Anda di halaman 1dari 22

1. THE ATOMIC MODEL.

PARTICLES AND WAVES


1. The Rutherford Nuclear Atom
2. The Bohr Hydrogen Atom
3. Particles and Waves

Problems

It will be assumed that the reader has some background in “modern physics.”
The topics of this chapter are somewhat arbitrary and are not meant to be a review
of that subject. By all means skip the first two sections if it suits you, but on no
account skip the third section. The first section, if passed over, should be returned
to in advance of the chapter on scattering in three dimensions. An understanding
of classical scattering greatly aids an understanding of quantum scattering.
From Rutherford came the picture of the atom as roughly like a little solar
system, with electrons orbiting a tiny nuclear sun that contains all the positive
charge and nearly all the mass. However, on classical principles Rutherford’s atom
seemed to be completely unstable; yet the world is made of extremely stable atoms.
Said Bohr, so much the worse for classical principles. He postulated that the single
electron bound in the simplest atom, hydrogen, could only exist in certain discrete
stable orbits; and he derived the properties of these orbits using the few hydrogen
spectral lines then known and assuming that in the limit of large orbits the electron
would radiate like a classical antenna. Although Bohr’s planetary orbits have been
replaced by the nebulous states of quantum mechanics, his model was a giant step
forward.
At the heart of quantum mechanics is the bizarre matter of wave-particle du-
ality. (Webster’s: “bizarre . . . involving sensational contrasts or incongruities.”)
Two modern examples are given here: The interference and diffraction of neutrons
by a double slit and by a sharp edge; and the interference of photons, one at a time,
in an interferometer. The discussion includes “interaction-free measurements,” a
quantum-mechanical way to see in the dark.
Along the way, there are occasional historical notes and quotes. In any physics
text, such notes give an extremely superficial idea of the messiness of what actually
happened. For a proper history, see M. Jammer, The Conceptual Development of
Quantum Mechanics (McGraw-Hill, New York, 1966) or A. Pais, Inward Bound
(Clarendon Press, Oxford, 1986). Much of these are difficult reading because they
require an understanding of the physics of the times, but they are full of interesting
and accessible passages.

c 2009, Charles G. Wohl, work in progress.


1
1·1. THE RUTHERFORD NUCLEAR ATOM
Solid angle—Figure 1(a) shows an arc of a circle with its center at O; the length of
the arc is s and the radius of the circle is r. The angle subtended by the arc from O—the
angle between the radial lines—is by definition θ = s/r radians. A complete circle about O
subtends 2πr/r = 2π radians. The angle subtended by any object in a plane from a point
O in the plane is the angle between the radial lines from O to the far sides of the object,
as seen from O; see Fig. 1(b). Symmetry is often useful: A side of a square subtends 2π/4
radians from the center of the square, and 2π/8 radians from an opposite corner.

(a) (b)

θ θ
r
O O
Figure 1. Defining angles in radians.

Solid angle is defined in an analogous way. Figure 2 shows an area A on the surface
of a sphere; the radius of the sphere is r and its center is at O. The solid angle subtended
by A from O is by definition Ω = A/r 2 ster adians (stereo = solid, in space; as in solid
geometry). The whole surface of the sphere subtends 4πr 2 /r 2 = 4π steradians from O.
An infinitesimal area dA on the surface of a sphere subtends an infinitesimal solid angle
dΩ = dA/r 2 . The solid angle subtended by an object in space from a point O is the fraction
of the 4π steradians surrounding O that is blocked off by the object. Symmetry is again
useful: A side of a cube subtends 4π/6 steradians from the center of the cube. What solid
angle does the side subtend from an opposite corner of the cube?

r
O

Figure 2. Defining solid angles in steradians.

Spherical coordinates—We shall often work with spherical coordinates, the variables
r, θ, and φ defined in Fig. 3. Rectangular and spherical coordinates are related by

x = r sin θ cos φ , y = r sin θ sin φ , z = r cos θ .

2
The ranges of the spherical variables are 0 ≤ r ≤ ∞, 0 ≤ θ ≤ π, and 0 ≤ φ ≤ 2π. The polar
angle θ is like latitude, except it is measured from the north pole (the positive z axis) to
the south instead of up and down from the equator; the azimuthal angle φ is like longitude,
except it is measured all the way around to the “east” instead of half way around east
and west from the Greenwich meridian. Note that r is always nonnegative. What are the
coordinates of the point antipodal to the point (r, θ, φ)?
z
P

r
θ

y
φ

x
Figure 3. Spherical coordinates.

In spherical coordinates, the infinitesimal area dA on the surface of a sphere of radius


r between θ and θ + dθ and between φ and φ + dφ is

dA = r 2 sin θ dθ dφ .

(There are also area elements that involve dr dθ and dr dφ. What are they?) The volume
element dV in these infinitesimal angular intervals and between r and r + dr is

dV = r 2 sin θ dr dθ dφ .

The surface area of a sphere of radius R and the volume of a sphere of this radius are
π 2π R π 2π
4
Z Z Z Z Z
A= R2 sin θ dθ dφ = 4πR2 , V = r 2 sin θ dr dθ dφ = πR3 .
θ=0 φ=0 r=0 θ=0 φ=0 3

Check these integrals.


The solid angle subtended by dA = r 2 sin θ dθ dφ from the origin is

dΩ = dA/r 2 = sin θ dθ dφ .

The scattering problems considered below will be symmetric about the z axis; that is,
the scattering angular distribution will not depend on φ. Integrating dΩ over φ gives
dΩ = 2π sin θ dθ; see Fig. 4(a). Equal intervals of dθ subtend solid angles differing by the
factor sin θ, which is largest at the equator. Another way to write dΩ = 2π sin θ dθ is

dΩ = 2π d(cos θ) .

