
Consider the case where v ¿ c. We use the approximation (1 + x)n ∼
= 1 + nx for small x to
expand the first term in the above formula. Then we obtain the expression

1
K = mc2 − m0 c2 ∼
= m0 v 2 , (74)
2

in agreement with classical mechanics.


Energy and momentum
Total energy and momentum are conserved in an isolated system, and the rest energy
of a particle is invariant. Hence these quantities are in some sense more fundamental than
kinetic energy and velocity. Let’s look into how the total energy, rest energy, and momentum
of a particle are related.
Total energy is
m0 c2
E=p . (75)
1 − v 2 /c2
Square of E is
m20 c4
E2 = . (76)
1 − v 2 /c2
Momentum is
m0 v
p= p . (77)
1 − v 2 /c2
We have
m20 v 2 c2
p2 c2 = . (78)
1 − v 2 /c2
When we subtract p2 c2 from E 2 , we obtain

2 m20 c4 − m20 v 2 c2
2 2
E −p c = 2 2
= m20 c4 . (79)
1 − v /c

Hence, we have the relation


E 2 = m20 c4 + p2 c2 . (80)

Massless particles
In classical mechanics, a particle must have rest mass in order to have energy and mo-
mentum. However, this requirement does not hold true in relativistic mechanics.
When m0 = 0 and v < c, we find from Eqs. (75) and (77) that E = p = 0. Such a
particle is meaningless. However, when m0 = 0 and v = c, we have E = 0/0 and p = 0/0,
indicating that E and p can have any values. Thus, massless particles may exist if they

20
travel with the speed of light. The relation between the total energy E and the momentum
p of a massless particle is
E = pc. (81)

An example of massless particles is the photon.

J. Velocity addition

Special relativity postulates that the speed of light c in free space has the same value for
all observers, regardless of their relative motion. Common sense tells us that if we throw a
ball forward at 10 m/s from a car moving at 30 m/s, the ball’s speed relative to the road
will be 40 m/s. What if we switch on the car’s headlights when its speed is v? The same
reasoning suggests that their light ought to have a speed of c + v relative to the road. But
this violates the above postulate. Common sense is not reliable in dealing with light or with
a body moving with a speed comparable to the speed of light. To get the correct results for
velocity addition, we must use the Lorentz transformation.
Consider something moving relative to both S and S 0 . An observer in S measures its
three velocity components to be

dx dy dz
Vx = , Vy = , Vz = . (82)
dt dt dt

Meanwhile, to an observer in S 0 they are

dx0 dy 0 dz 0
Vx0 = , Vy0 = , Vz0 = . (83)
dt0 dt0 dt0

By differentiating the Lorentz transformation equations for x0 , y 0 , z 0 , and t0 , we obtain


dx − vdt
dx0 = p
1 − v 2 /c2
dy 0 = dy
(84)
dz 0 = dz
dt − (v/c2 )dx
dt0 = p .
1 − v 2 /c2

21
So we have
dx
dx0 dx − vdt −v Vx − v
Vx0 = 0
= 2
= dt
dx
= ,
dt dt − (v/c )dx 1 − (v/c2 ) dt 1 − (vVx /c2 )
p dy
p p
0 dy 0 dy 1 − v 2 /c2 dt
1 − v 2 /c2 Vy 1 − v 2 /c2
Vy = 0 = = = , (85)
dt dt − (v/c2 )dx 1 − (v/c2 ) dxdt
1 − (vVx /c2 )
p dz
p p
0 2 /c2 1 − v 2 /c2
dz dz 1 − v V z 1 − v 2 /c2
Vz0 = 0 = = dt
= .
dt dt − (v/c2 )dx 1 − (v/c2 ) dx
dt
1 − (vVx /c2 )
Thus, the formulae for velocity addition are
Vx − v
Vx0 = ,
1 − (vVx /c2 )
p
0 Vy 1 − v 2 /c2
Vy = , (86)
1 − (vVx /c2 )
p
0 Vz 1 − v 2 /c2
Vz = .
1 − (vVx /c2 )
The inverse formulae are
Vx0 + v
Vx = ,
1 + (vVx0 /c2 )
p
Vy0 1 − v 2 /c2
Vy = , (87)
1 + (vVx0 /c2 )
p
Vz0 1 − v 2 /c2
Vz = .
1 + (vVx0 /c2 )
Consider a ray of light emitted in the moving frame S 0 in its direction of motion relative
to S. In this case, we have Vx0 = c. An observer in S will measure the speed
Vx0 + v c+v
Vx = 0 2
= = c. (88)
1 + (vVx /c ) 1 + (vc/c2 )
Thus both observers in S and S 0 find the same value for the speed of light.
Example
Spacecraft A is moving at a speed of 0.9c with respect to the earth. If spacecraft B is
to pass A at a relative speed of 0.5c in the same direction, what speed must B have with
respect to the earth.
Solution
Conventional mechanics says that the speed of B ought to be 1.4c, that is, larger than
the speed of light. However, according to the special relativity, the necessary speed of B is
only
Vx0 + v 0.5c + 0.9c
Vx = 0 2
= = 0.97c. (89)
1 + (vVx /c ) 1 + (0.5c)(0.9c)/c2 )
This speed is less than c.

22
II. WAVE PROPERTIES OF PARTICLES

A. De Broglie waves

A photon of light of frequency ν has the energy E = hν. According to the special relativity
theory, the energy E is related to the momentum p as E = pc. Hence, the momentum of a
photon is

p= . (90)
c
Here, h = 6.626 × 10−34 J s is the Planck’s constant and c = 2.998 × 108 m/s is the speed
of light in free space.
On the other hand, the wavelength of a photon is λ = c/ν. When we use Eq. (90), we
find the following relation between the wavelength and momentum of a photon:

h
λ= . (91)
p

The momentum p describes the particle property of the photon. The wavelength λ de-
scribes the wave property of the photon. Each photon has both particle and wave properties.
De Broglie suggested that Eq. (91) is completely general: It applies not only to photons
but also to material particles. A material particle, i.e., a moving body, can behave as it
has a wave nature. The matter waves are called the de Broglie waves. The wavelength of
a particle is given by Eq. (91) and is called the de Broglie wavelength. The greater the
particle’s momentum, the shorter its de Broglie wavelength.
The momentum of a particle of mass m and velocity v is p = mv. The de Broglie
wavelength is therefore given by
h
λ= . (92)
mv
In the above equation, m is the relativistic mass, which is related to the rest mass m0 as

m0
m= p . (93)
1 − v 2 /c2

The wave and particle properties of moving bodies can never be observed at the same time.
In some situations a moving body resembles a wave and in others it resembles a particle.
Which set of properties is most conspicuous depends on how its de Broglie wavelength
compares with its dimensions and the dimensions of whatever it interacts with.
To illustrate this statement, we show two examples.

23
Example 1
Find the de Broglie wavelength of a dust particle with a mass of 10−15 kg and a velocity
of 1 mm/s.
Solution
Since v ¿ c, we can let m = m0 . Hence

h 6.6 × 10−34 J s
λ= = = 6.6 × 10−16 m. (94)
mv (10−15 kg) × (10−3 m/s)
Assume that the diameter of the dust particle is 1 µm. Then, the de Broglie wavelength
of the dust particle is very small compared with its dimensions. Therefore, we do not expect
to find any wave aspects in its behavior.
Example 2
Find the de Broglie wavelength of an electron with a velocity of 107 m/s. The rest mass
of an electron is m0 = 9.1 × 10−31 kg.
Solution
Since v ¿ c, we can let m = m0 . Hence

h 6.6 × 10−34 J s
λ= = = 7.3 × 10−11 m. (95)
mv (9.1 × 10−31 kg) × (107 m/s)
The radius of the hydrogen atom is 5.3 × 10−11 m. The de Broglie wavelength of the
electron (with a velocity of 107 m/s) is comparable with the dimensions of atoms. Therefore,
the wave character of moving electrons is the key to understanding atomic structure and
behavior.

B. Wave function: de Broglie waves are waves of probability amplitude

In water waves, the quantity that varies periodically is the height of the water surface. In
sound waves, it is pressure. In light waves, electric and magnetic fields vary. What quantity
varies in the case of de Broglie matter waves?
The quantity whose variations make up matter waves is called the wave function and is
denoted by the symbol Ψ. This quantity is a function of space and time, i.e., Ψ = Ψ(r, t).
The value of the wave function Ψ at the particular point r = (x, y, z) in space at the time
t is related to the likelihood of finding the body there at the time. More precisely, |Ψ(r, t)|2 ,

24
the square of the absolute value of the wave function, is the probability of finding the body at
the point r at the time t. The value of the wave function Ψ(r, t) is the probability amplitude.
It can be negative, and is not an observable quantity.

C. Describing a wave

Consider a string stretched along the x direction. We shake the string at x = 0 up and
down along the y direction. Assume that the vibrations are harmonic in character. The
displacement of the string at x = 0 can be written as

y = A cos 2πνt. (96)

Here ν is the frequency of the vibrations and A is their amplitude.


A wave of vibrations propagates along the x direction with a wave velocity vp . Wave
formula: µ ¶
x
y = A cos 2πν t − . (97)
vp
The wave velocity vp is called the phase velocity. Since the wave velocity vp is given by
vp = λν, we have
³ x´
y = A cos 2π νt − . (98)
λ
The angular frequency is defined by the formula

ω = 2πν. (99)

The wave number k is defined by the formula

2π ω
k= = . (100)
λ vp

In terms of ω and k, the wave formula can be written as

y = A cos(ωt − kx). (101)

In the three-dimensional space, k becomes a vector k normal to the wave fronts and x is
replaced by the radius vector r. Then, kx is replaced by the scalar product k · r = kr.
In the case of de Broglie waves, the momentum of the particle is

p = h/λ = hk/2π = h̄k. (102)

25
Here, h̄ = h/2π = 1.054 × 10−34 J s is the reduced Planck’s constant.
Exercise 1:
A photon and a particle have the same wavelength. (a) Compare their linear momenta.
(b) Compare the photon’s energy and the particle’s total energy. (c) Compare the photon’s
energy and the particle’s kinetic energy.
Answer:
(a) They have the same linear momenta: p = h/λ.
(b) The photon’s energy is Eph = hν = hc/λ = pc. The particle’s total energy is
p
Ep = p2 c2 + E02 . Thus the photon’s energy is smaller than the particle’s total energy.
p
(c) The particle’s kinetic energy is K = Ep − E0 = p2 c2 + E02 − E0 < pc. Thus the
particle’s kinetic energy is smaller than the photon’s energy.
Exercise 2:
Show that the de Broglie wavelength of a particle of rest mass m0 and kinetic energy K
p
is λ = hc/ K(K + 2m0 c2 ).
Answer: Be definition, we have λ = h/p = hc/pc. To calculate pc, we use the kinetic
p p
energy K = E − E0 = p2 c2 + E02 − E0 = p2 c2 + m20 c4 − m0 c2 . The latter yields pc =
p
K(K + 2m0 c2 ). Hence, we obtain
p
λ = hc/ K(K + 2m0 c2 ). (103)
p
Note that, when v ¿ c, we have K = m0 c2 (1/ 1 − v 2 /c2 − 1) ¿ m0 c2 . Hence, we obtain

λ = h/ 2m0 K. This formula can be derived by another way using the nonrelativistic
formulae λ = h/p and K = p2 /2m0 .
Exercise 3:
Show that if the total energy of a moving particle greatly exceeds its rest energy, its de
Broglie wavelength is nearly the same as the wavelength of a photon with the same total
energy.
p p
Answer: For a particle, we have E = p2 c2 + E02 . Hence, pc = E 2 − E02 . When
E À E0 , we obtain pc ∼= E. In this case, the de Broglie wavelength of the particle is
λp = h/p ∼= hc/E. Meanwhile, for a photon, we always have E = pc and hence, λph =
h/p = hc/E. Thus λp ∼
= λph .

