100% found this document useful (7 votes)
2K views

Nuclear Physics Explained

Physics

Uploaded by

guruji 1
Copyright
© © All Rights Reserved
Available Formats
Download as PDF, TXT or read online on Scribd
100% found this document useful (7 votes)
2K views

Nuclear Physics Explained

Physics

Uploaded by

guruji 1
Copyright
© © All Rights Reserved
Available Formats
Download as PDF, TXT or read online on Scribd
You are on page 1/ 344

Topic Subtopic

Science Physics & Chemistry

Nuclear Physics
Explained
Course Guidebook

Professor Lawrence Weinstein


Old Dominion University
Published by
T he G r eat C ourses

Corporate Headquarters

4840 Westfields Boulevard | Suite 500 | Chantilly, Virginia | 20151‑2299


[phone] 1.800.832.2412 | [fax] 703.378.3819 | [ web] www.thegreatcourses.com

Copyright © The Teaching Company, 2018

Printed in the United States of America

This book is in copyright. All rights reserved. Without limiting the rights under copyright reserved
above, no part of this publication may be reproduced, stored in or introduced into a retrieval system,
or transmitted, in any form, or by any means (electronic, mechanical, photocopying, recording, or
otherwise), without the prior written permission of The Teaching Company.
Lawrence
Weinstein, PhD
Professor of Physics
Old Dominion University

L
awrence Weinstein is a Professor of Physics at Old Dominion
University (ODU) and a researcher at the Thomas Jefferson
National Accelerator Facility (Jefferson Lab). He received his
bachelor of science degree in Physics from Yale University and his
doctorate in Physics from the Massachusetts Institute of Technology.
Professor Weinstein’s research involves experimental nuclear physics,
using electron scattering to study the structure of the nucleus and the
structure of the proton, both inside and outside the nucleus.

Prior to arriving at ODU, Professor Weinstein conducted research at


MIT, performing experiments at the Bates Linear Accelerator Center.
His significant contributions include building major pieces of equipment
for Jefferson Lab, measuring how nucleons behave singly and in pairs in
nuclei, using an antimatter beam to measure the structure of the proton,
and discovering the correlation between nucleon pairing and bound
nucleon structure modification. Professor Weinstein’s current research
focuses on short-range correlations in nuclei and on the quark structure
of bound nucleons.

PROFESSOR BIOGRAPHY i
Professor Weinstein has won a plethora of teaching and research awards.
In addition to several awards from the ODU College of Sciences, he
received the ODU Teaching with Technology Award, was named
University Professor for his outstanding teaching, and received the
A. Rufus Tonelson Faculty Award from ODU, the George B. Pegram
Award for Excellence in Physics Education in the Southeast from the
American Physical Society, and the Virginia Outstanding Faculty
Award. In recognition of his research, Professor Weinstein was named an
Eminent Scholar, a distinction reserved for only 4% of the ODU faculty,
and was named a fellow of the American Physical Society. In addition,
he was elected chair of the 1600-member Jefferson Lab Users Group to
represent all the scientists performing research at Jefferson Lab.

Professor Weinstein is the author of 2 popular books—Guesstimation:


Solving the World’s Problems on the Back of a Cocktail Napkin (with
John Adam) and Guesstimation 2.0: Solving Today’s Problems on the
Back of a Napkin—about techniques for approximate solutions to
any problem. He and Guesstimation were featured in The New York
Times and National Geographic magazine. Professor Weinstein has
coauthored more than 200 publications in professional journals, with
more than 10,000 citations. He has given more than 110 professional
presentations, in addition to more than 75 talks and physics
demonstration shows for community groups, high schools, and
middle schools. u

ii Nuclear Physics Explained


TABLE OF CONTENTS

Introduction
Professor Biography . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . i
Disclaimer . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . v
Course Scope . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1

Guides
1 A Tour of the Nucleus and Nuclear Forces . . . . . . . . . . . . . . . . . . 4
2 Curve of Binding Energy: Fission and Fusion . . . . . . . . . . . . . . . . 14
3 Alpha, Beta, and Gamma Decay . . . . . . . . . . . . . . . . . . . . . . . 24
4 Radiation Sources, Natural and Unnatural . . . . . . . . . . . . . . . . . 35
5 How Dangerous Is Radiation? . . . . . . . . . . . . . . . . . . . . . . . . 44
6 The Liquid-Drop Model of the Nucleus . . . . . . . . . . . . . . . . . . . 55
7 The Quantum Nucleus and Magic Numbers . . . . . . . . . . . . . . . . 66
8 Particle Accelerators: Schools of Scattering . . . . . . . . . . . . . . . . . 77
9 Detecting Subatomic Particles . . . . . . . . . . . . . . . . . . . . . . . . 88
10 How to Experiment with Nuclear Collisions . . . . . . . . . . . . . . . . 98
11 Scattering Nucleons in Singles or in Pairs . . . . . . . . . . . . . . . . . . 107
12 Sea Quarks, Gluons, and the Origin of Mass . . . . . . . . . . . . . . . . 120

TABLE OF CONTENTS iii


13 Nuclear Fusion in Our Sun. . . . . . . . . . . . . . . . . . . . . . . . . . 131
14 Making Elements: Big Bang to Neutron Stars . . . . . . . . . . . . . . . 141
15 Splitting the Nucleus . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 154
16 Nuclear Weapons Were Never “Atomic” Bombs . . . . . . . . . . . . . . 164
17 Harnessing Nuclear Chain Reactions . . . . . . . . . . . . . . . . . . . . 175
18 Nuclear Accidents and Lessons Learned . . . . . . . . . . . . . . . . . . . 184
19 The Nuclear Fuel Cycle and Advanced Reactors . . . . . . . . . . . . . . 198
20 Nuclear Fusion: Obstacles and Achievements . . . . . . . . . . . . . . . . 210
21 Killing Cancer with Isotopes, X-Rays, Protons . . . . . . . . . . . . . . . 224
22 Medical Imaging: CT, PET, SPECT, and MRI . . . . . . . . . . . . . . . 235
23 Isotopes as Clocks and Fingerprints . . . . . . . . . . . . . . . . . . . . . 246
24 Viewing the World with Radiation . . . . . . . . . . . . . . . . . . . . . 258

Supplementary Material
The Periodic Table of the Elements . . . . . . . . . . . . . . . . . . . . . . . . 270
The Chart of Nuclides . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 274
Glossary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 276
Select Nuclear Physics Research Facilities . . . . . . . . . . . . . . . . . . . . 303
Timeline . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 305
Nobel Laureates . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 317
Bibliography . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 321
Image Credits. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 335

iv Nuclear Physics Explained


DISCLAIMER

This series of lectures is intended to increase your understanding of the principles


of nuclear physics. These lectures include demonstrations in the field of nuclear
physics, performed by an experienced professional. These experiments may
include dangerous materials and are conducted for informational purposes only,
to enhance understanding of the material.

WARNING: THE DEMONSTRATIONS PERFORMED IN THESE


LECTURES CAN BE DANGEROUS. ANY ATTEMPT TO PERFORM
THESE DEMONSTRATIONS ON YOUR OWN IS UNDERTAKEN AT
YOUR OWN RISK.

The Teaching Company expressly DISCLAIMS LIABILITY for any


DIRECT, INDIRECT, INCIDENTAL, SPECIAL, OR CONSEQUENTIAL
DAMAGES OR LOST PROFITS that result directly or indirectly from the use
of these lectures. In states that do not allow some or all of the above limitations
of liability, liability shall be limited to the greatest extent allowed by law.

DISCLAIMER v
vi Nuclear Physics Explained
NUCLEAR PHYSICS
EXPLAINED

Nuclear physics is all about the nucleus—which is only 0.0001 the size
of the atom but contains 99.9% of the atom’s mass—including the forces
that hold nuclei together or allow them to decay; the chart of nuclides,
which shows the myriad of stable and unstable nuclei; nuclear decay and
transmutation; energy release through fusion or fission; nuclear medicine;
and many other applications.

The curve of binding energy [LECTURE 2] shows the pattern by which


some elements can undergo fission to produce energy, while others
undergo fusion. Nuclei that are too heavy, or have too many protons or
neutrons, are unstable and emit radiation [LECTURE 3]. These radioactive
nuclei occur naturally all around us [LECTURE 4], but too much radiation
can cause sickness, cancer, or even death [LECTURE 5].

You will learn about nuclei by exploring models that predict their
features, starting with the simplest model, the liquid-drop model, and
adding quantum mechanics slowly, first with the Fermi gas model (which
also describes the largest nuclei of all, neutron stars) [LECTURE 6] and then
with the shell model (where protons and neutrons occupy shell-model
orbitals in the nucleus just like electrons in the atom) [LECTURE 7] to
describe nuclei more and more precisely.

COURSE SCOPE 1
You will then get a detailed inside look at the experimental techniques
used to study the nucleus by visiting a detector lab [LECTURE 9] and the
Thomas Jefferson National Accelerator Facility (Jefferson Lab) to tour
the mile-long accelerator and several of its massive experimental halls
[LECTURES 8 AND 10–12].

You will discover how accelerators smash high-energy particles into nuclei
[LECTURE 8] to study nuclear structure, to make new ultraheavy nuclei, or
to recreate the conditions of the early universe by making a quark-gluon
plasma. You will learn how scientists use particle detectors to detect the
high-energy particles from these subatomic collisions [LECTURE 9] and
perform nuclear physics experiments with spectrometers (combinations
of magnets and particle detectors) that can measure these particles
[LECTURE 10]. You will explore a series of experiments that show how
protons and neutrons behave in the nucleus [LECTURE 11] and how the
protons and neutrons are each made of the 3 standard quarks plus the
host of virtual quark-antiquark pairs (known as sea quarks) and gluons
that account for about 98% of all ordinary mass [LECTURE 12].

Nuclear fusion powers the universe. Our Sun gets its energy by fusing
hydrogen into helium [LECTURE 13]. Heavier stars can get energy by
fusing lighter elements into heavier ones, culminating in iron. Even
heavier nuclei are made in stars, in supernovas, and even in neutron-star
collisions [LECTURE 14]. Scientists have used the elusive neutrino to look
deep into our Sun and into supernovas to show that supernovas emit an
astounding 99% of their energy in the form of neutrinos.

The process of nuclear fission is used to release energy. You will discover
why certain isotopes fission [LECTURE 15], how nuclear weapons work
and how to produce or concentrate those fissile isotopes [LECTURE 16],
how to generate electricity from nuclear fission [LECTURE 17], and the
causes and consequences of nuclear accidents [LECTURE 18]. Better ways to
produce nuclear energy can improve safety and reduce nuclear waste with
possible advanced nuclear power plants, including pebble-bed and liquid

2 Nuclear Physics Explained


fluoride-thorium reactors [LECTURE 19]. The ideal power supply of the
future could be controlled nuclear fusion [LECTURE 20], converting heavy
hydrogen to helium at extremely high temperatures and pressures.

But nuclear physics is not just about fission and fusion. Radioactive
sources and particle beams are used to fight cancer with radiation, such
as at the Hampton University Proton Therapy Institute, where scientists
attack tumors with proton beams [LECTURE 21]. Nuclear techniques are
also used in medical imaging, including x-rays and CT scans, PET and
SPECT scans, and MRIs [LECTURE 22].

The many different isotopes of each element, both stable and unstable,
can be used to date or identify objects. You will learn how radioactive
isotopes such as carbon-14 are used to date archaeological finds and how
stable isotopes are used to study Paleolithic diets [LECTURE 23]. Scientists
also use radiation to view the world around us, from placing nuclear
physics experiments down oil wells to guide drilling, to scanning cargo
containers for suspicious materials [LECTURE 24].

By the time you complete this course, you will appreciate the
glorious complexity of the atomic nucleus and its remarkable range of
applications, from powering the universe and creating the elements to
identifying fake vanilla, detecting art forgeries, and curing cancer. u

COURSE SCOPE 3
01 A TOUR OF THE
NUCLEUS AND
NUCLEAR FORCES

T
 e nucleus of an atom consists of protons
h
and neutrons, which are held together by
the strong nuclear force. That force is
100 times stronger than the electric force,
which holds a positive nucleus and the
negative electrons together in the atom.
Nuclear physics studies how the energy
changes when nuclei are combined and
when they are taken apart. This lecture will
offer a whirlwind tour of nuclear physics;
these topics will be explored again, step by
step, later in the course. u

4
The Chart of Nuclides
wwThe periodic table [PAGE 270] is great for organizing atoms. Simply count
the number of protons, and there will be the same number of electrons
if it’s a neutral atom. Each element corresponds to a specific number
of protons.

wwBut each element has more than one possible nucleus. These isotopes,
what physicists call nuclides, have a different number of neutrons.
For example, carbon has 6 protons. Carbon-12, with 6 protons and
6 neutrons, is the most common isotope. It’s the most stable one.
Carbon-14 still has 6 protons, but with 8 neutrons, it is famously
unstable. But carbon has isotopes ranging from carbon-8, with only
2 neutrons, to carbon-23, with 17 neutrons.

wwChemists are justly proud of the periodic table, with more than
100 elements. But nuclear physicists are even more proud of another
table, organizing more than 3000 isotopes into the chart of nuclides
[PAGE 274]. It shows all of the known nuclei, and more are added all
the time. The number of protons is on the y-axis, while the number
of neutrons is on the x-axis, and the nuclei appear in a band that runs
from the bottom-left of the graph to the top-right.

Visit https://www.nndc.bnl.gov/chart/ to see a


comprehensive and interactive chart of nuclides.

wwAbout 250 isotopes are so stable that they never change on their own.
They inhabit the valley of stability, and 80 of the familiar elements
have at least 1 member—1 stable isotope.

Lecture 1
| A Tour of the Nucleus and Nuclear Forces 5
wwIn addition to the stable nuclides,
another 80 or more nuclides found on
Earth decay naturally, some very slowly.
From 1901, Ernest Rutherford and
Frederick Soddy began to notice parts
of what we now call the thorium decay
chain. They found that thorium,
element 90, spontaneously
decays into radium,
element 88. Then, radium
spontaneously decays to
radon, which is element 86,
with 86 protons. And this
process continues. ERNEST RUTHERFORD
1871–1937
wwRutherford had already named
the 3 types of decay radiation,
even before there was any evidence that
elements were being transmuted. So, he called this alpha decay, in
which a nucleus gives up 2 protons and 2 neutrons. In beta decay, a
nucleus emits an electron and 1 of the neutrons that emits that electron
turns into a proton. In gamma decay, the nucleus emits a high-energy
photon, which is a high-energy particle of light.

wwThe geology of our planet depends on this spontaneous transmutation


of elements. Very slow decays of uranium, thorium, and even an isotope
of potassium provide the heat that keeps the Earth’s core molten so
that the Earth remains geologically active and the continents keep
drifting. This happens because these isotopes have very long half-lives,
comparable to the age of the
Earth. Other isotopes decay
much more quickly. Frederick Soddy coined the
word “isotope,” and H. G. Wells
dedicated his story about
“atomic bombs” to him.

6 Nuclear Physics Explained


wwIsotopes used in nuclear medicine diagnosis have half-lives that range
from a few minutes to several days. After 1 half-life, half of the original
isotope has transmuted into a different isotope. After 10 half-lives,
99.9% of the original isotope has transmuted into a different isotope.
This was originally quite shocking. Atoms had been thought the
permanent, fundamental building blocks of the universe.

wwWhen Marie Curie introduced the notion of radioactivity, she was


describing properties of a new element that she had discovered:
thorium. But she also showed that atoms can decay and therefore are
not the fundamental unit of matter.

wwThe 20th century was the first century of nuclear physics, beginning
with Wilhelm Conrad Röntgen’s discovery of x-rays in 1899 and
Rutherford’s discovery of the nucleus in 1911. Since then, dozens
of nuclear physics Nobel Prizes have been awarded in both physics
and chemistry (PAGE 317). Nuclear physics
developed only recently because the
nucleus is so much smaller: The
nucleus is 10,000 times smaller
than the atom.

wwIndividual neutrons and


protons, called nucleons,
are even smaller than the
nucleus. They are about
10−15 meters, which is
1 femtometer, called
1 fermi, in honor of
nuclear physicist Enrico
Fermi. Quarks and gluons
are at least 1000 times
smaller than that.
MARIE CURIE
1867–1934

Lecture 1
| A Tour of the Nucleus and Nuclear Forces 7
wwNuclei are small, but they contain
99.9% of the atom’s mass. They are If the nucleus is the
made of protons and neutrons, and size of a tennis ball,
the number of protons determines a then the outermost
chemical element. Chemists generally electrons orbit 5
don’t care about the number of neutrons. football fields away.
Chemists average the total number of
protons plus neutrons into something
called the atomic weight. Protons provide the electric field that keeps
the electrons in orbit, but the electric field pushes protons apart. What
holds the nucleus together?

wwAt small scales, we encounter the strong nuclear force. This affects both
protons and neutrons equally, and we need both protons and neutrons
to make the nucleus stable. The strong nuclear force is holding it
together while the repulsive electric force is trying to push it apart, and
the nuclei that result come from the interplay between these 2 forces.

wwThis interplay changes as nuclei


Both nuclear physics get larger. Lighter nuclei have
and particle physics are about equal numbers of protons
subatomic, and both use and neutrons, but the heavier
accelerators. But while nuclei have up to 1.6 times
particle physics (also more neutrons to make up for
called high-energy physics) the electrical repulsion of all
studies the particles those protons.
themselves, nuclear physics
studies how the particles wwIf we leave the valley of stability,
combine—how protons and there are 2 possibilities. There
neutrons combine to form may be too many protons,
the nucleus and how quarks resulting in too much positive
and gluons combine to form charge, so the nucleus needs to
the protons and neutrons. get rid of protons. It does this

8 Nuclear Physics Explained


by beta-plus (+) decay, in which the proton decays to a neutron, plus
a beta-plus particle—which is a positively charged electron called a
positron, which is the antiparticle of the electron—and a neutrino.
Alternatively, a proton can combine with an electron to become a
neutron plus a neutrino.

wwThe other possibility is there may be too many neutrons, which means
there are not enough protons, so the nucleus needs to get more protons.
It does this by beta-minus (β−) decay, in which the neutron decays to a
proton, plus an electron, plus an antineutrino.

wwNuclear waste is basically a problem of excess neutrons in the fission


products compared with the same number of neutrons in the original
uranium or plutonium. As the fission products decay, they change
neutrons to protons and move up and to the left on the chart of
nuclides toward the valley of stability.

Studying Nuclei
wwNuclear physics studies where nuclei come from. The nuclei from
hydrogen to helium were made in the big bang. The nuclei from helium
to iron are made in stars, and heavier elements from iron all the way up
were made in supernovas and neutron-star collisions.

wwThere are 3000 nuclides on the chart of nuclides but only a few
hundred that we’ve detected on the Earth. The other 90% of the
known nuclides have already vanished from Earth because they
decayed away. Their half-lives were too short, and we only know about
them because we’ve been able to create them in a laboratory. Nuclides
no longer found on Earth may turn out to have been very important to
our very existence.

Lecture 1
| A Tour of the Nucleus and Nuclear Forces 9
wwWe study nuclei the same way a 5-year-old studies an alarm clock:
We hit it hard and see what comes out. To hit it hard, we need a
very small—subatomic—hammer. And we use a variety of large
accelerators to accelerate different types of hammers, such as
electrons, protons, and other nuclei. Then, we use large particle
detectors to measure the individual subatomic particles that emerge
from these collisions.

wwWe reconstruct the subatomic collisions using energy and momentum


conservation, just as we reconstruct car crashes to find out who was
at fault. The difference is collisions between cars damage the cars,
sometimes severely, while subatomic collisions can make new nuclides.
We can knock nucleons out of a target to make smaller nuclei, or we
can combine 2 nuclei to make larger ones.

wwWe study nuclei because we want to learn how the world works.
Understanding nuclear physics leads to using external beams (of
x-rays or protons) or internal radioactive isotopes for medical imaging,
medical treatment (such as killing cancers), and even industrial,
archaeological, and art analysis.

wwAtoms are made of electrons orbiting the nucleus. Nuclei are made of
nucleons—neutrons and protons—orbiting each other. Nucleons, like
electrons, obey a fundamental rule: 2 identical particles cannot occupy
the same state at the same time. This is called the Pauli exclusion
principle. Because of this principle, electrons
occupy orbitals, such as s-shell and p-shell
orbitals, in atoms. Protons and neutrons The average
occupy similar s-shells and p-shells in nuclei, American pays $2
just in a different order. per year for basic
nuclear physics
wwJust like atoms, nuclei can be in excited research for a
states and then deexcite by emitting a combination of
photon, visible light for an atom, or a gamma basic knowledge
ray for a nucleus. and useful
applications.
10 Nuclear Physics Explained
The Fundamental Forces of Nature
wwNuclear physics makes use of all 4 of the fundamental forces of nature:
gravity, electricity and magnetism, the weak nuclear force, and the
strong nuclear force.

1 It may be surprising that gravity is important for nuclei. It’s the most
important force at large scales, but on small scales, it’s the weakest
force. It’s 10−20 times as strong as the electromagnetic force. Gravity
affects things with mass. It’s crucial for neutron stars, which are
the biggest nuclei there are. They have about 10% protons and 90%
neutrons. Their mass is 1 to 2 times the mass of our Sun, but the
radius is only about 10 kilometers, or 6 miles.

2 Electricity and magnetism cause the protons in the nucleus to repel


each other and attract the electrons to the nucleus to form atoms.
The electromagnetic force extends over the entire nucleus and is
responsible for gamma radiation.

3 The weak nuclear force affects all things. You might think that this
is a misnomer, because the weak nuclear force is stronger than gravity
at very short ranges, but it is called the weak nuclear force not in
distinction to gravity, but in distinction to the strong nuclear force.
The weak nuclear force is carried by massive particles called the
W and Z bosons, which each have 100 times the mass of a proton.
But because they’re so massive, the weak force has the smallest range
of all the forces: 1/500 the radius of a proton. The weak force does not
help hold the nucleus together; instead, it can change protons into
neutrons, or vice versa, through beta decay if the nucleus is on the
slopes of the valley of stability.

4 The strong nuclear force is the most important of the forces for this
course. At a fundamental level, the strong force holds quarks together
into protons and neutrons. This same strong force also “leaks out” of
the protons and neutrons to hold the entire nucleus together. Quarks

Lecture 1
| A Tour of the Nucleus and Nuclear Forces 11
have colors—red, green, and blue—instead of charges; therefore,
this force is called quantum chromodynamics. It is carried by
particles called gluons and is very strong. The neutrons and protons
are made of 3 quarks each (mostly). The full quark structure of the
nucleon is very complicated: There are 3 quarks, but there are also
quark-antiquark pairs. The strong force is about a million times
stronger than the weak force.

wwAtoms change in 3 fundamental ways: alpha decay, when the strong


force can’t hold the nucleus together; beta decay, which converts
protons to neutrons, or vice versa, via the weak force; and gamma
decay, which is a release of excess electromagnetic energy.

Supplements

READINGS

Lilley, Nuclear Physics, chap. 1.


Mackintosh, Nucleus.
National Research Council, Nuclear Physics, Overview, p. 9–30.
Wilczek, “The Origin of Mass.”

QUESTIONS

1 How would the Earth be different if the half-lives of the longest-lived


radioactive uranium, thorium, and potassium isotopes were only
100 million years, rather than more than a billion years?

2 Suppose that the electrical force was weaker than it is today. How
would this change which heavier nuclei are stable? Would it have a
similar effect on lighter nuclei?

12 Nuclear Physics Explained


ANSWERS

1 If these half-lives were much shorter, then all the isotopes would have
decayed away in the 4.5 billion years since the Earth was formed. There
would be no radioactive isotopes left in the Earth’s core or mantle to provide
a continued source of heat to keep the interior of the Earth hot. The core
would have cooled more quickly and might have become cold enough to
solidify. Without a molten core, the Earth’s magnetic field would be much
smaller, and much more radiation would reach the surface in the form of
cosmic rays. Continued radioactivity inside the planet helps protect against
cosmic radiation.

2 If the electrical force were weaker, then there would be less electrical
repulsion between protons. Therefore, heavy nuclei would need fewer extra
neutrons to be stable. In the extreme case, where the electrical repulsion is
negligible, heavy nuclei would have equal numbers of protons and neutrons,
just like light nuclei. Weakening the electrical force would have much
less effect on light nuclei, because the effect increases as the square of the
number of protons.

Lecture 1
| A Tour of the Nucleus and Nuclear Forces 13
02 CURVE OF
BINDING ENERGY:
FISSION AND FUSION

H
 w much energy does it take to break
o
a nucleus into smaller pieces? This is
measured by the binding energy, which
shows up as a measurable change in the
mass of the nucleus due to E = mc2. The
highest binding energy corresponds to
the most stable nuclei. Iron and nickel are
the most stable. This is different than the
atomic density, where uranium is denser
than lead, which is denser than iron. Iron
is very stable and is the natural endpoint
for nucleosynthesis in stars, so there is a
lot of iron in the cores of rocky planets. u

14
Nuclear Sizes
wwThe repulsion between atoms, or atom-atom repulsion—the force
that keeps you from sinking through the floor—is due to the
electromagnetic force and quantum mechanics. It’s due to the Pauli
exclusion principle and electron orbitals.

wwSimilarly, there is repulsion between protons and neutrons (nucleons).