3
(A minus sign is dispensed with by integrating in the direction cos θ increases—from south
to north.) Equal intervals, large or small, of cos θ subtend equal solid angles; see Fig. 4(b).
Suppose you are measuring something in which all directions from an origin or all points on
the surface of a sphere are equally likely. Binning your data in equal intervals of θ will give
a sine distribution; binning your data in equal intervals of cos θ will give a flat distribution.
For this reason, cos θ is often used instead of θ in plotting data. Uniformity over the sphere
deserves a flat distribution.

(a) z (b) z

dθ θ = 60°
cos θ = 0.5
θ θ = 90°
cos θ = 0.0

θ = 120°
cos θ = −0.5

Figure 4. (a) The solid angle between θ and θ + dθ is dΩ = 2π sin θ dθ. (b) Cutting
a sphere at equal intervals of cos θ divides the pole-to-pole diameter into equal
lengths, the surface of the sphere into equal areas, and the 4π steradians about
the center into equal solid angles.

BB’s scattering off a bowling ball—Figure 5 shows a bowling ball, fixed at the
origin, of radius R. The z axis through the center of the bowling ball is the scattering axis.
A hail of BB’s (small round pellets), with an uniform flux of I BB’s per second per unit area
perpendicular to the scattering axis, is incident from the left, moving in the +z direction.
The perpendicular distance between the initial path of a particular BB and the scattering
axis is its impact parameter b. If b > R, the BB passes undeflected (the radius of a BB
is negligible compared to R). The total scattering cross section, σ(total), is the area on a
plane perpendicular to the scattering axis over which incident BB’s are scattered through
some angle. Clearly, σ(total) is πR 2 ; this is the cross-sectional area of the bowling ball.
The number of BB’s scattered out of the beam per second is the product of the incident
flux I and the total cross section:
number of BB0 s scattered per second = Iσ(total) = IπR 2 .
What is σ(total) if we take into account a small radius a of the BB’s?
If b < R, a BB bounces off the bowling ball. We shall assume elastic scattering, with
the angles of incidence and reflection, α, with respect to the normal at the surface of the
ball being equal. The change of direction of the BB—it was going this way, now it is going
that way—is the scattering angle θ. This is the spherical-coordinate polar angle (we have
tilted the cooordinate system of Fig. 3 on its side). From Fig. 5,
2α + θ = π and b = R sin α .

4
Thus the relation between the impact parameter b and the scattering angle θ is

b = R sin 21 (π − θ) = R cos(θ/2) ,

for b ≤ R. This is independent of the energy of the BB’s and of the azimuthal angle φ.

θ = scattering angle
α
BB α

b = impact parameter b R
α z
scattering axis
bowling ball

Figure 5. Scattering of a BB by a bowling ball.

As the impact parameter b = b(θ) decreases, the scattering angle θ increases. It follows
that the cross section for scattering through angles greater than some specified angle θ 0 is
simply the area of the circle of radius b(θ 0 ):

σ(θ > θ0 ) = π b2 (θ0 ) = πR2 cos2 (θ0 /2) ,

where b ≤ R. For example,

σ(θ > 0◦ ) = πR2 cos2 (0◦ /2) = πR2 = σ(total) ,

and
1
σ(θ > 90◦ ) = πR2 cos2 (90◦ /2) = πR2 .
2
Thus one-half of all the BB’s that are scattered are scattered through more than 90 ◦ .
The differential cross section dσ for impact parameters between b and b + db, where
b < R, is just the area of the ring of radius b and width db: dσ = 2π b db. BB’s incident
over this area will be scattered into angles between θ and θ + dθ, with θ related to b by
b = R cos(θ/2). Thus

dσ = 2π b db = 2πR cos(θ/2)(−R/2) sin(θ/2) dθ


R2 R2
=− 2π sin θ dθ = 2π d(cos θ) .
4 4
We have used sin θ = 2 sin 12 θ cos 12 θ. As shown earlier, the factor 2π d(cos θ) is the solid
angle dΩ between θ and θ + dθ. Thus

R2 R2
dσ = 2π d(cos θ) = dΩ .
4 4

5
Integrating over the 4π steradians of solid angle gives σ(total) = πR 2 again.
Since there is here no θ dependence in the coefficient of dΩ, the scattered BB’s scatter
equally in every direction. Suppose a small counter facing the bowling ball from a distance
large compared to R subtends dΩ, as seen from the origin. The rate at which it counts
scattered BB’s is equal to the rate at which BB’s cross the infinitesimal area dσ that maps
into dΩ:

R2
number of BB0 s scattered into the counter per second = Idσ = I dΩ .
4
This is proportional to dΩ and is independent of the direction to the counter: counters
subtending equal solid angles will count BB’s at (statistically) the same rate. (Rutherford
scattering will be very different.) Figure 6 shows dσ/d(cos θ) versus cos θ. Equal intervals
of cos θ subtend equal solid angles (see Fig. 4), and the distribution is flat.

2
dσ/d cos θ

2π R
4

0
−1 0 +1
cos θ
Figure 6. The angular distribution for BB’s scattering from a bowling ball.

Imagine a parallel beam of light incident upon a silvered spherical ball from which
light is totally reflected geometrically (angle of incidence = angle of reflection). Now walk
in a circular path centered on the ball and in the direction of increasing θ. Your eye (the
counter) subtends a constant infinitesimal solid angle from the ball; the spot on the ball
you see will change as you move, but the level of brightness you see will not.
A note: Nearly always in the literature, “differential cross section” means dσ/dΩ or
dσ/d(cos θ), not dσ. The units are still those of area, because Ω and cos θ are dimensionless.
Rutherford scattering—The most important scattering problem is that of a beam
of particles, each having charge q, mass m, and energy E, incident upon a fixed charge Q.
The force on q is the Coulomb force,

kqQ
F= r̂ .
r2

Here k ≡ 1/4π0 (we shall often use this shorthand), and r̂ is the unit vector from Q
toward q. The inverse-square law leads to hyperbolic orbits, bent away from Q for repulsive
scattering and toward Q for attractive scattering. Our results for the latter case would,
with the change of a few symbols, apply to the scattering of a comet swarm by a star.