26
D. Phase and group velocities of de Broglie waves

How fast do de Broglie waves travel? Since a de Broglie wave is associated with a moving
body, one may expect that this wave has the same velocity as that of the body. Let us see
if this is true.
We call the de Broglie wave velocity vp . To find vp , we can apply the usual formula

vp = νλ. (104)

The wavelength λ is simply the de Broglie wavelength, i.e.,

h
λ= . (105)
mv

To find the frequency ν, we equate the quantum expression E = hν with the relativistic
expression E = mc2 . Then we obtain

mc2
ν= . (106)
h

The de Broglie wave velocity is therefore


µ 2¶µ ¶
mc h c2
vp = νλ = = . (107)
h mv v

The wave velocity vp is the phase velocity. The particle velocity v is the group velocity.
Because the particle velocity v must be less than the velocity of light c, the de Broglie wave
velocity is always larger than c. To understand this result, we must look into the distinction
between phase velocity and group velocity.
We consider a harmonic wave

y = A cos(ωt − kx). (108)

The de Broglie waves associated with a moving body cannot be represented simply by a for-
mula resembling Eq. (108). Instead, the wave representation of a moving body corresponds
to a wave packet, or wave group. A wave group is a superposition of individual waves of
different wavelengths. The interference of the individual waves with one another results in
the variation in amplitude that defines the group shape. If the velocities of the individual
waves are the same, the velocity of the wave group is the common phase velocity. However,

27
FIG. 6: A wave group.

if the phase velocity varies with wavelength, an effect called dispersion, the different indi-
vidual waves do not proceed together. As a result, the wave group has a velocity different
from the phase velocities of the individual waves. This is the case with de Broglie waves.
As an example, we consider the case where the wave group consists of two waves that
have the same amplitude A but differ by a small amount ∆ω in angular frequency and a
small amount ∆k in wave number:

y1 = A cos(ωt − kx),
y2 = A cos[(ω + ∆ω)t − (k + ∆k)x]. (109)

The wave group is then given by

y = y1 + y2
1 1
= 2A cos (∆ω t − ∆k x) cos [(2ω + ∆ω)t − (2k + ∆k)x]. (110)
2 2

Since ∆ω ¿ ω and ∆k ¿ k, we find

1
y = 2A cos (∆ω t − ∆k x) cos(ωt − kx). (111)
2

The above equation represents a wave of angular frequency ω and wave number k whose
amplitude is modulated by an angular frequency ∆ω/2 and a wave number ∆k/2.
The effect of the modulation is to produce successive wave groups. The phase velocity vp
is
ω
vp = , (112)
k
and the velocity of the successive wave groups is

∆ω
vg = . (113)
∆k

28
When ω and k have continuous spreads instead of the two values in the above discussion,
the group velocity is

vg = . (114)
dk
We now use Eqs. (112) and (114) to calculate the phase and group velocities of de Broglie
waves. The angular frequency of the de Broglie waves associated with a body of rest mass
m0 moving with the velocity v is

2πmc2 2πm0 c2
ω = 2πν = = p . (115)
h h 1 − v 2 /c2

The wave number of the de Broglie waves is

2π 2πmv 2πm0 v
k= = = p . (116)
λ h h 1 − v 2 /c2

The group velocity of the de Broglie waves is

dω dω/dv
vg = = . (117)
dk dk/dv

It follows from Eqs. (115) and (116) that

dω 2πm0 v
= ,
dv h(1 − v 2 /c2 )3/2
dk 2πm0
= . (118)
dv h(1 − v 2 /c2 )3/2

Hence, we find
vg = v. (119)

Thus, the group velocity of the de Broglie waves is the velocity of the moving body.
The phase velocity of the de Broglie waves is, as found earlier,

ω c2
vp = = . (120)
k v

The fact that vp > c does not violate the special relativity theory because vp is the motion
of the phase of the wave group, not the motion of the individual waves that make up the
group, and consequently, not the motion of the body.
Exercise 1:
An electron has a de Broglie wavelength of 2 × 10−12 m. Find its kinetic energy and the
phase and group velocities of its de Broglie waves.

29
Solution: First, we calculate pc:

hc (6.6 × 10−34 )(3 × 108 )


pc = = = 1 × 10−13 J. (121)
λ 2 × 10−12

The rest energy of the electron is E0 = m0 c2 = (9.1 × 10−31 )(3 × 108 )2 = 0.8 × 10−13 J. The
kinetic energy of the electron is
q p
K = E − E0 = p2 c2 + E02 − E0 = (1 × 10−13 )2 + (0.8 × 10−13 )2 − 0.8 × 10−13
= 0.48 × 10−13 J. (122)

In the units of eV (1 eV= 1.6 × 10−19 J), we have K = 3 × 105 eV = 300 keV.
To find the electron velocity v, we use the formula

v mcv pc pc 1 × 10−13
= = = p = p = 0.78. (123)
c mc2 E p2 c2 + E02 (1 × 10−13 )2 + (0.8 × 10−13 )2

Hence, the group velocity is vg = v = 0.78 c and the phase velocity is vp = c2 /v = 1.28 c.
Exercise 2:
A proton and an electron have the same velocity. Compare the wavelengths and phase
and group velocities of their de Broglie waves.
Answer: (a) λ = h/p = h/mv. The electron has a smaller mass and consequently a longer
wavelength compared to those of the proton.
(b) Since vg = v, the de Broglie waves of the proton and electron have the same group
velocity.
(c) Since vp = c2 /v, the de Broglie waves of the proton and electron have the same phase
velocity.
Exercise 3:
(a) A proton and an electron have the same kinetic energy. (b) Compare the wavelengths
and phase and group velocities of their de Broglie waves.
p
Answer: (a) λ = h/p and pc = K(K + 2m0 c2 ) (or K ∼ = p2 /2m0 in nonrelativistic
considerations). The electron has a smaller mass than the proton. Therefore, we have
λelec > λproton .
p
(b) From K = E − E0 = m0 c2 (1/ 1 − v 2 /c2 − 1), we find
µ ¶2
v2 K
=1− 1− .
c2 K + m0 c2

30
Since melec < mproton , we have velec > vproton . From the nonrelativistic formula K ∼
= m0 v 2 /2,
we get the same conclusion. Since vg = v and vp = c2 /v, the electron has a larger group
velocity and a smaller phase velocity than the proton does.
Exercise 4:
Verify the statement that, if the phase velocity is the same for all wavelengths of a certain
wave phenomenon, the group and phase velocities are the same.
Answer: vp = ω/k and vg = dω/dk. If vp is a constant then vg = vp .
Exercise 5:
(a) Show that the phase group velocity of a particle of rest mass m0 and de Broglie
p
wavelength λ is vp = c 1 + (m0 cλ/h)2 . (b) Compare the group and phase velocities of an
electron with the de Broglie wavelength of 1 × 10−13 m.
p
Answer: (a) We have λ = h/p and p = m0 v/ 1 − v 2 /c2 . Hence, we find
p
v = c/ 1 + (m0 cλ/h)2 . (124)

Since vg = v and vp = c2 /v, we obtain


p
vp = c 1 + (m0 cλ/h)2 . (125)

(b) For m0 = 9.1 × 10−31 kg, h = 6.6 × 10−34 Js, c = 3 × 108 m/s, λ = 1 × 10−13 m, we find
1 + (m0 cλ/h)2 ∼
= 1.0016. Consequently, vp /vg = 1.0016, vp ∼ = 1.0008 c and vg ∼
= 0.9992 c.

31
FIG. 7: Scheme of the Davission-Germer experiment.

III. PARTICLE DIFFRACTION

A wave effect with no analog in the behavior of Newtonian particles is diffraction. In


1927, Davisson and Germer demonstrated that electron beams are diffracted when they are
scattered by the regular atomic arrays of crystal.
Davisson and Germer studied the scattering of electrons from a solid using an apparatus
sketched in Fig. 7. The energy of the electrons in the primary beam, the angle at which
they reach the target, and the position of the detector could all be varied.
Classical physics predicts that the scattered electrons will emerge in all directions with
only a moderate dependence of their intensity on scattering angle and even less on the energy
of the primary electrons. Using a block of nickel as the target, Davisson and Germer verified
these predictions. To prevent the crystal from being oxidized, the apparatus was kept in the
vacuum.
In the midst of their work an accident occurred that allowed air to enter their apparatus
and oxidize the metal surface. To reduce the oxide, the target was baked in a hot oven. After
this treatment, the target was returned to the apparatus and the measurements resumed.
Then the results were very different. Instead of a continuous variation of scattered electron
intensity with angle, distinct maxima and minima were observed whose positions depend on
the electron energy, see Fig. 8.
Two questions arise: What is the reason for this new effect? Why did it not appear until
the nickel target was baked?
De Broglie’s hypothesis suggested that electron waves were being diffracted by the target.

32
FIG. 8: Results of the Davission-Germer experiment.

This was realized when heating a block of nickel at high temperature causes the many
individual crystals in the target to form a single large crystal. This crystal acts like a set
of many planes of atoms, called Bragg planes. The spacing of the planes in this crystal is
denoted by d = 0.091 nm. The angle of incidence and scattering relative to the Bragg planes
is denoted by θ. The Bragg equation for maxima in the diffraction pattern is

nλ = 2d sin θ. (126)

When the kinetic energy of electrons is 54 eV, the angle is θ = 65◦ . For n = 1 and with
sin 65◦ = 0.906, the de Broglie wavelength λ is estimated to be

λ = 2d sin θ = 2 × 0.091 nm × sin 65◦ = 0.165 nm. (127)

Now we use de Broglie’s formula λ = h/mv to find the expected wavelength of the
electrons. The electron kinetic energy of 54 eV is small compared with its rest energy m0 c2
of 0.51 MeV, so we can ignore relativistic considerations. Therefore, the electron kinetic
energy is given by
mv 2
K= . (128)
2
Hence the electron momentum is
√ p
mv = 2mK = (2)(9.1 × 10−31 kg)(54 eV)(1.6 × 10−19 J/eV) = 4×10−24 kg m/s. (129)

The electron wavelength is therefore

h 6.6 × 10−34 Js
λ= = = 0.165 nm. (130)
mv 4 × 10−24 kg m/s

33
FIG. 9: The diffraction of the de Broglie waves by the target is responsible for the results of the
experiment.

Note that the realistic situation is more complicated than the above analysis. For ex-
ample, a complication arises from the fact that the energy of an electron increases when it
enters a crystal. Another complication is that there are several families of Bragg planes. The
interference between the waves diffracted from different families may prevent the observation
of maxima even when the Bragg condition is satisfied.
Exercise:
What effect on the scattering angle in the Davisson-Germer experiment does increasing
the electron energy have?
Answer: leads to a decrease of θ and an increase of the scattering angle ϕ.
Exercise:
A beam of 50-keV electrons is directed at a crystal and diffracted electrons are found at
an angle of 50◦ relative to the original beam. What is the spacing of the atomic planes of
the crystal? A relativistic calculation is needed for λ. Here m0 c2 = 0.5 MeV.
Answer: We have

hc
λ = p
K(K + 2m0 c2 )
(6.6 × 10−34 Js)(3 × 108 m/s)
= p
1.6 × 10−19 J/eV (50 × 103 eV)(50 × 103 eV + 2(0.5 × 106 eV))
(6.6 × 10−34 )(3 × 108 ) 19.8
= p m= × 10−11 m
1.6 × 10 −19 4
× [10 × 5 × (5 + 100)] 1.6 × 23
−11
= 0.5 × 10 m = 0.005 nm. (131)