Nucleon-nucleon repulsion is due to the Pauli exclusion principle
and quark orbitals. It gives us that the proton radius is a bit less than
1 fermi, or 10−15 meters.

wwNuclear sizes were first measured crudely by Ernest Rutherford in his


famous gold foil experiment with alpha particles, when he discovered
the nucleus. But he really just found that the nucleus is very small.
Now we can look at the diffraction patterns formed as we scatter
very-short-wavelength electrons from them.

wwHow does diffraction work? We pass a laser through a narrow aperture,


or slit, and look at the pattern of laser light on a screen. Diffraction
patterns are due to the wave interference of the light. The narrower the
slit, the wider the pattern. We can determine the width of the slit from
the wavelength of the light and the spread of the pattern on the screen.

wwHow does this apply to electrons? Electrons are particles. But they’re
quantum mechanical particles, which means that they are waves, too.
Like light, electrons travel as a wave but interact as a particle. Electrons
can be diffracted by nuclei, just as light is diffracted by narrow slits.

wwWe need a wavelength that is similar in size to the nucleus, which is


about 1 fermi. What energy is that? We need an energy of the electron
that’s 2000 times the mass energy of the electron. How can the energy
be greater than mc 2? If the electron’s not moving, then its mass energy
is mc 2. If it is moving, now it has both kinetic energy and mass energy.

Lecture 2
| Curve of Binding Energy: Fission and Fusion 15
And if the particle is moving at close to
Because of special
the speed of light (c), the total energy—
relativity and
mass plus kinetic—can be much greater
E = mc2, nuclear
than just the mass energy.
physics uses special
units for mass and
wwShort-wavelength, high-energy electrons
passing around nuclei make diffraction
for energy.
patterns similar to light passing through For energy, we
a narrow slit. And just like light passing use the electron
through a slit, we can determine the size volt (eV), which
of the nuclei by using the wavelength of is the charge of
the electrons, which we know, and the 1 electron or 1
width of the diffraction pattern, which proton passing
we can measure. The diffraction pattern through a potential
turns out to be narrower for heavier difference of 1 volt.
nuclei, which means that they’re bigger.
So, we’re going to
wwWe use the entire measured pattern, measure energy in
not just the minima, and some electron volts and
mathematical techniques to get the mass in electron
complete charge distribution—how volts divided by
much charge there is in each radius the speed of light
in the nucleus—and to figure out the squared—eV/c2—
precise charge radius. because E = mc2.

wwWhat do we learn from these nuclear Electron volts is


charge distributions and radii? All nuclei not always the best
have about the same density—number unit. Sometimes we
of protons and neutrons per cubic want thousands of
fermi—in the center. The difference is electron volts (keV),
how big the central region is. Nuclear millions of electron
sizes increase much more slowly than volts (MeV), or
the number of nucleons with mass. billions of electron
volts (GeV).

16 Nuclear Physics Explained


wwFor nuclei heavier than carbon, the volume increases as the number
of nucleons. The radius increases as the cube root of the number of
nucleons. This tells us that nucleons are incompressible. They stack like
marbles, and we can picture nuclei crudely as balls of marbles.

wwHow does this compare to atoms? Electrons, as far as we know, have


zero size. They’re point particles, so you can’t stack them like marbles.
The distance of the closest electron to the nucleus is approximately
1/Z (the number of protons). The distance of the farthest electron is
approximately constant, because the other electrons screen out all but
1 proton’s worth of that nuclear charge.

wwThe atomic size depends mostly on which column of the periodic


table the atom is in, rather than how many electrons it has. On the
other hand, the nuclear size depends primarily on mass, or the number
of nucleons.

Making Nuclei from Protons and Neutrons


wwNow let’s stick the marbles together and make nuclei from protons and
neutrons. Which nucleon properties are important?

§§ The mass of a proton or neutron is about 1 billion electron volts


(GeV) divided by the speed of light squared: 1 GeV/c 2. Technically,
the proton mass is 0.9383 GeV/c 2, while the neutron mass is slightly
more, 0.9396 GeV/c 2. It’s a small difference, but it is important for
nuclear decay.

§§ The spin, which is a kind of angular momentum, of a proton or


neutron is 1/2, and it can either be spin up (+1/2) or spin down (−1/2).
The Pauli exclusion principle tells us that we can’t have 2 protons in
the same state, but spin allows 2 protons or neutrons in each state:
1 up and 1 down.

Lecture 2
| Curve of Binding Energy: Fission and Fusion 17
§§ The radius of a proton or neutron is a bit less than 1 fermi—about
0.9 fermi.

wwThe nuclei are bound by the interplay of the very-short-range attractive


strong nuclear force—which has a range of only 1 or 2 fermis, so only
nearby nucleons feel the force—and the weaker long-range electrical
repulsion, which increases quadratically with proton number because
every proton repels every other proton, so the number of repelling
pairs is the number of protons times the number of protons minus 1:
Z(Z − 1).

wwThe strong force—sometimes described as a potential energy—is


not a fundamental force. It is determined from measurements:
proton-proton and proton-neutron collisions as well as the binding
energy of deuterium.

wwThe force between nucleons is very similar to the force between people:
When we’re far away from each other, there’s no interaction. When we
can get close, we can be attracted, and if we get too close, there can be
a very strong repulsion. This is a complicated interaction. But there
is a very small difference between the force on protons and the force
on neutrons—called the isospin terms—so we can treat protons and
neutrons almost identically.

wwNuclei prefer even numbers of protons and of neutrons, and they


also prefer relatively equal number of protons and neutrons. The
electromagnetic force is mostly the repulsion between the protons.

wwWith bigger nuclei, the nucleons feel that strong-force attraction only
from their nearest neighbors. But the protons feel that electromagnetic
repulsion from all the other protons, which means that bigger nuclei
have proportionately more neutrons and are also less stable.

18 Nuclear Physics Explained


Binding Energy
wwThe binding energy is the energy
Batteries change
released when the nuclei or atoms
mass as they charge—
are formed. It’s the energy that must
because they’re gaining
be supplied to disassemble the nuclei
energy (E = mc2), so
or atoms. For nuclei, the binding
they’re also gaining a
energy is the difference in mass energy
little bit of mass—but
between the total proton masses,
the difference is too
plus the total neutron masses, minus
small to measure.
the mass of the nucleus, times c 2:
[Zm(1H) + Nmn − m(nucleus)]c 2, so
we have E = mc 2. This is also true of
atomic and molecular binding energies, but those binding energies
are much less than the mass of the electron. Nuclear binding energies
are about 1% of the mass of the proton, so we see about a 1% change
in mass.

wwWe measure binding energy indirectly, using a mass spectrometer to


measure the mass precisely. We make an ion beam by stripping off the
electrons from the atom, because then it’s charged and much easier to
accelerate. Then, we pass the atom through a velocity selector, which
has an electric field and a magnetic field. The electric field makes the
ions want to go up, while the magnetic force makes the ions want to go
down. Only ions with just the right velocity—equal to the ratio of the
electric and magnetic fields—go straight. The next step is to pass these
ions with this specific velocity through a magnetic field.

wwThe magnetic field makes the ions go in circles, and by measuring the
radius of the circle, we can measure the mass of the particle. The mass
is proportional to the radius divided by the velocity: m ∼ R/v.

wwWe also use magnetic fields to steer particles in accelerators and to


measure the momentum of particles knocked out in nuclear collisions.

Lecture 2
| Curve of Binding Energy: Fission and Fusion 19
wwWe measure the radius of curvature by where the nucleus hits the
detector. Usually, we also measure a known element so that we can get
a more precise measurement. We make a spectrum and measure the
masses by the locations of the peaks in the spectrum, and we measure
the abundance of the different isotopes by the heights of those peaks.

wwWe can also use a mass spectrometer to separate isotopes—to get a pure
sample of an isotope.

wwNow that we can measure binding energies, we can plot the binding
energies of the most stable isotope of each element. We plot the binding
energy against the number of nuclei and get the curve of binding
energy. Note that there’s a very steep rise for light nuclei and a slow
decrease for heavy nuclei.

20 Nuclear Physics Explained


wwNuclei are made by smashing, or fusing, smaller nuclei—especially
protons and sometimes helium-4—together. The big bang made
hydrogen-2 (deuterium), helium-3, and helium-4. Helium, carbon,
nitrogen, and oxygen are made in small stars. Larger stars make the
elements up to iron, and supernovas or neutron-star mergers make the
heavier elements.

wwThe shape of the curve of binding energy is determined by the interplay


of electric repulsion and the strong nuclear force. The implication is
that energy can be released by fusing lighter elements into heavier ones
(up to iron). In the big bang, hydrogen was fused into heavy hydrogen
(hydrogen-2, or deuterium), helium-3, and helium-4, and fusion provides
the power source of stars. Energy can be released by fissioning heavier
elements into lighter ones (down to iron), but that also implies that
normal stars can’t make heavier elements. We need other mechanisms.

wwThe shallowness of the curve of binding


energy for very heavy nuclei means
that we get less energy per nucleon
from fission, and the steepness of
the curve of binding energy for
very light nuclei means that
we get much more energy
per nucleon from fusion.
Overall, the shape of the
curve of binding energy
determines how we can
get energy from nuclei,
whether fission from
heavy nuclei or fusion
from light nuclei.

The heaviest naturally abundant nucleus is uranium. Uranium-238


has a half-life of 4.5 billion years—about the age of the Earth.

Lecture 2
| Curve of Binding Energy: Fission and Fusion 21
Supplements

READINGS

Lilley, Nuclear Physics, chap. 1.


Mackintosh, Nucleus, chaps. 4, 6, and 8.
Wiringa, “Evolution of Nuclear Spectra with Nuclear Forces.”

QUESTIONS

1 We measure the nuclear binding energies of different isotopes by


measuring their masses precisely and then converting the mass
differences to the binding energies by using E = mc 2. Why don’t we
apply this same technique to measure the atomic (i.e., chemical)
binding energies of different elements?

2 How much more energy can we gain by fusing 1 gram of hydrogen


into helium than by fissioning 1 gram of uranium into lighter nuclei?
What about fusing 4 hydrogen nuclei into 1 helium nucleus versus
fissioning 1 uranium nucleus?

ANSWERS

1 Nuclear binding energies are much larger than atomic binding energies.
We can directly measure the masses of the different nuclei, from the proton
(the nucleus of the hydrogen atom) to uranium and beyond. The difference
between the mass of a nucleus and the total mass of the protons and
neutrons that comprise it is about 1%, which is easily measurable. Atomic
binding energies are thousands to millions of times smaller than nuclear
binding energies. The mass differences corresponding to these tiny energies
are difficult to impossible to measure directly.

22 Nuclear Physics Explained


2 We can release about 7 MeV of energy per nucleon by fusing hydrogen into
helium but only about 1 MeV per nucleon by fissioning uranium into lighter
nuclei. One gram of hydrogen is 1 mole, so it contains 6 × 1023 protons. One
gram of uranium contains a lot fewer uranium nuclei, but it also contains
about 6 × 1023 protons and neutrons. Therefore, fusing 1 gram of hydrogen
will release about 7 times more energy than fissioning 1 gram of uranium.

However, fusing 4 hydrogen nuclei into 1 helium nucleus will release


4 × 7 MeV = 28 MeV of energy, but fissioning 1 uranium nucleus will
release about 238 × 1 MeV = 238 MeV of energy.

Thus, nucleus for nucleus, uranium releases a lot more energy, but gram for
gram, hydrogen releases more energy.

Lecture 2
| Curve of Binding Energy: Fission and Fusion 23
03 ALPHA, BETA, AND
GAMMA DECAY

T
 e term “radiation” may sound scary, but
h
it refers to anything emitted—or radiated.
We really only worry about radiation that
breaks chemical bonds, called ionizing
radiation. Radiation in the broader sense
includes sound waves, gravitational waves,
fast-moving subatomic particles from
nuclear decay (alpha and beta particles and
gamma rays), cosmic rays (mostly muons,
the heavy cousins of the electron), particles
emitted at accelerators and nuclear
reactors, and all the other electromagnetic
waves that have lower energies than x-rays
and gamma rays. u

24
Electromagnetic Waves and Radioactivity
wwElectromagnetic waves are all basically the same; only the wavelength
varies. They all travel like waves and interact like discrete particles.
The energies of these particles, called photons, go as 1 over the
wavelength: E = hc/𝜆. Long-wavelength electromagnetic waves have
very-low-energy photons; short-wavelength electromagnetic waves have
very-high-energy photons.

wwElectromagnetic waves include radio waves, microwaves, and visible


light—all of which have very-low-energy photons. Ultraviolet radiation
has wavelengths from 10 to 400 nanometers; the typical ultraviolet
photon energy is greater than 3 electron volts. X-rays have very short
wavelengths on the order of nanometers; these photons have energies
from about 1 to 300,000 electron volts. Gamma rays have even smaller
wavelengths than x-rays; they have wavelengths that are shorter than
1/100 of a nanometer and energies above 300,000 electron volts.

The typical ultraviolet photon energy is greater than 3 electron


volts, while the typical energy of a chemical bond is only about 1
electron volt, so ultraviolet photons are energetic enough to break
chemical bonds and damage molecules. This is called sunburn.

Lecture 3
| Alpha, Beta, and Gamma Decay 25
wwWe worry about ionizing radiation. All radiation interacts in matter.
When ionizing radiation interacts, it deposits enough energy to break
chemical bonds. This weakens materials and damages DNA. Ionizing
radiation includes x-rays, gamma rays, ultraviolet light, and fast-moving
subatomic particles.

wwRadioactivity is measured in disintegrations per time. There are


2 units for this measurement: the metric system unit, which is the
becquerel, which is 1 disintegration per second; and the older unit that
is still in wide use, called the curie, which is 40 billion disintegrations
per second.

wwRadioactive materials emit radiation via nuclear decay. Radiation is


measured in particle flux: the number of particles per time, or the
number per area per time. That is measured with a Geiger counter,
which gives us counts per minute, which tells us the number of
disintegrations per minute.

wwRadiation is also measured in the amount that is absorbed in the


exposed material. This is measured in terms of energy, in joules per
kilogram. The unit for that is the gray, which is a metric system unit
that is equal to 1 joule absorbed in a
kilogram. The older unit, the rad, is
100 times smaller, so 100 rads would be Whereas x-rays were
1 joule per kilogram of deposited energy. called “x” because
they were new and
wwThere is not much energy in a joule unknown, gamma
per kilogram. It is enough energy rays were called
to lift 1 kilogram (a few pounds) by “gamma,” the third
10 centimeters (or 4 inches). If it were letter of the Greek
turned to heat, it would be less than a alphabet, because
1000th of a degree. But it can break a lot they were the third
of chemical bonds, so it can do a lot of type of radiation
damage to the molecules in that material. discovered given off
by radioactive decay.

26 Nuclear Physics Explained


wwDifferent kinds of radiation have different biological effects. To
measure the biological effects, we need to convert grays and rads,
which are in joules per kilogram, to rems and Sieverts. We multiply
the number of rads by a factor to get rems or the number of grays by a
factor to get sieverts. The factor we use for beta and gamma radiation
is 1. Alpha particles have a factor of 20; they deposit a lot of energy in
a very small distance, so they break a lot of chemical bonds in a small
region. Neutrons and protons are somewhere in between.

The banana equivalent dose is the amount


of radiation that you get from eating a
banana. It is 1/10 of a microsievert, or
about 10 microrems. This is not used
seriously, but it is used to show that
there is radiation in the environment and
that it does give us radiation.

Nuclear Decay
wwHalf of the nuclei in a sample decay in 1 half-life. This is a statistical
process; it’s impossible to predict which specific nuclei will decay. If
we just have 1 nucleus, we can’t even tell when it will decay. A short
half-life means that the nuclei are decaying very quickly and the
material is very radioactive, but it is not going to stay radioactive for
long. If there is a long half-life, the material is not very radioactive and
will decay slowly; it will stay radioactive for a much longer time.

Lecture 3
| Alpha, Beta, and Gamma Decay 27
wwDifferent isotopes have different half-lives. Nuclei that have too
many protons or too many neutrons are moving away from the valley
of stability.

wwThere are 3 main types of nuclear decay: alpha, beta, and gamma.
All of these are emitted by radium and its decay products. They all
behave differently in a magnetic field. When you pass them through a
magnetic field, alpha particles are distributed one way, beta particles
are deflected the other way, and gamma rays are not deflected—they
go straight through. This tells us that alpha and beta particles have
opposite charges and gamma rays are uncharged. Fission is a completely
different process, and it’s much rarer.

wwThe alpha particle, which is the nucleus of a helium atom, consists


2 protons and 2 neutrons very tightly bound together. That means
that when you emit an alpha particle, because it conserves protons and
neutrons, the total number of protons and the total number of neutrons
have to be the same before and after. The daughter nucleus will have
2 fewer protons and 2 fewer neutron, so it will move 2 down and 2 to
the left, on the chart of nuclides [PAGE 274].

wwWhy does the alpha particle decay and not emit a proton? Heavy
nuclei are bound by about 8 million electron volts per nucleon, so to
emit a proton, you have to find about 8 million electron volts. That’s
difficult. The alpha particle is already bound by 7 million electron
volts per nucleon, so it is much easier to find the energy to emit an
alpha particle.

wwAlpha decay is due to the competition between electric repulsion and


strong force attraction. Alpha decay conserves charge, so there are
the same number of positive and negative charges before and after. It
also conserves energy, so the difference in the binding energy, or mass
energy, goes into the kinetic energy of the fragments.

28 Nuclear Physics Explained


wwThe difference in the binding energy is the mass of the parent nucleus,
minus the mass of the daughter, minus the mass of the alpha particle,
times c 2: Q = (m A − mB − m𝛼)c 2. The bigger this difference is, the
shorter the half-life. These differences range from 4 to 10 million
electron volts, giving us half-lives ranging from 10 gigayears to
100 nanoseconds.

wwAlpha decay also conserves momentum. There are only 2 particles in


the final state: the daughter nucleus and the alpha particle. They have
to have equal and opposite momenta. The alpha particle carries most
of the kinetic energy, so it only has a single energy from a given decay.
This energy is used to measure the nuclear mass differences between
the parent nucleus and the daughter nucleus.

wwAlpha decay is due to quantum tunneling. It’s classically forbidden


because the alpha particle is attracted to the center of the nucleus, and
the electrical repulsion keeps it inside the nucleus. The alpha particle is
quickly moving around and hitting the barrier of the nucleus at about
1021 times per second until it quantum mechanically tunnels out.

wwInversely, if we aim an alpha particle at a nucleus and it doesn’t have


enough energy to overcome the electrical repulsion between the
positively charged alpha particle and the positively charged nucleus,
then it has to quantum mechanically tunnel in. And that, too, is
unlikely. It takes a while to happen.

wwHow does tunneling work? We have a region where the wave function
is decreasing exponentially because there is so much repulsion between
the alpha particle and the rest of the protons, but we don’t have the
corresponding strong force attraction. The probability that the alpha
particle is in the forbidden region decreases by a factor of 2 every
0.5 fermi. So, a small change in the energy of the alpha particle can
make a huge change in the half-life of the nucleus.

Lecture 3
| Alpha, Beta, and Gamma Decay 29
wwOn the chart of nuclides, where the number of protons is plotted
vertically and the number of neutrons is plotted horizontally, the stable
isotopes are in black. The isotopes shown in yellow decay by alpha
decay. They are heavier and have more protons.

wwBeta radiation is caused by the weak nuclear force. There are 2 kinds of
beta decay. If the particle is to the right or below the valley of stability,
then it’s going to have too many neutrons. The neutrons are going
to decay to a proton, an electron, and an antineutrino. This is called
beta-minus decay. As a neutron turns into a proton, the particle will
move diagonally up and to the left on the chart of nuclides.

wwIf, on the other hand, the particle started with too many protons—
meaning that it’s to the left or above the valley of stability on the
chart of nuclides—then the proton is going to decay to a neutron plus
a positron and a neutrino. This is called beta-plus decay or positron
emission. This particle will move diagonally down and to the right on
the chart of nuclides as the proton turns into a neutron.

wwAnother way to change protons to neutrons is by electron conversion.


When there are too many protons, a proton and an electron can
combine via the weak force to make a neutron plus a neutrino. This
particle moves one box diagonally down and to the right, and the total
number of protons and neutrons is unchanged.

wwThe weak force conserves energy, momentum, charge, total number


of protons and neutrons (but can change neutrons to protons), and
number of electrons. In fact, it conserves the number of electron-like
things called neutrinos that count like electrons. The number of
electrons plus the number of neutrinos minus the number of positrons
minus the number of antineutrinos is unchanged.

30 Nuclear Physics Explained


wwThe positron, or beta-plus particle, is the antiparticle of the electron,
and it leaves the atomic weight unchanged. It makes an electron or a
positron. If it makes an electron, it has to make an antineutrino, and if
it makes a positron, it has to also make a neutrino.

wwA neutrino has a tiny mass and no charge. Its existence was inferred
from the fact that the energies of beta decay have a continuous
spectrum of energy, unlike the alpha particles from alpha decay, that
just have 1 energy. The alpha particle has a single energy because there
are only 2 particles in the end: the daughter particle and the alpha
particle. The electrons have a continuous spectrum because there have
to be at least 3 particles involved. That third missing particle was
hypothesized to be the neutrino. It was discovered a few decades later.
We can also use the maximum energy of the electron to measure the
nuclear binding energies.

wwWeak decay is described by Fermi theory. There is no tunneling


involved; it is just very weak. The probability has to do with the
weakness of the weak force plus the similarity of the initial and final
states of the nucleus.

wwGamma rays, or photons, are due to the electromagnetic force. There


is no change in the number of neutrons, the number of protons, or the
total number of nucleons. Most alpha and beta decays leave an excited
daughter nucleus, which deexcites via gamma emission. The energies
range from about 0.1 to 10 million electron volts (10 MeV).

wwThere are discrete energies that gamma rays give off that correspond
to specific nuclei and specific nuclear states in those nuclei. This is just
like an atom (when an electron changes orbit, it emits photons of a very
specific wavelength of colors) or a nucleus (when a neutron or proton
changes its orbit, it emits photons of a very specific energy).

Lecture 3
| Alpha, Beta, and Gamma Decay 31
wwBlocking alpha and beta radiation from entering your body is not
difficult and makes a big difference. Gamma radiation is always much
more difficult to shield against. You have to have a barrier, such as lead,
to stop the gamma rays or they will hit you.

Supplements

READINGS

Jorgenson, Strange Glow.


Lilley, Nuclear Physics, chap. 3.
Mackintosh, Nucleus, chap. 2.
Rhodes, The Making of the Atomic Bomb.

QUESTIONS

1 Radon is harmful, but


how does it decay? The
isotope commonly found
in basements and mines is
radon-222 (with 86 protons
and 136 neutrons), which
decays with a half-life of
3.8 days to polonium-218
(with 84 protons and
134 neutrons). Which decay
process is responsible? And
how difficult is it to shield
against the decay products?

2 Radon-223 (with 86 protons and 137 neutrons) decays by beta-minus


decay. What is its daughter nucleus?

32 Nuclear Physics Explained


3 Fluorine-18 has 9 protons
and 9 neutrons. It decays by
beta-plus decay. What is its
daughter nucleus?

4 Which is more dangerous:


an isotope with a short
half-life (e.g., carbon-11, with
a half-life of 20 minutes) or
with a long half-life (e.g.,
carbon-14, with a half-life of
almost 6000 years)?

ANSWERS

1 Because polonium-218 has 2 fewer protons and 2 fewer neutrons than


radon-222, the decay occurs by emitting an alpha particle. Alpha particles
have a very short range and are easily blocked by a layer of paper or
dead skin. However, because radon is a gas, it can be inhaled, or it can
contaminate a groundwater drinking supply. Once the radon is inside the
body, the alpha particles will damage living tissue. Polonium-218 itself
decays (with a half-life of 3 minutes) to lead-214, which also decays, leading
to additional alpha and beta radiation. Radon-222 is the most common
isotope of radon (the one that can accumulate in basements and in mines)
because it is part of the uranium-238 decay chain.

2 The beta-minus decay will convert 1 neutron to 1 proton, emitting


an electron (the “beta minus”) and an antineutrino. Therefore, the
daughter nucleus will have 1 more proton and 1 fewer neutron, giving
87 protons, 136 neutrons, and the same 223 total nucleons. This nucleus is
francium-223.

Lecture 3
| Alpha, Beta, and Gamma Decay 33
3 The beta-plus decay will convert 1 proton to 1 neutron, emitting a positron
(the “beta plus”) and a neutrino. This will give a daughter nucleus with
8 protons and 10 neutrons. This nucleus is oxygen-18. Fluorine-18 is used
for PET (positron-emission tomography) scans in nuclear medicine.

4 The material with the short half-life will be much more radioactive to begin
with. For example, carbon-11 will be 100 million times more radioactive.
However, the material with the short half-life will decay much faster,
and its radioactivity will decrease much faster. For example, after 1 hour
(3 half-lives), there will be 23 = 8 times less carbon-11. After 1 day, there will
be 824 = 1021 times less carbon-11, so it will be negligibly radioactive. The
carbon-14 will be 100 million times less radioactive than the carbon-11, but
it will remain radioactive for 100 million times longer.

34 Nuclear Physics Explained


RADIATION
SOURCES, NATURAL
04
AND UNNATURAL

T
 ere are 3 main sources of radiation:
h
terrestrial radiation, which comes from
the decay of uranium and thorium, so is
primarily natural; cosmic rays, which come
from particles such as muons; and medical
radiation, such as x-rays, CT scans, and
nuclear medical procedures. The average
dose of radiation that we get is about
600 millirem, or 6 millisieverts, per year.
About half of that is natural background,
about half is medical, and a few percent
come from consumer products. About
2/3 of the natural background radiation
comes from radon in the air. The other
1/3 comes from food and drink, including
bananas; terrestrial radiation, usually from
the uranium in the granite around us; and
cosmic rays, which increase with altitude. u

35
The presence and
intensity of radiation
can be measured with
a Geiger counter.

Terrestrial Radiation
wwTerrestrial radiation comes from the decay of uranium and thorium.
Uranium-238 has a 4.5-billion-year half-life and decays by alpha
decay. Thorium-232 has a 14-billion-year half-life and also decays by
alpha decay.

wwUranium and thorium are present in many rocks. There is about


6 parts per million of thorium and 3 parts per million of uranium,
so that gives us about 10 or 20 grams of thorium or uranium in every
cubic meter of rock. These are about the only stable elements that are
heavier than lead. They’re not really stable, but they’re stable enough—
they have half-lives of billions of years. Their concentration is about
a microcurie per cubic meter, which means we get very low radiation
levels from thorium and uranium in granite.