6
Since the Coulomb force is of infinite range, any particle, no matter how large its impact
parameter, will be scattered through some angle. Thus the total cross section for scattering
from a “bare” charge Q—the area perpendicular to the scattering axis over which incident
particles are scattered through some angle—is infinite. This is not an infinity that should
cause much anxiety.
Figure 7 shows a trajectory. The total energy of an incident particle is the sum of its
kinetic and potential energies:

1 1 kqQ 1
E= mv02 = mv 2 + = mv00 2 .
2 2 r 2
Here v0 and v00 are the speeds long before and long after the particle passes Q, and v is the
speed at some intermediate time when the particle is at a distance r from Q. Since we are
taking Q to be fixed, v00 = v0 . If q and Q are of like sign, then a particle that is heading
directly at Q (b = 0) is running straight up a potential-energy hill and will stop and reverse
direction when
1 kqQ
E = mv02 = ,
2 D
where
kqQ kqQ
D≡ = 1 .
E 2
mv02
D is the distance of closest approach for head-on repulsive scattering, and D ≡ |kqQ|/E
will be a useful parameter even for the attractive case.

β θ
z
Q
Figure 7. A hyperbolic trajectory for Rutherford scattering of like charges.

In Fig. 7, the component of velocity of q perpendicular to the radius vector r from Q


to q is r dβ/dt, where β is the angle between r and the negative z axis. The magnitude
of the angular momentum L about Q is |r × p| = mr 2 dβ/dt. The angular momentum is
initially mv0 b, and it is conserved. Thus L = mv0 b = mr 2 dβ/dt, a relation we shall need
in a moment.
An incident particle scattered through an angle θ acquires a component of momentum
transverse to the z direction given by

∆px = ±mv0 sin θ .

7
(The polar angle θ is nonnegative, no matter which way the incident particle is bent from
the z direction; in Fig. 7, ∆px will have the sign of qQ.) We can calculate ∆p x as a function
of b. The component of F in the transverse direction at an arbitrary point along the path
of q is
kqQ
Fx = 2 sin β .
r
Since Fx = dpx /dt, the change of px during the scattering is
Z +∞ Z +∞
kqQ
∆px = Fx dt = sin β dt .
−∞ −∞ r2
There are three variables here, r, β, and t. However, we found above that L = mv 0 b =
mr 2 dβ/dt, from which we can write dt = r 2 dβ/v0 b. As t ranges from −∞ to +∞, β sweeps
from 0 to π − θ. Changing the variable of integration from t to β, we get
Z π−θ
kqQ r2
∆px = sin β dβ
0 r2 v0 b
kqQ   kqQ
= cos 0 − cos(π − θ) = (1 + cos θ) .
v0 b v0 b
The convenient cancellation of factors of r 2 only occurs because F is here proportional to
r −2 .
Equating the expressions for ∆px at the start and the end of the preceeding paragraph
and solving for b, we get
|kqQ| (1 + cos θ) D
b= 2 = cot(θ/2) .
mv0 sin θ 2

In the last step, we used 1 + cos θ = 2 cos 2 12 θ and sin θ = 2 sin 21 θ cos 12 θ. And, again,
D ≡ |kqQ|/E. The relation b = 21 D cot 21 θ may look much like the BB-bowling ball relation
b = R cos 12 θ, but cot 12 θ is a very different function from cos 12 θ. Figure 8 is a construction
that shows the relation between b and θ (see Prob. 3).
As the impact parameter b decreases, the scattering angle θ increases. The cross section
for scattering through angles greater than θ 0 is
πD 2
σ(θ > θ0 ) = π b2 (θ0 ) = cot2 (θ0 /2) .
4
For example, σ(θ > 90◦ ) = πD 2 /4. The differential cross section dσ is obtained in exactly
the same way as for the BB-bowling ball example: we evaluate dσ = 2π b db using our
relation between b and θ. From d(u/v) = (v du − u dv)/v 2 , we get d cot 21 θ = − 12 dθ/ sin2 12 θ.
Then
D2 dθ
dσ = 2π b db = −2π cot(θ/2) .
8 sin2 (θ/2)
Using sin θ = 2 sin 21 θ cos 21 θ and dΩ = 2π sin θ dθ = −2π d(cos θ), we get
2
D2

dΩ kqQ dΩ
dσ = 4 = 4 .
16 sin (θ/2) 4E sin (θ/2)

8
ote
ympt
s
ing a
outgo
incoming asymptote

out
goi
ng
b

asy
like charges

m pto
te
D/2 D/2 z
Q
unlike charges b

incoming asymptote

Figure 8. A construction that illustrates b = 12 D cot(θ/2). The vertical lines are a


distance D/2 from the scattering center Q—in front or in back of Q as the charges
are of like or unlike sign. The incoming and outgoing asymptotes of the hyperbolic
orbits make equal angles with the lines from Q.