34
The spacing between the atomic planes is
λ 0.005 nm 0.005 nm
d= = ◦
= = 2.76 × 10−3 nm. (132)
2 sin θ 2 sin 65 2 × 0.906
Exercise:
A beam of 5.4-keV electrons is directed at a crystal and diffracted electrons are found at
an angle of 50◦ relative to the original beam. What is the spacing of the atomic planes of
the crystal? We can ignore relativistic considerations.
Answer: We have
h
λ = √
2m0 K
6.6 × 10−34 Js
= p
(2)(9.1 × 10−31 kg)(5.4 × 103 eV)(1.6 × 10−19 J/eV)

6.6 × 10−34 m 6.6 × 10−34 m


= √ = = 1.65 × 10−11 m = 0.0165 nm. (133)
15.7 × 10−46 4 × 10−23
The spacing between the atomic planes is
λ 0.0165 nm 0.0165 nm
d= = ◦
= = 9.1 × 10−3 nm. (134)
2 sin θ 2 sin 65 2 × 0.906
Particle in a box
The wave nature of a moving particle leads to some remarkable consequences when the
particle is restricted to a certain region of space instead of being able to move freely.
The simplest case is that of a particle in a box. We assume that the particle can move
only along one direction of the box, bouncing back and forth between the walls. We assume
that the walls are infinitely hard, so the particle does not lose energy each time it strikes a
wall. We also assume that the velocity of the particle is sufficiently small that we can ignore
relativistic considerations.
From a wave point of view, a particle trapped in a box is like a standing wave. The
possible de Broglie wavelengths of the particle are therefore determined by the width L of
the box. The longest wavelength is λ = 2L, the next is λ = L, then λ = 2L/3, and so forth.
The general formula is
2L
λ= (n = 1, 2, 3, . . . ). (135)
n
Because mv = h/λ, the restrictions on λ imposed by the box width L are equivalent to
limits on the momentum of the particle and, in turn, to limits on its kinetic energy. The

35
kinetic energy of the particle is
mv 2 h2
K= = . (136)
2 2mλ2
Since the permitted wavelengths are λn = 2L/n and the particle has no potential energy in
this model, the permitted energies of the particle are

n2 h2
En = (n = 1, 2, 3, . . . ). (137)
8mL2

Each permitted energy is called an energy level. The integer number n that specified an
energy level En is called its quantum number.
We can draw three general conclusions:
1. A trapped particle cannot have an arbitrary energy, as a free particle can. The energies
are quantized (discrete) and can be characterized by a quantum number.
2. A trapped particle cannot have zero energy. The zero energy means v = 0 and
therefore λ = ∞. There is no way to trap a wave with an infinite wavelength in a box.
3. Because the Planck constant h is very small–only 6.63 × 10−34 J s – quantization of
energy is conspicuous only when m and L are also small. This is why we are not aware
of energy quantization in our own experience. The smaller the confinement, the larger the
energy required for confinement.
If a particle is confined into a rectangular volume (a three-dimentional box), the permitted
energies are
(n21 + n22 + n23 )h2
En1 n2 n3 = (n = 1, 2, 3, . . . ). (138)
8mL2
Exercise: The lowest possible energy of a particle in a box is 1 eV. What are the next
two higher energies the particle can have?
Answer: 4 eV and 9 eV.

36
FIG. 10: (a) A narrow de Broglie wave group. (b) A wide wave group.

IV. UNCERTAINTY PRINCIPLE

Look at the wave group of Fig. 6. The particle that corresponds to this wave group can
be found anywhere within the group at a given time. The probability of finding the particle
is given by |Ψ|2 .
When the wave group is narrower, the particle’s position can be specified more precisely,
see Fig. 10(a). However, the wavelength of the waves in a narrow packet is not well defined.
The reason is that the range of wavelengths of individual waves is large. This means that,
since λ = h/mv, the particle’s momentum is not a precise quantity. If we make a series of
momentum measurements, we will find a broad range of values.
When the wave group is wider, the particle’s wavelength can be specified more precisely,
see Fig. 10(b). Therefore, the momentum can be measured more precisely. However, since
the wave group is wide, the position of the particle is not well defined. If we make a series
of position measurements, we will find a broad range of values.
Thus we have the uncertainty principle (discovered by Heisenberg in 1927):
It is impossible to know both the exact position and exact momentum of an object at the
same time.
We present a mathematical expression for this principle. A moving body corresponds to
a single wave group, not a series of them. An isolated wave group is a superposition of an
infinite number of wave trains with different frequencies, wave numbers, and amplitudes, see

37
FIG. 11: An isolated wave group is the result of superposing an infinite number of waves with
different wave lengths.

Fig. 11.
At a certain time t, when Ψ(x) is a real function, the wave group Ψ(x) can be represented
by the Fourier integral Z ∞
Ψ(x) = g(k) cos kx dk. (139)
0

More precisely and more generally, we have the formula


Z ∞
Ψ(x) = g(k)eikx dk. (140)
−∞

The function g(k) describes the amplitudes and wavelengths of the harmonic waves that
contribute to the wave group. The narrower the wave group, the broader the range of
wavelengths involved, see Fig. 12.
We assume that the wave function Ψ(x) is spread in the interval ∆x, and that the
Fourier transform g(k) is spread in the interval ∆k. The spreads ∆x and ∆k are defined as
the standard deviations of x and k, respectively. These spreads are related to each other.
Due to the properties the Fourier transformation, we always have

1
∆x ∆k ≥ . (141)
2

To understand Eq. (141) qualitatively, we consider a simple example. Assume that


g(k) consists of only three individual components. The wave vectors of these individual
components are k0 , k0 − ∆k/2, and k0 + ∆k/2. Their amplitudes are proportional to 1, 1/2,
and 1/2, respectively. We then have
½ ·µ ¶ ¸ ·µ ¶ ¸¾
1 ∆k 1 ∆k
Ψ(x) = g0 cos(k0 x) + cos k0 − x + cos k0 + x
2 2 2 2
· µ ¶¸
∆k
= g0 cos(k0 x) 1 + cos x . (142)
2

38
FIG. 12: The wave functions and Fourier transforms for (a) a pulse, (b) a wave group, (c) a wave
train, and (d) a Gaussian distribution.

The above function is maximum at x = 0 and goes to zero at x = ±2π/∆k. The width of
this function is ∆x = 4π/∆k. Thus we have
1
∆x ∆k = 4π ≥ . (143)
2
If the wave function Ψ(x) is a Gaussian function, then the Fourier transform g(k) is also
a Gaussian function, and we have ∆x ∆k = 1/2. Indeed, we take
µ ¶
x2
Ψ(x) = N exp − 2 . (144)
4a
The width of Ψ(x) is ∆x = a. We find
¡ ¢
g(k) = N 0 exp −a2 k 2 , (145)

which is a Gaussian function. We can prove this with the help of the formula
Z ∞ √
−a2 k2 π − x22
e cos kx dk = e 4a . (146)
−∞ a
The width of g(k) is ∆k = 1/(2a). Thus we have ∆x ∆k = 1/2.
The Fourier transform g(k) is related to the probability of finding the particle with the
momentum p = h̄k. More precisely, |g(k)|2 is proportional to the probability for the particle
to have the momentum p = h̄k. When we use the relation ∆p = h̄∆k and Eq. (141), we
find

∆x ∆p ≥ . (147)
2
The above equation is a mathematical expression of the uncertainty principle. The spreads
∆x and ∆k are the measures of the uncertainties in the position and momentum, respectively,

39
of the particle. Equation (147) says that the product of the uncertainty ∆x in the position
of an object at some instant and the uncertainty ∆p in its momentum component in the x
direction at the same time is equal or greater than h̄/2.
If we arrange the particle in such a way that ∆x is small, then ∆p will be large. If we
reduce ∆p in some way, then ∆x will be large.

Average value and standard deviation

Consider an observable A. Assume that the probability for A to take the value α is
described by the probability density ρ(α). The mean value of A is
Z ∞
hAi = αρ(α) dα. (148)
−∞

We introduce the notation Z ∞


hF (A)i = F (α)ρ(α) dα. (149)
−∞
In particular, we have Z ∞
n
hA i = αn ρ(α) dα. (150)
−∞
The standard deviation ∆A of A is defined by the formula

(∆A)2 = h(A − hAi)2 i = hA2 i − hAi2 . (151)

Compatible observables

If two observables A and B are compatible, they can be described by a joint probability
density ρ(α, β). We can define
Z ∞ Z ∞
hF1 (A)F2 (B)i = F1 (α)F2 (β)ρ(α, β) dαdβ. (152)
−∞ −∞

Proof of the uncertainty principle

Consider the position x and the momentum p of a particle. In quantum mechanics, x


and p are not compatible, that is, the averages hxpi and hpxi cannot be defined in the
conventional way. Moreover, we have

hxpi 6= hpxi. (153)

40
More precisely, we have
hxpi − hpxi = ih̄. (154)

Define δx = x − hxi and δp = p − hpi. It follows from the above equation that

hδxδpi − hδpδxi = ih̄. (155)

As known, CC ∗ ≥ 0. For any real variable ξ, we always have

h(δx + iξδp)(δx − iξδp)i ≥ 0. (156)

We calculate the left-hand side and find

h(δx + iξδp)(δx − iξδp)i = aξ 2 + bξ + c, (157)

where

a = hδp2 i

b = −i(hδxδpi − hδpδxi)

c = hδx2 i. (158)

Since aξ 2 + bξ + c ≥ 0 for any real ξ, the following condition should be satisfied:

b2 ≤ 4ac. (159)

On the other hand, we have a = (∆p)2 , b = h̄, and c = (∆x)2 . Hence, we find

h̄2 ≤ 4(∆p)2 (∆x)2 , (160)

that is,

∆x∆p ≥ . (161)
2

Uncertainty principle from the particle approach

Suppose we look at an electron using light of wavelength λ. Each photon of this light
has the momentum h/λ. We can see the electron only if one of these photons bounces off
the electron. The electron’s original momentum will be changed. The exact amount of the

41
change ∆p cannot be predicted but will be of the same order of magnitude as the photon
momentum h/λ. Consequently, the uncertainty in the electron’s momentum is

h
∆p ≈ . (162)
λ

On the other hand, light is a wave phenomenon as well as a particle phenomenon. We


cannot determine the position of the electron with an accuracy better than the wavelength.
Consequently, we have
∆x ≥ λ. (163)

Combining Eqs. (162) and (163) gives

∆x∆p ≥ h. (164)

This result is consistent with the formula ∆x∆p ≥ h̄/2.

Uncertainty principle for energy and time

Another form of the uncertainty principle concerns energy and time.