36 Nuclear Physics Explained


wwWith thorium and uranium, there is a long chain Radiation
of alpha and beta decays—turning neutrons to from cosmic
protons and moving down and to the right on the rays is 50%
periodic table—that includes radon and terminates greater in
in lead for both. In both decay chains, about Denver than
50 million electron volts of energy is released in the at sea level.
form of radiation.

wwThe radioactive decay of uranium and thorium powers Earth’s rock


cycle, moves continents, and helps the core stay molten. How do we
know this? First, we’ve dug deep holes, called bore holes, into the
ground and measured how the temperature changes, and from that,
we calculate that the core emits about 46 terawatts, or trillion watts,
of power. In addition, we can calculate the expected decays from the
abundances of those nuclei.

wwWe can also detect geoneutrinos from uranium and thorium decay,
which tells us that there’s about 20 terawatts of uranium and thorium
decay—specifically, uranium-238 and thorium-232. There are also
uranium-235 and potassium-40 decay, which give 4 more terawatts of
energy, and 22 terawatts come from the residual heat of the core. The
potassium-40 in Earth’s core is what helps keep it molten and creates a
magnetic field, which is crucial for cosmic-ray shielding.

wwWhat makes radon special? Why do we get so much of our radiation


from radon when it’s just one of the uranium decay products? The
other decay products are solids and stay put in the original rock. Radon
is a noble gas—so it doesn’t combine with anything—and it can
migrate and concentrate in mines and in your basement.

wwIn addition, because it’s a gas, we inhale it, which means that its alpha
particles can do damage. The average dose of radiation from radon in
the United States is about 230 millirem per year. Also, radon dissolves
in water, so we can inhale radon gas when showering. And the decay
products of radon become dust and then become inhalable.

Lecture 4
| Radiation Sources, Natural and Unnatural 37
wwThe concentration of radon varies widely with region and geology. It is
greater in hilly and mountainous terrain. It’s also 8 times less dangerous
for nonsmokers than for smokers. And it’s not present in submarines.

wwHow else do uranium and thorium irradiate us? The decay products—
the elements in the uranium and thorium decay chains—in rock emit
gamma rays. Alpha and beta particles can’t escape the rock, but we do
get about 21 millirem per year average in the United States. This is
widely variable and ranges from about 10 to 100 millirem per year.

Cosmic Rays
wwDid you know that hundreds of cosmic rays pass through you every
second? If you hold out your hand, there’s 1 cosmic ray passing through
your hand every second.

wwCosmic rays start out as high-energy nuclei—mostly protons (90%),


but some alpha particles (9%)—that hit the upper atmosphere. These
particles react with air nuclei, making showers of particles, including
pions, protons, and neutrons.

wwThe protons, neutrons, and pions interact via the strong nuclear force,
so they interact a lot. And they don’t reach the ground. We don’t care
about the neutrinos; they’re not going to interact. The particles that are
left are muons, which come from the decay of pions and are the heavy
cousins of the electron. They have a positive charge and a negative
charge, but they have a very short lifetime: only 2.2 microseconds, or
millionths of a second.

wwIn addition, when these high-energy cosmic rays interact with the
atoms in the atmosphere, they can change them. For example, they
can interact with nitrogen-14 and turn it into carbon-14. They can also
make other radioisotopes in the atmosphere.

38 Nuclear Physics Explained


wwThe muons are the only ones that reach the ground. How many muons
are there? There are 104 per square meter per second, which is about
1 per hand per second, mostly coming down from above. For our whole
body, we get about 104 per second, which is comparable to what we’d
get if we held a microcurie radioactive source in our hand.

wwWhy do we care that the Earth has a magnetic field? It deflects cosmic
rays from hitting the atmosphere. That means that fewer of these
high-energy charged particles actually hit the Earth’s atmosphere.
Instead, they form the Van Allen radiation belts, which are about
600 to 60,000 kilometers from the Earth and are due to cosmic rays
being trapped by the magnetic field. The Van Allen belts are dangerous
to humans and satellites.

wwAt the North and South Pole, the magnetic field is almost vertical,
which means that cosmic rays can reach the surface of the Earth much
more easily in these areas than at the equator. That’s why we have the
aurora; the high-energy charged particles excite atoms in the air and
make those colors.

wwAstronauts that are outside the Earth’s magnetic field see flashes of
light associated with cosmic rays passing through their eyes. We don’t
know exactly how this works, but it’s definitely seen.

The amount of medical


radiation has increased
dramatically, although it
is variable from person
to person. In 1987, only
15% of the radiation we
received was from medical
purposes, such as CT scans
and x-rays; by 2017, this
number was close to 50%.

Lecture 4
| Radiation Sources, Natural and Unnatural 39
Other Radiation Sources
wwUranium fission, not just decay, creates very radioactive by-products.
Radiation exposure from bomb test fallout and from nuclear power
plants is less than 1% of our exposure today. Bomb test fallout decays
and leaves the atmosphere, and nuclear power plants in normal
operation emit almost no radiation.

wwSome consumer products contribute to our radiation exposure. For


example, smoke detectors contain americium, which is an alpha
emitter used to ionize the air to help us measure whether there’s smoke
present. And alpha particles have a very short range; they don’t escape
the container.

wwAnother source of radiation is gaseous tritium light sources. Tritium,


which is heavy hydrogen (with 1 proton and 2 neutrons), has about
a 10-year half-life. A gas tube containing tritium is coated on the
inside with a phosphorescent chemical. (These are the same kinds of
phosphors that were used in old cathode ray television sets.) The beta
decay from tritium (electron) hits phosphor, which emits light. It’s
only an 18-kilovolt electron, so it can’t penetrate the tube. The color
of the light has nothing to do with the radiation; it’s due to the choice
of the phosphor. It’s used in exit lighting, watch dials, compasses, and
gun sights.

wwThere are also some old products that contain radiation. Orange
Fiestaware glaze, used on dinnerware, contains uranium oxide. The
uranium for this was confiscated in World War II, so it stopped
being made during the war. In 1959, producers switched to depleted
uranium, so it’s much less radioactive, but it still contains uranium. The
old product is much more radioactive.

wwRadiation used to be good for us—or that’s what we thought—so


it was used in a number of ways. From the 1920s to the 1950s, shoe
fluoroscopy was used in shoe stores. X-rays were shined through shoes

40 Nuclear Physics Explained


and feet, onto a fluorescent screen, which was much less sensitive than
film. A shoe fluoroscope had 3 viewing ports so that the salesperson
and customer could directly see how well the shoe fit: how the bones
of the feet fit inside the shoe. The problem is, for a 20-second view,
you got a radiation dose of about 10 rem—a huge dose. There were no
reported injuries to customers, but there was at least one shoe model
who needed her leg amputated.

wwIn addition, radium-infused carbon dioxide cartridges could be used to


make soda at home. There was also radithor, which is radium dissolved
in water. A case of bottles sold for 10 times the cost of radium.

wwRadium is known as a bone seeker, or an element that accumulates


in the bones when added to the body. It was thought to be beneficial,
and warnings were disregarded. Fortunately, about 95% of the radium
products were frauds; only the wealthy could afford the real stuff.

wwRadioactivity was also used to power pacemakers. They used the


heat from radioactive decay to generate an electric current using a
thermocouple. They used plutonium-238, an isotope of plutonium with
a half-life of 90 years. It emits a 5-million-electron-volt alpha particle
and puts out about half a watt per gram of plutonium. Radiation
can’t penetrate the pacemaker, so there’s no danger from the radiation
leaving the pacemaker.

wwThe plutonium pacemaker was used because it never needs new


batteries, because the half-life of plutonium is longer than the
half-life of the patient. It used a lot of plutonium: 5 curies per patient.
Plutonium was used in pacemakers from 1973 to 1988, when people
realized that cremating radioactive pacemakers would be bad and when
10-year lithium batteries became available.

wwThe same technology is still used to power remote lighthouses and


NASA deep space missions. Pioneer, Voyager, Cassini, and Viking all
used radioisotope thermal generators to generate long-term power.

Lecture 4
| Radiation Sources, Natural and Unnatural 41
Supplements

READINGS

Bolus, “NCRP Report 160 and What It Means for Medical


Imaging and Nuclear Medicine.”
Jorgenson, Strange Glow.
National Council on Radiation Protection and Measurements,
Report No. 160.
National Research Council, Health Risks from Exposure to Low
Levels of Ionizing Radiation.
Rhodes, The Making of the Atomic Bomb.
The KamLAND Collaboration, “Partial Radiogenic Heat Model
for Earth Revealed by Geoneutrino Measurements.”

QUESTIONS

1 The average background radiation received in the United States is


about 600 millirems (mrem) per year. Why is this average number
misleading for calculating your own exposure?

2 How much more radiation would we receive if we ate all of our meals
on prewar orange Fiestaware?

ANSWERS

1 Half of this average number (300 mrem/year) comes from medical radiation
used for diagnostic purposes. If you do not have any x-rays, CT scans, or
PET or SPECT scans, then your radiation dose is only 300 mrem/year.
If you do get a CT, PET, or SPECT scan, then you can receive from
100 to 1000 mrem per scan. In addition, 2/3 of nonmedical radiation
(200 mrem/year) comes from radon (mostly radon trapped in basements).
This also varies dramatically with location—from almost nothing to

42 Nuclear Physics Explained


1000 mrem, depending on soil composition and basement ventilation. Thus,
if you live in a low-radon area and did not receive medical radiation, your
yearly radiation dose could be as a low as 100 mrem.

2 When we measured the radioactivity of the Fiestaware plate in the


video lecture, we saw that the Geiger counter counted about 100 times
faster than for cosmic rays. We receive about 30 mrem per year from
cosmic rays. We will get about 100 times more than that from the
Fiestaware, but only the parts of our body close to the plates will get
irradiated (about 20%), and we are only near it during meals (i.e., about
10% of the time). Thus, the whole-body radiation dose would be about
30 mrem/year × 100 × 0.2 × 0.1 = 60 mrem/year. If we also drink from an
orange Fiestaware teacup, we have to include the very-short-ranged beta
radiation, which would give an additional yearly dose of 400 mrem to the
lips and 1200 mrem to the fingers. For more information, see
http://www.orau.org/ptp/collection/consumer%20products/fiesta.htm.

Lecture 4
| Radiation Sources, Natural and Unnatural 43
05 HOW DANGEROUS
IS RADIATION?

R
a diation is scary because it’s invisible
and its effects are invisible. People were
equally scared of electricity 100 years
ago; newspapers avidly reported every
accidental electrocution from the tangle
of dangling wires that ran across city
streets. Like electricity, radiation is a very
useful tool. Unlike electricity, if radiation
escapes, it can harm a lot of people at
the same time. Fortunately, it only hurts
people in rare circumstances. u

44
How Particles Interact with Matter
wwRadiation damages cells by knocking loose the electrons that form
chemical bonds and by breaking apart molecules. The energies of these
particles are measured in thousands to millions or billions of electron
volts. The chemical bond energies are measured in only electron volts.

wwCharged particles interact with the atomic electrons in the material


they pass through, exciting them to higher-energy states or knocking
them loose. Electrons, beta-minus particles, are very light. They knock
out fewer atomic electrons in one place, so it’s easier for molecules to
recombine or for the body to repair them.

wwHeavier particles, such as alpha particles or protons, knock a lot of


electrons loose. They’re short-ranged, so they deposit a lot of energy in
a short distance, ionizing a lot of atoms or molecules in the same place.

wwNeutrons aren’t charged. High-energy neutrons lose energy through


collisions. A high-energy neutron could elastically scatter with a proton
in one of the nuclei and knock the protons out so that they transfer
their energy to a proton. The proton can then transfer its energy to
atomic electrons. Low-energy neutrons can combine with a proton
in a hydrogen nucleus to make a deuterium nucleus and give off a
2-million-electron volt photon: a gamma ray.

wwPhotons, such as x-rays and gamma rays, interact discretely—in


individual collisions. They can transfer a lot of energy to a single
electron through a few processes.

1 The photon can collide with 1 electron. The photon bounces off
of the electron in a process called Compton scattering, which is
dominant in tissue in the body.

2 The photon can be absorbed on an atom, followed by emitting an


electron from it. This is called the photoelectric effect.

Lecture 5
| How Dangerous Is Radiation? 45
3 If the photon is very high in energy, it can turn into an electron plus
a positron in a process called pair production. This is important at
higher energies for particle detectors called shower counters. The
energetic electrons that were knocked out then interact with the
atomic electrons.

wwWhich type of radiation is worse? There is a biological waiting factor


that depends on the ionization density: How many atoms and molecules
are disrupted in a tiny space? The more energy that is deposited in a
small region leads to a lot more damage in that region, and that is much
more difficult to repair. The electrons and the photons, or gamma rays,
that knock out other electrons have a very low weight. They have a
weight of just 1. But low-energy electrons, beta rays, are stopped by the
skin and give something called beta burns because they deposit all of
their energy there. Gamma rays just keep going and going.

wwThe other particles have higher biological weighting factors. Protons


have a weighting factor of about 5. The weighting factor for neutrons
is between 5 and 20, depending on their energy. And alpha particles,
because they deposit so much energy in such a small space, have a
weighting factor of 20. Multiply the dose in rads by the weighting
factor to get rems. Or, to get radiation in metric system units, multiply
the absorbed dose in grays by the correction factor to get sieverts.

Chemical Bonds
wwWhich chemical bonds are important? DNA is the big important
molecule that’s critical to the functioning of our cells. The other bonds
are important indirectly. Free radicals can be made. If some radiation
interacts with a water molecule, it can knock an electron loose, creating
a positively charged water molecule ion and an electron. Then, the
electron gets absorbed in another water molecule, making a negatively
charged water molecule. These water molecule ions can disassociate to

46 Nuclear Physics Explained


form OH and H free radicals. There are also some other free radicals:
HO2 and H2O2. And oxygen enhances this effect. These free radicals
can then react with and disrupt other molecules, including DNA.

wwThe DNA in dividing cells is the most vulnerable because DNA in a


regular cell is all coiled up. But when the cell divides, the DNA uncoils
and is exposed. The cells that divide quickly are in lymph nodes,
sperm, bone marrow, and the intestine. Children also have much more
cell division during growth.

wwThe damage does not always kill the cell. There are many other ways
that cells can be damaged, besides radiation, so they have repair
mechanisms. There are 2 possibilities: The repair can succeed—now
the cell is fine—or the repair can fail. If the repair fails, there are again
a few possibilities: cell death, which only becomes a problem if many
cells die at the same time; somatic effects, where something changes in
the cell and can lead to cancer; and genetic effects, which are very rare.

wwNot all radiation is dangerous. Only ionizing radiation—which


can ionize atoms and break the chemical bonds—is dangerous.
Nonionizing radiation doesn’t break chemical bonds. This includes
radiation from cell phones. There have been some contradictory case
control studies, but 2 large cohort studies both found that cell phones
had no effect on brain cancer. More importantly, there has been no
change in the brain cancer incidents in the United States in 40 years as
cell phone usage has grown dramatically.

wwPower lines emit 60 hertz of


The World Health electromagnetic radiation. These are
Organization lists cell very-low-frequency radio waves. There
phones as a “possible were some scary magazine articles
carcinogen,” but that about radiation from power lines, but
list also includes aloe, there’s no real evidence of any damage.
coffee, and talc—so Large-scale studies show that there is,
it’s a very low bar. in fact, no damage from power lines.

Lecture 5
| How Dangerous Is Radiation? 47
wwA microwave typically puts out about 1 kilowatt of power, which in
1 second just raises body temperature by a hundredth of a degree
Fahrenheit. On the other hand, 1 kilowatt of ionizing radiation could
kill a person in 1 second. So, a microwave is 1000 times less dangerous
than ionizing radiation. Although, if it’s on too long, the heat can
be dangerous.

Doses of Radiation
wwWith ionizing radiation, we can get the radiation all at once, which
is an acute dose, or we can get it over a period of time, which is a
chronic dose.

wwA single large dose of ionizing radiation is called acute radiation


syndrome. It kills lots of cells all at once, especially the
fast-dividing ones.

§§ If the dose is less than 100 rem, or 1 sievert, there are generally
no symptoms.

§§ From 100 to 200 rems, or 1 to 2 sieverts, there is typically nausea and


vomiting and fatigue, and some white blood cells are killed.

§§ From 200 to 600 rem, or 2 to 6 sieverts, there are the same


symptoms—although more severe—plus headache, confusion, fever,
hair loss, hemorrhage, and infections. And it leads to a 50% death
rate in about 4 to 6 weeks.

§§ From 600 to 800 rem, or 6 to 8 sieverts, there are the same symptoms
plus diarrhea and major system failure. About 95% of people will die
without care, and 50% to 100% die even with care.

48 Nuclear Physics Explained


§§ Above 800 rem is fatal. But that’s more than 1000 times the average
yearly dose all in a very short period of time.

wwTo treat radiation poisoning, we first decontaminate the outside of the


patient by changing their clothing and washing off any radiation on
the outside by showering them. Then, we can try to decontaminate
the inside of the patient. This includes chelation therapy for heavy
metals as well as “acceleration of the metabolic cycle of the radionuclide
by isotope dilution.” If you’re exposed to radioactive iodine, take
a potassium iodide tablet so that the nonradioactive iodine from
the potassium iodide will replace some of the radioactive iodine.
If you’re exposed to tritium, drink lots of beer to flush out the
radioactive hydrogen.

wwWhat should you do if a nuclear bomb or a dirty bomb goes off?

§§ A dirty bomb is a conventional explosive that distributes specific


isotopes from something like a stolen radioactive source. Not much
radiation is released; it’s more scary than harmful. You should cover
your mouth and nose, go inside, remove potentially contaminated
clothes and bag them, and shower.

§§ If a nuclear bomb goes off, there is a lot of direct radiation and fallout
from the fission isotopes. The wind can carry radioactive fallout large
distances. Assuming that you survive the blast and the fire effects,
which are much worse than the radiation, there will have been lots of
neutrons and gamma radiation. First, find shelter; go inside as soon
as possible. Then, wash off any fallout by taking a shower. Remove
potentially contaminated clothes and bag them. Only take iodine
tablets if recommended by the health authorities, because iodine
tablets have side effects and you can overdose on them. Finally,
manage the symptoms: Treat shock, give blood transfusions, give
fluids, and give antiemetics to reduce nausea and vomiting.

Lecture 5
| How Dangerous Is Radiation? 49
wwIf the radiation doesn’t kill or sicken the person quickly, then it is
considered a chronic dose. For example, radioactive iodine-131 from
nuclear fallout concentrates in the thyroid and can cause cancer. The
half-life of the iodine is only 8 days, and after about 10 half-lives, or
80 days, it’s pretty much all gone. Thyroid cancers have been seen in
nuclear bomb survivors and Chernobyl victims.

wwLeukemia is another common radiation-caused cancer. It is seen in


islanders near Pacific atoll bomb tests. It’s slightly elevated among
radiation workers, and it’s also seen in Hiroshima survivors. Unlike
thyroid cancer, it’s not clear if leukemia is caused by a specific isotope.

wwBy conservative estimates, there’s about a 0.05% extra chance in a


lifetime per rem of radiation received of developing extra cancers. With
100 rem of extra radiation as a lifetime dose, that implies a 5% extra
chance of getting cancer.

wwWhat about the long-term effects from the more than 500
above-ground bomb tests that were ended around 1963? There are
some local effects where the bomb fallout was the densest. There is
more carbon-14 in the atmosphere, but that’s in parts per trillion, and
this carbon-14 change is actually a useful dating tool.

wwWe have good data at high doses of radiation, but there’s poor data at
low doses of radiation. There are big uncertainties in the effects at low
doses because we expect the effects to be small, and therefore they are
difficult to measure.

wwMutations are caused by damage to sperm or egg cells, show up in the


next generation, and are rarely beneficial. About 100 rem doubles the
natural mutation rate in mice. There are no measurable mutations
above the natural background rate of mutations in nuclear bomb
survivors’ children. One reason for that is there were too few survivors
who received high doses to see an effect. Fortunately, there’s not
enough data on humans, so we use the mouse data.

50 Nuclear Physics Explained


wwRadiation dose limits only apply to power plants and other nonmedical
sources. The average background radiation of 0.6 rem per year—half
of which comes from natural sources and half of which comes from
medical procedures—is not included. Medical radiation is not included
because the benefit of the test should dramatically outweigh the harm
done from receiving the extra radiation.

wwThe radiation worker limits are 50 millisieverts, or 5 rem, per year.


Limits for the general public are 1 millilsievert, or 0.1 rem, per year. It
took a while to realize that radium is harmful because the effects were
so delayed and because people had to correlate the different cases of
radium poisoning to figure out their common cause.

The EPA standard for homes is that if a smoker lives for 75


years in the same home, there is a 6% elevated lifetime cancer
risk. The EPA says that radon is the number 2 cause of lung
cancer, after smoking. And if a smoker has radon exposure, it’s
10 times more dangerous to them than for a nonsmoker.

Lecture 5
| How Dangerous Is Radiation? 51
How much life do we lose for each activity
that we do?
§§ Smokers will lose 2400 days, or about 6.5
years, of life.

§§ People who are 20% overweight will lose about


1000 days of life, which is almost 3 years.

§§ People who work in agriculture, construction,


and mining are more susceptible to accidents
and will lose somewhere between half of a year
and a year of life.

§§ Driving, on average, costs each of us about


half of a year of life.

§§ If we get an extra 300 millirem per year of


radiation throughout our lives, that will cost us
10 times less than driving, or about 15 days
of life.

wwOther occupational exposures include early radiation scientists, such as


Thomas Edison, Henri Becquerel, and Marie Curie; dentists, who used
to hold x-ray film in their hands; and miners, who were exposed to
radon down in the mines. There is also radon in homes.

wwThere are studies of nuclear accident survivors and of Hiroshima and


Nagasaki survivors. A prospective long-term cohort study called the
Life Span Study, which was originally intended to look for the extra
mutations caused by the nuclear bombings, found that there was only
5% more cancer among survivors who were exposed to radiation. There
were no measurable genetic effects—no mutations—above the natural
incidence of cancer.

52 Nuclear Physics Explained


wwHuman cancer risk is very variable, and low-dose radiation effects are
small. There are 2 opposite ideas of low-dose effects. The first is the
radiation hormesis hypothesis, which is that small doses of radiation
stimulate cellular repair mechanisms and are therefore good for you.
This is not widely accepted. The second is the linear no-threshold
hypothesis, which assumes that there is no “safe” dose of radiation. We
don’t know if this is reasonable, but we use this hypothesis because it’s
the most conservative one for radiation regulation.

Supplements

READINGS

Jorgenson, Strange Glow.


National Research Council, Health Risks from Exposure to Low
Levels of Ionizing Radiation.
Rhodes, The Making of the Atomic Bomb.

QUESTIONS

1 Could an astronaut traveling to Mars die from exposure to cosmic


radiation during the trip? On Earth, the radiation dose from
cosmic rays is roughly 30 millirems (mrem) per year, but in space
an astronaut would receive 1000 times more radiation, or 30 rem
per year.

2 A 1-rem CT scan is estimated to increase our chance of premature


death by 0.05%. How many miles would we need to drive to incur
the same risk?

Lecture 5
| How Dangerous Is Radiation? 53
ANSWERS

1 A trip from the Earth to Mars will take about 6 months. The radiation dose
from cosmic rays on Earth is only about 30 mrem/year because most of the
cosmic rays are deflected by the Earth’s magnetic fields and most of the
remainder are blocked by the Earth’s atmosphere. A mission to Mars will
travel both outside the atmosphere (like the International Space Station)
and beyond the protection of the Earth’s magnetic fields. At 30 rem/year,
an astronaut will receive 15 rem in the 6-month voyage. This is well below
the 100-rem threshold for acute radiation syndrome but will increase the
astronaut’s lifetime probability of getting cancer by about 2%. The biggest
unknown is the effect of solar storms, which can emit large quantities of
radiation and could be quite dangerous.

2 Driving is dangerous. There is 1 extra death for every 100 million miles
driven. To get an extra 0.05% of a death (5 × 10 −4), we would need to drive
(5 × 10 −4)(108) = (5 × 104) miles, or 50,000 miles. That is about 4 years
of driving.

54 Nuclear Physics Explained


THE LIQUID-DROP
MODEL OF THE
06
NUCLEUS

T
 e nucleus of the atom behaves
h
approximately like a droplet of water.
Individual protons and neutrons are like
solid marbles, but enough marbles together
can flow like water—just like flowing
sand. The liquid-drop model, which was
developed in the late 1930s, describes a
lot of the nuclear masses. The curve of
binding energy just includes the most stable
isotope for each element. This lecture will
offer a way to describe the masses of all
the isotopes. u

55
Isotopes and Isotones
wwThe simplest atom is a hydrogen atom. It has 1 electron interacting
with 1 proton. They interact via the electric and magnetic forces.
The hydrogen atom has lots of excited states corresponding to lots of
emission and absorption lines. We measure the energies of those lines,
and that gives us lots of information about how the electron and proton
interact with each other.

wwThe simplest nontrivial nucleus has 1 proton interacting with


1 neutron. This is the deuteron, sometimes called deuterium, and it’s
the isotope of hydrogen. The deuteron only has one state—the bound
state—so we can’t learn quite as much about the nucleon-nucleon
interaction from studying the deuteron. There are no stable nuclei with
2 protons or with 2 neutrons.

wwWhat can we learn from the deuteron? We can learn about the binding
energy, which is measured using a mass spectrometer. We compare
the mass of deuterium with the mass of hydrogen. We can also look at
the reaction of when a neutron hits a proton at low energies; they stick
together to form a deuteron and give off a gamma ray. We can also
look at the inverse reaction: We can aim gamma rays at deuterium and
detect the neutron and proton that come out.

wwWe find that the deuteron is barely bound. The binding energy is
2.2 million electron volts, which means that the potential energy of
the system is very large and the kinetic energy is almost as big. When
you add the 2 together, the positive kinetic energy almost equals
the negative potential energy. This is very different in an atomic or
gravitational system, where the potential energy is about twice as big
as the kinetic energy. The radius of deuterium is also fairly large, at
2.1 fermis, or femtometers.