Figure 9 is a plot of 1/ sin 4 (θ/2) versus cos θ. Multiplied by 2πD 2 /16, the curve is
dσ/dΩ. Unlike the BB-bowling ball example, Rutherford scattering is very sharply peaked
in the forward direction. It is also energy dependent, since D 2 is proportional to 1/E 2 .
The nuclear atom—In 1909, Hans Geiger (who later invented the Geiger counter)
and Ernest Marsden (an undergraduate), working in Ernest Rutherford’s laboratory at
Cambridge University, were studying the scattering of a collimated beam of α particles by a
thin gold foil. The α particles came from the decay of radon. They were surprised—in fact,
astonished—that a small fraction of the α’s were scattered through large angles. Rutherford
later said, “It was almost as incredible as if you fired a 15-inch shell at a piece of tissue
paper and it came back and hit you.”
Here is why they were astonished. A few years earlier, J.J. Thompson, the discoverer
of the electron, had proposed a provisional model of the atom in which the positive charge
and nearly all the mass of the atom fill its volume, and the electrons are somehow embedded
in this “like plums in a pudding.” Suppose the positive charge Q is uniformly distributed
throughout a sphere of atomic dimensions, and for the moment ignore the electrons. Outside
the sphere, the electric field E of Q is the same as that of a point charge, but in going inward
from the surface the field decreases linearly to zero at the center. It is then easy to show
that, given the energy of the incident α’s, the electric field E is nowhere remotely large
enough to scatter the α’s through large angles—but scatter through large angles some few
of them do. Now suppose Q is in a sphere with a radius ten times smaller than that of
the atom; then the maximum E is 100 times larger than before. In fact the radius of a
nucleus is more than 104 times smaller than its atom, so that the maximum E is more than
108 times larger than in the Thompson model. Rutherford derived the angular distribution
for a point Q (see Fig. 9), and this agreed with the results of the experiment. Thus was
born the nuclear atom—the word “nucleus” being borrowed from cell biology. For how the

9
nucleus was shown to be not vanishingly small, see Prob. 4.

100

80

60
1/sin 4 (θ/2)

40

20

0
−1.0 −0.5 0.0 0.5 1.0

cos θ
4
Figure 9. A plot of 1/ sin (θ/2) versus cos θ. The curve reaches 400 at cos θ = 0.9
and 4 × 104 at cos θ = 0.99. Multiplied by 2πD 2 /16, the curve is dσ/d(cos θ).

1·2. THE BOHR HYDROGEN ATOM


Instabilities—The world around us is made of extremely stable atoms, but it seemed
to be impossible to construct a model of an atom made of a tiny, massive positive nucleus
surrounded by electrons held in orbit by the Coulomb force. Either other, unknown forces
were in play, or classical physics was inadequate for the task.
Consider a static model. Obviously, a single electron set down at rest near a nucleus
will fall straight into it. And two electrons set down on opposite sides of a nucleus with
charge 2e will suffer the same fate. In fact, if there are only the mutual electrostatic forces,
the only stable, static arrangement of any number of point charges is one in which they are
sitting on top of one another or have flown off to infinity. To see this, suppose there exists
a stable, static arrangement of Z electrons occupying distinct positions in space around a
nucleus of charge Ze. If an electron at position P is nudged in any direction, it must, for
stability, be pushed back toward P by the electrostatic forces due to all the other charges.
This requires that the electrostatic field due to the other charges be radially outward in
every direction from P . Now draw a little Gaussian surface around P . There is a net flux
of field out of the surface, and therefore, by Gauss’s law, there is a positive charge at P .
This is contrary to the assumption that the charges occupy distinct positions in space.
What about a dynamic model? Systems of planets held in orbit about stars by that
other inverse-square-law force, gravity, are certainly stable. But an electron circling a
nucleus acts as a little antenna and radiates its energy away and spirals into the nucleus

10
in a tiny fraction of a second. To show this, we need a result from electromagnetism: The
rate at which an accelerating charge e radiates away energy E is

dE 2 ke2 2
=− a ,
dt 3 c3

where k ≡ 1/4π0 , c is the speed of light, and a is the acceleration of the charge; the minus
sign is because E decreases with time. See, for example, E.M. Purcell, Electricity and
Magnetism, 2nd ed. (McGraw-Hill, New York, 1985), App. B.
We begin with an electron in a circular orbit around a nucleus of charge Ze, and
assume (as may be shown) that as the electron radiates it follows a nearly circular, spiral
path inward. From F = ma, where F = kZe 2 /r 2 is the Coulomb force and a = v 2 /r is the
centripetal acceleration, we have

kZe2
mv 2 = mar = F r = .
r

This lets us write the total energy, kinetic plus potential, of the orbiting electron as

1 kZe2 kZe2
E =K+V = mv 2 − =− .
2 r 2r

The energy is negative because the zero of energy is the electron at rest at infinity. To find
how r varies with t, we take the time derivative of E, put it into the equation for the rate
at which the electron radiates, and use a = F/m = kZe 2 /mr 2 , to get
2
kZe2 dr 2 ke2 2 2 ke2 kZe2

dE
= = − a = − .
dt 2r 2 dt 3 c3 3 c3 mr 2

Rearranging this, we get


3 m2 c3 2
dt = − r dr .
4 k 2 Ze4
Integrating to get the time τ it takes for the radius to fall from some initial value r 0 to any
much smaller value, we have
1 m2 c3 3
τ' r .
4 k 2 Ze4 0
For Z = 1 and r0 = 10−10 m, the approximate size of an atom, we get τ ' 10 −10 s. Check
the numbers.
Bohr’s model—Niels Bohr had two things to explain—the stability of atoms, and the
frequencies of light they emit and absorb. He succeeded for the simplest atom, hydrogen,
and for other single-electron atoms such as singly ionized helium. The only way forward
was to break the rules of classical physics. His new rules were: (1) An electron bound to
a nucleus can only exist in a denumerable set of orbits having energies E n , where E1 <
E2 < E3 < · · · < 0. (2) A transition between these orbits is accompanied by the emission
or absorption of radiation of frequency ν given by hν = E n0 − En , where h is Planck’s
constant.