Consider the measurement of the energy E emitted during the time interval ∆t in an
atomic process. Assume that the energy is in the form of electro-magnetic waves. The
energy is E = hν. Therefore, the uncertainty in energy is

∆E = h∆ν. (165)

To measure the frequency ν, we account the number of waves N for the interval ∆t and
divide this number by the time interval, that is, ν = N/∆t. Assume that the uncertainty in
number of waves in the wave group is one. Then, the uncertainty in frequency is

1
∆ν ≥ . (166)
∆t

It follows from Eqs. (165) and (166) that

∆E ∆t ≥ h. (167)

A more rigorous treatment gives



∆E ∆t ≥ . (168)
2

42
Thus the product of the uncertainty in an energy measurement and the uncertainty in the
time at which the measurement is made is greater than or equal to h̄/2.
Consider a conservative system. For this system, the greater the energy uncertainty, the
more rapid the time evolution. More precisely, if ∆t is a time interval at the end of which
the system has evolved to an appreciable extent and if ∆E denotes the energy uncertainty,
∆t and ∆E satisfy the relation

∆E ∆t ≥ . (169)
2
The above equation is a mathematical expression of the uncertainty principle for energy and
time.
The proof is given below. Consider a wave packet. The energy uncertainty ∆E is asso-
ciated with the momentum uncertainty ∆p via the formula
dE
∆E = ∆p. (170)
dp
Since E = h̄ω and p = h̄k, we have
dE dω
= = vg . (171)
dp dk
Hence
∆E = vg ∆p. (172)

Now the characteristic evolution time ∆t is the time taken by this wave packet to pass a
point in space. If ∆x is the spatial extension of the wave packet, we have
∆x
∆t = . (173)
vg
From this we obtain

∆E∆t = ∆p∆x ≥ . (174)
2
Example 1
A measurement establishes the position of a proton with an accuracy of ±1 × 10−11 m.
Find the uncertainty in the position of the proton 1 second later. The rest mass of a proton
is m0 = 1.672 × 10−27 kg. Assume v ¿ c.
Solution
The uncertainty in the proton’s position at t = 0 is ∆x0 = 1 × 10−11 m. According to
Eq. (147), the uncertainty in its momentum at this time is

∆p ≥ . (175)
2∆x0

43
Since v ¿ c, the momentum is p = mv = m0 v. Therefore, we have ∆p = m0 ∆v. Hence, the
uncertainty in the proton’s velocity is

∆v ≥ . (176)
2m0 ∆x0
After the time t, the position of the proton cannot be known more accurately than
h̄t
∆x = t∆v ≥ . (177)
2m0 ∆x0
Hence ∆x is inversely proportional to ∆x0 . This means that the more we know about the
proton’s position at a given time, the less we know about its later position.
The value of ∆x at t = 1 s is
(1.054 × 10−34 J s) × (1 s)
∆x ≥ −27 −11
= 3.15 × 103 m. (178)
2 × (1.672 × 10 kg) × (1 × 10 m)
Exercise: (a) Discuss the prohibition of E = 0 for a trapped particle in a box in terms of
the uncertainty principle. (b) How does the minimum momentum of such a particle compare
with the momentum uncertainty required by the uncertainty principle if we take ∆x = L?
Answer: (a) Since the particle is trapped in the box, ∆x is not infinite. Therfore, ∆p
cannot be zero and consequently p cannot be zero. This is why the particle cannot have
E = 0. (b) If we take ∆x = L then ∆p ≥ h̄/2∆x = h̄/2L. On the other hand, the first
permitted value of λ is 2L. Therefore, the minimum momentum is pmin = h/λ = h/2L =
πh̄/L > h̄/2L. Thus the minimum momentum of the trapped particle is larger than the
momentum uncertainty required by the uncertainty principle.
Exercise: Compare the uncertainties in the velocities of an electron and a proton in a
small box.
Answer: Take ∆x = L. Then (∆p)min = h̄/2∆x = h̄/2L. Since p = mv, we have
(∆v)min = h̄/2mL. Hence, the uncertainty in the velocity of the electron is larger than that
of the proton.
Exercise: Verify that the uncertainty principle can be written as ∆L∆θ ≥ h̄/2, where L
is the angular momentum and θ is the angular position.
Answer: Consider the rotational motion of a particle along a circle of radius a. We
have L = mvr and θ = x/r. Hence ∆L = mr∆v = r∆p and ∆θ = ∆x/r. Therefore,
∆L∆θ = ∆p∆x ≥ h̄/2.
Exercise: A hydrogen atom is 5.3 × 10−11 m in radius. Use the uncertainty principle to
estimate the minimum kinetic energy an electron can have in this atom.

44
Answer: Here we have ∆x = 5.3 × 10−11 m. The uncertainty in momentum is

h̄ ∼
∆p ≥ = 1 × 10−24 kg m/s. (179)
2∆x

The momentum of the electron must be at least comparable to its uncertainty. Consequently,
the kinetic energy of the electron is

p2 (1 × 10−24 )2
K= ≥ J ≥ 5 × 10−19 J ∼
= 3 eV. (180)
2m (2)(9.1 × 10−31 )

45
V. ATOMIC SPECTRA

When an atomic gas is excited by passing an electric current through it, the emitted
radiation has a spectrum which contains specific wavelengths only.
Each element has a characteristic line spectrum.
The number, strength, and exact wavelengths of the lines in the spectrum of an element
depend on temperature, pressure, the presence of electric and magnetic fields, and the motion
of the source.
Spectroscopy is therefore a useful tool for analyzing the composition and the state of a
source.

A. Spectral series

It has been experimentally found that the spectral lines of an element fall into sets called
spectral series.
For hydrogen, the Lyman series contains the wavelengths given by the formula
µ ¶
1 1 1
=R − with n = 2, 3, 4, . . . , (181)
λ 12 n2

and is in the ultraviolet region (400–10 nm). Here, R = 1.097 × 107 m−1 is the Rydberg
constant.
The Balmer series contains the lines
µ ¶
1 1 1
=R − with n = 3, 4, 5, . . . , (182)
λ 22 n2

and is in the visible region (800–400 nm).


In the infrared region (from 800 nm to 1 mm), three series have been found. They are
µ ¶
1 1 1
Paschen: =R − with n = 4, 5, 6, . . . , (183)
λ 32 n2
µ ¶
1 1 1
Brackett: =R − with n = 5, 6, 7, . . . , (184)
λ 42 n2
and µ ¶
1 1 1
Pfund: =R 2
− 2 with n = 6, 7, 8, . . . (185)
λ 5 n
The existence of spectral lines and series poses a test for any theory of atomic structure.

46
FIG. 13: Spectral series of hydrogen.

Exercise: What is the shortest wavelength present in the Brackett series of spectral lines?
Answer: The Brackett series is 1/λ = R(1/42 − 1/n2 ) with n = 5, 6, 7, . . . . The shortest
wavelength in this series is λ = 16/R.
Exercise: What is the shortest wavelength present in the Paschen series of spectral lines?
Answer: The Paschen series is 1/λ = R(1/32 − 1/n2 ) with n = 4, 5, 6, . . . . The shortest
wavelength in this series is λ = 9/R.

B. The Bohr atom

The first theory of the atom to meet with any success was put forward in 1913 by Niels
Bohr. This theory can be formulated in terms of the de Broglie waves as shown below.
Consider an electron in orbit around a hydrogen nucleus. The de Broglie wavelength of
this electron is
h
λ= . (186)
mv

47
To determine v, we recall that the centripetal force is

mv 2
Fc = . (187)
r

This force is provided by the electric force

1 e2
Fe = . (188)
4π²0 r2

The condition for a stable orbit is Fc = Fe , i.e.,

mv 2 1 e2
= . (189)
r 4π²0 r2

The electron velocity is therefore found to be

e
v=√ . (190)
4π²0 mr

Hence, the orbital electron wavelength is


r
h 4π²0 r
λ= . (191)
e m

We assume that the motion of the electron in the hydrogen atom is analogous to the
vibrations of a wire loop. We know that, in a wire loop, the loop’s circumference is an
integer number of the wavelength of the resonant mode. Therefore, we assume that an
electron can circle a nucleus only if its orbit contains an integer number of the de Broglie
wavelength. Thus, the condition for orbital stability is

nλ = 2πr, with n = 1, 2, 3, . . . (192)

The integer n is called the quantum number of the orbit. When we substitute Eq. (192)
into Eq. (191), we find that the radii of the orbits are

h2 ²0
r = rn = n2 , with n = 1, 2, 3, . . . (193)
πme2

The radius of the innermost orbit is called the Bohr radius of the hydrogen atom and is
denoted by a0 :
h2 ²0
a0 = r1 = = 5.292 × 10−11 m. (194)
πme2
Here we have used the parameters e = 1.6×10−19 C, ²0 = 8.85×10−12 F/m, m0 = 9.1×10−31
kg, and h = 6.6 × 10−34 Js.

48
FIG. 14: Vibrations of a wire loop.

The other radii are given in terms of a0 by the formula

rn = n2 a0 , with n = 1, 2, 3, . . . (195)

Exercise 1: In the Bohr model, the electron is in constant motion. How can such an
electron have a negative energy?
Answer: The potential energy of the electron interacting with the proton is U =
−e2 /4π²0 r, a negative value. The kinetic energy of the electron is K = mv 2 /2 = e2 /8π²0 r.
Since the absolute of the potential energy is larger than the kinetic energy, the total energy
E = K + U of the electron is negative.
Exercise 2: Derive a formula for the speed of an electron in the nth orbit of a hydrogen
atom using the Bohr model.
Answer: In the Bohr model, rn = n2 a0 , where a0 = h2 ²0 /πme2 is the Bohr radius. From
the condition for orbital stability nλ = 2πr, we have λ = 2πna0 . From the de Broglie
relation λ = h/mv, we obtain v = h̄/nma0 .
Exercise 3: An electron at rest is released far away from a proton and moves toward the

proton. (a) Show that the de Broglie wavelength λ is proportional to r, where r is the
distance of the electron from the proton. (b) Find λ when r = a0 . How does this compare
with the wavelength of an electron in the ground-state Bohr orbit? (c) In order for the
electron to be captured by the proton to form a ground-state hydrogen atom, energy must
be lost by the system. How much energy?
Answer: (a) It follows from the energy conservation law E = K + U = 0 that K =

49
p p
p2 /2m = −U = e2 /4π²0 r. Hence p = e m/2π²0 r. Therefore, λ = h/p = 2π²0 h2 r/me2 =

π 2a0 r.

(b) When r = a0 , we have λ = π 2 a0 . The wavelength of an electron in the ground-state
Bohr orbit is λ0 = 2πa0 . As we can see, λ < λ0 .
(c) The energy to be released is −E1 = −U/2 = (1/8π²0 )(e2 /r1 ) = e2 /8π²0 a0 =
me4 /8²20 h2 = 2.18 × 10−18 J = 13.6 eV.
Exercise 4: Compare the uncertainty in the momentum of an electron confined to a region
of linear dimension a0 with the momentum of an electron in the ground-state Bohr orbit.
Answer: ∆p ≥ h̄/2a0 . According to exercise 2, the momentum of an electron in the
ground-state Bohr orbit is p1 = mv1 = h̄/a0 . Thus the uncertainty in the momentum of the
electron is less than the momentum of the electron in the ground-state Bohr orbit.