56 Nuclear Physics Explained


wwAngular momentum measures how much an object rotates. It is one of
the basic conserved quantities in the universe: We know that charge,
energy, and momentum are conserved. Angular momentum is also one
of the fundamental attributes of particles.

wwThere are 2 types of angular momentum. The first is spin, which


is the intrinsic angular momentum that belongs to a particle. Most
elementary particles spin like a top, and the spin is quantized, which
means that it can only have certain specific values. The proton and the
neutron each have a spin of 1/2. Spin can either be down, which means
that it’s spinning clockwise, or up, which means that it’s spinning
counterclockwise. The nucleon-nucleon interaction (the force between
2 nucleons) depends on the alignments of their spins—whether their
spin is in the same direction or in opposite directions.

wwThe second type is orbital angular momentum. The nucleons can orbit
each other like the Moon orbits the Earth. Orbital angular momentum
is also quantized; it can only have values of 0, 1, 2, etc. The total
angular momentum is the orbital angular momentum combined with
spin angular momentum.

wwHow does spin affect the shape of the deuteron? The proton and
neutron spins are aligned—they’re both spinning in the same
direction—so the total spin is 1/2 plus 1/2, which is equal to 1. The total
angular momentum of the deuteron is 1. That means that there are
2 possibilities: either orbital angular momentum is 0 (s state), where the
neutron and the proton are apart from each other; or orbital angular
momentum is 2 (d state), where the neutron and the proton are orbiting
each other.

wwWhat about 2 protons or 2 neutrons? Nucleons are particles that


can’t occupy the same state; they obey the Pauli exclusion principle.
A neutron is different from a proton, so each can have the same spin,

Lecture 6
| The Liquid-Drop Model of the Nucleus 57
such as spin up. But if you have 2 protons (or 2 neutrons), those are
the same particles, so they have to have different spin. The problem is
that the nucleon-nucleon force is slightly weaker for different spin. The
2 protons repel each other, and this makes them more unbound than
2 neutrons because it’s more difficult for 2 protons to form a nucleus
than 2 neutrons.

wwHow are heavier nuclei affected by changing the number of protons (Z)
and neutrons (N)? We have isotopes with the same number of protons
and different neutrons. Isotopes with the same number of protons are
the same element. As we add neutrons, the nuclei become less bound
and less stable and have shorter half-lives. Those extra neutrons are
going to beta-decay to a proton and electron and an antineutrino. If
there are way too many neutrons, then they don’t wait around for beta
decay; instead, they drip off.

wwIf instead we subtract neutrons, the nuclei also become less bound and
less stable and have shorter half-lives. But in this case, the protons will
beta-plus decay to a neutron, a positron, and a neutrino. Or the proton
might absorb an electron—called electron conversion—to become a
neutron plus a neutrino. And just like with neutrons, if there are way
too many protons, they don’t wait around for beta decay and instead
just drip off.

wwWhen we add or subtract protons, we have an


isotone, which has a fixed number of neutrons
Z represents
and a different number of protons and therefore
the number of
is a different element. protons and
N represents
the number
of neutrons. Z
comes from the
German word
zahl, which
means number.

58 Nuclear Physics Explained


Charting Protons and Neutrons
wwThe chart of nuclides [PAGE 274] shows all of the combinations of
protons and neutrons. There is a very narrow band of stable nuclei.
Often there is just 1 stable isotope, especially for odd-even nuclei, and
frequently there are no stable isotones. There is also a narrow band of
known nuclei. There is a lot of space on the chart that is not filled; this
represents nuclei that are either not known or, in most cases, impossible.

wwWith light nuclei, the number of neutrons is about equal to the number
of protons. This is represented on the chart of nuclides by a 45° line,
starting at the bottom left and moving up and to the right. With heavy
nuclei, there are more neutrons than protons—the ratio is about 1.5—
so that 45° line bends over. This is because we need more neutrons to
offset the repulsion among all the protons.

wwAlso on the chart of nuclides are magic numbers that show us


where there are more stable isotopes or isotones. There are very few
stable nuclei with an odd number of protons and an odd number of
neutrons, ranging from deuterium, with 1 proton and 1 neutron, up to
tantalum-180.

wwThere are no stable nuclei with more than 82 protons—that’s lead.


Bismuth-209 is close, but it’s unstable; the half-life is greater than the
age of the universe. Is there an island of stability, or relative stability, at
some larger magic number that we haven’t discovered yet?

wwThe stable nuclei form a valley of stability. The nuclei decay toward
the valley. If the nuclei have too many protons, then they’re going to
beta-plus decay down toward the valley. If there are too many neutrons,
then they are going to beta-minus decay down toward the valley. And if
there are way too many protons or neutrons, then those excess protons
or neutrons just drip off. These are the drip lines. We know where the
proton drip lines are, but we don’t know where the neutron drip lines
are. And there are still more nuclei to discover.

Lecture 6
| The Liquid-Drop Model of the Nucleus 59
Systematizing Binding Energy
wwHow can we describe the curve of binding energy and the isotopes
and isotones? We know that the binding energy per nucleon is almost
constant for heavy nuclei. For all nuclei heavier than carbon, the
binding energy is about the same: 7 to 8 MeV. Therefore, we can use
a volume term proportional to the number of nucleons, which will be
some number times the number of nucleons: aVA.

wwThis implies that each nucleon is only attracted to its closest neighbors.
Note that this is not true for atoms, because the binding energies for
electrons in atoms is not constant. It increases proportional to the
square of the number of protons, not just the number of protons.

wwWhat about the nucleons on the surface? Nucleons are attracted to their
nearest neighbor. Surface nucleons have fewer neighbors, so they’re less
bound, just like in a water droplet. The effect is bigger in smaller nuclei
because smaller nuclei have more surface relative to volume.

wwThis is not precise for light nuclei; it does not give us the wiggles and
peaks in the curve of binding energy, but it gives us the general shape.
This is similar to surface tension in liquids: Water droplets bead up to
minimize their surface area. The same thing happens with nuclei.

wwThe surface area of a sphere, 4πr 2, is a radius squared. The radius of


a nucleus is about the cube root of the number of nucleons, so the
radius squared is a number of nucleons to the 2/3 power: A2/3. We’ll add
a negative term (some number) times the number of nucleons to the
2/3 power to get the surface energy: −asA2/3.

wwWe also want to include the electrical repulsion between the protons
because that makes a nucleus less bound. Every proton repels every
other proton, so if we have Z protons, each of those protons has (Z − 1)
other protons, so we’ll multiply Z times (Z − 1) times another number:
−ac Z(Z − 1).

60 Nuclear Physics Explained


wwWe’re going to add more terms to fit the data better. These are not
motivated by just water droplets; they are added to describe the
behavior of the nuclei that we can see. And we can tell that nuclei
prefer to have the same number of protons and neutrons, so we’re
going to take the number of neutrons minus the number of protons
and square it to deal with the problem of having too many neutrons
or too many protons. This is divided by the total number of nucleons:
–asym(N − Z)2/A. This is less important for stable nuclei with increasing
atomic number, but it describes the valley of stability very nicely.

wwThere is also a pairing term. Recall that there are very few stable
odd-odd nuclei. It turns out that nuclei prefer to have an even number
of protons and an even number of neutrons, so we’ll add a term that
gives a bonus to even-even nuclei (+ap A –3/4), a 0 for even-odd nuclei, and
a penalty for odd-odd nuclei (–ap A –3/4).

wwWe now have 5 parameters, and we’re going to fit them to the binding
energy data to give us the best description of the data:

Energy = volume − surface − proton


repulsion − asymmetry + pairing

E = aV A − as A2/3 − ac Z(Z − 1) − asym(N − Z)2/A + pairing term

wwNucleons inside the volume of the nucleus all interact with their nearest
neighbors, which gives us the volume term. Nucleons on the surface
have fewer neighbors and are therefore less bound. Every proton repels
every other proton, giving a negative term proportional to the square of
the number of protons. Nuclei prefer to have equal numbers of protons
and neutrons, and this gives us the asymmetry term. Lastly, nuclei
prefer to have even numbers of protons and of neutrons.

wwThe surface term explains a dramatic rise in the curve of binding energy
for light nuclei. The electrical repulsion and asymmetry terms explain
the slow decrease in the curve of binding energy for heavy nuclei.

Lecture 6
| The Liquid-Drop Model of the Nucleus 61
wwWe’ve now systematized all of the binding energies we’ve already
measured and can predict the binding energies of more asymmetric
nuclei. But the asymmetry and pairing terms are made up and do
not come from a liquid-drop model. Instead, those terms come from
quantum mechanics. We’ll add the quantum mechanics a little at
a time.

The Fermi Gas Model


wwLet’s start with the Fermi gas model, which is
the least amount of quantum mechanics we
can add. We’re going to confine the nucleons
to the nucleus but ignore all of the
details of the nucleon-nucleon
interaction. We’re going to
include the Pauli exclusion
principle, which tells us that
we can only have one particle
in each state at each time.

wwSo, 1 state is going to be


a box of a certain size in
ENRICO FERMI
space and in momentum, 1901–1954
and the size of that in
space times momentum is h-bar
(ℏ): ΔxΔp = ℏ. And h-bar is Planck’s constant (h) divided by 2π, or
10−34 joule-seconds, or 200 fermis times a million electron volts divided
by the speed of light: 200 fm MeV/c. The momentum is the mass times
the velocity (nonrelativistically): p = mv. We also have to include the
relativistic contraction factor: p = mv𝛾. We count protons and neutrons
separately because they’re separate particles.

62 Nuclear Physics Explained


wwSo, in 1 dimension, we have protons in a box of size x, where x is the
size of the nucleus. If we have 1 proton, it’s going to have momentum
h-bar divided by the size: p = ℏ/x. If we have Z protons, the momentum
will be the number of protons times h-bar divided by the size:
pFermi = Zℏ/x. If we have N neutrons, the momentum will be the
number of neutrons times h-bar divided by the size: pFermi = Nℏ/x. It’s
actually half that, because we can put 2 protons in each box: spin up
and spin down.

wwThe kinetic energy of the proton or neutron is 1/2 the mass times
velocity squared, or momentum squared divided by twice the mass:
½mv2 = p2/2m.

wwThis explains the preference for equal numbers of protons and


neutrons. If we have the same number of protons and neutrons, they
have the same Fermi momentum. But if we have 1 more proton and
1 fewer neutron, then the neutron momentum is a little less and the
proton momentum is a little more, but the energy is the square of the
momentum, so the total energy increases.

wwHow do we generalize this to 3 dimensions? Reality is not


1-dimensional. In 3 dimensions, we will fill each dimension separately.
So, the box is now going to have a size of x times y times z: xyz. We
will assume that we have the same Fermi momentum ( pFermi ) in
each direction.

wwThe number of protons we stack in the x direction is the Fermi


momentum times the box size divided by h-bar: Nx = pFermix/ℏ. We do
the same in the y direction and in the z direction. So, the total number
is 2 (for spin) times the number in x, times the number in y, times
the number in z: N = 2NxNyNz. This means that the total number
of protons will be 2 times the volume in momentum-space times the
volume in space-space divided by the modified Planck’s constant
(h-bar) cubed: 2 times (momentum volume) times (volume)/ℏ3.

Lecture 6
| The Liquid-Drop Model of the Nucleus 63
wwActually, we’re going to use spheres and not cubes, so the Fermi
momentum in a nucleus is going to go as the density (number of
nucleons per volume) raised to the 1/3 power: pFermi ∼ density1/3. The
Fermi momentum for nuclear density is 270 MeV/c, which gives us a
kinetic energy for the nucleon at that Fermi momentum of 37 million
electron volts. Because the binding energy is about 8 million electron
volts, the potential energy of attraction is about −45 million electron
volts (it’s negative because of the attraction; you have to put in energy
to pull them apart).

wwBy using the quantum mechanics in the Fermi gas model, we have
explained the asymmetry term in the liquid-drop model.

Supplements

READINGS

Henley and Garcia, Subatomic


Physics, chap. 16.
Lilley, Nuclear Physics, sections
2.1–2.2.
Mackintosh, Nucleus, chap. 6.

QUESTIONS

1 Why doesn’t uranium-238 decay


by beta-plus or beta-minus decay?

2 Which should take more energy: to knock a proton out of the center
of a nucleus or to knock a proton out from the surface of a nucleus?

64 Nuclear Physics Explained


ANSWERS

1 Even though uranium-238 is unstable, it has a half-life of almost 5 billion


years and lies in the valley of stability. It is an even-even nucleus. If it
decayed by beta decay, either beta-plus or beta-minus, it would become
odd-odd, because 1 of the protons would be become a neutron (or vice
versa). Even-even nuclei are more stable than odd-odd nuclei because of the
pairing term.

2 Protons in the center of the nucleus are more bound, because there are more
nucleons around them that attract them. Protons on the surface of a nucleus
have fewer neighboring nucleons to attract them and are therefore less
tightly bound. Therefore, it will take less energy to knock out a proton from
the surface of a nucleus.

Lecture 6
| The Liquid-Drop Model of the Nucleus 65
07 THE QUANTUM
NUCLEUS AND
MAGIC NUMBERS

A
lthough nuclei can ring like a bell or spin
like a top, the quantum structure of
the nucleus is remarkably similar to the
quantum structure of the atom, with
single nucleons instead of electrons in
s-shell, p-shell, d-shell, etc., orbitals. The
atomic shell model works because it has
the nucleus to provide a central force, the
electromagnetic force is weak, and the
electrons are point particles. In nuclei, on
the other hand, the only central attraction
is provided by the other protons and
neutrons, the force is very strong, and
the nucleons have substructure—they are
made of quarks—that can be distorted by
these forces. It’s amazing that the shell
model works for nuclei at all. u

66
The Atomic Shell Model
wwThe liquid-drop model describes nuclear binding energies and the
valley of stability. The Fermi gas model adds some quantum mechanics
to explain why nuclei with similar numbers of protons and neutrons are
more bound. These models do not explain everything we know about
the nucleus.

wwIn chemistry, there is an atomic shell model. There are peaks in the
energy it takes to remove an electron at the magic numbers—2, 10,
18, 36, and 54—corresponding to the noble gasses [PAGE 270]: helium,
neon, argon, krypton, etc. These peaks form a quantum pattern that is
evidence for the atomic shell model, with electrons orbiting the nucleus
in s-shells, p-shells, d-shells, etc.

wwSimilarly, there is a nuclear shell model, with its own set of magic
numbers: 2, 8, 20, 28, 50, and 82.

wwIf we put the energy needed to remove 2 protons on a plot of the


number of protons on the vertical axis versus the total number of
nucleons on the horizontal axis—the chart of nuclides—then we see
slightly darker horizontal bands for the proton magic numbers of 8,
20, 28, 50, and 82. If instead we put the energy needed to remove
2 neutrons on the same plot, then we see slightly darker diagonal bands
for the neutron magic numbers of 28, 50, 82, and 126. This quantum
pattern is evidence for the nuclear shell model.

wwNuclei with magic numbers of neutrons or protons—or both—are


particularly stable and abundant. For example, tin (element 50,
with 50 protons) has more stable isotopes than any other element.
Other elements with magic numbers include helium-4, oxygen-16,
calcium-40, and lead-208.

Lecture 7
| The Quantum Nucleus and Magic Numbers 67
wwTo explain these magic numbers, we have to go beyond the Fermi gas
model, which includes some quantum mechanics, such as the Pauli
exclusion principle, but doesn’t have any details of the nucleon-nucleon
interaction. We need to include some details of the nucleon-nucleon
interaction to explain these magic shell numbers.

wwLet’s start by looking at the atomic shell model, in which the electrons
orbit around the nucleus in the mean field, the average force due to
the nucleus plus all the other electrons. The positively charged nucleus
provides a central potential, or force, diluted by the average of all the
other electrons. Electrons are point particles; they have no structure
and can’t be distorted. And the electromagnetic interaction is weak.

wwA potential is another way to describe a force. By “potential,” we mean


a potential energy curve. The force is due to changes in the potential
energy with location. A hill is a gravitational potential energy curve.
The force is downhill and depends on steepness. We use whichever
description—either potential energy or force—makes it easier to solve a
particular problem.

wwThe innermost electron “sees” all the charge in the nucleus (Z protons).
The innermost electron orbits much closer to the nucleus and is much
more tightly bound. In a neutral atom, the outermost electron “sees” a
net charge of just +1 (the effect of all of the protons plus the rest of the
electrons). The rest of the electrons “screen,” or neutralize, the rest of
the nuclear charge.

The Nuclear Shell Model


wwHow can we make a shell model work for nuclei? The protons and
neutrons orbit around the other nucleons in the mean field due to
the other nuclides, but there is no center. Furthermore, protons and
neutrons are composite particles (they are made of quarks) and therefore

68 Nuclear Physics Explained


can be modified by the force; they interact by the complicated strong
nuclear force and are pushed apart by the electromagnetic force. Despite
all of these complications, the shell model works anyway—mostly.

wwLet’s use a simple example to show how we calculate the proton and
neutron states. We assume that the average potential looks like the
nucleons are attached to the center of the nucleus by a simple spring:
The more you stretch the spring, the stronger the force. The force
increases linearly with distance: F = −kx. This gives us a potential that
increases as the distance squared, so the potential energy is the square
of the distance from the center: 1/2kx 2.

wwThis is very artificial, because if the potential looks like this, then the
nucleons can never escape the nucleus. But the advantage is that it’s
easy to calculate, at least for physicists.

wwWhen we make wave functions—that’s the quantum mechanical


description of the particles—in this simple spring model, we get
2 important quantum numbers that describe the distribution of a given
proton or neutron.

wwThe first quantum number is the orbital angular momentum (L),


which tells us what shell the particle is in: s-shell, p-shell, d-shell, etc.
In other words, L tells us how rapidly the particle orbits the nucleus:
The bigger the orbital angular momentum, the more rapidly the proton
orbits the nucleus. For a given value of L, there are 2L + 1 substates.

§§ If L = 0, which is the s-shell, there is 1 substate. The s-shell is


spherically symmetric, so it looks like a sphere.

§§ If L = 1, which is the p-shell, there are 3 substates. The p-shell has


2 lobes.

§§ If L = 2, which is the d-shell, there are 5 substates. The d-shell has


4 lobes.

Lecture 7
| The Quantum Nucleus and Magic Numbers 69
§§ It keeps getting more complicated from here.

wwOrbital angular momentum is similar to spin, which can be up


or down. If L = 1, orbital angular momentum can be up, sort of
horizontal, or down.

wwThe second quantum number is n, which is the principle quantum


number. It can be 1, 2, 3, etc., and tells us how many wiggles there are
in the wave function as it moves away from the center.

§§ If n = 1, there are no wiggles, and it’s the lowest energy.

§§ If n = 2, there is 1 wiggle and a bit more energy.

§§ If n = 3, there are 2 wiggles and even more energy.

§§ This trend continues.

70 Nuclear Physics Explained


wwThese are very similar to the electron The labels s, p,
orbitals. When we put electrons in a d, and f came
nucleus, we start with the 1s shell, followed from people
by 2s and 2p; then 3s and 3p; then 4s, 3d, looking through
and 4p. spectrographs at
lines. There were
wwThe nuclear orbitals have the same names “sharp” lines for
and shapes, but they’re in a different order. s, “principal” lines
The nuclear order is 1s, 1p, 1d, 2s, 1f, 2p, for p, “diffuse”
etc. And the orbitals go beyond s, p, d, f to lines for d, and
g, h, i, k ( j is skipped because it can look “fine” lines for f.
too similar to i and cause confusion).

wwThe magic numbers are also not the same. The magic numbers
for nuclei are 2, 8, 20, 28, 50, 82, etc., for protons, corresponding
to helium, oxygen, calcium, nickel, tin, and lead. Neutron
numbers—2, 8, 20, etc.—correspond to different isotopes.

wwThe chart of nuclides is frequently marked by horizontal and vertical


lines showing the location of the magic numbers. But the magic numbers
become less important as we go farther from the valley of stability.

wwWhat do these quantum numbers do? We can combine the principle


quantum number and the orbital quantum number to tell us where the
major shells are. The energy (lambda, or Λ) defines major shells and is
equal to 2n + L − 2. The number of states in the shell is (Λ + 1)(Λ + 2).

§§ Let’s start with lambda = 0. The only way we can make lambda = 0 is
if we have a principle quantum number of 1 and an orbital quantum
number of 0—that’s the 1s-shell. There are 2 neutrons and 2 protons,
which is helium-4.

§§ To make lambda = 1, we have principle quantum number 1 and


orbital quantum number 1, which is the 1p-shell. There are 6 neutrons
and 6 protons in addition to the s-shell, which is oxygen-16.

Lecture 7
| The Quantum Nucleus and Magic Numbers 71
§§ If we have lambda = 2, there are 2 ways to make it: with n = 2 and
L = 0 (the 2s-shell) or with n = 1 and L = 2 (the 1d-shell). There are
12 neutrons and 12 protons, which is calcium-40.

§§ Lambda = 3 gives us a 2p-shell or a 1f-shell, with 20 neutrons and


20 protons. That gives us magic numbers of 40, but we don’t have
magic numbers there, so it doesn’t work.

§§ The model works for the first 3 orbitals—1s, 1p, 2s/1d—all the
way up to calcium, but then it breaks down. We’re going to need
a more realistic—more complicated—interaction to describe the
higher shells.

wwAngular momentum interacts with itself. There are 2 types of angular


momentum: spin and orbital angular momentum. The spin of the
protons and neutrons is 1/2; the orbital angular momentum depends on
the shell—s-shell, p-shell, d-shell, f-shell, etc.—and corresponds to 0,
1, 2, 3, etc. The spin-orbit force, then, depends on the relative direction
of the spin and the orbital parts.

ORBITAL SHELL s p d f g h i k

ANGULAR MOMENTUM 0 1 2 3 4 5 6 7

wwThe 1p shell can have total angular momentum, which is equal to the
orbital plus or minus (because direction matters) the spin: 1 ± 1/2, which
is either 1 + 1/2, which gives us 3/2, or 1 − 1/2, which gives us 1/2. So,
we’re changing the direction between the orbital angular momentum:
3/2 where the spins and the orbital angular momentum are aligned and
1/2 where they point in the opposite direction. Those 2 states previously
had the same energy.

72 Nuclear Physics Explained


wwThe p3/2 state moves to a lower energy while the p1/2 state moves to a
higher energy, and the energy of the shell splits. The bigger the angular
momentum, the more splitting there is. So, the s-shell is unsplit, the
p-shell is split a little, the d-shell is split more, and the f-shell is split so
much that there is a shell energy gap between those 2 states. This gives
us the correct magic numbers at the energy gaps in the shell structure.

The order of the proton orbitals in nuclei are very different


from the order of the electron orbitals in atoms [PAGE 272].

wwWhy was adding this extra term—the spin orbit term—so important?
It describes how the binding energies change at specific magic
numbers, just like the binding energies change in atoms in noble gasses.
It tells us how the energies of the shells split into subshells. It predicts
the ground-state spins of nuclei 1 nucleon away from a closed shell as
well as the excited states of these nuclei. Now we can describe nuclei in
terms of their constituent protons and neutrons.

wwHow does this shell model relate to nuclear abundances? It turns out
that even-even nuclei are more tightly bound and therefore are more
common than nuclei with an odd number of protons, and nuclei with
magic numbers are more abundant than their nearest neighbors. We
can also use the shell model to accurately calculate nuclear densities.

wwWhat are the limits of the shell model? It was created to describe
stable nuclei, so interactions among many nucleons are relatively
more important for weakly bound or unbound nuclei. And the magic
numbers vanish for very unstable nuclei, with too many protons or too
many neutrons. It’s unclear what will happen in super-heavy nuclei,
where there’s a delicate balance between the short-range attraction and
the long-range electrical repulsion.

Lecture 7
| The Quantum Nucleus and Magic Numbers 73
Quantum Vibrations and Rotations
wwAn entire nucleus can vibrate like a bell, where the whole nucleus is
involved, and it’s not just single nucleon excited states.

wwIn giant dipole resonance, the protons oscillate opposite to the


neutrons. This is common in all medium to heavy nuclei. And as
the nucleus gets bigger, the energy of this state decreases. It’s about a
14-million-electron volt excited state in aluminum and about half of
that in lead.

wwThere are also quadrupole resonances, where the nuclei oscillate


between being slightly cigar-shaped (prolate) and slightly
Frisbee-shaped (oblate).

wwThen there are monopole resonances, where the whole nucleus expands
and contracts.

wwWhat does it mean when we find a bunch of nuclear excited states with
evenly spaced levels? In the spring model, the excited states of particles
on springs are evenly spaced. So, if we see a bunch of levels that are
evenly spaced, then we’re looking at vibrations that we can describe as
2 masses connected by a spring. So, a set of evenly spaced levels tells us
that the nucleus is vibrating.

wwWe can explain other, uneven-level spacing with more quantum


numbers. Those uneven levels can tell us that the nucleus is rotating.
The energy for classical rotation is 1/2𝜏𝜔2, where 𝜏 is the rotational
inertia of the thing that’s rotating (mass times distance from the axis of
rotation squared) and 𝜔 is the rotational speed (measured in rotations
per second). The angular momentum, which is the rotational inertia
times the rotational speed (L = 𝜏𝜔), is conserved.