11
At the time, only a few hydrogen spectral lines were known, and Johann Balmer had
found an empirical formula for them. Bohr was guided by these experimental data, and by
the idea that in the limit of large orbits the electron ought to radiate like a classical antenna.
Bohr’s original derivation involved the energies directly, but the result was equivalent to
the famous quantization of angular momentum, L n = nh̄, where h̄ ≡ h/2π. This is usually
just pulled out of a hat, and we shall use that hat too (but see Prob. 6).
The quantization of the angular momentum leads to quantization of the speed, the
radius, and the energy of an orbit. Consider a single electron in a circular orbit around a
nucleus with charge Ze, where Z = 1 for hydrogen, Z = 2 for singly ionized helium, etc.
From F = ma again, we have
kZe2 v2
= m ,
r2 r
or
kZe2 = mv 2 r = (mvr)v = Lv ,
where L = mvr is the angular momentum. With L n = nh̄, we have

Z ke2
vn = .
n h̄

Dividing by the speed of light c, we have

vn Z
βn ≡ = α,
c n

where α ≡ ke2 /h̄c ' 1/137 is called the fine-structure constant (its value is usually given
in this inverse form). As will be seen, α pervades atomic physics. It is a dimensionless
constant, and as such its value is independent of the units we use; it is the same number,
just as is π, in advanced civilizations throughout the Universe. In hydrogen, the largest
speed the electron has is c/137, when n = 1. Thus nonrelativistic dynamics is a good
approximation. Eventually, we shall worry about a small relativistic correction.
Next, r. From Ln = mvr = nh̄, we have

nh̄ n2 h̄2 n2
rn = = = a0 ,
mvn Z mke2 Z

where a0 ≡ h̄2 /mke2 ' 0.529 × 10−10 m is called the Bohr radius. The radii of the Bohr
orbits increase as n2 .
Finally, and most importantly, E. This is, again,

1 kZe2 kZe2
E =K+V = mv 2 − =− .
2 r 2r

Using the quantized rn , we get

Z 2 mk 2 e4 Z2
En = − = E1 ,
n2 2h̄2 n2

12
where |E1 | = mk 2 e4 /2h̄2 ' 13.6 eV. Two other ways to write |E1 | are
2
h̄2 mke2 h̄2 1

|E1 | = = ,
2m h̄2 2m a20

and 2
ke2

1 1 2 2
|E1 | = mc2 = α mc ,
2 h̄c 2
where mc2 ' 0.511 MeV is the rest-mass energy of the electron. That |E 1 | is much smaller
than mc2 indicates again that nonrelativistic dynamics is a good approximation.
Figure 10(a) shows as horizontal lines some of the bound-state energy levels. An infinite
number of levels crowd in just below E = 0. Any positive energy is allowed, as is indicated
by the shading.
The hydrogen spectrum—What is experimentally observed is not the energy levels
but the frequencies of electromagnetic radiation emitted or absorbed. The second part of
Bohr’s model is that the spectrum comes from transitions between the energy levels, with
the frequencies ν being given by ∆E = hν:
 
2 1 1
hν = En0 − En = Z |E1 | 2
− 02 .
n n

Radiation is emitted when an electron drops down to a lower energy level and is absorbed
when it is bumped up to a higher level; and energy is conserved. Figure 10(a) shows as
vertical lines some of the transitions; their lengths are proportional to the frequencies of
the radiations. Transitions with the n = 1 level as the lower-energy state are in the Lyman
series; those with the n = 2 level as the lower state are in the Balmer series; and so on.
Figure 10(b) shows some of the frequencies; the scale is set so that the Lyman frequencies
match up with the upper energy levels involved in Lyman transitions. Transitions to or from
positive-energy states are suggested by shading. The only frequencies visible to our eyes
are the first few in the Balmer series; see Fig. 10(b). A major success of the model was to
predict the Lyman and other previously unseen series. This of course required instruments
that see at frequencies at which we cannot.
The Bohr model was a giant step toward a full understanding of atoms, but it is not
the correct theory. It is hardly a complete theory; it could not even handle an atom with
more than one electron. However, by breaking the rules of classical physics, it did show that
Planck’s constant would play a role in the correct theory; it did quantize bound states of
hydrogen and, aside from fine details, find the same levels eventually obtained with quantum
mechanics; it did lead to rules for quantizing other bound systems; and it did properly relate
the frequency spectra of electromagnetic radiation to the energy differences between levels.
Insofar as it goes, Fig. 10 is fully quantum mechanical.

13
ν (1015 Hz)
(a) (b)
E (eV) n

−0.85 4

3
−1.51 3

Lyman
Paschen

−3.40 2
Balmer

2
1
Balmer

visible
Paschen

−13.6
0

1
Lyman
Figure 10. (a) The bound-state energy levels of hydrogen, with some transi-
tions. (b) The radiation spectrum. The frequencies are proportional to energy
differences. Shading indicates transitions to or from the continuum of unbound
(positive-energy) states.

1·3. PARTICLES AND WAVES

Planck, Einstein, Compton, de Broglie—A cavity in a block of material contains


electromagnetic radiation. The energy density of the radiation depends sharply on the
absolute temperature T of the walls of the cavity. The energy distribution of the radiation
as a function of frequency and T is found by measuring the radiation that emerges from
a small hole in the wall. Such measurements, which again require instruments that see at