C. Energy levels and spectra

The total electron energy E is the sum of its kinetic energy


mv 2
(196)
2
and potential energy
e2
− . (197)
4π²0 r
This means that
mv 2 e2
E= − . (198)
2 4π²0 r
Due to Eq. (189), we find
e2
E=− . (199)
8π²0 r
Different permitted orbits have different energies: Due to Eq. (189), we find
e2
En = − . (200)
8π²0 rn
When we use Eq. (193) for rn , we see that
me4 1 E1
En = − 2 2 2
= 2, with n = 1, 2, 3, . . . (201)
8²0 h n n
Here, E1 = −2.18 × 10−18 J = −13.6 eV.
The energies specified by Eq. (201) are called the energy levels of the hydrogen atom.
These levels are all negative. This means that the electron does not have enough energy to

50
escape from the nucleus. An electron in the hydrogen atom can have only these energies
and no others.
The lowest level E1 is called the ground state, and the higher levels E2 , E3 , E4 , . . . are
called excited states.
As the quantum number n increases, the level energy En tends to 0. In the limit of
n = ∞, we have E∞ = 0, that is, the electron is no longer bound to the nucleus to form an
atom. The work needed to remove an electron from its ground state is called its ionization
energy. It is equal to −E1 . In the case of hydrogen, the ionization energy is 13.6 eV.

D. Origin of line spectra

According to the Bohr model, electrons cannot exist in an atom except in certain specific
energy levels. When an electron in an excited state drops to a lower state, the lost energy is
emitted as a single photon of light. If the energy of the initial state of the electron is Ei and
the energy of the final state of the electron is Ef , then the energy of the emitted photon is

hν = Ei − Ef . (202)

Here, ν is the frequency of the photon. According to Eq. (201), we have


à !
1 1
Ei − Ef = E1 − 2 , (203)
n2i nf

where ni is the quantum number of the initial state and nf is the quantum number of the
final state. Therefore, we have
à !
E1 1 1
ν=− 2
− 2 . (204)
h nf ni

Since λ = c/ν, we find the formula for the spectrum:


à !
1 E1 1 1
=− − 2 . (205)
λ hc n2f ni

The Lyman series corresponds to the case where nf = 1:


µ ¶
1 E1 1 1
Lyman (nf = 1): =− − , with n = 2, 3, 4, . . . (206)
λ hc 12 n2
µ ¶
1 E1 1 1
Balmer (nf = 2): =− − , with n = 3, 4, 5, . . . (207)
λ hc 22 n2

51
FIG. 15: Transitions between energy levels.

µ ¶
1 E1 1 1
Paschen (nf = 3): =− − , with n = 4, 5, 6, . . . (208)
λ hc 32 n2
µ ¶
1 E1 1 1
Brackett (nf = 4): =− − , with n = 5, 6, 7, . . . (209)
λ hc 42 n2
µ ¶
1 E1 1 1
Pfund (nf = 5): =− − , with n = 6, 7, 8, . . . (210)
λ hc 52 n2
The Rydberg constant is calculated to be
E1 me4
R=− = 2 3 = 1.097 × 107 m−1 . (211)
hc 8²0 h c
Thus, the theoretical results of the Bohr model are in agreement with the experimental
results.
Example 1
Find the longest wavelength present in the Balmer series of hydrogen.
Solution
In the Balmer series, the quantum number of the final state is nf = 2. The wavelengths
of the lines in this series is given by the formula
µ ¶
1 1 1
=R − , with n = 3, 4, 5, . . . (212)
λ 22 n2

52
The longest wavelength corresponds to n = 3:
µ ¶
1 1 1
=R − = 0.139 R. (213)
λ 22 32
Hence, we obtain
1 1
λ= = 7 −1
= 6.56 × 10−7 m = 656 nm. (214)
0.139 R 0.139 × (1.097 × 10 m )
Exercise 1: When radiation with a continuous spectrum is passed through a volume of
hydrogen gas whose atoms are all in the ground state, which spectral series will be present
in the resulting absorption spectrum?
Answer: Lyman series: 1/λ = R(1 − 1/n2 ).
Exercise 2: A proton and an electron, both at rest initially, combine to form a ground-
state hydrogen atom. A single photon is emitted in this process. What is its wavelength?
Answer: The energy of the ground state is E1 . The wavelength of the emitted photon is
1/λ = −E1 /ch = R = 1.097 × 107 m−1 . Hence λ = 0.912 × 10−7 m = 91.2 nm.
Exercise 3: How many different wavelengths would appear in the emission spectrum of
hydrogen atoms initially in the n = 5 state?
Answer: 4 lines (n = 5 → n0 = 1, 2, 3, 4) for the downward transitions with ni = 5.
However, there are 10 lines for the downward transitions with ni = 5, 4, 3, 2.
Exercise 4: An excited hydrogen atom emits a photon in returning to the ground state.
Derive a formula for the quantum number ni of the initial state in terms of λ and R. Use
this formula to find ni for a 102.55-nm photon.
p
Answer: Lyman series: 1/λ = R(1 − 1/n2 ). Hence we find n = 1/ 1 − 1/λR.
p
For a 102.55-nm photon, we find n = 1/ 1 − 1/[(102.55 × 10−9 )(1.097 × 107 )] ==
p √ √
1/ 1 − 1/1.125 = 1/ 1 − 0.889 = 1/ 0.111 = 1/0.333 = 3.
Exercise 5: Doppler effect An excited atom of mass m and initial speed v emits a photon
in its direction of motion. Assume that v is small compared to c, so that relativistic con-
siderations are not required. If h̄ω/2mc2 ¿ 1, use the requirement that linear momentum
and energy must both be conserved to show that the frequency of the photon is higher by
∆ω/ω0 ∼= v/c than it would have been if the atom had been at rest.
Answer: We have two equations

mv = h̄k + mv 0 ,
mv 2 mv 02
Ei + = Ef + h̄ω + . (215)
2 2

53
Introduce the notation Ei − Ef = h̄ω0 . Elimination of v 0 from the above two equations gives
µ ¶
v h̄ω
h̄ω0 = h̄ω 1 − + . (216)
c 2mc2

Since h̄ω/2mc2 ¿ 1, we neglect this term. Consequently, we obtain

ω0 ∼
ω= = ω0 (1 + v/c). (217)
1 − v/c

Then, we have ∆ω/ω ∼


= v/c.

54
VI. CORRESPONDENCE PRINCIPLE

Quantum physics is very different from classical physics in the microworld. However,
quantum physics must give the same results as classical physics in the macroworld where
classical physics is valid. We show that this basic requirement is true for the Bohr model of
the hydrogen atom.
We calculate the frequency of radiation from a hydrogen atom.
a) Classical picture. According to the electromagnetic theory, the radiation frequency of
a hydrogen atom is equal to the frequency of the evolution of the electron rotating around
the nucleus, or to an integer multiple of this frequency. According to Eq. (190), the speed
of the electron is
e
v=√ . (218)
4π²0 mr
Hence, the evolution frequency of the electron is
v e
f= = √ . (219)
2πr 2π 4π²0 mr3
The radius of an stable orbit is given in terms of the quantum number n by Eq. (193) as
h2 ² 0
2
r=n . (220)
πme2
Therefore, the evolution frequency of the electron is
me4 E1 2
f= 2 3 3
=− . (221)
4²0 h n h n3
b) Quantum picture. According to the Bohr model, when the electron drops from an
orbit ni to an orbit nf , a photon is emitted. The frequency of the emitted photon is given
by Eq. (204) as à !
E1 1 1
ν=− 2
− 2 . (222)
h nf ni
Under what circumstances should the Bohr atom behave classically. Assume that the
electron orbit is so large that we might be able to measure it directly. In this case, quantum
effects ought to be negligible. An orbit 10 µm across, for instance, meets this specification.
The quantum number of this orbit is n = 435, a large number.
Let’s write n = ni and n − p = nf . With this notation,
µ ¶
E1 1 1 E1 2np − p2
ν=− − = − . (223)
h (n − p)2 n2 h n2 (n − p)2

55
When n is much larger than p, we can use the approximations 2np−p2 ≈ 2np and (n−p)2 ≈
n2 . Hence, we find the photon frequency
E1 2p
ν=− . (224)
h n3
When p = 1, the photon frequency ν is exactly the same as the frequency of rotation f of
the orbital electron given by Eq. (221). Multiples of the rotation frequency f are radiated
when p = 2, 3, 4, . . . . Hence, both quantum and classical pictures of the hydrogen atom give
the same results in the limit of very large quantum numbers.
The requirement that quantum physics give the same results as classical physics in the
limit of large quantum numbers was called by Bohr the correspondence principle.
Since the de Broglie electron wavelength is λ = h/mv, we can rewrite the condition for
orbital stability nλ = 2πr, see Eq. (192), in the form
nh
mvr = . (225)

Exercise 1: Of the following quantities, which increase and which decrease in the Bohr
model as n increases: Frequency of evolution, electron speed, electron wavelength, angular
momentum, potential energy, kinetic energy, total energy.

Answer: Frequency of evolution f ∝ v/r ∝ 1/n3 decreases, electron speed v ∝ 1/ r ∝
1/n decreases, electron wavelength λ ∝ r/n ∝ n increases, kinetic energy K ∝ v 2 ∝ 1/n2
decreases, potential energy U ∝ −1/r ∝ −1/n2 increases, total energy E ∝ −1/n2 increases.
Exercise 2: Show that the frequency of the photon emitted by a hydrogen atom in going
from the level n+1 to the level n is always intermediate between the frequencies of evolution
of the electron in the respective orbits.
Answer: Photon frequency is
· ¸ · ¸
E1 1 1 |E1 | 1 1
ν= − = − . (226)
h (n + 1)2 n2 h n2 (n + 1)2
The frequencies of revolution is
E1 2 |E1 | 2
fn = − 3
= . (227)
h n h n3
Since
2 1 1 2
3
< 2− 2
< 3, (228)
(n + 1) n (n + 1) n
we have fn+1 < ν < fn .

56
FIG. 16: Transitions between two energy levels in an atom.

VII. THE LASERS

The laser is a device that produces a light beam with some remarkable properties:
1. The light is nearly monochromatic.
2. The light is coherent, with the waves all exactly in phase with each another. An
interference pattern can be obtained not only by placing two slits in a laser beam but also
by using beams from two separate lasers.
3. The beam is well collimated. It spreads out very little, even over long distances.
4. The beam is extremely intense, much more intense than any other source.
The key to the laser is the presence in many atoms of one or more excited energy levels
whose lifetimes may be 10−3 s or more instead of the usual 10−8 s. Such long-lived states
are called metastable (temporally stable).
Consider two energy levels E0 and E1 . There are three kinds of transition involving these
levels. If the atom is initially in the lower state E0 , it can be raised to E1 by absorbing a
photon of energy E1 − E0 = hν. This process is called induced absorption. If the atom is
initially in the upper state E1 , the atom can drop to E0 by emitting a photon of energy hν.
This process is called spontaneous emission.
Einstein in 1917 pointed out that there is a third possibility, called induced emission,
in which an incident photon of energy hν causes a transition from E1 to E0 . In induced
emission, the radiated light waves are exactly in phase with the incident ones, so the result is
an enhanced beam of coherent light. Induced emission has the same probability as induced
absorption. In other words, when a photon is incident to an atomic system, the probability
of emission of an initially excited atom is equal to the probability of absorption of an initially
unexcited atom.
A three-level laser

57
photons

FIG. 17: The principle of the laser.

The simplest kind of lasers is a three-level laser. A three-level laser uses an ensemble of
atoms (or molecules) that have a metastable state E1 with the energy hν above the ground
state E0 and a still higher excited state E2 that decays to the metastable state E1 , see Fig.
17.
We need to create a situation where more atoms are in the metastable state than in the
ground state. Then, we shine light of frequency ν on the ensemble. There will be more
induced emissions from atoms in the metastable state than induced absorption by atoms
in the ground state. The result will be an amplification of the original light. This is the
principle of the operation of the laser.
In an atomic system, the ground state is normally occupied to the greatest extent. The
term population inversion describes an ensemble of atoms in which the majority are in an
excited state.
There are many ways to produce a population inversion. One of them is optical pumping.
In this method, an external light source is used. Some photons of this source have the right
frequency to raise ground-state atoms to an excited state that decays spontaneously to the
desired metastable state.