74 Nuclear Physics Explained


wwQuantum rotations have quantized energy and angular momentum.
We look for energy levels that—instead of being evenly spaced,
like with vibrations (1, 2, 3, 4, etc.)—are the square of the angular
momentum (L2): 1, 22 = 4, 32 = 9, etc.

wwBut quantum mechanical spheres can’t rotate because there is no


difference between different orientations. That means that magic
nuclei, such as oxygen-16 and lead-208, can’t rotate.

wwCigars and pancakes (prolate and oblate spheroids) can rotate and
typically have a set of energy levels that is spaced as the angular
momentum squared (L2)—these rotational states. So, if we see a set of
rotational states, that tells us that the nucleus is not spherical.

wwHow do we measure these rotational bands? Practically, we measure the


cascade of gamma rays as each rotational state decays down to the next
one, and the energies correspond to the energy differences between the
states. This lets us measure the rotational inertia of the nucleus and
tells us how deformed it is. And we discover that the energy levels don’t
exactly match this pattern because the faster the nucleus rotates, the
more deformed it gets and the more rotational inertia it has.

Supplements

READINGS

Henley and Garcia, Subatomic Physics, chaps. 17–18.


Lilley, Nuclear Physics, sections 2.3–2.5.
National Research Council, Nuclear Physics, p. 30–56.

Lecture 7
| The Quantum Nucleus and Magic Numbers 75
QUESTIONS

1 How are nuclear and atomic structure similar? How do the proton
and neutron orbitals in a nucleus compare with the electron orbitals
in an atom?

2 Elements with filled atomic shells (i.e., with magic numbers of


electrons), such as helium, neon, and argon, react very differently
than other elements. Why don’t elements with filled nuclear
shells (i.e., with magic numbers of protons and neutrons), such as
oxygen-16, calcium-40, and lead-208, react very differently than
other elements?

ANSWERS

1 The protons and neutrons in the nucleus occupy orbitals that are very
similar to the electron orbitals around an atom, although 10,000 times
smaller. The big differences are that the protons and neutrons fill their
orbitals in a different order than the electron orbitals, and the energy gaps
(the magic numbers) are in very different locations.

2 The chemical reactivity of an element is determined by its electronic


structure, because chemical reactions are driven by sharing electrons.
The nuclear structure of an isotope has almost no effect on its chemical
reactivity. Isotopes with filled nuclear shells are more stable and are thus
more likely to be produced in nuclear reactions.

76 Nuclear Physics Explained


PARTICLE
ACCELERATORS:
08
SCHOOLS OF
SCATTERING

I
 this lecture, you will learn about the
n
techniques that are used to develop and test
models of the nucleus. These new techniques
will allow us to go beyond the bulk properties
of nuclei—mass, spin, binding energy—and look
more closely at how the individual nucleons
behave. This lecture will focus on how and why
we accelerate the subatomic particles that
we then scatter from nuclei. Together, beams
of heavy ions, radioactive ions, and electrons
show us where nuclei come from, the most
extreme examples of what a nucleus can be,
and the internal structure of the nucleus. u

77
Scattering Particles and Studying Reactions
wwBefore 1911, the atom was seen as a
plum pudding—a uniform blob
with the electrons interspersed.
In that model, massive alpha
particles passing through matter
would only be slightly deflected.

wwIn 1909, Hans Geiger


and Ernest Marsden, who
were working in Ernest
Rutherford’s lab, aimed alpha
particles from radium decay
at very thin gold or platinum
foils. This was one of the first
scattering experiments; they didn’t
accelerate the alpha particles. To detect the
particles, they counted light flashes on small zinc sulfite screens from
alpha particles hitting the screen. The alpha particles scattered in all
directions. Most of them went straight, some of them bounced off at
small angles, and 1 out of 8000 bounced backward.

wwRutherford interpreted this backscatter as due to a very large force on


the alpha particles. The force was due to the electrical repulsion from
the nucleus. The electrical repulsion gets much bigger as the distance
gets smaller and smaller. That meant that there needed to be a large
amount of charge in a very small region and a large amount of mass in
that region so that it didn’t get pushed out of the way. This scattering
experiment didn’t determine whether the central charge of the nucleus
was negative or positive. That came later.

78 Nuclear Physics Explained


wwThis experiment showed the existence of tiny, massive atomic nuclei. It
was the first experiment to scatter subatomic particles from the nucleus.
All nuclear physicists are doing today are fancier and fancier versions of
Rutherford’s experiment.

wwWhy do we need higher energies—and accelerators to provide them?


The smaller the wavelength of our probe, the better we can see things.
You can’t see features smaller than the wavelength of what you’re
looking with. With photons, the wavelength is Planck’s constant
times the speed of light divided by the energy: 𝜆 = hc/E. But photons
also have momentum, and that is the energy divided by the speed of
light: p = E/c. That means that the wavelength of a photon is Planck’s
constant divided by the momentum: 𝜆 = h/p.

wwThis is also true for particles. Just like photons travel as waves and
interact as particles, particles travel as waves and interact as particles
in quantum mechanics. Their wavelength is also equal to Planck’s
constant divided by their momentum: 𝜆 = h/p. Planck’s constant (h)
multiplied by the speed of light (c) is 1200 MeV fermis. That means
that if we want a wavelength of 1 fermi, we need a momentum of about
1000 MeV/c. Geiger and Marsden’s alpha particles had a momentum of
only about 200 MeV/c, so they could see that the nuclei were in a small
region, but they couldn’t see any details about the nucleus.

wwWe also want higher energies to study different (inelastic) reactions:


excite the nucleus to different states, knock particles out of the nucleus,
and make new nuclei. There are 3 main types of large accelerators.
There are electron accelerators, such as at the Thomas Jefferson
National Accelerator Facility (Jefferson Lab). There are also ones that
accelerate heavy ions. And there are 2 main types of those. At high
energy, accelerators such as the Relativistic Heavy Ion Collider in New
York and the Large Hadron Collider in Switzerland and France study
the quark-gluon plasma.

Lecture 8
| Particle Accelerators: Schools of Scattering 79
Thomas Jefferson National Accelerator Facility
(Jefferson Lab)

wwAt lower energies, accelerators such as the Facility for Rare Isotope
Beams or Van de Graaff generators accelerate heavy ions to make
unusual nuclei and to make nuclei in unusual excited states. We need
this information to understand stellar nucleosynthesis—how stars
make nuclei.

80 Nuclear Physics Explained


wwWhy do we use electron scattering? Electrons are point particles; they
don’t have a complicated structure. Plus, they interact via the electric
and magnetic forces, which are well understood.

The famous Van de Graaff generator is a powerful electrostatic


accelerator of protons or ions. And while it’s used in some
research accelerators, you can also see it in science museums,
usually for hair-raising demonstrations.

Lecture 8
| Particle Accelerators: Schools of Scattering 81
Accelerating and Colliding Particles
wwHow do electrons interact with nuclei? There are several possibilities:

1 The electron can elastically bounce off the nucleus, leaving the
nucleus unchanged. In this case, we measure the charge distribution
of the nucleus.

2 The electron can excite the nucleus to a specific shell-model or


rotational state, and we can measure the energy, spin, etc., of
that state.

3 The electron can excite the nucleus to a giant resonance, where the
protons and neutrons oscillate back and forth.

4 The electron can interact with single nucleons or quarks, giving us


information about the nucleon or quark distribution in the nucleus.

wwWe also hit the nucleus with other particles, such as protons or other
nuclei. We do that to try to understand the nucleon-nucleon force and
to make new nuclei. Also, the proton and other ions interact strongly,
so we can learn about the strong force.

wwWe don’t use neutrons that often because they don’t live long enough—
only about 15 minutes. They’re very difficult to accelerate and steer,
because they have no charge. So, we predominantly use protons
and ions.

wwHow do we accelerate protons and ions?


The first technique is called the cyclotron, One of the biggest
which is a large magnet with a circular pole cyclotrons is the
face. There are 2 parts of it—2 halves that 88-Inch Cyclotron
look like Ds—and an alternating voltage is at Berkeley.
applied to the gap between the 2 Ds.

82 Nuclear Physics Explained


wwWe inject an ion near the gap, and the voltage accelerates the ion
toward the other D. The magnetic field then bends the ion trajectory
around in a circle, and the radius of that circle is the mass times the
velocity divided by the charge and the magnetic field: R = mv/qB.
As the velocity increases, the radius gets bigger, but the time it takes
to go around the half circle doesn’t change. By the time the particle
comes back to the gap, the voltage has reversed, and the voltage now
accelerates the particle the other way. Then, this process is repeated.
The bigger the cyclotron, the more energy we get. Today, cyclotrons are
mostly used to accelerate protons for proton cancer therapy.

wwThe problem with cyclotrons is they have a maximum useful energy.


There are 2 causes of this. First, the magnets just get too big. Second,
the cyclotron relies on the fact that the proton takes the same amount
of time for each of its semicircles—whether it’s a small one at low speed

Lecture 8
| Particle Accelerators: Schools of Scattering 83
or a big one at high speed—and that timing gets messed up when
the proton starts traveling too close to the speed of light and special
relativity complicates things.

wwWhile a cyclotron has a fixed magnetic field and lets the radius motion
of the proton get bigger and bigger, a synchrotron has a fixed radius
and increases in magnetic field. A synchrotron is a ring of bending
magnets—plus focusing magnets and accelerating cavities and the
same radio frequency cavities as electron accelerators. As the particle
accelerates and gets closer to the speed of light, we increase the
magnetic field to match.

wwAt Fermilab, they extract the high-energy beam periodically. At the


European Organization for Nuclear Research (CERN), they have
2 counter-rotating beams where they collide particles. They strip
electrons from protons or other nuclei, accelerate them first in a linear
accelerator and then in 2 synchrotrons to boost up the energy. Then,
they inject them into the ring of the Large Hadron Collider and
accelerate them up to 12 trillion electron volts. A huge radius is needed
to achieve these huge energies, because it’s limited by the maximum
magnetic field that we can make with electromagnets. That’s why
CERN’s Large Hadron Collider stretches for tens of miles.

CERN’s Large Hadron Collider

84 Nuclear Physics Explained


wwWith these accelerated particles, we measure how protons and neutrons
interact by elastically scattering protons from other protons and from
neutrons. We also use protons to study the nucleus. We aim protons
at a target containing an interesting isotope and count the number
of particles emitted at a certain angle and energy. We measure cross
sections, which are the number of particles that are emitted at a
certain angle and energy divided by the number of beam particles
hitting the target and the areal density (number of nuclei per square
centimeter) of the target. We measure cross sections in units called
barns which are 10−24 cm2, or 10−28 m2. A barn is a measure for the
effective area of the target for the reaction we’re interested in.

wwWe use barns to measure the probability of different interactions. For


example, we might aim a proton beam at a calcium-48 target and
measure the protons that come out with elastic scattering, leaving the
calcium-48 nucleus unchanged. Or we might measure the protons that
come out, leaving the calcium excited to a specific state. Or we might
measure the neutrons that come out, leaving a different nucleus behind.

wwWhat about bigger collisions? When a nucleus hits another nucleus,


there is a large electric repulsion energy, which is the charge of the first
nucleus times the charge of the second nucleus divided by the distance
between them. This is typically about 10 million electron volts.

wwAt energies slightly higher than this barrier, the cross sections are
large—about 0.1 barn. There are big changes to those nuclei. These are
big collisions. We can transfer up to 20 nucleons from one nucleus to
the other, or maybe they could even stick together. We could transfer
huge amounts—50 units—of angular momentum. When we’re doing
this, we’re studying nuclei in extreme conditions. We’re looking for
new isotopes, ultraheavy elements, and ultrahigh spin. We’re looking at
nuclei far from the valley of stability.

Lecture 8
| Particle Accelerators: Schools of Scattering 85
Making New Elements and Isotopes
wwHow do we make new elements? For example, we can accelerate
calcium-48 to a tenth of the speed of light, but because it has such a
heavy mass, it’s actually not that much energy—only about 10 million
electron volts for each nucleon. Then, we can collide it with one of the
heaviest nuclei that’s stable enough to make a target of: americium-241.
We often vary the speed of the projectile and target. Then, because
americium has 95 protons and calcium has 20 protons, we look for
element-115 in a mass spectrometer—with a much lower velocity, about
0.02 times the speed of light—and measure its alpha decay with a
time of, in this case, about 0.2 seconds. Element-115, moscovium, was
discovered in Dubna and confirmed in a lab called GSI in 2013.

wwHow do we make and then study new isotopes? We smash nuclei


into other nuclei, use a mass spectrometer to separate out the desired
new isotope to study, reaccelerate these nuclei, smash them into other
targets, and see what comes out.

Supplements

READINGS

Borel, “Making New Elements.”


FRIB Users Organization, FRIB.
Henley and Garcia, Subatomic Physics, chap. 2.
Lilley, Nuclear Physics, sections 6.8 and 4.6.
National Research Council, Nuclear Physics, p. 30–56 and 80–104.

86 Nuclear Physics Explained


QUESTIONS

1 What would be the best particle and energy to collide with a proton
to study the distribution of quarks inside the proton?

2 What would be the best particle and energy to collide with uranium
to make an ultraheavy nucleus?

ANSWERS

1 The best particle would be an electron, because it has no structure itself.


The best energy would be as high as possible, to get the smallest-possible
wavelength to resolve the smallest structures in the proton. The smaller the
wavelength, the more detail we see in the proton.

2 The heaviest known nucleus as of 2017 was oganesson, with 118 protons. If
we wanted to make element 119, then we would want to accelerate a nucleus
with 119 − 92 = 27 protons, or cobalt. Because ultraheavy elements have a
large neutron excess, we would want an isotope with a lot more neutrons
than protons. However, the only stable isotope of cobalt is cobalt-59, with
only 32 neutrons. If instead we use nickel, we can use nickel-64, with
28 protons and 36 neutrons. We want to use the lowest-possible energy that
will allow the cobalt or nickel projectile nucleus to barely overcome the
electrical repulsion and fuse with the uranium target nucleus.

Lecture 8
| Particle Accelerators: Schools of Scattering 87
09 DETECTING
SUBATOMIC
PARTICLES

P
a rticle detectors see the subtle traces
left behind as high-energy particles pass
through them. In this lecture, you will learn
how to measure individual attributes of
single particles. u

88
High-Energy Particles
wwTo detect high-energy particles, we use the tiny amount of energy they
leave behind as they pass through matter. Charged particles interact
with atomic electrons. They excite and ionize them. Then, we can
either collect the ionized electrons as an electrical signal or detect the
light emitted as the electrons recombine with their atoms or deexcite.

wwWhat can we learn from the amount of energy left behind by these
charged particles? The slower the particle moves, the more it interacts,
and the more energy it leaves behind: energy ∼ 1/v2. Particles with
more charge deposit more energy: energy ∼ z2.

wwSurprisingly, the mass of the particle doesn’t matter. Protons,


deuterons, and even tritium nuclei deposit the same amount of energy
as each other. It only depends on the velocity and the charge. It also
depends on the material they’re passing through. They’ll deposit
the most energy per gram of hydrogen, and then helium, and then
everything else, because it depends on the number of electrons in
that material.

wwWith neutral particles, we wait until they hit something and then
detect the charged particles that are knocked out. High-energy
photons—for example, gamma rays between 0.1 and 10 million
electron volts—are about a million times more energetic than the
photons from the Sun. These high-energy photons can be absorbed on
an atom and knock an electron loose. This is called the photoelectric
effect. Or they can bounce off an electron in the atom, and the electron
will recoil.

wwWith the photoelectric effect, we get all of the energy of the gamma
ray. When the electron recoils, we get just some of the energy of the
gamma ray. Really-high-energy photons, above 10 million electron
volts, will make an electromagnetic shower.

Lecture 9
| Detecting Subatomic Particles 89
wwNeutrons travel through material until they hit a nucleus and knock
out one or more protons. The typical interaction distance is about
30 centimeters (about 1 foot) of plastic (or water or people), or about
4 centimeters (about 1.5 inches) of iron, or about 2.5 centimeters (about
1 inch) of lead. We then detect the energy left behind by the protons
that were knocked out by the neutrons as they travel.

Scintillator Detectors
wwTo detect such a tiny flash of light, we use scintillator detectors, which
detect the light from the ionized or excited electrons that were knocked
loose in the material when they recombine or deexcite. These are
higher-tech versions of the zinc-sulfide screens used by Hans Geiger
and Ernest Marsden.

wwThere are 2 main types


of scintillator detectors:
inorganic crystals or
organic scintillators.
Inorganic crystals, such
as sodium iodide or
germanium, have really
good energy resolution
for measuring the
energies of gamma
rays—high-energy
photons. But they’re
much too slow for some
purposes; they can
take up to a millionth
of a second to produce
the light.
Scintillator at Jefferson Lab

90 Nuclear Physics Explained


wwTo have a much faster scintillator, we use organic scintillators that are
common plastics that have been doped with a special material. When
a high-energy charged particle passes through the plastic, it deposits
some energy, which goes into jiggling the atomic electrons. When those
atomic electrons recombine, they give off tiny flashes of light.

wwThe light is emitted in all directions. Light emitted within 45° travels
toward the photomultiplier tube on the end. Light at larger angles
escapes the scintillator. If the scintillator is longer than it is wide, the
light totally internally reflects from the edges as it travels to the tube.
Then, the light hits a photocathode on the tube and knocks electrons
loose (about 1 electron per 4 photons).

wwThe electrons are then amplified by the photomultiplier tube and


emerge as an electronic signal. The electrons from the tube are then
accelerated by the voltage difference from the tube to the first dynode,
or the first part of the multiplying structure, where they knock more
electrons loose. This process is repeated for 6 or 8 or 12 dynodes,
causing an avalanche of electrons, which is finally accelerated to the
anode, where the signal is read out.

Lecture 9
| Detecting Subatomic Particles 91
Geiger Counters and Wire Chambers
wwA Geiger counter can be used to detect radiation—such as alpha, beta,
and gamma radiation and cosmic rays—by chirping when radiation
passes through the sensitive part of it. It consists of a gas-filled tube at
ground and a thin wire that runs down the center of the tube and is at
positive high voltage. A charged particle or a photon passes through the
tube and knocks electrons loose from the atoms of gas. Those electrons
drift toward the central wire, getting amplified as they go, resulting in
a big electrical signal and a click of the wire.

wwThe great thing about Geiger counters is they let you count radiation
and hear how much radiation there is. The problem is they can’t count
that quickly. It can’t really count more than 1000 or a few thousand
times per second. And in modern physics experiments, we need
detectors that can count hundreds of thousands or even millions of
times per second. And it only covers a very small area—just the area of
the tube.

92 Nuclear Physics Explained


wwWire chambers are bigger, faster, better Geiger-type counters. Where
a Geiger counter has 1 wire passing through a cylinder, a wire chamber
contains several field wires at negative high voltage and 1 sense wire
at positive high voltage. The field wires function like the cylinder. A
charged particle passes through the gas and knocks electrons loose, and
the electrons drift toward the sense wire. When they get close enough
to it, the electric field is big enough in between collisions with the gas
molecules, and the electron gains enough energy that it can knock
another electron out of the gas molecule, resulting in an avalanche and
a detectable signal.

wwThere are 2 differences between Geiger counters and wire chambers.


First, the amplification is not as big in a wire chamber, so we don’t
get a click or chirp; instead, we get a signal we have to read out by
computer. But because we don’t get that huge signal, we can have a
lot more of them, so the process can go a lot faster. Second, instead
of having a tube with a wire in it, wire chambers have several field
wires surrounding a sense wire. A full-sized chamber might have a few
thousand sense wires and a few thousand field wires around them, so
we can cover huge areas.

Lecture 9
| Detecting Subatomic Particles 93
wwTo get more information, we can place scintillators after wire chambers
to measure the arrival time of charged particles. Scintillators can
measure the arrival time of a charged particle very precisely—to better
than a nanosecond. Specifically, a scintillator can measure how long
it took for the signal to get from where the charged particle passed
through the wire chamber to the sense wire.

wwElectrons are drifting to the sense wire very slowly. The drift time is
the difference between the arrival time of the charged particle and
the sense wire signal time. This is proportional to the drift distance,
which is how far away the charged particle passed from the sense
wire. Drift distance can be measured to a fraction of a millimeter, so
we can measure the track of the charged particle passing through the
wire chamber very precisely. And we can do this over a huge area and
thousands—or tens or hundreds of thousands—of times per second.

Shower Counters
wwTo measure the energy of a particle, we use a shower counter. It
measures the energy of electrons or photons. And just as we can convert
mass to energy—an electron and an antielectron combine to produce
energy—here we can reverse the process. A high-energy photon comes
in, and if it passes through something heavy, such as lead, some of the
time it’s going to pair-produce. In other words, instead of an electron
and an antielectron combining and annihilating to produce photons,
the photon will pair-produce to make an electron and a positron (the
antiparticle of the electron).

wwThe other thing that could happen is an electron or a positron passing


through the lead will radiate a photon. If you start with a photon,
after passing through a certain amount of lead, you end up with an
electron and a positron. If you start with an electron, you end up with

94 Nuclear Physics Explained


an electron and a photon. If you start with a positron, you end up with
a positron and a photon. We started with 1 particle; now we have 2.
Eventually, this process results in a whole shower of particles.

wwThe electrons and positrons end up depositing all of their energy in


the scintillant. By measuring this energy, we can reconstruct the initial
energy of a photon or electron.

Cherenkov Counters
wwNothing can go faster than the speed of light in vacuum. But the speed
of light in material is slower than the speed of light in vacuum. The
speed of light in water is 30% slower; the speed of light in air is 0.1%
smaller. So, if a particle traveling through that material is going faster
than the speed of light in the material, it gives off an electromagnetic
boom, or a tiny flash of light.

wwThe blue glow in nuclear waste pools comes


from these flashes of light, called Cherenkov
If something, such light or Cherenkov radiation, after Pavel
as a supersonic Cherenkov. Cherenkov counters can
airplane, is going measure the velocity of a particle. There is
faster than the a threshold Cherenkov counter that detects
speed of sound that flash of light—that electromagnetic
in air, it makes boom—if the particle traveling through the
a sonic boom. Cherenkov counter is faster than the speed
of light in the material. There are also ring
imaging Cherenkov counters that measure
the size of the cone of light emitted by the
particle. The faster the particle moves, the
wider the cone is.

Lecture 9
| Detecting Subatomic Particles 95
Cherenkov radiation glowing in Idaho National Laboratory’s
Advanced Test Reactor core

Supplements

READINGS

Henley and Garcia, Subatomic Physics, chap. 4.


Leo, Techniques for Nuclear and Particle Physics Experiments.
Lilley, Nuclear Physics, chap. 6.

96 Nuclear Physics Explained


QUESTIONS

1 Why is it so much easier to detect high-energy charged particles than


neutral particles, such as neutrons and photons (gamma rays)?

2 How could Geiger and Marsden detect alpha particles with their
naked eyes?

3 Why do we use photomultiplier tubes to multiply and amplify the


signals emitted by scintillators?

ANSWERS

1 This is because charged particles interact with large numbers of the electrons
in the material that they pass through. They ionize some atoms, knocking
electrons loose, and they excite other atoms, exciting electrons to a higher
energy state. We can either collect and amplify those knocked-out electrons
to make a detectable electric signal, or we can collect the light emitted by
the deexcitation of the excited electrons. Neutral particles do not interact
with those atomic electrons. For neutral particles to be detected, they need
to have a hard collision with an atom or an atomic electron and knock out
a charged particle. We can then detect the charged particle that is knocked
out by the photon or neutron.

2 The alpha particles hit a zinc-sulfide screen, which gives off light when hit.
Alpha particles interact more strongly with material and deposit much more
energy than electrons, so they produce much more light when they hit a
zinc-sulfide screen. The human eye is also extremely sensitive and can detect
very faint light flashes when completely adapted to the dark.

3 We use photomultiplier tubes for 2 reasons: to amplify very faint signals that
could not have been seen with the naked eye and to convert the flash of light
into an electrical signal that we can record in a computer.

Lecture 9
| Detecting Subatomic Particles 97
10 HOW TO
EXPERIMENT
WITH NUCLEAR
COLLISIONS

T
 is lecture is about reconstructing nuclear
h
collisions. To measure the momentum
and the type—such as electron, proton,
pion—of the knocked-out particles after the
collision, large magnets are combined with
particle detectors to make spectrometers.
The particles are passed through large
magnets, which bend the trajectories of the
particles to determine their momentum.
Then, the positions of the particles—and
hence their momentum—are measured
with detectors such as wire chambers. The
type of each particle is measured using
detectors such as scintillators, Cherenkov
counters, and shower counters. u

98
Spectrometers
wwSpectrometers detect and measure properties of the particles that are
knocked out in the collision between an electron and the nucleus,
for example. A spectrometer needs to have 2 things: a dipole magnet
to spread out the momentum of the particles and detectors to detect
those particles. Lower-momentum particles traveling through the
dipole magnet will be bent more; higher-momentum particles will be
bent less. The detectors will be able to measure the positions of those
particles to determine their momentum.

Spectrometers in experimental
Hall C at Jefferson Lab measure
what happens when electrons
collide with nuclei.