14
frequencies at which our eyes cannot, were quite good by the end of the nineteenth century.
However, no one could derive the experimental results from theory.
In 1900, Max Planck made a guess at the mathematical equation that would fit the
spectral distribution, and then tried to derive that equation. He was only able to do so by
making a radical asumption: that exchanges of energy between the radiation in the cavity
and the atoms in the walls of the cavity occurred in “quanta” of magnitude ∆E = hν,
where ν is the frequency and h is a new constant the value of which was determined by
fitting the spectral distribution. “At that time, the new postulate served one and only one
purpose: to derive—if one may call it that—his radiation law . . . ” (A. Pais, p. 133; see
p. 1 of this chapter). Thus was born the constant that rules over all of quantum mechanics.
And a very small constant it is indeed: h = 6.626 × 10 −34 J s.
No one, Planck included, had any idea whether his postulate had deep physical signifi-
cance or was just a mathematical trick, awaiting some future explanation. In fact, scarcely
anyone paid any attention to Planck’s result. The next step, by Albert Einstein in 1905,
concerned another obscure phenomenon, the photoelectric effect, in which light incident
upon a clean metallic surface can eject electrons. Measurements were poor, but two things
were known: (1) Weak blue light might eject electrons from a given metal, whereas intense
red light would not. This was puzzling because the energy delivered by an electromag-
netic wave is proportional to the intensity, independent of the frequency. (2) If electrons
are ejected, they begin to be ejected as soon as light strikes the surface, even with weak
light. This was puzzling because the energy of the incident wave is spread over the whole
wave front, and it was hard to see how weak radiation could be concentrated instantly on
a particular electron.
Einstein’s solution was to take Planck’s postulate a step farther. Suppose, he said,
that Planck’s exchanges of energy are quantized because the energy in the electromagnetic
radiation is itself quantized, in packets with energies E = hν. In this view, light is a rain
of “photons” (as we now call the light quanta). Einstein’s equation could not be simpler.
The maximum kinetic energy, Kmax , of an ejected electron will be

Kmax = hν − φ ,

where φ is the minimum energy needed to eject an electron from the metal (φ varies from
metal to metal, but is of order 2 eV). If hν < φ, then no electrons are ejected. Blue photons
might succeed in ejecting electrons where red ones would not because the frequency of blue
light is larger than that of red light: ν blue > νred .
This was a truly revolutionary idea—a century of experiments on interference and
diffraction and polarization had shown that light is a wave, not a rain of particles. In fact,
in 1913 none other than Planck himself, together with three other eminent physicists, in
a letter of recommendation for Einstein, wrote: “That he may sometimes have missed the
target in his speculations, as, for example, in his hypothesis of light quanta, cannot really
be held too much against him . . . ” Nor should their disbelief be held too much against
them—the first experiment to accurately confirm the relation K max = hν − φ was not done
until 1915. And not until 1923, when Arthur Holly Compton explained the scattering of
x-ray photons by electrons by treating the photons as particles, was the dual nature of
light—wave and particle—finally uneasily accepted by (nearly) everyone.

15
Compton’s kinematical calculations treated the x-ray photons as particles having en-
ergy E and momentum p, where

E = hν and p = h/λ ,

where λ = c/ν is the wavelength of the light. Notice how particle aspects E and p on the
left sides of the equations are related to wave aspects ν and λ on the right. The equation
for p comes from the relativistic relation E 2 = p2 c2 + m2 c4 . The photon is massless, so
p = E/c = hν/c = h/λ.
In 1923, Louis de Broglie advanced another absurd idea: If waves sometimes exhibit
particle-like behavior, perhaps particles sometimes exhibit wave-like behavior. And he
proposed that the relation between the particle and wave aspects was again

p = h/λ .

This gave a new way to look at the Bohr quantization rule for angular momentum: the orbits
are those for which an integral number of wavelengths fit round a circle. If 2πr = nλ, where
n = 1, 2, . . . , then 2πr = nh/p, or L = rp = nh/2π = nh̄: Bohr’s rule. Furthermore, de
Broglie’s hypothesis provided an explanation for patterns already beginning to be observed
for electron beams scattered by crystals: the electrons are diffracted by the regular array
of atomic sites in the crystal. Experiments soon showed that the scattering by crystals of x
rays and of electrons of the same wavelength gave the same diffraction patterns.
Interference and diffraction of neutrons—Neutrons are particles. Aside from
ordinary hydrogen and the rare isotope 3 He, all the stable nuclei have at least as many
neutrons as protons. And a neutron has slightly more mass than a proton plus an electron
(a free neutron decays to a proton, an electron, and a neutrino). Thus, aside from any
hydrogen content, a least half the mass of every object we see around us comes from the
neutrons it contains. Neutrons are particles!
Figure 11 shows an experimental set-up used to investigate the wave nature of neutrons;
see R. Gähler and A. Zeilinger, “Wave-Optical Experiments with Very Cold Neutrons,”
Am. J. Phys. 59 (1991) 316. A beam of neutrons from a nuclear reactor is collimated by
slits S1 and S2 and directed upon a quartz prism. The prism fans out neutrons according to
wavelength, just as an ordinary prism fans out light (but light would be bent upward, not
downward; the neutrons speed up in the prism). Slit S 3 then selects neutrons in a narrow
band of wavelength and diffracts them toward the object slit S 5 . The slits in S5 , shown
magnified in Fig. 12, are about 22 µm wide, and the center-to-center distance is about 126
µm—a tiny arrangement, but just visible with the naked eye. The slit widths and separation
can be varied; the wire can be removed to make a single slit of variable width; and a jaw
can be removed to leave a single sharp edge. The pattern produced by the object slit is
observed at the far right by “stepping” the scanning slit S 4 across the pattern and counting
behind the slit at each step for some interval of time.

16
Figure 11. Neutrons from a reactor are collimated by slits S 1 and S2 , refracted by
a quartz prism, selected on momentum by S 3 , and directed upon the object slit
S5 . The patterns produced are measured at S 4 . Note the changes of scale in the
figure. From R. Gähler and A. Zeilinger, Am. J. Phys. 59 (1991) 316.

Figure 12. The object slit arrangement at S 5 . From R. Gähler and A. Zeilinger,
Am. J. Phys. 59 (1991) 316.

