You are on page 1of 23

Chapter 39

Particles Behaving
as Waves
PowerPoint® Lectures for
University Physics, Thirteenth Edition
– Hugh D. Young and Roger A. Freedman

Copyright © 2012 Pearson Education Inc.


Goals for Chapter 39
• To study the wave nature of electrons
• To examine the evidence for the nuclear model of
the atom
• To understand the ideas of atomic energy levels
and the Bohr model of the hydrogen atom
• To learn the fundamental physics of how lasers
operate
• To see how the ideas of photons and atomic
energy levels explain the continuous spectrum of
light emitted by a blackbody
• To see how the Heisenberg uncertainty principle
applies to the behavior of particles

Copyright © 2012 Pearson Education Inc.


Introduction
• At the end of the 19th century light was regarded as a wave
and matter as a collection of particles. Just as light was found
to have particle characteristics (photons), matter proved to
have wave characteristics.
• The wave nature of matter allows us to use electrons to make
images (such as the
one shown here of
viruses on a bacterium).

Copyright © 2012 Pearson Education Inc.


De Broglie waves
• The physicist de Broglie hypothesized, from an argument for symmetry in
Physics, that the relationships E = hf = hc/ and p = h/ for photons also
apply to electrons. Thus electrons should have wave characteristics.
• The de Broglie hypothesis was confirmed by the discovery by Bell Labs
physicists Davisson and Germer in 1927 that electrons undergo diffraction
when interacting with crystalline matter, just like x rays do (see Figure 39.4).
• Note: Because electrons have mass and do
not travel at the speed of light, they DO
NOT obey the relation f = c.
• Kinetic energy for non-relativistic particles
can be written 2
p
K  12 mv 2 
2m
• If we accelerate an electron through a
potential difference V, it gains a kinetic
energy eV, so we2
can write:
p h
K  eV   p  2meV 
2m 
• Thus, the wavelength of an electron   h Diffraction, as usual, obeys
accelerated through potential V is 2meV d sin   m
Copyright © 2012 Pearson Education Inc.
Electron Diffraction
• Example 39.1: In an electron-diffraction experiment, an electron is
accelerated by a potential difference of 54 V and a maximum in the
diffraction pattern occurs at 50°. X-ray diffraction shows the atomic spacing
in the target to be d = 2.181010 m. Find the electron wavelength.
h 6.626 1034 J  s
   1.67 1010 m  0.17 nm
2meV 2(9.1 1031 m)(1.602 1019 C)(54 V)

• Alternatively, we could use the diffraction information:


  d sin   (2.18 1010 m) sin(50)  1.7 10 10 m

• Accelerating to a higher energy would yield a smaller wavelength for the


electron, which suggests using electrons a probes for imaging at high
resolution.

Copyright © 2012 Pearson Education Inc.


Electron microscopy
• Example 39.3—An electron microscope.
One can use magnets as focusing elements,
and create an “optical” system for
electrons. Note, the blue cones in the
diagram are not ray paths like we used for
geometric optics.
• What voltage is needed to accelerate
electrons to a wavelength of 10 pm
(roughly 50,000 times smaller than visible
light)?

h2
V
2me 2
(6.626 1034 J  s) 2

2(9.11031 m)(1.602  1019 C)(1.0 10 11 m) 2
 15000 V

Copyright © 2012 Pearson Education Inc.


Atomic line spectra
• As we saw last week, if He gas is sealed in a glass tube and heated to form a
hot gas, the light emitted by the He atoms includes only certain discrete
wavelengths. The spectrum of this light (“line spectrum”) is different for
different elements. Nineteenth-century physics had no explanation for this.

Copyright © 2012 Pearson Education Inc.


The nuclear atom
• Rutherford probed the structure of the atom
by sending alpha particles at a thin gold foil.
Some alpha particles were scattered by large
angles, leading him to conclude that the
atom’s positive charge is concentrated in a
nucleus at its center.
• Example 39.4.

Copyright © 2012 Pearson Education Inc.


Size of a Gold Nucleus
• Example 39.4: Rutherford Scattering
• An alpha particle (charge 2e) is aimed directly at a gold nucleus (charge 79e).
What is the minimum kinetic energy of the alpha particle in order to approach
within 5.0x10-14 m of the center of the gold nucleus? Assume the gold nucleus
remains at rest.
The alpha particle must overcome the electrostatic repulsion of the gold
nucleus, which can be characterized by the potential energy at the point of
closest approach
1 qq0
U
4 0 r
(2)(79)(1.602  1019 C) 2
 9.0 10 N  m /C
9 2 2

5.0 10 14 m


 7.3 1013 J  4.6 MeV

By energy conservation, the incoming kinetic energy must be at least this high.
By changing the incoming kinetic energy, the atom can be probed at different
depths, which allows the size of the nucleus to be determined.

Copyright © 2012 Pearson Education Inc.