Lecture 10
| How to Experiment with Nuclear Collisions 99
wwDrift chambers are used to detect and measure particle positions. Each
drift chamber consists of many very fine wires sandwiched between
2 very thin aluminized Mylar foils with gas in between and high
voltage on the wires. When the particle that we want to detect passes
through the gas, it knocks electrons loose. The high voltage causes
those electrons to drift toward the wire and then to be amplified, so
we measure an electrical signal. By determining which wire saw the
signal and how long it took those electrons to drift from where they
were knocked out of the gas to the wire, we can measure the position
of the particle that passed through the drift chamber to a fraction
of a millimeter. We can trace a particle’s trajectory back through the
magnetic field to know its momentum and angle as it left the target.

wwOnce we know the momentum of the particle, we need to figure out


what type of particle it is (electron, pion, proton, etc.). We need other
detectors to do this. The first one is a scintillator detector, which gives
off a tiny flash of light when a high-energy charged particle passes
through it. We detect that tiny flash of light and can measure its time
to within a fraction of a billionth of a second. That helps tell us the
speed of the particle.

wwAfter the particle passes through the scintillator, the next detector is a
Cherenkov counter, which gives off a tiny flash of light when a particle
traveling through it is traveling faster than the speed of light. The
Cherenkov counter helps tell us whether the particle was an electron or
something else.

wwThe last detector is an electromagnetic calorimeter that also helps


distinguish between electrons and other particles. An electron that
enters the calorimeter will start an electromagnetic shower and will
deposit a lot of its energy in the calorimeter. Other particles traveling
through the calorimeter will deposit much less energy, and from the
amount of energy deposited, that will also help tell us whether it’s an
electron or something else.

100 Nuclear Physics Explained


wwTypically, we want to detect the electron that bounced off the nucleus
and some other particle. To determine the identity of the other particle,
we use the relative time that the particle arrived in 2 spectrometers.
The 2 types of spectrometers that are used to look at the particles
coming out of collisions are small-aperture spectrometers and
large-acceptance spectrometers.

wwThe small-aperture spectrometer can typically detect 1 particle


at a time, and the particle has to go in a particular angle and in
a particular range of momentum. When we want to detect more
particles or over a wider range of angles or of momentum, we use a
large-acceptance spectrometer.

wwA small-aperture spectrometer just needs a magnetic field in a limited


region to analyze the particles that go into it. A large-acceptance
spectrometer needs a magnetic field over a huge volume because it’s
detecting particles over a wide range of angles and of momentum. That
means that the magnetic field is going to be weaker than the magnetic
field in a small-aperture spectrometer, which also means that our
ability to measure momentum won’t be as good in a large-acceptance
spectrometer. We also need bigger detectors with large-acceptance
spectrometers because the detectors have to cover a much wider area.

wwSo, why would we use a large-acceptance spectrometer? We might be


studying heavy ion collisions, where a lead nucleus hits a lead nucleus
and dozens or hundreds or thousands of particles come out in all
directions and we want to detect all of them. Or we might be at an
electron scattering lab and an electron hits a nucleus and we want to
detect more than 2 particles. Alternatively, we might want to do a
number of experiments at the same time; a range of different scientists
can take the same data with the same beam on the same target and
analyze it for different reactions.

Lecture 10
| How to Experiment with Nuclear Collisions 101
wwWith a small-aperture spectrometer, only the particles going out at a
small angle in a narrow range of momentum make it from the target
to the detectors. That means it can use a really intense beam and a
really thick target and make lots of nuclear collisions and just pick out
the ones that we are interested in, because only a tiny fraction of those
particles make it up to the detectors. (The fact that a small-aperture
spectrometer can use a very intense beam also means that it can
measure small things very precisely, which is a great benefit.) With a
large-acceptance spectrometer, we’re looking at just about all of the
particles knocked out in the collisions, so we have to turn the beam
intensity way down to be able to handle everything.

wwJefferson Lab has 2 large-acceptance spectrometers: one in


Experimental Hall D and one in Experimental Hall B known as the
CLAS, or the CEBAF Large Acceptance Spectrometer.

CEBAF Large Acceptance


Spectrometer

102 Nuclear Physics Explained


wwWith these spectrometers, the electron beam hits a target in the center
of the solenoid, which provides a magnetic field. Every now and then,
one of the electrons hits one of the nuclei in the target. It bounces off
and knocks other particles out. Lower-momentum particles go in all
directions. If they go in large angles—from 45° to 135°—we detect
them in the detectors of the central detector and measure their position
and time to get their momentum and what kind of particle they are.

wwHigher-energy particles go forward. They go through 3 layers


of drift chambers before, in the middle, and after the magnetic
field from the superconducting torus magnet. Then, they pass
through the Cherenkov counters, the scintillators, and the
electromagnetic calorimeters.

wwUsing all of this information, we can measure the momentum of each


particle in the event and we can measure what kind of particle it is—
whether it is an electron, a pion, a kaon, or a proton. And we can use
that information to completely reconstruct all of the particles emitted
in an electron-nucleus collision.

How Do We Do an Experiment?
wwTo do an experiment, we first have to come up with an idea. There
has to be some question that we want to answer. For example, how
does the motion of protons in the nucleus depend on the number of
neutrons? To answer this question, we have to figure out numerous
things, including what we can measure, what we already know, why
this is important, how much beam time it takes to measure this, which
experimental hall to use, what beam energy we want to use, and how
much time it will take to do the measurement. Then, we write all of
this up in an experimental proposal.

Lecture 10
| How to Experiment with Nuclear Collisions 103
wwNext, we defend the proposal to the Program Advisory Committee,
which is a committee at Jefferson Lab with outside nuclear physicists
who read and discuss proposals, listen to presentations, talk
about the proposals with the people proposing them, and decide
which experiments get beam time at Jefferson Lab. Only about
1/3 of experiments are actually approved by the Program Advisory
Committee. An experiment can range from requiring 4 days
of beam time to 200 days of beam time, depending on what is
being measured.

wwOnce the experiment is approved, the real work starts. We do many


computer simulations to figure out what we’re going to see and exactly
how we want to measure it. We make a detailed run plan: Where do
we want to place the spectrometers? What momentum particles do we
want to measure? How long do we need to measure in each location?
How many events do we want to accumulate in each location? If
necessary, we’ll build new detectors and new targets.

wwWe’re now ready to actually perform the experiment. We turn the


beam on, we turn the detectors on, and the electron beam hits a target.
Hundreds or thousands of times—maybe 10,000 times—per second,
an electron in the beam hits a nucleus in the target. We get particles in
our spectrometers.

wwWe want to measure specific events. To do that, we set up an


experimental trigger that tells the computer to read out specific events
from the detectors. The computer takes an event, such as a collision,
and writes it to tape, accumulating not just thousands or millions,
but sometimes billions, of events. The computer then reads out all of
the detectors.

104 Nuclear Physics Explained


wwThe experiment doesn’t stop. Beam time is a precious resource, so
the beam runs 24 hours a day, 7 days a week. There are at least a few
people in the experimental control room all the time. After a week, or a
month, or 6 months of taking data, we have to analyze the data.

wwFirst, we have to calibrate all of the detectors. When we read out the
detectors, we get signals, such as time signals. We have to convert the
time signals we get from the drift chambers to positions in the drift
chamber, which have to be converted to momentum and angles of
particles. We get pulsate information: How big was the signal from the
Cherenkov counter? We have to convert that to particle information:
Was it an electron or wasn’t it an electron?

wwAfter months spent calibrating the detectors, we can finally analyze


the data. We use information about the particles, such as energy
and momentum, to reconstruct what happened before and during
the collision.

wwDuring the experimental process, we have to make sure that we don’t


fool ourselves. As scientists, we need to be very careful: If we know
the answer that we want to get from an experiment, we’re more likely
to get that result, so we have to set up safeguards to make sure that
doesn’t happen.

wwOnce we have analyzed the data and have the result, we have found out
something new about the universe. We know something that nobody
else in the world knows yet. The final step is to publish our results and
go on to do the next experiment.

Lecture 10
| How to Experiment with Nuclear Collisions 105
Supplements

READINGS

Grupen, Particle Detectors.


Jefferson Lab, “Jefferson Lab Virtual Tour.”

QUESTIONS

1 Why do we need bigger spectrometers to measure


higher-energy particles?

2 What determines the order in which we put detectors


in spectrometers?

ANSWERS

1 We measure the momentum of a charged particle by measuring how much


its trajectory deflects when it travels through a magnetic field. The radius of
the curvature of the particle’s trajectory is proportional to its momentum. A
more energetic particle has higher momentum and therefore a larger radius
of curvature. If we double the momentum of a particle, then we need to
double the distance it travels in a magnetic field to make it deflect the same
amount. Therefore, we need bigger spectrometers with bigger magnets to
measure higher-energy particles.

2 We put the detectors that measure particle positions first (e.g., wire
chambers) so that the measured position is unaffected by particle
interactions in other detectors. We then put the detectors with least material
next so that the particles are most likely to pass through undisturbed and
interact again with subsequent detectors. Thus, we put Cherenkov counters
next, which contain only gas and very thin mirrors, and then relatively thin
plastic scintillators, and then thick calorimeters.

106 Nuclear Physics Explained


SCATTERING
NUCLEONS
11
IN SINGLES
OR IN PAIRS

I
 this lecture, you will learn how we scatter
n
electrons from the nucleus to learn more
about it. Specifically, you will learn about 3 kinds
of experiments: where we detect the scattered
electron, where we detect the scattered
electron and the proton it knocks out, and
where we detect the scattered electron,
the proton it knocks out, and a second
proton or neutron. u

107
Detecting the Scattered Electron
wwIn electron-scattering experiments, the electrons come down the
beam pipe, hit a target in the middle of the scattering chamber, and
then bounce off in all directions. The electrons are detected in the
high-resolution spectrometer in Hall A at Jefferson Lab, which has a
few quadrupoles to focus the electrons. Then, the dipole bends the
electrons up to the particle detectors in the shielding hut at the top
of the spectrometer. The higher-momentum electrons bend less; the
lower-momentum electrons bend more. We have now detected the
electron momentum in the high-resolution spectrometer.

wwTo do an experiment, we select the beam energy we want, the target—


for example, if we are studying carbon, we would use a thin slice of
graphite —and the angle of the spectrometer, which determines how
hard the electrons we’re detecting hit a carbon nucleus.

scattered electron (e–)


electron scattering
angle
incident electron (e)

momentum transferered
to nucleus (q)

108 Nuclear Physics Explained


wwFor example, if the spectrometer is in a very forward angle, then the
electron enters and scatters just a tiny bit, and it doesn’t hit the nucleus
very hard. On the other hand, if we put the spectrometer at a much
larger angle, then the electron has to come in and bounce off at a large
angle. In that case, it transfers much more momentum to the nucleus.
The momentum transfer (q) is what we’re interested in.

wwNext, we want to study how much energy the electron transfers to the
nucleus. Where does the energy go? The energy goes into 2 things: the
kinetic energy of the recoiling particles and the excitation energy of the
nucleus. How much is the minimum amount of energy we can transfer?

wwThe electron hits the carbon nucleus, bounces off, and transfers
momentum. This is kind of like a ping pong ball hitting a tennis ball:
The tennis ball is going to recoil; it’s going to be moving.

wwKinetic energy is 1/2 times mass times velocity squared: KE = 1/2mv2.


But because we’re dealing with momentum, kinetic energy is 1/2 times
the momentum transfer squared divided by the mass of the nucleus:
KE = p2/2m A = q2/2m A .

wwThe mass of the nucleus is very big, so we’re not transferring a lot
of energy to the nucleus when the electron just hits it and it recoils
elastically. If we transfer more energy than that, then we can make the
nucleus ring like a bell—we can excite the nucleus to excited states,
which are the shell-model states. If we provide even more energy to the
nucleus, we can make the nucleus vibrate as a whole—which are giant
resonances. If we transfer even more energy, we can knock a single
proton or neutron out of the nucleus.

wwWhat happens if the electron, instead of scattering from an entire


carbon nucleus, just scatters from a single proton? If we put in a
hydrogen target and look at the electrons scattering from the proton,
the electron comes in and we detect it at a certain angle, know the
momentum transfer, and therefore know the kinetic energy (1/2 times

Lecture 11
| Scattering Nucleons in Singles or in Pairs 109
the momentum transfer squared divided by the mass). The mass of
1 proton is 12 times smaller than the mass of carbon, so 12 times as
much energy is transferred.

wwThe elastic peak is all the way on the left of the energy transfer
diagram, and the discrete resonances and giant resonances are to the
right of the elastic peak. There is also another peak corresponding to
scattering from a proton or a neutron in the nucleus.

110 Nuclear Physics Explained


wwWe want to know the number of electrons that scattered elastically,
that scattered to each of those different discrete states that excited the
nucleus to a giant resonance, or that scattered elastically from a proton
or a neutron in the nucleus.

wwHowever, when an electron scatters from a proton or a neutron in the


nucleus, instead of scattering from a proton in hydrogen, that proton
or neutron in the nucleus is moving, so we can use information about
what the quasielastic peak (because it’s not quite elastic) looks like for
electron scatter and for proton in carbon, versus what the elastic peak
looks like, when the electron just scatters from a proton in hydrogen.

wwWhen the number of electrons is plotted, we see that they transfer


different amounts of energy to the nucleus. Instead of a very sharp
peak, corresponding to electrons scattering from protons that are
moving, we see a broader peak. The peak is also shifted a little because
the protons and neutrons are bound in the nucleus—there’s a binding
energy. The peak is broad because the protons or neutrons in the
nucleus are moving. How does that happen?

wwIf you hit a proton, you transfer momentum to the proton, and the
proton is already moving in that direction. In that case, you’re going to
end up with a much faster proton; it will have higher momentum. The
kinetic energy of that proton is its momentum squared divided by twice
the mass: q2/2m. You can transfer a lot more energy to it.

wwOn the other hand, if the proton is traveling toward the electron,
then the electron transfers the momentum and you end up with a
slower-moving proton. This means that when it comes out of the
nucleus, the momentum squared divided by twice the mass (q2/2m)
is much smaller. You’re transferring less energy. And that’s going to
broaden the peak.

Lecture 11
| Scattering Nucleons in Singles or in Pairs 111
wwThe simplest model of protons and neutrons moving in the nucleus is
the Fermi gas model, in which these particles move with all possible
momenta—from 0 up to a maximum. If the proton or neutron is
moving away from the electron, then it will have a greater momentum
when it leaves the nucleus—more energy. The momentum transfer plus
the Fermi momentum gives the maximum momentum, which is the
maximum energy the electron can transfer.

wwOn the other hand, if the electron transfers its momentum and the
proton is moving exactly opposite of the electron, then the proton
momentum is the momentum transfer minus its initial momentum.
And if it’s moving with the Fermi momentum, that gives us the
minimum kinetic energy of the proton or neutron and the minimum
energy transfer.

wwWhat about for protons and neutrons in between? If they are moving
perpendicular to the momentum transfer, then they end up with
a momentum that is at a different angle. There are many protons
and neutrons that come out with about the same momentum as the
momentum transfer.

wwVery few electrons transfer the minimum possible energy, and very few
electrons transfer the maximum possible energy. But many electrons
in the middle transfer the average kinetic energy. And that gives us the
shape of the quasielastic peak. It’s rounded like a parabola.

wwIf we know the probability for an electron to scatter from an individual


proton and from an individual neutron, then the total area of the
quasielastic peak should correspond to the total probability for an
electron to scatter from those individual protons and neutrons.

wwThe Fermi gas model does a very good job of describing the data all
the way from a really light nucleus—such as lithium, with only 6 or
7 protons or neutrons—all the way up to lead, with 208 protons
and neutrons.

112 Nuclear Physics Explained


Detecting the Scattered
Electron and the Proton
wwIn addition to detecting the electron, we want to detect the proton, too.
Detecting the neutron is usually too difficult, so we measure protons
and assume that neutrons do what protons do. Protons are charged
particles, which means that we can use spectrometers and detectors to
measure them.

The first coincidence experiment—in which the electron and


the proton were detected in coincidence—was done with
zinc-sulfide screens, which gave off little flashes of light
when a particle hit it. Physicists looked at those screens in a
darkened room, with dark-adapted eyes, to see those flashes.
To make sure that the particles were detected at the same
time and to avoid having the physicists’ results influence
each other, the physicists had a clicker that was connected
by a wire to somebody in another room, who wrote down
when each of the clickers were pushed, indicating that a
particle was detected.

wwTo detect the electron and the proton, there are 2 high-resolution
spectrometers in Jefferson Lab’s Hall A. The electron hits the target—
for example, carbon—and we position the electron spectrometer to
detect the electrons that bounce off of it at a particular angle and
position the proton spectrometer at the angle where the electron
transfers momentum. Alternatively, we could do the same experiment
in Hall B, with the large-acceptance spectrometer, which will detect
the proton and the electron in the same spectrometer at the same time.

Lecture 11
| Scattering Nucleons in Singles or in Pairs 113
wwHow do we know that the electron and the proton came from the
same interaction? We measure the difference in the arrival time of the
electron in its spectrometer and the arrival time of the proton in its
detectors. For each event, we plot the time difference of the 2 particles.
When the electron and the proton came from the same event, there is a
large spike on the time-difference spectrum.

wwNext, we want to reconstruct the proton’s initial state. To do this, we


plot the proton energy on the vertical axis and the electron energy on
the horizontal axis. We make this new quantity called the missing
energy, which is the energy transfer minus the proton kinetic energy:
Emiss = Ebeam–E’ − Tp, where Tp is the proton energy, E’ is the electron
energy, and Ebeam is the beam energy.

wwThe difference between the momentum that the electron hit the proton
with and the momentum that the proton comes out of the nucleus with
is the missing momentum, and it is very close to the momentum that
the proton had before we hit it.

114 Nuclear Physics Explained


Detecting the Scattered Electron,
the Proton, and a Second Nucleon
wwDistributions of missing momentum are beautifully well predicted
and calculated. Energy distributions are also very well predicted and
calculated. But when we compare the number of protons that we see
with the number of events that the calculations predict that we should
see, we only see 2/3 the number of protons that we expect.

wwTo look for the missing protons, we move the proton spectrometer
to a larger angle, and then to a larger angle, and so on, and look for
protons with larger and larger initial momentum. Then, we make a
plot for each of those spectrometer angles—each of those proton initial
momenta—and look at the missing energies.

wwWe see many events where there are very large missing energies, which
means that we have to be knocking out more than 1 proton or neutron.
To find those extra knocked-out nucleons, we need another detector.
We can either use a large-acceptance spectrometer that detects all of the
particles at once, or we can put a third large-acceptance spectrometer—
in addition to the electron spectrometer and the proton spectrometer—
in Hall A to detect the extra nucleon.

wwWe need to be able to detect both protons and neutrons because we


don’t know what’s carrying away that energy and momentum, so
we use a medium acceptance detector—a magnetic spectrometer
called BigBite—and instrument it to detect protons. We put layers
of scintillator detectors at the back, detect the protons that are bent
through the dipole, measure their momentum in position with the
scintillators, and figure out what their momentum and angle was.

Lecture 11
| Scattering Nucleons in Singles or in Pairs 115
wwBut we also have to detect neutrons. How do we do that? Because
neutrons aren’t affected by the magnetic field, they don’t interact very
much. That’s a disadvantage, but it’s also an advantage because it
means that we can put a neutron detector behind BigBite. The protons
enter BigBite and go upward to the detectors, while the neutrons go
straight through and hit the neutron detectors behind. The neutron
detectors are in the Hall A Neutron Detector (HAND).

116 Nuclear Physics Explained


wwHow do we reconstruct the momentum of the neutron? We can’t use the
magnetic field. We have to use timing. We know what time the electron
and the proton left the target; we know what time the neutron was
detected at the target. The time difference between when the interaction
occurred and when we detected the neutron tells us how long it took the
neutron to get from the target to the detector. We know the time and the
distance, from which we can measure the velocity.

wwTo find out which particles are carrying away the momentum and
energy, we place BigBite, with HAND behind it, in the direction of
the missing momentum. And it turns out that the missing momentum
is always carried by a single proton or neutron. This is very surprising,
because that momentum could’ve been carried by the nucleus as a
whole or by a bunch of protons and neutrons. The fact that it’s carried
by a single proton or neutron tells us that we’ve got a proton-neutron or
a proton-proton pair, where they’re moving at very high momentum in
respect to one another, the proton is knocked out, and the rest of the
momentum is carried away by the second proton or neutron.

wwAnother thing that’s surprising is that more than 90% of the time, that
second nucleon is a neutron, and about 5% of the time, the second
nucleon is a proton. This tells us that about 80% of the nucleons are at
low momentum in shell-model orbitals but that the other 20% are in
pairs that have high momentum.

wwThe fact that these pairs have high momentum tells us that they’re
at a short distance from each other. This tells us 2 things about these
short-range pairs: Their density is much higher because they’re much
closer to each other, and the quarks (which protons and neutrons are
made up of) are overlapping with each other.

wwThe shell model is still a useful approximation; it still describes about


75% of the protons and neutrons in the nucleus. But in the 21st century,
we’ve learned that short-range correlations are a very important
addition that need to be included.

Lecture 11
| Scattering Nucleons in Singles or in Pairs 117
Supplements

READINGS

Wilson, “Electron-Scattering Experiments Resolve Short-Range


Correlations among Nucleons.”

QUESTIONS

1 Can we make sure that an electron scatters elastically from a nucleus,


like we can make sure that a billiard ball scatters elastically from
another billiard ball?

2 Why are more scattering experiments done on carbon than on


oxygen, even though oxygen is a doubly magic nucleus?

ANSWERS

1 No. Electrons interact through quantum mechanics, and therefore we


can only predict what will happen probabilistically. When we collide an
electron with a nucleus, we cannot determine ahead of time whether it will
scatter elastically with the entire nucleus, excite the nucleus to a specific
excited state, scatter quasielastically with 1 nucleon in the nucleus, etc. All
we can do is aim a beam of electrons at a slice of material and then select
the electrons we are interested in afterward. For example, we can detect
electrons that scatter quasielastically by placing our electron spectrometer
so that it only detects electrons that scatter at a particular angle and a
particular momentum.

118 Nuclear Physics Explained


2 This is because it is much easier to make a target containing carbon than a
target containing oxygen. A carbon target can be a thin slice (0.1 millimeter)
of graphite, which is pure carbon and has a high enough melting point that
it is unaffected by the electron beam passing through it. In addition, it is a
solid target with a constant thickness. A pure oxygen target would have to
be either a gas (and therefore too thin) or a liquid (and therefore cooled to
about −180°C or −300°F). A solid target containing oxygen would have to
contain other nuclei (e.g., beryllium oxide). At Jefferson Lab, physicists used
a complicated waterfall target that involved using the hydrogen in the target
as a calibration target.

Lecture 11
| Scattering Nucleons in Singles or in Pairs 119
12 SEA QUARKS,
GLUONS, AND THE
ORIGIN OF MASS

D
id you know that 99.9% of our mass—in
fact, 99.9% of the entire visible mass of
the universe—comes from the protons
and neutrons in the nucleus of the atom.
But only 1% of that mass comes from the
masses of the 3 up and down quarks that
make up the proton and the neutron. The
famous Higgs boson explains mass, but it
only explains the mass of those 3 up and
down quarks. Half of that missing mass
comes from antimatter. u

120
Studying the Proton and the Neutron
wwThe proton is easier to study than the neutron because it’s charged
and it’s stable. How do we know that the proton is not an elementary
particle? An elementary particle is something like an electron or a
quark that has zero size and no structure (as far as we know).

wwComposite particles are made of elementary ones. They have structure


and size. Protons, neutrons, and particles called pions are made of
quarks. Nuclei are made of protons and neutrons, which are made of
quarks. They all have structure and size.

wwAll particles have a wavelength, which is Planck’s constant divided by


the momentum: 𝜆 = h/p. The higher the momentum, the smaller the
wavelength. This affects quantum behavior, such as tunneling and
diffracting, but it’s different from the inherent size of the object.

wwHow do we know that the proton is not a point particle like the
electron, with zero size and no structure? The proton has a bigger
magnetic field than expected from just its spin. The discovery of
this was awarded the 1943 Nobel Prize in Physics to Otto Stern.
But we also measure the size of the proton directly. We use the same
technique—diffraction patterns in elastic electron scattering—to
measure the proton radius as we use to measure the nuclear radius.

wwWe scatter 188-million-electron volt electrons from protons at the W.


W. Hansen Experimental Physics Laboratory (HEPL) at Stanford. If
we plot how many electrons bounced off at different angles—the cross
section against the angle—there is a curve on the point proton, which
is the reference cross section if the proton had zero size. But the data
is smaller than that point proton curve at large angles. And we know
from diffraction that if the slit is bigger, the cross section decreases
faster. The bigger the slit, the narrower the pattern. The fact that the
data decreased faster tells us that the proton has size. This earned the
1961 Nobel Prize in Physics for Robert Hofstadter.

Lecture 12
| Sea Quarks, Gluons, and the Origin of Mass 121
wwHow big is the proton? We can measure the electron-proton cross section
and divide that by the point proton cross section to get the form factor. A
point proton would have a form factor of 1 at all angles. The form factor
decreases with the angle of scattering: The bigger the angle, or the higher
the energy of the electrons, the faster the form factor decreases. That tells
us that the radius of the proton is about 0.8 fermi, or 0.8 × 10−15 meters.

wwWhy do we care about the size of the proton? The size is related to
the strength of the quantum chromodynamic (QCD) force that holds
the quarks together: If the proton is bigger, the force is weaker; if the
proton is smaller, the force is stronger.

wwThe mass of the proton is also related to the strength of the QCD
force: A stronger force would probably give us a bigger proton mass; a
weaker force would probably give us a smaller proton mass. The strong
force—the force between protons and neutrons—derives from the
QCD force between the quarks.