Figure 13 shows the patterns produced by the double slit and by a single edge. The
paper cited also shows single-slit diffraction patterns. The curves through the data points
are calculated from the geometry of the set-up and wave optics, using de Broglie’s λ = h/p;
the only free parameter is the total number of counts. With p = mv (the neutrons here are
nonrelativistic), the wavelength is given by
3.96 × 10−7
λ= ,
v
where λ is in meters and v is in m/s. The neutrons are made slow (“cold”) by passing
them through a bath of liquid deuterium at 25 K before they enter the set-up. Otherwise

17
λ would be too small to observe interference effects in a set-up of (barely) macroscopic
dimensions. As it is, the wavelength selected is about 2 nm, much smaller than the 400–700
nm of visible light. The interference minima in Fig. 13(top) are not zero mainly because
there is some spread of wavelengths. Otherwise, the interference is exactly as it would be
for light. Neutrons are waves!

Figure 13. The interference pattern produced by two slits, and the diffraction pat-
tern produced by a single sharp edge. The only free parameters in the fitted curves
are the total numbers of counts. From R. Gähler and A. Zeilinger, Am. J. Phys. 59
(1991) 316.

This discussion—the explanation—has been in terms of waves, but neutrons are par-
ticles! The neutrons are produced in the fission of uranium atoms in the reactor. The
rate at which neutrons pass through the set-up is such that usually one neutron has been
counted before the next one that will reach the counter is even born in the decay of another
uranium atom. Thus the patterns are produced one neutron at a time. If the scanning
slit and counter were replaced by a pixelated detector covering the whole back area, each

18
neutron would arrive at a single pixel, not spread out. Surely a particle that always arrives
at some one place must on its way to contributing to the double-slit pattern pass through
only one slit or the other. But if that is the case, how can the neutron then even know that
the other slit exists? But just as surely, the pattern built up requires the interference of
something that passes through both slits—the two-slit interference pattern is not remotely
like the superposition of two one-slit patterns. Neutrons are waves! Neutrons are particles!
We are driven to conclude that neutrons (and all other “particles”) have a dual nature,
being neither purely particle nor wave. We have a theory—quantum mechanics—that pre-
dicts the results of all such experiments. But our brains can’t fully get round the duality
of waves and particles. As Richard Feynman said, “Nobody understands how it can be
like that.” Chapter 1 of Vol. III of The Feynman Lectures is all about this duality: “[It] is
impossible, absolutely impossible, to explain [the duality] in any classical way . . . ” Everyone
should read that chapter.

Interference of photons—Figure 14 shows a Mach-Zehnder interferometer. At the


lower left (ignore the “down-converter” for a moment), a parallel beam of light of wavelength
λ is incident upon a 50-50 beam splitter—half of the intensity of the light is reflected and
half is transmitted. Each half beam is then reflected from a mirror toward a second beam
splitter, where the beams are recombined. In general, some of the light will go to detector
C1 and some will go to C2. We might simply observe the illumination of screens placed
at positions 1 and 2; or we might hook up the detectors to loudspeakers and compare the
loudness of two tones. To see circular fringes on a screen at C1 or C2, we might introduce
a diverging lens in front of the first beam splitter; to see linear fringes, we might tilt one of
the mirrors. But here we do neither.
C4 C2
or
irr
M

C1
Beam splitter

C3

Down-
converter Beam splitter
or
irr
M

UV light
Figure 14. A Mach-Zehnder interferometer.

Suppose that the distance to C1 is the same by either path, or that these lengths differ
by an integral number of wavelengths. Then all the light goes to C1. This is because each
path to C1 involves a reflection from a mirror and a reflection from one beam splitter and a
transmission of the other. Since exactly the same things happen to the two beams (although

19
the order of happenings is different), they arrive at C1 in phase and interfere constructively.
The happenings along the two paths to C2 are not the same.
The basic physics of an interferometer and a double-slit experiment is the same, but
the interferometer set-up can be arranged to collect the constructive interference at one
detector and the destructive interference at another, whereas the double-slit experiment
spreads varying degrees of interference—constructive and destructive—over a plane.
We can reduce the intensity of the incoming beam to the point where only one photon
is in the interferometer at a time. We can know when a photon enters the interferometer
by placing an “ultraviolet down-converter,” shown in Fig. 14, at the entrance. This device
“splits” an ultraviolet photon into two visible photons. Let the two path lengths to C1 be
equal or differ by an integral number of wavelengths; and set the detectors to “click” when
a photon arrives. Then when C3 clicks so does C1.
We can ask which path a particular photon takes to C1 by placing a counter, C4, in,
say, the upper path (see Fig. 14). About half the time a photon enters the system, we get
a click at C4: but then we get no click at either C1 or C2. The photon took the upper
path but was absorbed—we can’t both record the arrival of the photon at C4 and have it
continue undisturbed on its way. The other half of the time, we get a click at either C1 or
C2. The photon took the lower path, but now it is equally likely to get to C1 and C2—the
upper path is now blocked by C4, and the interference has been destroyed by our looking
to see which path was taken. But now we have the same kind of question we asked about a
neutron in the double-slit experiment: Since the photon only ever shows up at one place or
another, how can it somehow take—or even know—about two paths, as interference (with
C4 absent) seems to require? Is light particles or is it waves?
No one has been able to devise an arrangement in which we can know both which path
the photon took and get interference. To get interference, we must not interfere. If we look,
the photon takes either the upper path or the lower path. If we don’t look, it somehow
knows about both paths. It follows the advice of the philosopher Yogi Berra: “When you
come to a fork in the road, take it.”
Seeing in the dark—Interest in the interferometer arrangement just discussed was
enlivened by A.C. Elitzur and L. Vaidman, in “Quantum Mechanical Interaction-Free Mea-
surements,” Found. Phys. 24 (1993) 987. Imagine a stock of bombs, each one having a
trigger sensor so sensitive that the slightest touch or even a single photon striking it will
explode the bomb. However, some of the sensors have fallen out, and a photon reaching
where the sensor should be will pass through undisturbed. Can the “good” bombs—those
with sensors in place—be sorted out and saved? With the present arrangement, the answer
is: about one-third of them.
In total darkness, we place a bomb in our interferometer, with the sensor—or where
the sensor should be—in the upper path. The sensor of a good bomb blocks the upper path;
the sensor of a bad bomb is absent and the upper path is open. We send in a photon, and
three things can happen: (1) The bomb explodes. As Elitzur and Vaidman put it, “The
bomb was good.” The photon took the upper path. An explosion occurs with a probability
of 50% when the bomb is good. (2) C2 clicks. Thi means that the bomb is good: the
loss of constructive interference at C1 means that the upper path is blocked by a sensor.
C2 clicks with a probability of 25% when the bomb is good. We quickly prevent any more
photons from entering the interferometer, because the next one might take the upper path