58
FIG. 18: The ruby laser.

Why are three levels needed? Suppose that there are only two levels. The pumping will
produce not only upward but also downward transitions. When half of the atoms are in
each state, the rate of induced emission is equal to the rate of induced absorption, so the
atomic ensemble cannot ever have more than half its atoms in the metastable state. Due to
the lack of population inversion, laser amplification cannot occur.
A practical three-level laser
The first successful laser is the ruby laser. It is based on the three energy levels in the
chromium ion Cr3+ . A ruby is a crystal of aluminum oxide, Al2 O3 , in which some of the Al3+
ions are replaced by Cr3+ ions. A Cr3+ ion has a metastable level whose lifetime is about
0.003 s. In the ruby laser, a xenon flash lamp excites the Cr3+ ions to a higher level from
which they fall to the metastable level by losing energy to other ions in the crystal. Photons
from the spontaneous emission of some Cr3+ ions are reflected back and forth between two
mirrors at the ends of the ruby rod, stimulating other excited Cr3+ ions to radiate. After a
few microseconds, the result is a large pulse of monochromatic, coherent red light from the
partly transparent end of the rod.
The length of the rod is made to be an integer multiple of half-wavelength, so the radiation
trapped in its forms an optical standing wave. Since the induced emissions are stimulated
by the standing wave, the generated waves are all in phase with it.
Exercise 1: For laser action to occur, the medium used must have at least three energy
levels. What must be the nature of each of these levels? Why is three the minimum number?
Answer: One level is the ground state, another one is an excited state, the last one is an
metastable (long-lived) state.
If there is only two levels, one cannot achieve positive population inversion. Optical
pumping produces not only upward but also downward transitions. Therefore, in a two-

59
level system, the population of the upper level cannot be larger than the population of the
lower level.
Exercise 2: A ruby laser emits 1-J pulses of light whose wavelength is 694 nm. What is
the minimum number of Cr3+ ions in the ruby?
Answer: The energy of a photon is h̄ω = hc/λ = (6.63 × 10−34 )(3 × 108 )/(694 × 10−9 ) =
2.87 × 10−19 J. The number of photons in each pulse is N = 1/(2.87 × 10−19 ) = 3.5 × 1018 .
This is the minimum number of Cr3+ ions required for such laser pulses.

60
VIII. QUANTUM MECHANICS

The fundamental difference between classical (or Newtonian) and quantum mechanics
lies in what they describe.
In classical mechanics, the future history of a particle is completely determined by its
initial position and momentum together with the forces that act on it. In the everyday
world, i.e. the world of macroscopic bodies, all these quantities can be determined well
enough for the predictions of classical mechanics to agree with observations.
The uncertainty principle suggests that the nature of an observable quantity is different
in the world of microscopic particles. It is impossible to determine exactly the future history
of a particle because its initial state cannot be established with sufficient accuracy. This
world requires quantum mechanics.
The quantities whose relationships quantum mechanics explores are probabilities. For
example, instead of asserting that the radius of the electron’s orbit in the ground-state
hydrogen atom is 5.3 × 10−11 m, as the Bohr theory does, quantum mechanics states that
this is the most probable radius. In a suitable experiment most trials will yield different
values, either larger or smaller, the value most likely to be found will be 5.3 × 10−11 m.
Classical mechanics is an approximate version of quantum mechanics. The certainties
in classical mechanics are illusory, and their apparent agreement with experiment occurs
because ordinary objects consist of so many individual atoms that departures from average
behavior are unnoticeable.
Wave function
The quantity with which quantum mechanics is concerned is the wave function Ψ of a
body. The wave function Ψ itself has no physical interpretation. However, the square of its
absolute value, |Ψ(x, t)|2 , evaluated at a particular place x at a given time t, is proportional
to the probability of finding the body there at that time. The probabilities of all other
characteristics of the body, such as linear momentum, angular momentum, and energy, can
also be established from Ψ.
The problem of quantum mechanics is to determine Ψ for a body when its freedom of
motion is limited by the action of external forces.
Wave functions are usually complex, with both real and imaginary parts. Every complex

61
wave function can be written as
Ψ = A + iB, (229)

where A and B are real functions.


The complex conjugate is
Ψ∗ = A − iB. (230)

The probability density is


|Ψ|2 = ΨΨ∗ = A2 + B 2 . (231)

Thus the probability density P ∝ |Ψ|2 is always a positive real number, as required.
We now establish certain requirements for Ψ.
Since |Ψ|2 is proportional to the probability of finding the body at a point, the integral
of |Ψ|2 over all space, Z ∞
|Ψ|2 dV (232)
−∞

is proportional to the probability of finding the body somewhere in space. Therefore, this
integral must be finite. It cannot be 0 or ∞. If this integral is zero, the particle does not
exist. If this integral is ∞, it cannot not have the meaning of a probability.
It is convenient to have |Ψ|2 be equal to the probability density P of finding the particle.
If so, then it must be true that Z ∞
|Ψ|2 dV = 1 (233)
−∞

because Z ∞
P dV = 1 (234)
−∞

A wave function that satisfies Eq. (234) is said to be normalized. Every acceptable wave
function can be normalized by multiplying it by an appropriate constant.
Besides being normalizable, Ψ must be single-valued since P can have only one value at
a particular place and time. Ψ must also be continuous. On the other hand, the momentum
components are proportional to the partial derivatives ∂Ψ/∂x, ∂Ψ/∂y, ∂Ψ/∂z. Therefore,
these derivatives must be finite, continuous, and single-valued. Only wave functions with
all these properties can yield physically meaningful results when used in calculations, so
only such well-behaved wave functions are admissible as mathematical representations of
real bodies.
To summarize:

62
1. Ψ must be single-valued and continuous everywhere.
2. ∂Ψ/∂x, ∂Ψ/∂y, ∂Ψ/∂z must be single-valued and continuous everywhere.
3. Ψ must be normalizable, which means that Ψ must go to 0 as x, y, or z tends to ±∞
R
in order that |Ψ|2 dV over all space be a finite constant.
The above rules are not always obeyed by the wave functions of particles in model situ-
ations that only approximate actual ones.
Exercise: Which of the wave functions in Fig. 5.14 of the textbook cannot have physical
significance in the interval shown? Why not?
Answer: (b), not single-valued; (c), derivative not continuous; (d), derivative not finite
and not continuous; (f), not continuous.

63
IX. SCHRÖDINGER EQUATION

Wave equation
Consider a wave whose variable quantity is y. Assume that this wave propagates in the
x direction with the speed v. Then, the propagation of the wave is described by the wave
equation
∂ 2y 1 ∂ 2y
= . (235)
∂x2 v 2 ∂t2
All the solutions of the above equation must be of the form
³x´
y =F t± , (236)
v

where F is any function that can be differentiated. The solutions F (t − x/v) represent waves
traveling in the +x direction, and the solutions F (t + x/v) represent waves traveling in the
−x direction.
Consider the wave equivalent of a free particle moving in the +x direction. This wave is
described as an undamped, monochromatic harmonic wave

y = Ae−iω(t−x/v) . (237)

Time-dependent Schrödinger equation


In quantum mechanics, a particle is described by a wave function Ψ. We assume that
the wave function of a free particle moving along the +x direction is specified by

Ψ = Ae−iω(t−x/v) . (238)

When we use the relations ω = 2πν and v = λν, we have

Ψ = Ae−2πi(νt−x/λ) . (239)

In terms of the total energy E = hν = 2πh̄ν and the momentum p = h/λ = 2πh̄/λ, the
wave function of the free particle is

Ψ = Ae−(i/h̄)(Et−px) . (240)

We differentiate Eq. (240) twice with respect to x. Then we obtain

∂ 2Ψ p2
= − Ψ. (241)
∂x2 h̄2

64
Differentiating Eq. (240) once with respect to t gives

∂Ψ iE
= − Ψ. (242)
∂t h̄

Assume that Eqs. (241) and (242) are valid for an arbitrary particle. Assume that the
velocity v of the particle is small compared to the light velocity c. Then, the total energy
E of the particle is the sum of its kinetic energy p2 /2m and its potential energy U (x, t):

p2
E= + U (x, t). (243)
2m

Substituting Eq. (243) into Eq. (242) and using Eq. (241), we obtain the equation

∂Ψ h̄2 ∂ 2 Ψ
ih̄ =− + U Ψ. (244)
∂t 2m ∂x2

This is the time-dependent Schrödinger equation in one dimension.


In three dimensions, the time-dependent Schrödinger equation is
µ ¶
∂Ψ h̄2 ∂ 2 Ψ ∂ 2 Ψ ∂ 2 Ψ
ih̄ =− + + + U Ψ. (245)
∂t 2m ∂x2 ∂y 2 ∂z 2

Here, the potential energy U is a function of x, y, z, and t.


The Schrödinger equation is a basic principle. It must be postulated. It cannot be derived
from other principles of physics.
Linearity and superposition
An important property of the Schrödinger equation is that it is linear in the wave function
Ψ. This means that the equation has terms that contain Ψ and its derivatives but no terms
independent of Ψ and no terms that involve higher powers of Ψ and its derivatives.
As a result, a linear combination of solutions of the Schrödinger equation is also itself a
solution. If Ψ1 and Ψ2 are two solutions, then

Ψ = a1 Ψ1 + a2 Ψ2 (246)

is also a solution. Here, a1 and a2 are constants. Thus the wave functions Ψ1 and Ψ2 obey
the superposition principle that other waves do. Consequently, interference efects can occur
for wave functions just as they can for light, sound, water, and electromagnetic waves.
Exercise: Prove that the Schrödinger equation is linear by showing that

Ψ = a1 Ψ1 + a2 Ψ2 (247)

65
is also a solution of Eq. (244) if Ψ1 and Ψ2 are themselves solutions.
Expectation values
The wave function of a particle contains all the information about the particle. How to
extract information from the wave function? As an example, let us calculate the expectation
value hxi of the position of a particle that is described by the wave function Ψ(x, t) in one
dimension.
To make the procedure clear, we first consider a number of identical particles. Assume
that there are N1 particles at x1 , N2 particles at x2 , and so on. The average position is then
given by P
N1 x 1 + N2 x 2 + N3 x 3 + · · · Ni x i
x̄ = = P . (248)
N 1 + N2 + N3 + · · · Ni
When we are dealing with a single particle, we must replace the number Ni of particles at
xi by the probability Pi that the particle be found in an interval dx at xi . This probability
is
Pi = |Ψi |2 dx, (249)

where Ψi is the wave function evaluated at x = xi . Making the substitution Ni = Pi =


|Ψi |2 dx and changing the summations to integrals, we find that the expectation value of the
position of the single particle is
R∞
x|Ψ|2 dx
hxi = R−∞
∞ . (250)
|Ψ|2 dx
−∞
R∞
We say that Ψ is a normalized wave function if −∞ |Ψ|2 dx = 1. When Ψ is normalized, the
expectation value for position is
Z ∞
hxi = x|Ψ|2 dx. (251)
−∞

In more general, consider an arbitrary quantity G(x) that depends on the position x but
not on the momentum p. The expectation value of this quantity is calculated as
Z ∞
hG(x)i = G(x)|Ψ|2 dx. (252)
−∞

The expectation value hpi for momentum cannot be calculated this way because, accord-
ing to the uncertainty principle, no such function as p(x) can exist. Indeed, if we specify x,
so that ∆x = 0, we cannot specify a corresponding p since ∆x∆p ≥ h̄/2. The same problem
occurs for the expectation value hEi for energy.