122 Nuclear Physics Explained


wwLet’s compare the electric and magnetic sizes of the proton. When
an electron interacts with a proton, it can either interact electrically,
with the charge of the proton, or magnetically, with charges moving
in the proton. The electric interaction tells us about the position of
the charges; the magnetic interaction tells us about the motion of
the charges.

wwPrior to 1998, measurements of the ratio of the electric to the magnetic


form factor showed that it is flat as the electrons scatter to bigger and
bigger angles, meaning that the electric and magnet sizes of the proton
were the same. But starting in 1998, we found out that the ratio is not
flat. Are the electric and magnetic sizes different? Are there relativistic
effects? This was a huge surprise. This elastic scattering of particles
from other particles—the kind that Ernest Rutherford did with alpha
particles—is still interesting.

Lecture 12
| Sea Quarks, Gluons, and the Origin of Mass 123
wwFree neutrons decay, so we can’t just make a target of neutrons to put in
an accelerator. Instead, we measure scattering from deuterium, which
has 1 proton and 1 neutron, and subtract scattering from hydrogen,
which is what we get from scattering a proton. But we have to account
for the fact that the proton in deuterium is moving around.

wwWe expect a similar neutron and proton structure. Both of them


contain 3 quarks: The 3 quarks in the proton are up, up, down; the
3 quarks in the neutron are up, down, down. The charge is different,
but the electromagnetic force is much weaker than the QCD force, and
the difference in mass of the neutron and proton is only about 0.1%.

wwBut the neutron has no charge. How can we scatter electrons from it?
It has no total charge, but the quarks in it are charged, so—just like a
neutral atom has positive protons and negative electrons with a total
charge of zero—the neutral neutron has 1 up quark with a positive
charge of +2/3 and 2 down quarks with a negative charge of −1/3 each.
The total charge on the neutron is 0.

wwWhat’s the charge distribution of the proton and the neutron? The
neutron charge distribution is much smaller than the proton. The
neutron charge distribution is positive at small radius and negative
at large radius. The neutron looks kind of like there’s a proton in the
middle, circled by a negative pion.

124 Nuclear Physics Explained


Excited States
wwThe nucleus has excited states. Does the proton have excited states,
too? Just like with the nucleus, where we plot the cross section versus
the energy, here we plot the cross section versus the missing mass. We
know the energy of a particle relativistically: energy squared is equal
to its momentum squared plus its mass squared. That means that the
mass squared is equal to the energy squared minus the momentum
squared, so the energy is equal to the mass of the proton plus however
much energy we gave it. The momentum of this excited state is equal
to however much momentum we gave the proton when we bounced an
electron off it. This tells us the excited mass of the proton.

wwWe look at the cross section, which is the probability of something


happening, and plot it versus the mass of whatever we scattered from.
We see bumps in the cross section corresponding to proton excited
states—just like we see bumps when we scatter electrons from the
nucleus corresponding to nuclear excited states. Unlike in the nuclear
case, where the excited states are really narrow, for the proton, the
excited states are really wide in energy.

Lecture 12
| Sea Quarks, Gluons, and the Origin of Mass 125
wwCan we see the quarks inside the nucleus? We need better spatial
resolution, shorter wavelengths, and higher-energy electrons to answer
this question.

wwThe Stanford Linear Accelerator Center (SLAC), which was built in


1962 and became operational in 1966, is a 2-mile-long accelerator
that is in a straight line because electrons lose energy when they go in
circles. Small-aperture spectrometers, similar to the ones at Jefferson
Lab, were used to detect electrons and other particles knocked out in
high-energy collisions.

wwIt turns out that the cross section has no bumps for larger masses. How
does that cross section change when you transfer more momentum
to the nucleus (when the angle gets bigger)? The probability of elastic
scattering decreases dramatically as more momentum is transferred (to
the nucleus or to the proton) or as the electron scatters at a bigger angle.

wwWhen we do elastic scattering at very large excitation energies of the


proton, the cross section is flat. That tells us that instead of scattering
from an extended object (so that you get diffraction patterns), there is
scattering from point particles in the proton. This tells us that there are
quarks in the proton. This discovery earned the 1991 Nobel Prize in
Physics for Henry Kendall, Richard Taylor, and Jerome Friedman.

Quarks and Gluons


wwThere are up and down quarks (more exotic particles have strange,
charm, top, and bottom quarks). The spin of the quarks is 1/2. They
have a fractional charge: The up quark is charged 2/3 while the down
quark is charged −1/3. They have 3 colors (red, green, and blue); this
is the QCD version of charge. The antiquarks have anticolors; each
antiquark has 1 anticolor.

126 Nuclear Physics Explained


wwThere are 2 ways to combine colors to make the color neutral. We can
add 3 quarks in a proton or neutron (red plus green plus blue gives
white), or we can combine a quark and an antiquark in a pion (i.e.,
green quark and antigreen antiquark).

wwHow do 3 quarks make a proton? A proton is up, up, down; a neutron


is up, down, down. What about other combinations of quarks? We
can look at possible combinations of up, down, and strange quarks:
1 up, 1 down, 1 strange makes
a particle called a lambda [Λ];
1 down, 1 down, and 1 strange
makes a particle called the sigma
[Σ−]; and 1 down, 1 strange, and
1 strange makes a particle called a
cascade [Ξ −].

wwQuarks were invented to explain


the multiplicity of these particles.
Nobody has ever seen a quark.

wwHow are these 3 quarks (the up, the up, and the down) distributed
in the proton? There are more up quarks than down quarks in the
proton—which is good—but there are an infinite number of quarks in
the nucleus. How can this be? What are these extra quarks (beyond the
up, up, down) in the proton?

wwThe QCD force is carried by gluons between the quarks. Quarks are
always exchanging gluons to keep in touch. There are always gluons
in the proton or neutron. But quark-antiquark pairs can only exist for
short periods of time. These are called virtual particles. So, the proton
is made up of 2 up quarks and 1 down quark, plus gluons that help
keep them in touch, plus virtual quark-antiquark pairs.

Lecture 12
| Sea Quarks, Gluons, and the Origin of Mass 127
wwHow many quarks are there in the proton? It depends on the
wavelength (the resolving power) of the electron or muon we are using
to study it. The shorter the wavelength, the more detail we see—more
quark-antiquark pairs and more gluons.

wwThere can be an infinite number of quarks and antiquarks, as long as


they carry very little momentum, because the momentum fraction has
to add up to 1. So, we distinguish between the valence quarks—which
are the 2 up quarks and the down quark that make it a proton—and
the sea quarks, which are the quark-antiquark pairs.

wwHow do we distinguish valence quarks from sea quarks? We have to


use other reactions to measure antiquarks. Antiquarks can only be sea
quarks. The total number of quarks comes from the 3 valence quarks
plus all the sea quarks.

wwThere are exactly twice as many up quarks as down quarks in the


valence quarks—which we know are 2 up quarks and a down quark.
But what kind of sea quarks do we have? There are lots of up and down
quarks in the sea, and they have a few million electron volts each.
There are some strange quarks, with 100 million electron volts. There
are very few charm quarks and almost no bottom quarks.

wwBut the quarks only provide half of the momentum of the proton. The
rest of the proton momentum is carried by the gluons.

wwThe total mass of the valence quarks—the 2 up quarks plus 1 down


quark—is only about 10 million electron volts, so it’s only a few percent
of the mass of the proton. That means that about 99% of each proton’s
mass is in the energy of the virtual sea quarks and gluons. About 50%
comes from sea quarks and about 50% comes from the gluons. This
comes from the details of QCD. Nucleons are 99.9% of the atomic
mass, which means that 99% of our mass comes from quark-antiquark
pairs and gluons. The biggest question in nuclear physics is why the

128 Nuclear Physics Explained


nucleus is made of protons and neutrons and not just quark soup. Why
do protons and neutrons keep their identities in the nucleus? There are
2 approaches to looking at nucleons in the nucleus:

1 Protons and neutrons are strongly interacting, often overlapping


particles, so of course they are modified, with a different structure
when they are in the nucleus.

2 The average proton binding energy is 8 million electron volts, which


is less than 1% of the mass. How will that affect an experiment
using 80,000-million-electron volt electrons to measure proton
structure? Of course, we won’t see a difference.

wwThe European Muon Collaboration (EMC) discovered that quarks in


heavier nuclei move more slowly than quarks in light nuclei, and the
heavier the nucleus, the bigger the effect. This means that the nucleons
in nuclei are modified.

wwWhat causes the EMC effect? When the size of the EMC effect
for different nuclei is plotted against the probability of finding a
fast-moving nucleon pair in the nucleus (short-range correlated
pair), there is an almost perfect correlation. But correlation doesn’t
imply causation.

wwAre nucleons modified by being in a fast-moving short-distance pair?


We know that at short distance, the nucleon-nucleon force is stronger.
Experiments are being planned at Jefferson Lab to measure this.

Surprisingly, 99% of the proton and neutron


mass—and therefore our mass—comes from
the energy of a cloud of virtual particles.

Lecture 12
| Sea Quarks, Gluons, and the Origin of Mass 129
Supplements

READINGS

Collins, “Parton Distribution Functions.”


Feltesse, “Introduction to Parton Distribution Functions.”
Friedman, “Deep Inelastic Scattering.”
Kendal, “Deep Inelastic Scattering.”
National Research Council, Nuclear Physics, p. 104–123.
Taylor, “Deep Inelastic Scattering.”
Wilczek, “The Origin of Mass.”

QUESTIONS

1 Why do we say that a proton or a neutron is made up of 3 quarks


when each can contain an infinite number of quarks?

2 Are quarks point particles, or are they composed of


still-smaller particles?

ANSWERS

1 The proton and neutron are made up of 3 quarks plus a large number of
quark-antiquark pairs. The 3 valence quarks combine to provide the charge
and spin of the proton (up, up, down) and neutron (up, down, down). Each
quark-antiquark pair has a total charge of zero and a total spin of zero.
However, the quark-antiquark pairs do contribute to the mass of the proton
and neutron.

2 We don’t know. So far, there is no evidence that quarks have size or that they
are composed of smaller particles. However, physicists are still looking.

130 Nuclear Physics Explained


NUCLEAR FUSION
IN OUR SUN
13

O
 r Sun is a fusion-powered nuclear
u
reactor, providing the Earth
with 100,000 terawatts of solar
power. This provides direct power for
photosynthesis in plants and for solar
panels, indirect power for wind and
hydroelectricity, and the energy stored
in coal, oil, and gas. This lecture will
pull together our knowledge of nuclear
masses, forces, decays, and reactions
and apply that to our favorite nuclear
reactor: the Sun. u

131
Stellar Formation and Solar Characteristics
wwStars form from the gravitational collapse of interstellar gas clouds,
which are mostly hydrogen. The gravitational energy of the collapse,
as the hydrogen falls inward, goes to radiation and to heating up the
gas. About half of the energy is radiated out; the other half heats up
the gas. If the mass is more than 0.08 of the mass of the Sun, then
the temperature gets hot enough for fusion, and the hot cloud of gas
becomes a star.

wwSolar structure is determined by a number of competing


physical processes:

§§ Pressure balance: The weight of the gas above is balanced by the


pressure of the gas below. As you go deeper in the Sun, the weight
becomes bigger, so the pressure becomes bigger—just like water in
the ocean or air in the atmosphere.

§§ Temperature gradient: All the power is generated in the core, but it


is radiated outward from the surface, so the energy flows from hot to
cold, which means that the core is much hotter than the surface.

§§ Power balance: When the star is in steady state, just like the Sun,
all of the power generated in the core is radiated from the surface.
This power is carried away by photons as mostly visible light. It takes
hundreds of thousands of years for each photon to reach the surface
from the core.

wwHow does a photon (gamma ray) reach the surface? The gas is very
opaque, and the photon is repeatedly absorbed and reemitted. The
reemission is random. This is called a random walk. The distance
traveled by the photon increases with the number of steps, but it
increases very slowly.

132 Nuclear Physics Explained


wwThe Sun is 93 million miles, or 150 million kilometers, away.
The diameter of the Sun is 1.5 million kilometers. The radius is
0.7 million kilometers. To cover this distance, a photon needs
1 trillion squared, or 1024, random steps. At this rate, it takes a photon
about 100,000 years to reach the surface.

wwAll of these collisions thermalize the photons. This means that the
spectrum of the photons depends on the temperature. It gives us a
blackbody spectrum of light. This means that the distribution of
photons and energy depends only on the temperature. It tells us the
surface temperature, but it doesn’t tell us anything about how the
energy is made.

wwThe radius of the Sun is 7000 kilometers. The mass is


2 × 1030 kilograms. The temperature on the surface of the Sun is about
6000 Kelvin. The central temperature is about 16 million Kelvin,
which corresponds to an energy of the particles in the Sun of about
1500 electron volts. This number is much less than the nuclear binding
energies of 8 million electron volts per nucleon.

In 1920, precise
measurements of
atomic masses led
Sir Arthur Eddington
to suggest the
possibility of nuclear
fusion in stars.

SIR ARTHUR EDDINGTON


1882–1944

Lecture 13
| Nuclear Fusion in Our Sun 133
wwWhere does the Sun’s energy come from? Could it come from chemical
energy, the energy stored in molecules? Chemical energy is about
100 million times less than the energy stored in nuclei. This could
only power the Sun for about 10,000 years. What about gravitational
energy? This is the energy released by all the mass falling inward.
Even gravitational energy could only power the Sun for about
1 million years.

The Power of Nuclear Fusion


wwThe Sun’s power comes from nuclear fusion. The Sun is mostly
hydrogen and some helium. The curve of binding energy shows that
if we fuse lighter elements to form heavier ones, it releases energy. The
biggest energy gain comes from fusing 4 protons into helium-4; almost
1% of the mass of the hydrogen is converted into energy when it fuses
to helium.

wwHow do we make this happen? We need a high enough temperature so


that the protons can fuse, and we need enough density so that there are
enough protons to fuse. Nuclear fusion only happens in the core of the
star, where the temperatures and densities are high enough.

wwHow do stars fuse protons to helium, and where do the neutrons


come from? Let’s start with the protons. Two protons can combine to
form helium-2, which is extremely unstable and most of the time just
falls apart again. Some of the time, helium-2 decays to a deuteron, a
positron, and a neutrino.

wwThe total energy of 2 protons going to the deuterium is twice the mass
of the proton, minus the mass of the deuteron, minus the mass of the
positron, or about half a million electron volts. If we include the energy

134 Nuclear Physics Explained


we get from the positron finding an electron and annihilating and
subtract the energy carried off by the neutrino, the total energy gained
is about 1.2 million electron volts.

wwBut this is very unlikely to happen, for a few reasons. The protons
repel each other electrically. The energy barrier is about 1.6 million
electron volts. The protons move with thermal energy. The central
temperature of about 16 million Kelvin means that they each have an
energy of about 1500 electron volts. The proton kinetic energy is about
1000 times less than the electrical repulsion energy.

wwThe protons get close enough to fuse by quantum tunneling—just


like alpha particles tunneling out of an unstable nucleus. In this case,
the energy is 1000 times too low, so they need to tunnel a distance of
about 1000 times the proton radius. This is very unlikely. This is why
we need very high temperatures: As the temperature increases, the
probability of fusion increases dramatically.

wwSo, 2 protons fuse into helium-2, which is very unstable. It decays by


the strong decay into 2 protons almost instantly, but a tiny fraction of
helium-2 decays by the weak decay to a deuterium nucleus, a positron,
and a neutrino. So, it takes a very long time for 2 protons to make a
deuteron. Fortunately, there are a lot of protons in the Sun, so this
reaction does occur. Once there’s a deuteron, it gets hit by another
proton, becomes a helium-3, and emits a photon very quickly. So, there
is a bottleneck in making the deuterons.

wwWhat happens to helium-3? There are 2 branches. In the PP-I


chain (PP stands for proton-proton), which happens 69% of the
time, a helium-3 and a helium-3 hit each other, combine, and make
helium-4 and 2 protons. The net result of 6 protons combined and
ended up as a helium-4, 2 protons, 2 positrons, 2 neutrinos, and
2 photons. So, there’s about 24.7 MeV, plus the positron annihilation
energy, minus the neutrino escape energy, leaving us with about
26 MeV.

Lecture 13
| Nuclear Fusion in Our Sun 135
wwThe other 31% of the time, the helium-3 hits a helium-4 and makes a
beryllium-7 plus a photon. This branch splits again. In the PP-II chain,
which happens 30.9% (out of 31%) of the time, the beryllium-7 beta
decays to a lithium-7, a positron, and a neutrino. The lithium hits a
proton and becomes 2 helium-4 nuclei. The net result is 4 protons,
which makes a helium-4 plus 2 positrons, 2 neutrinos, and 2 photons.
This makes about 24.7 MeV of energy, plus the energy of the positron
annihilation, minus the energy the neutrinos carry off with them,
giving us about 26 MeV. This second neutrino has 0.8 MeV.

wwIn the PP-III chain, which happens 0.1% of the time, the
beryllium-7 absorbs a proton, becomes boron-8, and gives off a
photon—a gamma ray. The boron-8 then immediately decays into
2 helium-4 nuclei plus a positron and a neutrino. So, the net result
is the same energy: 4 protons make a helium-4 nucleus, 2 positrons,
2 neutrinos, and 3 gamma rays. The same energy goes in, but the
neutrinos are carrying off a lot more energy. The net result is 19.3 MeV.
This second neutrino has a lot of energy, about 7.2 MeV. It’s only 0.1%
of the total fusions, but it’s important for detecting solar neutrinos.

wwHow do we know this? We can measure reaction rates and cross


sections in accelerators. This is difficult, because the rates are tiny at
these very low energies. We need to measure at larger energies and
extrapolate down to the lowest energies. We can calculate the reaction
rates in cross sections using nuclear models and then use these cross
sections to help build solar models.

wwHow can stars avoid the bottleneck with The Sun converts 4
2 protons having to wait a really long megatons of mass
time to go to a deuterium, a positron, to energy every
and a neutrino? Stars can use carbon, second, and 500
nitrogen, or oxygen as a catalyst. This is megatons per second
called the CNO cycle. of hydrogen is
converted to helium.

136 Nuclear Physics Explained


wwWhich is better: the PP How Fusion Affects
chain or the CNO cycle? It Stellar Lifetimes
depends on the temperature
of the star. Carbon has a lot Bigger stars have much higher
more electrical repulsion internal temperatures. The
because it has more charge, core temperature can be up
so the protons need more to 10 times higher (instead
energy to penetrate and of 16 million Kelvin, it can be
tunnel into the carbon 150 million Kelvin). The higher
nucleus. Our Sun gets more temperature means that
energy from the PP chain, it’s much easier for protons
but if the mass were greater, to quantum mechanically
the star would get more tunnel. There is a much
energy from the CNO cycle. higher reaction rate. As the
temperature increases from,
wwBoth the proton-proton for example, 20 to 100 million
and CNO processes are Kelvin, the reaction rate
slow. The proton-proton increases by a factor of about
to deuterium is especially 1010. Bigger stars burn a lot
slow, even in the Sun. This faster and die younger.
is why the Sun can shine
for billions of years. Fusion Our Sun’s lifetime is estimated
reactors on Earth will at about 15 billion years. Our
never work this way; they Sun is only 5 billion years old
need to consume their fuel now, which means that it’s
much faster. fairly young.

Detecting Solar Neutrinos


wwHow do we know that there is nuclear fusion in the Sun?

§§ The Sun is primarily hydrogen and helium, which we know


from spectroscopy.

Lecture 13
| Nuclear Fusion in Our Sun 137
§§ Nuclear fusion is the only known source with enough energy to last
the Sun’s lifetime, because the strong force is so much greater than
the electromagnetic force.

§§ Simulations of the Sun, including known nuclear physics from Earth,


describe the Sun’s surface characteristics well. These models tell us
several things about the inside of the Sun: that the density is about
150 tons per cubic meter at the center (almost 10 times the density of
gold), the pressure is about 100 billion atmospheres at the center of
the Sun, and the temperature is about 16 million Kelvin.

wwBut we still want direct evidence of nuclear processes from the solar
interior. Maybe we can detect some of the solar neutrinos.

wwThere are 1038 neutrinos emitted by the Sun every second. By the time
these neutrinos reach the Earth, they are spread out over a sphere whose
radius is the distance from the Earth to the Sun. This means that
there are about 1015 neutrinos for every square meter per second, which
means that about 1 quadrillion neutrinos are passing through each of
us every second. Fortunately for us, they rarely interact.

wwAll 3 proton-proton chains give low-energy neutrinos (less than


0.5 MeV) from the proton-proton going to a deuterium, a positron,
and a neutrino. The PP-II chain gives a 0.8 MeV neutrino from the
beta decay of beryllium-7 to lithium-7. The PP-III chain gives a 7 MeV
neutrino from boron-8 going to 2 helium-4 nuclei, plus a beta, plus a
neutrino. That only happens 1 in 1000 times. But those high-energy
neutrinos are the ones we want to detect.

wwHow do we detect these elusive neutrinos? One way is to make a


detector with a lot of chlorine-37 (fortunately, natural chlorine is
about 1/4 of chlorine-37). Then, we look for argon-37 atoms made by
the reaction where a neutrino hits chlorine-37 and basically inverse

138 Nuclear Physics Explained


beta-decays to make argon and a positron. This is sensitive to neutrinos
with energy greater than about 0.8 MeV, and the cross section increases
rapidly with energy.

wwHow much chlorine do we need? The cross section is about 10−50 square
meters for each chlorine-37 nucleus, and we have 1015 neutrinos
per square meter per second, so the interaction probability for
1 chlorine-37 nucleus (assuming that 15% of the neutrinos have
0.8 MeV or more) is 10−36 interactions per second. That means that we
need 1036 chlorine nuclei to measure 1 neutrino per second.

wwThe solar neutrino unit (SNU) gives 1 interaction per 1036 target nuclei
per second, which is about 60 million tons of chlorine. Ray Davis
conducted a very difficult experiment that continued for 25 years and
still only saw about 1/3 of the expected number of neutrinos from the Sun.

wwFrom this and other neutrino experiments, particle physicists have


learned that neutrinos are weird and can oscillate from one kind
of neutrino to another, that our simulations of the Sun predict the
measured neutrino energy distribution, and that our Sun is indeed
powered by nuclear fusion.

Supplements

READINGS

“Fusion,” Nobelprize.org.
Henley and Garcia, Subatomic Physics, section 19.3.
LeBlanc, An Introduction to Stellar Astrophysics, sections 6.1–6.5.
Mackintosh, Nucleus, chap. 9.
Rosen, “Ray Davis.”
Thomson (Lord Kelvin), “On the Age of the Sun’s Heat.”

Lecture 13
| Nuclear Fusion in Our Sun 139
QUESTIONS

1 How does a star maintain a stable core temperature?

2 Why does increasing the core temperature of stars by a factor of only


a few increase their power output by a factor of millions or billions?

ANSWERS

1 If the temperature decreases, then the core contracts. This means that the
protons are closer together so that the nuclear fusion rate increases. This
releases more energy, increasing the temperature. On the other hand, if the
temperature increases, then the core expands. This increases the distance
between protons, decreasing the nuclear fusion rate. This releases less energy,
decreasing the temperature.

2 This situation is very similar to the half-life of alpha decay (lecture 3). Alpha
particles have to quantum mechanically tunnel out of the nucleus to decay.
Protons have to quantum mechanically tunnel into the other proton to fuse.
When the energy of the alpha particle increased, the distance that it had to
tunnel out through the region of repulsion decreased. This exponentially
increased the probability of decay, decreasing the half-life exponentially. If a
star has a core temperature that is twice as large as that of the Sun, then its
protons have twice the kinetic energy. This decreases the distance they have
to tunnel in through the region of repulsion, which exponentially increases
the probability of tunneling, exponentially increasing the proton-proton
fusion rate.

140 Nuclear Physics Explained


MAKING
ELEMENTS:
14
BIG BANG TO
NEUTRON STARS

T
 e overall name for how our universe
h
creates nuclei is nucleosynthesis, and
there are 4 main processes: big bang
nucleosynthesis, which made hydrogen and
helium; fusion in ordinary stars, which made
helium, carbon, and oxygen; cosmic-ray
fission, in which fast protons collide with
carbon to make nuclei skipped over in stars;
and explosive processes, with supernovas
or colliding neutron stars. u

141
The Big Bang
wwIn 1929, Edwin Hubble observed that The biggest
the spectral lines of stars in galaxies are explosion ever was
redshifted, and the redshift increases making hydrogen
linearly with distance. The raisin bread and helium in the
model is used to explain this. If you think big bang, which
of raisin bread rising—or expanding—each was unlike any
of the raisins is moving apart from each chemical or nuclear
other. The farther a raisin is from another bomb explosion.
one, the faster they’re moving apart.

wwUsing this model as a guide, we can extrapolate backward to the big


bang, about 14 billion years ago, when the universe was very hot and
exploding, but cooled as it expanded. The farther back in time we go,
the smaller, denser, and hotter the universe is.

wwWhat do these insanely high temperatures mean? Temperature is the


average kinetic energy of a particle. This is 10,000 Kelvin, which is a
kinetic energy of 1 electron volt. And that’s the average; it’s actually a
wide distribution of energies. If the temperature is greater for a particle
of twice its mass times the speed of light squared (2mc 2), then it can
make particle-antiparticle pairs.

wwIf the temperature is greater than the binding energy of a system,


then the particles don’t stay bound. If the temperature is greater than
10 electron volts, atoms fall apart. If the temperature is greater than
1 million electron volts, nuclei fall apart. And if the temperature is
greater than 100 million electron volts, protons and neutrons start to
fall apart. “Freeze out” is the term for when the temperature drops
below these points and the particles are now stable.