The failure of classical physics
• Rutherford’s experiment suggested that
electrons orbit around the nucleus like a
miniature solar system. However,
classical physics predicts that an orbiting
electron would emit electromagnetic
radiation and fall into the nucleus (see
Figure 39.14). So classical physics could
not explain why atoms are stable.
• Niels Bohr explained atomic line spectra
and the stability of atoms by postulating
that atoms can only be in certain discrete
energy levels, and cannot further decay
from a minimum energy called the
ground level. The atom is said to be in
the ground state.

Copyright © 2012 Pearson Education Inc.


Atomic energy levels
• When an atom makes a transition from one energy level to a lower level, it
emits a photon whose energy equals that lost by the atom (see Figure 39.16
at lower left).
• An atom can also absorb a photon, provided the photon energy equals the
difference between two energy levels (see Figure 39.17 at lower right).
When an atom is in a higher energy level, it is said to be in an excited state.
• Energy levels for soduim

Copyright © 2012 Pearson Education Inc.


• The same gas that, when heated, gives emission lines at certain wavelengths
will also show the same spectrum in absorption (i.e. it is able to remove
photons from the incoming light by absorbing them and raising the energy
level of the atom to an excited state).

Copyright © 2012 Pearson Education Inc.


The Bohr model of hydrogen
• Bohr explained the line spectrum of hydrogen
(see Figure 39.25 below) with a model in which
the single hydrogen electron can only be in
certain definite orbits.
• In the nth allowed orbit, the electron has
quantized orbital angular momentum nh/2
• Recall that angular momentum is L  r  p
• For a circular orbit, the magnitude of
momentum for the nth orbit is
L  rn mvn  n
(note that this means h has units of momentum,
check it!).
• Bohr’s argument is rather complicated, but with
it he was able to explain quantitatively the
spectral lines of hydrogen.

Copyright © 2012 Pearson Education Inc.


Classical Explanation for Quantized Momentum
• One way to get a (rather misleading) physical intuition
about the quantized nature of atomic orbits, consider the
wavelength of an electron in various orbits.
h h
n  
p mvn
• It makes sense that electrons, due to their wave nature,
ought to require orbits that are an integral number of
wavelengths, as if they were standing waves. If we require
nh
2 rn  nn 
mvn
then rearrangement gives Bohr’s result rn mvn  n .
• Also, since the electrostatic force must supply the
centripetal force to keep the electron in its circular orbit,
we have 1 e 2 mvn 2
F 
4 0 rn 2
rn
• Combining these two equations allows us to solve
simultaneously
2 2
for rn and v2n: 2
nh 1 e h
rn   0 ; vn  a0   0 (Bohr radius)
 me 2  0 2nh  me2
Copyright © 2012 Pearson Education Inc.
Scale Drawing of Orbits and Standing Waves

Copyright © 2012 Pearson Education Inc.


Hydrogen spectrum in more detail
• The line spectrum at the bottom of the slide is not the entire spectrum of
hydrogen; it’s just the visible-light portion, involving transitions to the n = 2
energy state. Here, n is the principle quantum number.
• Hydrogen also has series of spectral lines in the infrared and the ultraviolet.
En  K n  U n
1 e2
 mv 
1 2
n
4 0 rn
2

1 me 2
 2 2 2
 0 8n h
hcR

n2
me 4
R  2 3 (Rydberg constant)
8 0 h c
1  1 1 
 R 2  2 
  nL nU 

Copyright © 2012 Pearson Education Inc.


Hydrogen-like atoms
• Numerical value of the Rydberg constant is: R  1.097 107 m 1
• If a higher-charge nucleus has only a single electron, the same pattern of
energy levels exists, but one must take account of the nuclear charge Z:
hcZ 2 R
En   (energy levels for single electron atom)
n2
• Note that all bound energies are negative, and approach zero as n gets larger.
• Here are the energy levels for hydrogen (H) and singly-ionized helium (He +).

• The simple picture does NOT work


for atoms with more than one
electron!
• Note, we have not taken account of
the finite mass of the nucleus. To be
correct, we should use the reduced
mass in place of the electron mass
m1m2
 (reduced mass)
m1  m2

Copyright © 2012 Pearson Education Inc.


Distribution of States
• Absorption of radiation vs. collisional excitation.
• Spontaneous emission vs. stimulated emission.
• Atoms spontaneously emit photons of frequency
f when they transition from an excited energy
level to a lower level. Excited atoms can be
stimulated to emit coherently if they are
illuminated with light of the same frequency f.
This happens in a laser (Light Amplification by
Stimulated Emission of Radiation).
• One can relate the temperature of a gas to the
distribution of energy states in the gas using the
Boltzmann equation. The ratio of atoms in state
n to atoms in the ground state is given by
nn Ae  En / kT
  E1 / kT
 e  ( En  E1 )/ kT
n1 Ae

Copyright © 2012 Pearson Education Inc.