66
Example
A particle limited to the x axis has the wave function Ψ = ax between x = 0 and x = 1;
Ψ = 0 elsewhere.
(a) Find a for which Ψ is a normalized wave function.
(b) Find the probability that the particle can be found between x = 0.45 and x = 0.55.
(c) Find the expectation value hxi of the particle’s position.
Solution
(a) The constant a satisfies the normalization condition

Z Z ¯1
1 1 3¯
x ¯ a2
1= |Ψ|2 dx = a2 x2 dx = a2 ¯ = . (253)
0 0 3¯ 3
0

Hence we find a = 3.
(b) The probability is
Z Z ¯0.55
x2 0.55 3¯
x ¯
|Ψ|2 dx = a2 x2 dx = a2 ¯ = 0.0251a2 = 0.0753. (254)
x1 0.45 3¯
0.45

(c) The expectation value for position is


Z Z ¯1
1 1 4¯
x ¯ a2 3
x|Ψ|2 dx = a2 x3 dx = a2 ¯ = = . (255)
0 0 4¯ 4 4
0

Steady-state Schrödinger equation


Note that the wave function of a free particle (U = 0) can be written as

Ψ = Ae−(i/h̄)(Et−px) = Ae−iEt/h̄ eipx/h̄ = ψe−iEt/h̄ , (256)

where ψ depends on position but not on time.


Consider the case where the particle is not free. Assume that the potential U is a function
of x but is independent of time. Consider the situation where the wave function Ψ can be
written in the form
Ψ = ψe−iEt/h̄ , (257)

where ψ is independent of time. This situation is called the steady state. Substituting Eq.
(257) into Eq. (244), we find

h̄2 −iEt/h̄ ∂ 2 ψ
Eψe−iEt/h̄ = − e + U ψe−iEt/h̄ . (258)
2m ∂x2

67
Dividing the above equation by the common exponential factor gives

∂ 2 ψ 2m
+ 2 (E − U )ψ = 0. (259)
∂x2 h̄

The above equation is called the steady-state Schrödinger equation in one dimension. In
three dimensions, the steady-state Schrödinger equation is

∂ 2 ψ ∂ 2 ψ ∂ 2 ψ 2m
+ 2 + 2 + 2 (E − U )ψ = 0. (260)
∂x2 ∂y ∂z h̄

An important property of the steady-state Schrödinger equation is that, if it has one or


more solutions for a given system, each of these solutions corresponds to a specific value of
the energy E. That is how energy quantization appears in quantum mechanics.
Eigenvalues and eigenfunctions
The values of energy En for which the steady-state Schrödinger equation can be solved
are called eigenvalues or eigenenergies. The corresponding wave functions ψn are called
eigenfunctions or eigenstates. An example of a set of eigenenergies is the discrete energy
levels of the hydrogen atom:

me4 1
En = − , with n = 1, 2, 3, . . . (261)
32π 2 ²20 h̄2 n2

68
FIG. 19: An infinite potential.

X. PARTICLE IN A BOX

The simplest quantum-mechanical problem is that of a particle trapped in a box with


infinitely hard walls. Consider a particle that is restricted to traveling along the x axis
between x = 0 and x = L by infinitely hard walls. This situation can be described by the
potential energy U that is infinite on both sides of the box and is a finite constant on the
inside. Because the particle cannot have an infinite amount of energy, it cannot exist outside
the box. Hence, the wave function ψ is zero for x ≤ 0 and x ≥ L. We will calculate the
steady-state wave function ψ for the inside of the box.
Inside the box, U is a constant. For convenience, we take U = 0 for the inside of the box.
Then, the steady-state Schrödinger equation (259) becomes
d2 ψ 2m
+ 2 Eψ = 0. (262)
dx2 h̄
The total energy E = K +U = K is positive because so is the kinetic energy K. The general
solution of the above equation is
√ √
2mE 2mE
ψ = A sin x + B cos x. (263)
h̄ h̄
Here, A and B are coefficients to be evaluated.
The boundary conditions are ψ(0) = 0 and ψ(L) = 0. From ψ(0) = 0, sin 0 = 0 and

cos 0 = 1, we find B = 0. The condition ψ(L) = 0 will be satisfied if sin 2mEL/h̄ = 0,
that is, if √
2mE
L = nπ, with n = 1, 2, 3, . . . . (264)

Hence, the energy of the particle can have only certain values
n2 π 2 h̄2
En = , with n = 1, 2, 3, . . . . (265)
2mL2

69
The above values are the eigenenergies of the particle. They constitute the energy levels of
the particle.
Wave function of a particle in a box
Since B = 0, the wave function of the particle inside a box is

2mEn
ψn = A sin x. (266)

Using Eq. (265) for En , we find
nπx
ψn = A sin . (267)
L
It is easy to verify that the above wave function satisfies the steady-state Schrödinger equa-
tion and the boundary conditions ψ(0) = ψ(L) = 0. Note that Eq. (267) is valid only for
0 ≤ x ≤ L. Outside the box, we have ψn = 0. We also note that the spatial derivative ψn0 of
the wave function (267) is not continuous at x = 0 as well as at x = L. The reason is that
the box with infinitely hard walls is just an approximate model, not a realistic situation (see
the next section).
To find the constant A, we calculate the integral of |ψn |2 from x = −∞ to x = +∞. We
find
Z +∞ Z L Z L Z L µ ¶
2 2 2 2 nπx A2 2nπx
|ψn | dx = |ψn | dx = A sin dx = 1 − cos dx
−∞ 0 0 L 2 0 L
2
AL
= . (268)
2
To satisfy the normalization condition
Z +∞
|ψn |2 dx = 1, (269)
−∞

we choose r
2
A= . (270)
L
Thus the normalized wave functions of the particle are
r
2 nπx
ψn = sin , with n = 1, 2, 3, . . . . (271)
L L
The normalized wave functions ψ1 , ψ2 , and ψ3 together with the probability densities
|ψ1 |2 , |ψ2 |2 , and |ψ3 |2 are plotted in Fig. 20.
As seen, ψn = 0 at x = 0 and x = L. At a particular place in the box, the probability
of finding the particle may be very different for different quantum numbers. For example,

70
FIG. 20: Wave functions and probability densities of a particle in a box with rigid walls.

|ψ1 |2 has its maximum value in the middle of the box, while |ψ2 |2 = 0 there. In other words,
a particle in the lowest energy level of n = 1 is most likely to be in the middle of the box,
while a particle in the next higher state of n = 2 is never there. Classical physics, of course,
suggests the same probability for the particle being anywhere in the box.
Example 1
Consider a particle trapped in a box of with L. Find the probability that the particle
can be found between 0.45 L and 0.55 L for the ground and first excited states.
Solution

71
FIG. 21: Wave functions and probability densities of a particle in a box with rigid walls.

This part is one tenth of the box width and is centered on the middle of the box. Clas-
sically we would expect the particle to be in this region 10 percent of the time.
However, quantum mechanics gives different predictions depending on the quantum num-
ber. In fact, the probability of finding the particle between x1 and x2 is
Z x2 Z µ ¶x
2 2 x2 2 nπx x 1 2nπx 2
Px1 x2 = |ψn | dx = sin dx = − sin
x1 L x1 L L 2nπ L x1
x2 − x1 1 nπ(x2 + x1 ) nπ(x2 − x1 )
= − cos sin . (272)
L nπ L L
For x1 = 0.45 L and x2 = 0.55 L, we obtain
1 1 nπ
Px1 x2 = − cos nπ sin
10 nπ 10

1 sin
= + (−1)n+1 10
.
10 nπ
For the ground state, which corresponds to n = 1, we have

Px1 x2 = 0.198 = 19.8%. (273)

This is about twice the classical probability.

72
For the first excited state, which corresponds to n = 2, we have

Px1 x2 = 0.0065 = 0.65%. (274)

This low probability is consistent with the fact that |ψ2 |2 = 0 at x = 0.5 L.
Note that, when n is very large, we have Px1 x2 → (x2 − x1 )/L. This result is in agreement
with classical physics, as required by the correspondence principle.
Example 2
Consider a particle trapped in a box of with L. Find the expectation value hxi of the
position of the particle.
Solution

Z L Z L
2 2 nπx
hxi = x|ψn | dx = x sin2 dx
0 L 0 L
· 2
¸L
2 x x sin(2nπx/L) cos(2nπx/L)
= − −
L 4 4nπ/L 8(nπ/L)2 0
L
= . (275)
2
This result means that the average position of the particle is the middle of the box. It
reflects the symmetry of the system. There is no conflict with the fact that |ψn |2 = 0 at L/2
for the states of n = 2, 4, 6, . . . because hxi is an average, not a probability.
Exercise 1: Consider a particle in a box. Show that as n → ∞, the probability of
finding the particle between x and x + ∆x is ∆x/L and so is independent of x, which is the
classical expectation.
Answer: According to Eq. (272), the probability of finding the particle between x and
x + ∆x is
Z x+∆x Z µ ¶x+∆x
2 2 x+∆x 2 nπx x 1 2nπx
Px,x+∆x = |ψn | dx = sin dx = − sin
x L x L L 2nπ L x
∆x 1 2nπ(x + ∆x) 1 2nπx
= − sin + sin . (276)
L 2nπ L 2nπ L
As n → ∞, the last two terms vanish. Hence, we obtain in this limit the classical result
Px,x+∆x = ∆x/L.
Exercise 2: Consider a particle in a 3D cubic box of length L. The wave function is
given by
µ ¶3/2
2 n1 πx n2 πy n3 πz
ψn1 n2 n3 = sin sin sin , with n1 , n2 , n3 = 1, 2, 3, . . . . (277)
L L L L

73
Find En1 n2 n3 !
Answer: Using the steady-state Schrödinger equation, we find

(n21 + n22 + n23 )π 2 h̄2


En1 n2 n3 = . (278)
2mL2

Exercise 3: What is the relation between the ground-state energy of a 1D box and that
of a 3D box with the same length L?
(3D) (1D)
Answer: E111 = 3E1 .
Exercise 4: Consider a particle in its ground state in a 3D cubic box of length L. What
is the probability of finding the particle in the volume defined by 0 ≤ x, y, z ≤ L/4?
Answer: Look at Eq. (272). In the 1D case, the probability of finding the particle in the
volume defined by 0 ≤ x ≤ L/4 is
1 1
P = − . (279)
4 2π
In the 3D case, the probability of finding the particle in the volume defined by 0 ≤ x, y, z ≤
L/4 is
µ ¶3
1 1
P = − . (280)
4 2π

74
FIG. 22: A finite potential well.

XI. FINITE POTENTIAL WELL

Potential energies are never infinite in the real world. Therefore, the box with infinitely
hard walls has no physical counterpart. However, potential wells with barriers of finite
heights certainly do exist.
Consider a potential well with square corners that is U high and L wide and contains
a particle whose energy E is less than U . According to classical mechanics, when the
particle strikes the sides of the well, it bounces off without entering regions I and III. In
quantum mechanics, the particle also bounces back and forth, but now it has a probability
of penetrating into regions I and III even though E < U .
In regions I and III, the steady-state Schrödinger equation is

d2 ψ 2m
+ 2 (E − U0 )ψ = 0. (281)
dx2 h̄

We rewrite the above equation as

d2 ψ
2
− a2 ψ = 0, (282)
dx

where p
2m(U0 − E)
a= > 0. (283)

The solutions to Eq. (282) are

ψI = Aeax + Be−ax ,

ψIII = Ceax + De−ax . (284)

Both ψI and ψIII must be finite everywhere. Since e−ax → ∞ as x → −∞ and eax → ∞ as

75
FIG. 23: Wave functions and probability densities of a particle in a finite potential well

x → ∞, the coefficients B and C must be zero. Hence we have

ψI = Aeax ,
ψIII = De−ax . (285)

These wave functions decrease exponentially inside the barrier at the sides of the well.
Inside the well, the steady-state Schrödinger equation is the same as Eq. (262), that is,
d2 ψ 2m
+ 2 Eψ = 0. (286)
dx2 h̄
The solution is again √ √
2mE 2mE
ψII = G sin x + F cos x. (287)
h̄ h̄
In the case of a well with infinitely high barriers, we found that F = 0 in order that ψ = 0
at x = 0 and x = L. Here, because ψII = A at x = 0 and ψII = DeL at x = L, both the sine
and cosine components in the solution (287) are possible.
For either solution, both ψ and dψ/dx must be continuous at x = 0 and x = L. The
wave functions inside and outside each side of the well must have the same values and the
same slopes where they join. When these boundary conditions are taken into account, we
find that exact matching occurs only for certain specific values En of the particle energy.