20
and explode the bomb. (3) C1 clicks. We can’t tell whether the bomb is good or not. For a
good bomb this happens with a probability of 25%; for a bad bomb it always happens. We
send in another photon. If the bomb is good, eventually we either blow up the apparatus
or we get a click at C2. But if C1 clicks for, say, twenty photons in a row, we become
convinced that the sensor is missing and go on to test the next bomb.
We can identify all the bad bombs: each one gives a long series of clicks at C1. And
we can save one-third of the good bombs, because the ratio of clicks at C2 to explosions
is 25-to-50 when the bomb is good. Getting a click at C2 tells us that the upper path is
blocked, even though the photon didn’t take that path. We know that the sensors are there
even though no photon touches the bombs we save. With a more complicated arrangement,
it is in theory even possible to save all of the good bombs, while touching none of them. See
P. Kwiat, H. Weinfurter, and A. Zeilinger, “Quantum Seeing in the Dark,” Sci. Amer. 275
(November 1996) 72.

PROBLEMS

1. Some solid angles.


Find the solid angle subtended by:
(a) Each face of a cube from a corner of the cube.
(b) Iraq from the center of the Earth.
(c) The Sun or the Moon from the Earth (each has an angular diameter of about 1/2 ◦ ).
(d) The curved surface of a cylinder of length L and radius R from the center of an
end.

2. A bull’s-eye pattern.
Consider Rutherford scattering. Imagine looking in along the scattering axis. Draw circles
centered on the axis having radii equal to impact parameters for scattering through 30, 60,
90, 120, and 150◦ . Use a scale of 4 cm for an impact parameter equal to D, the distance of
closest approach for head-on repulsive scattering. What would the radius on your bull’s-eye
pattern be for scattering through 1 ◦ ?
Make a table giving the cross sections for scattering in the intervals between 30 and
60 , between 60 and 90◦ , etc., as multiples of D 2 .

3. More Rutherford scattering.


(a) Show that the construction given in Figure 8 does indeed illustrate the relation
1
b = 2 D cot(θ/2) for Rutherford scattering.
(b) Let σ(θ > 90◦ ) be the cross section for Rutherford scattering through angles greater
than 90◦ . Find the angle θ1 such that σ(θ > θ1 ) = 2σ(θ > 90◦ ).
(c) Let dσ(90◦ ) be the differential cross section for scattering at 90 ◦ into a counter that
subtends the solid angle dΩ. Find the angle θ such that dσ(θ) = 2 dσ(90 ◦ ).

21
4. Where Rutherford breaks down.
(a) For Rutherford scattering, show that the distance of closest approach of q to Q is
given in terms of b and D by
 2 1/2
D D
r0 = + b2 ± ,
4 2
where the ± is for charges of like or unlike sign.
(b) Inside the scattering nucleus, the electric field is no longer that of a point charge
Q. For charges of like sign, this will not affect the scattering angular distribution as long
as D is greater than the nuclear radius R. For D = R/2, find the angle θ 0 past which the
angular distribution won’t be Rutherfordian.
(c) A swarm of comets, all having the same velocity v 0 , approaches a star having a
radius R. Find the “capture cross section,” the area of the (distant) swarm front within
which comets will strike the star.

5. Quantizing the oscillator.


Use Bohr’s rule, Ln = nh̄, to find the quantized speeds, radii, and energies for circular
orbits of a particle of mass m bound to the origin by the Hooke’s-law force, F(r) = −Kr.
(Quantum mechanics gives energies lower by h̄ω/2; see Chap. 5.)

6. A classical limit.
Classically, a charged particle moving in a circular orbit radiates at the frequency of the
motion. In Bohr’s hydrogen atom, the frequency of radiation for transitions between adja-
cent orbits is ν = (En+1 − En )/h. Show that in the limit of large n this frequency is equal
to the frequency of the motion.

7. Energies in electron volts.


(a) The range of visible wavelengths of light is about 400 to 700 nm. What range of
photon energies is this, in eV? (This range is worth remembering.)
(b) Show
√ that the de Broglie wavelength of an electron having kinetic energy K is
λ = 1.23/ K nm, where K is in eV. Find the corresponding relation for a proton.

8. Standing waves in a box.


(a) Find the quantized energies of a nonrelativistic particle of mass m in a one-
dimensional box of length L by requiring that the round-trip distance 2L be an integral num-
ber of de Broglie wavelengths. (Quantum mechanics gives the same energies; see Chap. 2).
(b) Do the same for a photon. Now the relation between E and p is E = pc.

9. The neutron wavelength.


Determine the wavelength of the neutrons in the experiment discussed in Sec. 3 using the
dimensions of the apparatus given in Figs. 11 and 12 and in the text and:
(a) the spacing of the interference fringes in Fig. 13(top).
(b) the spacing of the fringes in Fig. 13(bottom). Begin by finding where the geometric
shadow of the edge would be. (You might need to look at an optics text.)

22

Anda mungkin juga menyukai