Example: Hydrogen Lines in the Solar Spectrum
• Say we heat hydrogen to 6000 K (the temperature of the Sun’s surface).
What is the ratio of atoms in the first excited state (n = 2) compared to the
ground state (n = 1)?

• The value of (E2 – E1) is


 1 1
E2  E1   hcR  2  2 
2 1 
3hcR
  0.75(6.626 10 34 Js)(3 108 m/s)(1.097 107 m 1 )
4
 1.64 1018 J
• The value of kT is:
kT  (1.38  1023 J K 1 )(6000 K)
 8.28  1020 J
• Thus
n2
 e  ( E2  E1 )/ kT  e 19.8  2.5  109
n1

• This is a very small number, but nevertheless the Sun has deep absorption
lines in the Balmer series (due to transitions from even higher energy states).
Copyright © 2012 Pearson Education Inc.
Continuous spectra and blackbody radiation
• A blackbody is an idealized case of a hot, dense
object. Such objects show a continuous spectrum
rather than spectral lines. The reason is that many-
many photon interactions scatter and shift the
wavelengths, spreading them into a characteristic
shape as shown at right, called the blackbody
spectrum.
• The three characteristics of a blackbody spectrum
are:
1. The intensity grows with temperature as I = T 4,
where s = 5.67108 W m2 K4 is the Stefan-
Boltzmann constant.
2. The peak shifts according to mT = 2.90 10 m-K
(called the Wien displacement law)
3. The shape of the distribution I() is invariant (after
scaling for the above two factors).
• Classical physics could not explain the shape of the 2 ckT
I ( ) 
blackbody spectrum. Rayleigh had an idea that 4
explained the lower-frequency part: Consider normal
modes
Copyright of Education
© 2012 Pearson wavesInc.in a box, with energy equipartion kT.
Continuous spectra and blackbody radiation
• Rayleigh’s formula works well for long
wavelengths, but “blows up” at short wavelengths,
which was referred to as the ultraviolet catastrophe.
Planck provided an explanation by assuming that
atoms in the blackbody have evenly-spaced energy
levels, so that at shorter wavelengths (higher
energies) there were fewer “states” for the photons
to be in. Planck’s formula was entirely empirical:
2 hc 2
I ( )  5  hc /  kT (Planck blackbody radiation law)
 e  1
• This formula explains all three of the characteristics
of the blackbody spectrum in the previous slide.
• It was this empirical quantization of photon energy
that Einstein showed was actually true with his
explanation of the photoelectric effect.

Copyright © 2012 Pearson Education Inc.


Continuous spectra and blackbody radiation
• Example 39.7: Light from the Sun
To a good approximation, the Sun is a blackbody with temperature 5800 K.
(a) At what wavelength does the Sun emit most strongly?
To answer this, we use the Wien displacement law
2.90 10 3 m  K
  0.500  106 m  500 nm
5800 K
(b) What is the total radiated power per unit surface area?
For this, we use the Stefan-Boltzmann law
I   T 4   5.67 108 W  m 2  K 4  (5800 K) 4  6.42 107 W  m 2  64.2 MW/m 2
• Example 39.8: A slice of sunlight
Find the power per unit area radiated from the Sun’s surface in the
wavelength range 600.0 to 605.0 nm.
We need an integral over the Planck curve. Because this is such a narrow
range, we can approximate by assuming the intensity is nearly constant:
2 2 hc 2
I   I ( )d    2  1  where    2  1  / 2
1

5
e  hc /  kT

1

• Plugging in all of the numbers, we get I = 0.39 MW/m2.


Copyright © 2012 Pearson Education Inc.
The Heisenberg Uncertainty Principle revisited
• The Heisenberg uncertainty principle for momentum and position applies to
electrons and other matter, as does the uncertainty principle for energy and time.
This gives insight into the structure of atoms.
• Example 39.9: The uncertainty principle: position and momentum.
h2
Consider an electron bound to the ground-state orbit, a0   0  5  10 11
m
 me 2

According to the Heisenberg uncertainty principle, the momentum uncertainty is


 1.055 1034
px   10
 1.055  1024 kg m/s
2x 10
This is associated with an energy uncertainty:
 1.055 1024 kg m/s 
2
 px 
2

E   31
 6.11 1019 J  3.81 eV
2m 2(9.110 kg)

• Example 39.10: The uncertainty principle: energy and time.


A sodium atom remains in its first excited state for about 1.6108 s before
decaying to the ground state and emitting a photon of  = 589 nm. What is the
uncertainty in energy of that state? What is the spread of wavelengths of the
photon emitted?  1.055 1034 8 E 2
E   8
 2.1 10 eV    0.0000059 nm
2t 2(1.6 10 ) hc
Copyright © 2012 Pearson Education Inc.

You might also like