76
We shows this energy quantization in the below. We rewrite the solution as

ψI = Aeax ,
ψII = G sin kx + F cos kx,
ψI = De−ax , (288)

where p √
2m(U0 − E) 2mE
a= , k= . (289)
h̄ h̄
The boundary conditions ψI (0) = ψII (0) and ψI0 (0) = ψII0 (0) give

F = A,

Gk = Aa, (290)

0
The boundary conditions ψIII (L) = ψII (L) and ψIII (L) = ψII0 (L) give

G sin kL + F cos kL = De−aL ,

Gk cos kL − F k sin kL = −Dae−aL . (291)

Substituting Eqs. (290) into Eqs. (291), we obtain

A(a sin kL + k cos kL) − Dke−aL = 0,

A(a cos kL − k sin kL) + Dae−aL = 0. (292)

The above have nonzero solutions for A and D only if

ae−aL (a sin kL + k cos kL) + ke−aL (a cos kL − k sin kL) = 0, (293)

that is, if
2ak 2(a/k)
tan kL = = . (294)
k 2 − a2 1 − (a/k)2
Since tan kL = 2 tan(kL/2)/[1 − tan2 (kL/2)], two possible cases are:

kL a kL k
tan = or tan =− . (295)
2 k 2 a

We set √
√ 2mU0
k0 = a2 + k2 = . (296)

77
L L L L L

FIG. 24: Graphic solutions for the energies of the bound states of a particle in a square potential
well.

(1) The case


kL a
tan = (297)
2 k
leads to
¯ ¯
¯ kL ¯
¯cos ¯ = k,
¯ 2 ¯ k0
kL
tan > 0. (298)
2

In deriving the above equations we used the formula cos2 x = 1/(1 + tan2 x)
(2) The case
kL k
tan =− (299)
2 a
leads to
¯ ¯
¯ kL ¯
¯sin ¯ = k,
¯ 2 ¯ k0
kL
tan < 0. (300)
2

In deriving the above equations we used the formula sin2 x = tan2 x/(1 + tan2 x)
In Fig. 24, we illustrate the graphic solutions for the energies of the bound states of a
particle in a square potential well. When U0 is finite, k0 is finite. In this case, the number
of solutions is finite. The number of solutions increases with increasing potential height U0 .
Note that the wavelengths that fit into the well with a finite height are longer than those
for the well with an infinite height. Due to this fact, the corresponding momenta are lower,
and therefore the energy levels En are lower for each n than they are for a particle in an
infinite well.

78
XII. TUNNEL EFFECT

In the previous section, the walls of the potential well were of finite height but they were
assumed to be infinitely thick. As a result the particle was trapped forever though it could
penetrate the walls.
We now look at the situation where the potential barrier has a finite height as well as a
finite width. What we will find is that the particle has a certain probability—not necessarily
great, but not zero either—of passing through the barrier and emerging on the other side.
The particle lacks the energy to go over the top of the barrier, but it can nevertheless tunnel
through it. The higher the barrier and the wider it is, the less the chance that the particle
can get through.
Consider a beam of identical particles all of which have the kinetic energy E. The beam
is incident from the left on a potential barrier of height U0 > E and width L. On both sides
of the barrier the potential is zero. In these regions, the steady-state Schrödinger equation
is
d2 ψI 2m
+ 2 EψI = 0,
dx2 h̄
d2 ψIII 2m
+ 2 EψIII = 0. (301)
dx2 h̄
The solutions to these equations are

ψI = Aeik1 x + Be−ik1 x , (302)

ψIII = F eik1 x + Ge−ik1 x . (303)

Here √
2mE p 2π
k1 = = = (304)
h̄ h̄ λ
is the wave number of the de Broglie waves that represent the particles outside the barrier.
The incoming wave is described by the term

ψI+ = Aeik1 x . (305)

This wave corresponds to the incident beam of particles in the sense that |ψI+ |2 is their
probability density. The flux of the particles that arrive at the barrier is proportional to

|ψI+ |2 . (306)

79
The reflected wave is described by the term

ψI− = Be−ik1 x . (307)

The total wave on the left side of the barrier is

ψI = ψI+ + ψI− . (308)

On the right side of the barrier there can only be a wave

ψIII+ = F eik1 x (309)

that propagates in the +x direction at the group velocity vIII+ since region III contains
nothing that could reflect the wave. Thus, we have G = 0 and

ψIII = ψIII+ = F eik1 x . (310)

The flux of the particles that transmit through the barrier is proportional to

|ψIII |2 . (311)

Hence, the transmission probability T for a particle to pass through the barrier is the ratio

|ψIII |2 |F |2
T = = . (312)
|ψI+ |2 |A|2

Classically T = 0 because a particle with E < U0 cannot exist inside the barrier. Let’s
consider the quantum mechanical case.
In region II, the steady-state Schrödinger equation is

d2 ψII 2m d2 ψII 2m
+ (E − U 0 )ψ II = − 2 (U0 − E)ψII = 0. (313)
dx2 h̄2 dx2 h̄
Since U0 > E, the solution is
ψII = Ce−k2 x + Dek2 x . (314)

Here, p
2m(U0 − E)
k2 = . (315)

Note that ψII does not oscillate and therefore does not represent a moving particle. However,
the probability |ψII |2 is nor zero, so there is a finite probability of finding a particle within
the barrier. Such a particle may emerge into region III or may return to region I.

80
FIG. 25: Upper part: Reflection of a particle from a potential barrier in classical mechanics. Lower
part: Reflection and transmission of a particle from a potential barrier in quantum mechanics.

Applying the boundary conditions


In order to calculate the transmission probability T we have to apply the appropriate
conditions to ψI , ψII , and ψIII . As discussed earlier, both ψ and its derivative ∂ψ/∂x must
be continuous everywhere.
The boundary conditions at x = 0 are

ψI (0) = ψII (0),


dψI (0) dψII (0)
= . (316)
dx dx
The boundary conditions at x = L are

ψII (L) = ψIII (L),


dψII (L) dψIII (L)
= . (317)
dx dx
We substitute ψI , ψII , and ψIII from Eqs. (302), (303), and (314) into the above equations.
This yields

A + B = C + D,

ik1 (A − B) = −k2 (C − D), (318)

and

Ce−k2 L + Dek2 L = F eik1 L ,


−k2 (Ce−k2 L − Dek2 L ) = ik1 F eik1 L . (319)

81
FIG. 26: At each wall of the barrier, the wave functions inside and outside it must match up
perfectly, which means that they must have the same values and slopes there.

Solving Eqs. (318) for A and B, we find


µ ¶ µ ¶
1 ik2 1 ik2
A = 1+ C+ 1− D, (320)
2 k1 2 k1
µ ¶ µ ¶
1 ik2 1 ik2
B = 1− C+ 1+ D. (321)
2 k1 2 k1
Solving Eqs. (319) for C and D, we find
µ ¶
1 ik1
C = 1− F e(ik1 +k2 )L ,
2 k2
µ ¶
1 ik1
D = 1+ F e(ik1 −k2 )L . (322)
2 k2
Substituting Eqs. (322) into Eq. (320) yields
· µ ¶¸ · µ ¶¸
A 1 i k2 k1 (ik1 +k2 )L 1 i k2 k1
= + − e + − − e(ik1 −k2 )L . (323)
F 2 4 k1 k2 2 4 k1 k2
We assume that U0 À E. Then, we have k2 /k1 À k1 /k2 and consequently
k2 k1 ∼ k2
− = . (324)
k1 k2 k1
We also assume that the barrier is wide enough for ψII to be severely weakened between
x = 0 and x = L. This means that k2 L À 1 and therefore

ek2 L À e−k2 L . (325)

Hence Eq. (323) can be approximated by


µ ¶
A 1 ik2
= + e(ik1 +k2 )L . (326)
F 2 4k1

82
It follows from the above equation that
µ ¶
|A|2 1 k2
= + 22 e2k2 L . (327)
|F |2 4 16k1
We recall that the transmission probability is
|F |2
T = . (328)
|A|2
Therefore, we have
|F |2 16
T = 2
= 2
e−2k2 L . (329)
|A| 4 + (k2 /k1 )
From the definition of k1 , Eq. (304), and k2 , Eq. (315) we see that
µ ¶2
k2 U0 − E U0
= = − 1. (330)
k1 E E
This quantity varies much less with E and U0 than does the exponential. Therefore, a
reasonable approximation of the transmission probability is

T = e−2k2 L . (331)

Exercise 1: Electrons with energies of 1 eV and 2 eV are incident on a barrier 10 eV


high and 0.5 nm wide. (a) Find their respective transmission probabilities. (b) How are
these affected if the barrier is doubled in width?
Answer:
(a) For the 1-eV electrons
p
2m(U0 − E)
k2 =
p h̄
(2)(9.1 × 10−31 kg) × [(10 − 1) eV] × (1.6 × 10−19 J/eV)
=
1.054 × 10−34 J s
= 1.54 × 1010 m−1 . (332)

Since L = 0.5 nm = 5 × 10−10 m, we have 2k2 L = (2)(1.54 × 1010 )(5 × 10−10 ) = 15.4. Hence,
the approximate transmission probability is

T1 = e−2k2 L = e−15.4 = 2.05 × 10−7 . (333)

Thus one 1-eV electron out of 4.9 million can tunnel through the 10-eV barrier on average.
For the 2-eV electrons a similar calculation gives T2 = 5.15 × 10−7 . The probability T2 is
over twice the probability T1 .

83
(b) If the barrier is doubled in width to 1 nm, the transmission probabilities become

T10 = 4.2 × 10−14 and T20 = 26.5 × 10−14 . (334)

Clearly, T is more sensitive to the width of the barrier than to the particle energy here.
Exercise 2: An electron and a proton with the same energy E approach a potential
barrier whose height U0 is greater than E. Do they have the same probability of getting
through? If not, which has the greater probability?
Answer: the electron has the smaller mass, the smaller wave number k2 =
p
2m(U0 − E)/h̄ inside the barrier, and consequently the greater transmission probability
T = e−2k2 L .

84

You might also like