142 Nuclear Physics Explained


Lecture 14
| Making Elements: Big Bang to Neutron Stars 143
wwThe big bang timetable shows what happened as the temperature
dropped. We know this from modeling that incorporates the 4 forces of
the standard model.

§§ If the temperature was greater than 1019 billion electron volts, it is


unknown what happened. We need theories of quantum gravity that
we don’t have yet.

§§ When the temperature had dropped to 1014 billion electron volts, that
was the time of cosmic inflation, when space expanded dramatically
by an incredible rate.

§§ When the temperature had dropped to 100 billion electron volts,


then the forces—such as the strong force and the weak force—
separate. At that time, there was a seething mass of electrons, quarks,
gluons, etc., and it was the time of quark-gluon plasma.

§§ When the temperature had dropped to 100 million electron volts (or
10 billion Kelvin) at about 1 microsecond after the big bang, quarks
started clumping into protons and neutrons, which froze out. There
are a lot more protons than neutrons because the proton has a smaller
mass than the neutron, so it was more energetically favorable to make
protons. All the protons (hydrogen nuclei) that exist today were made
back then; they are primordial.

§§ At about 3 minutes after the big bang, the temperature had dropped
to about 100,000 electron volts. That’s when nucleosynthesis began.
Nuclei were now stable enough, and these are the kinds of reactions
we can measure in accelerators.

§§ About 30 minutes after the big bang, the temperature had dropped
to 10,000 electron volts (or 100 million Kelvin), which is too cold
for nucleosynthesis.

144 Nuclear Physics Explained


§§ At 300,000 years after the big bang, the temperature had dropped to
3000 Kelvin, or about 0.3 electron volts. That’s when electrons and
nuclei joined to form atoms.

§§ At 14 billion years after the big bang, the temperature of the universe
is 2.7 Kelvin, or 2.7° above absolute zero, which is the temperature of
the interstellar vacuum today.

Nucleosynthesis
wwTo make nuclei from neutrons and protons, the neutron has to hit a
proton and form a deuterium nucleus (heavy hydrogen) and give off a
gamma ray. The neutrons have not yet decayed to protons because the
lifetime of a free neutron is 15 minutes. The deuterium binding energy
is only 2.2 MeV, so the temperature has to be much less than that or
the deuterium nuclei will fall apart.

wwDeuterium was made in big bang nucleosynthesis. If there is deuterium


in stars, it gets destroyed because the stars are too hot.

wwThere are many subsequent reactions. A proton can hit the deuteron,
making helium-3 plus a gamma ray. (There are more protons around,
but it’s more difficult for a proton to fuse with a deuteron.) A neutron
can hit a deuteron, making tritium (even heavier hydrogen). A neutron
could hit a helium-3 to make helium-4. A proton could hit a tritium to
make helium-4. Or a deuterium could hit helium-3 to make a proton
plus helium-4.

wwThere can also be heavier reactions. Tritium can hit helium-4, making
lithium-7. Helium-3 could hit helium-4 and make beryllium-7, which
then beta-decays to lithium-7. Or a proton could hit lithium-7, split

Lecture 14
| Making Elements: Big Bang to Neutron Stars 145
it up, and make 2 alpha particles—2 helium nuclei. This is very
complicated. It depends on the temperature, density, and abundances.
But almost all of the neutrons end up in helium-4.

wwWhy didn’t heavier elements form? Heavier nucleosynthesis is almost


blocked by the fact that there are no stable nuclei with 5 nucleons.
There are also no stable nuclei with mass 8. This is because helium is
so tightly bound itself, because it’s at a magic number. There’s also a
relatively low density in the big bang, so a rare helium-4 would need
to hit a very rare lithium-7 to make, for example, boron-11, which is
even heavier.

wwLet’s pull together all of these processes and use them to model
what happened during big bang nucleosynthesis. We will have
1 parameter—the proton and neutron density at the start of big bang
nucleosynthesis—from which we will predict the abundances of
helium-4, helium-3, deuterium, and lithium-7.

wwThere was 25% helium-4 because almost all of the neutrons ended up
in helium-4. There was a 7-to-1 ratio of protons to neutrons, which is
equivalent to a 14-to-2 ratio, which gives us 12 protons plus a helium-4,
and that means that a quarter of the mass is helium-4. In addition,
there were about 30 parts per million deuterium, about 10 parts per
million helium-3, and about 300 parts per trillion of lithium.

wwThere is some conflict with the lithium number. If the universe had
been denser—if there were more matter—then there would be less
deuterium and less helium-3, because those would have been made into
helium-4. And there would have been more lithium-7, because there
would have been more chances for helium-3 to hit helium-4 and make
beryllium-7, which would beta-decay to lithium-7.

146 Nuclear Physics Explained


wwHeavier elements weren’t made in the big bang. When a star exhausts
its hydrogen, it gravitationally shrinks—because it has lost its fuel
source—and becomes hotter and hotter, until it finds a new power
source. That new power source is fusing helium-4.

wwHelium-4 burning in stars starts at a temperature of 100 million


Kelvin. We need enough helium-4 so that 2 things can happen:
2 helium-4s can come together and make an excited state of
beryllium-8, but that lasts for about 10−15 seconds, so we need enough
helium so that a third helium can come along, hit the beryllium-8, and
make an excited state of carbon-12, plus a gamma ray.

wwThat excited state of carbon-12 was predicted by Fred Hoyle to make it


much more likely that 3 helium nuclei could come together at the same
time to make carbon production possible. This is called the triple-alpha
reaction, because 3 helium come together to make carbon-12 plus
3 gamma rays (from the decay of beryllium-8 and from the carbon
excited states).

wwThis produces a lot less energy than 4 protons going to helium; in fact,
the mass fraction is 10 times smaller. It also lasts for much less time
in the lifetime of a star than the proton-proton chain. We need lots of
helium-4 in one place for a long time to make enough carbon, so this
didn’t happen in the big bang. Once we have carbon, helium-4 plus
carbon can make oxygen-16 plus a gamma ray. Also, helium-4 can
hit the carbon and make oxygen-15, where a neutron is released, or
nitrogen-15, where a proton is released.

wwThis explains carbon and oxygen, but how did the in-between nuclei
get made? The carbon and oxygen were made in a star and then
expelled into the void at the end of the star’s lifetime. High-energy
protons—cosmic rays—hit some of those carbon and oxygen nuclei
and broke them down into smaller nuclei. Some of the lithium and
almost all of the beryllium and boron in the universe was made
this way.

Lecture 14
| Making Elements: Big Bang to Neutron Stars 147
wwWhen a star exhausts its helium, then it
Bigger stars are
gravitationally shrinks and becomes hotter
bluer and more
until it finds a new power source. If the
luminous; smaller
star is big enough, which means more
stars are redder
than 8 times the mass of the Sun, then
and less luminous.
carbon burning begins in the core. For this
process, we need a temperature of more than
1 billion Kelvin, and it lasts for only about
600 years. Two carbon nuclei come together to make an excited state of
magnesium-24, which then decays—to neon-20 plus an alpha particle,
or to sodium-23 plus a proton, or to magnesium-23 plus a neutron. It’s
an endothermic reaction, but it makes neutrons.

wwWhen the star exhausts its carbon, the core now contains neon, oxygen,
and magnesium. Again, it has lost its power source, so it gravitationally
shrinks and gets hotter until it finds a new power source. If the star
is big enough, then neon burning in the core starts. The temperature
needs to be much more than 1 billion Kelvin. Neon-20 plus helium
makes magnesium-24, and this process keeps going.

wwOxygen burning requires a higher temperature because it’s more stable.


It’s a closed shell, so it needs a higher temperature to fuse.

How long does it take for a star to go from proton-proton


fusion to silicon burning? For a 25-solar-mass star (it needs
to be heavy to progress through all of the stages):
BURNS PRODUCES CORE TEMPERATURE TIME
H He 6 × 107 7 × 106 years
He C, O 2 × 108 5 × 105 years
C O, Ne, Mg 9 × 108 600 years
Ne O, Mg, Si 1.7 × 109 0.5 years
O Si, S 2.3 × 109 6 days
Si Fe 4 × 109 1 day

148 Nuclear Physics Explained


wwSilicon burning involves repeated fusion with helium to make heavier
and heavier nuclei. And in the process, it makes all of the nuclei from
hydrogen up through iron and nickel.

Neutron Stars and Supernovas


wwWhat will happen to our Sun after proton burning? This process
exhausts the protons in the core, which becomes denser and hotter.
Proton burning continues in the shell, but the envelope of the star
expands to about the radius of the Earth and cools. The star leaves the
main sequence and becomes a red giant. Our Sun is small, so it lives
longer; it will stay a red giant for about a billion years. When the core
becomes hot enough, there is helium burning in the core and hydrogen
burning in the shell, and carbon accumulates in the core of the star.

wwHow will our Sun end? Our Sun is not big enough to burn the
carbon, so the inert carbon core won’t fuse. The core will continue
to contract. The radius of the star will keep growing, and it will
become a red supergiant. Some of the neutrons made in the star
can be absorbed to slowly make heavier elements, in what is called
s-process nucleosynthesis. But there isn’t enough mass to start carbon
burning. The core will become a white dwarf. The envelope will be
expelled to become a planetary nebula, although the material won’t
disperse widely.

wwA really big star, with a mass of more than 10 times the mass of the
Sun, ends up with an iron core and lots of burning and inert shells.
Iron can’t fuse and release energy, so nothing heavier than iron can
be made in normal stars. There is an energy crisis—it makes more
and more iron. The core is inert; it’s not producing power. The inert
core contracts, and the temperature increases. The energy from the
gravitational collapse increases the temperature.

Lecture 14
| Making Elements: Big Bang to Neutron Stars 149
wwWhen the core temperature gets high enough, the iron in the central
core is broken down into alpha particles, which are then broken down
into their constituent protons and neutrons. Only the very central iron
in the core is destroyed. It absorbs energy, cools the core, contracts
faster, and heats up more.

wwThen, because it’s so dense, protons and electrons combine to make


neutrons plus neutrinos. This is called neutronization. It’s driven by
the Pauli exclusion principle and explained by the Fermi gas model.
The protons absorb the electrons, and the electron pressure is reduced.
It speeds the core collapse. On the chart of nuclides, the nuclei move
dramatically down and to the right, and we end up with a neutron star.

wwThe core then collapses very rapidly, at 0.2 times the speed of light.
But the neutrons obey the Pauli exclusion principle. The core has a
maximum density for the neutrons, which is about 3 times the nuclear
density. If the core mass is too high, it’s going to make a black hole,
and there won’t be a supernova. But if the core mass is not too high,
then the core forms a hot neutron star with a temperature of about
100 billion Kelvin.

wwThe core hits that neutron-star surface and bounces back, making a
shockwave through the envelope. Explosive nucleosynthesis occurs in
the outside layers as the shockwave passes through. Lots of neutrons
absorb on nuclei to make new nuclei. This is r-process nucleosynthesis.
There is some beta decay that converts some of the neutrons to protons,
but it’s very rapid, and lots of very-short-lived nuclei are involved.
Heavy elements are made and ejected, but the shockwave of the star
stalls quickly.

wwHigh temperatures lead to lots of particle-antiparticle production.


Normal particles (electrons and antielectrons) reinteract and annihilate.
But neutrino-antineutrino pairs can escape. They carry away about

150 Nuclear Physics Explained


10% of the rest mass; 99% of the
A core-collapse
energy in a supernova is released as
supernova releases
neutrinos. The pressure from these
2 × 1047 joules of
neutrinos is what drives the explosion.
energy. It’s as bright
Without neutrinos, the supernova
as a galaxy—1010 times
can’t explode—even though neutrinos
interact through the weak nuclear
the luminosity of the
force. The outer layers are violently
Sun—for a few days.
ejected, and we end up with a
But this is still much
core-collapse supernova. less than the energy
given off in neutrinos.
wwThe core collapse turns the electrons
and protons in iron into neutrons.
The gravitational potential energy of the core collapse makes the
explosion energy, which creates nuclei much heavier than iron. The
evidence of this is we can model the process, tested by looking at the
light curve and spectra of the supernova and the detection of neutrinos
from supernova 1987A, which was detected in 1987 in the Large
Magellanic Cloud.

wwAfter a supernova, what’s left is a neutron-star core. This is the biggest


nucleus of all, with between 1 and 3 times the mass of the Sun. While
the core of a neutron star is made almost entirely of neutrons, there are
lots of heavy nuclei on the neutron-star surface, where the pressure is
much less.

wwHow could some of those heavy nuclei escape to become the Earth’s
heavy elements? We need the inspiral and merger of a binary
neutron-star system, which is 2 neutron stars orbiting around each
other. Tidal forces can rip material from the surface of the smaller
neutron star and eject it into the galaxy. Or gravitational energy
can make neutrinos, and the pressure from the neutrinos can knock
material off of the surface.

Lecture 14
| Making Elements: Big Bang to Neutron Stars 151
Supplements

READINGS

American Association of Variable Star Observers, “Supernova


1987A.”
Drake, “In a First, Gravitational Waves Linked to Neutron Star
Crash.”
Ellis and Schram, “Could a Nearby Supernova Explosion Have
Caused a Mass Extinction?”
Fields, “When Stars Attack!”
Henley and Garcia, Subatomic Physics, sections 19.1, 19.2, and 19.4.
LeBlanc, An Introduction to Stellar Astrophysics, sections 6.6–6.14.
Mackintosh, Nucleus, chaps. 9–10.
White, “Big Bang Nucleosynthesis.”
——— , “Origin of the Light Elements.”

QUESTIONS

1 Where does the energy come from to make all of the elements heavier
than helium?

2 How do we know that some of the heavy elements are made in


neutron-star collisions?

152 Nuclear Physics Explained


ANSWERS

1 The energy to make the elements from helium to iron comes from the
curve of binding energy. As lighter elements fuse to form heavier elements,
energy is released. The energy to make the heavier elements comes from
gravitational energy, either from the collapse of an ultraheavy star after it
runs out of fuel or from the collision of 2 neutron stars. As the ultraheavy
star collapses, its layers fall inward and gain energy. Part of this energy
goes into fusing nuclei into heavier elements than iron, which are then
distributed into interstellar space as the star becomes a supernova. Similarly,
when neutron stars collide, we can think of it as one falling into the other
one, gaining a lot of energy. While the interior of the neutron star is almost
entirely neutrons, there are a lot of normal nuclei on its surface. Some of
the energy gained from falling together makes heavier elements and causes
material from the neutron stars to be ejected.

2 Prior to 2017, we relied on computer models of supernovas and neutron-star


collisions, which told us that supernovas could not produce all the nuclei
we see in the correct abundances. In 2017, scientists saw the gravitational
waves from 2 neutron stars falling into each other. When they focused
optical telescopes at that location, they saw absorption lines in the light
corresponding to heavy elements such as gold. This was the first direct
evidence that heavy elements were made in neutron-star collisions.

Lecture 14
| Making Elements: Big Bang to Neutron Stars 153
15 SPLITTING THE
NUCLEUS

N
UCLEAR FISSION is the process whereby
a nucleus splits into 2 smaller nuclei. This
is very rare in nature. It happens when
an appropriate nucleus absorbs a neutron. Only a
few very large, even-numbered elements fission.
Uranium is the only element with a naturally
occurring isotope that can be fissioned. Uranium
is relatively abundant. It’s heavy enough to fission
but stable enough to still exist. It’s in the middle of
the first so-called island of stability on the chart
of nuclides [PAGE 274]. There are more protons, so
there is more electrical repulsion, which makes it
less stable. u

154
Fissioning Uranium James Chadwick
discovered the
wwThere are several isotopes of uranium.
neutron in 1932,
The most common one is uranium-238,
with 92 protons and 146 neutrons. It
after which scientists
has a half-life of 4.5 billion years, about
looked to see what
the age of the Earth. Uranium-235, they could do with it.
with the same number of protons Just 2 years later,
but 143 neutrons, has a half-life of Enrico Fermi
0.7 billion years. Uranium-234, with discovered neutron-
92 protons but only 142 protons, has a inducted radioactivity,
half-life of 0.25 million years, so there which earned him the
is no uranium-234 left over from when 1938 Nobel Prize
the Earth was formed. in Physics.
wwUranium-238 has the same half-life Otto Hahn and Fritz
as the Earth, so about 1/2 of the Strassman discovered
uranium-238 nuclei have decayed since uranium fission in
the Earth was formed. Uranium-235 has 1939, which earned
a half-life that is 7 times less than the Hahn the Nobel Prize
age of the Earth, so all but 1/128 of the in Chemistry in 1944.
uranium-235 nuclei have decayed.

wwWe expect the relative abundance of uranium-235 to be about 1.5%,


but the actual abundance is 0.7%. We assume that uranium-235 and
uranium-238 were created in equal amounts in a supernova and then
formed relatively quickly, within about a billion years into our planet.

wwThere is a long decay chain for uranium.

§§ Uranium-238 alpha-decays to thorium-234, with a half-life


of 5 billion years. This decays very quickly by beta decay to
protactinium-234. Then, it beta-decays again to uranium-234,
giving off an electron and an antineutrino, in about a minute. Then,
it decays to thorium-230 by alpha decay, to radium-226 by alpha

Lecture 15
| Splitting the Nucleus 155
decay, to radon-222 by alpha decay, etc. It ends up at lead-206, with
82 protons and 206 nucleons, having given off 8 alpha particles,
6 electrons, 6 antineutrinos, and about 50 million electron volts
of energy.

§§ The uranium-235 decay chain is very similar. It also ends up at


lead, but it ends up at lead-207, having given off 7 alpha particles,
4 electrons, 4 antineutrinos, and also about 50 million electron volts
of energy.

wwMaking uranium-235 fission instead of just decaying lets us liberate


a lot more energy without waiting for 700 million years. We do this
by hitting uranium-235 with a neutron. It absorbs the neutron and
becomes uranium-236. This makes the charged nuclear droplet
oscillate, just like in the liquid-drop model. If the oscillations get too
big, then electrical repulsion makes it fission into smaller nuclei. It ends
up in 2 big pieces—called fission fragments —plus 2 or 3 neutrons.
Those fission fragments carry away 170 million electron volts of kinetic
energy. Then there are neutrons and gamma rays.

wwThe reason that uranium-235 fissions but not uranium-238 is because


uranium-235 has an even number of protons and an odd number
of neutrons. When uranium-235 absorbs a neutron, it becomes
uranium-236, with an even number of neutrons and an even number
of protons. That even-even uranium-236 is more tightly bound
than uranium-235, and the binding energy difference goes into the
excitation energy. The excitation energy for uranium-235 is 6.5 million
electron volts, which is enough to make it fission.

wwWhen uranium-238 absorbs a neutron, it


becomes uranium-239, which is even-odd Even uranium-235
and is less tightly bound per nucleon than only spontaneously
uranium-238. The excitation energy of fissions 10−10 of
uranium-239 is only 4.8 million electron the time.

156 Nuclear Physics Explained


volts, which is not quite enough to fission uranium-238. If we hit it
with a higher-energy neutron (with more than 1 million electron volts),
that will make uranium-238 fission.

wwThe energy released by fission comes from the curve of binding energy.
Uranium has 7.6 million electron volts per nucleon of binding energy.
But the fragments with smaller mass have more binding energy per
nucleon—8.5 million electron volts. That means there is a difference of
about 1 million electron volts per nucleon. Because uranium has about
200 nucleons, about 200 million electron volts will be released.

wwWhere does this fission energy go?

§§ 170 MeV goes to the fission fragments.

§§ 5 MeV goes to the neutrons.

§§ 7 MeV goes to the gamma rays.

§§ 27 MeV comes from delayed radioactivity. (Energy from the decays


of the fission fragments is split up among the beta particles, with
8 MeV; the neutrinos, with 12 MeV; and the gamma rays, with
7 MeV).

wwWhich nuclei does uranium-235 fission into? There are many


possibilities, but it typically fissions into 2 nuclei that don’t have equal
mass; one is a bit heavier and the other is a bit lighter. The fission
fragments are statistically distributed, with the lighter mass fragments
typically between 90 and 100 nucleons and the heavier mass fragments
typically between 130 and 150 nucleons.

wwWhy are the decay products so radioactive? They have too many
neutrons. Heavier nuclei need more neutrons to reduce the effects of
the proton-proton repulsion. Fission makes lighter fragments with too
many neutrons for their mass.

Lecture 15
| Splitting the Nucleus 157
wwHow do we achieve a self-sustaining chain reaction? A neutron plus
uranium-235 makes fission fragments plus 2 or 3 more neutrons.
If exactly 1 of those 2 or 3 neutrons goes on to fission another
uranium-235 nucleus, then it’s critical, and fission continues at exactly
the same rate.

wwBut if less than 1 neutron goes on to fission another


uranium-235 nucleus, then it’s subcritical, and the fission rate
will steadily decrease. If more than 1 neutron goes on to fission a
uranium-235 in the next generation, then it’s super critical, and the
fission rate increases. You might get an explosion. The number we refer
to here is the neutron multiplication factor.

wwA few things can happen to those extra neutrons: They can be absorbed
by a uranium-238 nucleus, be absorbed by a uranium-235 nucleus
without causing it to fission and instead emit a gamma ray, or can
escape entirely if there is not enough uranium.

Enriching Uranium
wwHow can we increase the neutron multiplication factor? We can have
more uranium so that the neutron is less likely to escape. We can
make the uranium more enriched so that the neutron is more likely
to hit a uranium-235 nucleus than a uranium-238 nucleus. We can
reduce the neutron energy (moderate the neutrons). This greatly
increases the uranium-235 fission cross section and decreases the
uranium-235 nonfission cross section, or the absorption cross section.
Then, we can remove the other materials that can absorb neutrons.

wwHow do we enrich uranium? Chemical methods don’t work.


Uranium-235 and uranium-238 react almost identically, so we have to
use physical methods that exploit the small mass difference between
them (about 1%). This is very difficult.

158 Nuclear Physics Explained


wwWhat enrichment levels do we need? We’re starting with 0.7%
uranium-235. For fuel, we need it enriched to between 3% and 5%. For
bombs, we need it enriched to 90%. Low-enriched uranium means that
the enrichment level is less than 20%, meaning that it’s less than 20%
uranium-235 and the rest is uranium-238. High-enriched uranium is
anything above that. We’ll use the same process to make low-enriched
uranium or high-enriched uranium, but just with a different
arrangement. It’s much easier to enrich uranium if we start with 20%
to get up to 90% than if we start at 0.7% to go to 20%.

wwThere are a number of uranium separation techniques, including


thermal diffusion, gaseous diffusion, electromagnetic (where you use
a kind of mass spectrometer), centrifuges, and lasers. The Manhattan
Project used the first 3 techniques together to achieve more than 82%
uranium-235 enrichment for the Little Boy gun-type nuclear bomb.

wwFor these techniques, we need the uranium to be able to move freely,


which means that we need it in the form of a gas. There is only
1 molecule that is gaseous at even close to room temperature: uranium
hexafluoride. It’s solid at room temperature and pressure but boils at
57° Celsius (134° Fahrenheit), so above that temperature, it’s a gas.
Fluorine only has 1 stable isotope, and this makes separations simpler
because any difference in mass between 2 molecules of uranium
hexafluoride is due to the uranium and not to the fluorine.

wwHow do we separate uranium electromagnetically? It’s a variant of


a mass spectrometer that is optimized for throughput rather than
for precision. In a typical mass spectrometer, we want to measure
masses precisely, but here we want to separate 2 molecules with
2 different masses.

wwWe ionize uranium hexafluoride gas and accelerate the molecules


through a potential difference so that they have the same amount of
kinetic energy. Then, we pass them through a magnetic field. The
radius of curvature in that magnetic field is mass times velocity, divided

Lecture 15
| Splitting the Nucleus 159
by charge times magnetic field: R = mv/qB.
Therefore, a 1% difference in mass, if
they have the same velocity, gives us a 1%
different in the radius of curvature. This
means that the uranium-235 will be bent
in a smaller circle than the uranium-238,
and we can use that to separate the
uranium-235 and the uranium-238.

wwIn addition to electromagnetic separation, there is gaseous diffusion, in


which we apply pressure to force the uranium hexafluoride gas through
a porous membrane. All of the uranium hexafluoride molecules have
about the same average kinetic energy, which is 1/2mv2. If the mass
is smaller, the velocity is bigger. There is about a 0.5% change in
the average velocity. That means that the uranium-235 hexafluoride
diffuses slightly faster than the uranium-238 hexafluoride, which
means that the uranium hexafluoride that passes through the tiny holes
in the membrane is slightly enriched in uranium-235, while the gas
left behind is slightly depleted. We need many stages of this gaseous
diffusion to get decent enrichment.

wwCentrifuges spin very rapidly. Think of a cylinder spinning on its axis


very rapidly. The acceleration experienced by a particle on the outside
of that cylinder is the centripetal acceleration, forcing it inward.
Centripetal acceleration is the square of the rotational speed times the
radius: 𝜔2r. This acceleration provides a huge force.

wwThe centrifuge fights against the thermal


motion of the particles. The thermal
Centrifuges were
motion randomizes the material. The
invented in 1934
centrifuge makes heavier material move
by Jesse Beams at
to the outside. We need to spin a big
the University of
centrifuge as fast as possible to improve
Virginia to separate
the separation. The maximum speed will
chlorine isotopes.

160 Nuclear Physics Explained


be about 500 meters per second, or 1000 miles
per hour. So, we need really strong materials and
really good bearings.

wwTo optimize centrifuges, we heat the bottom to


generate a flow of gas up the inner wall and down
the outer wall. We increase the height of the
centrifuge to take advantage of this. The lighter
material—the uranium-235 hexafluoride—will
collect at the inside and the top. The heavier
material will collect at the outside and the bottom.

wwWhen we’re centrifuging uranium, we feed in the


uranium hexafluoride gas continuously. There are 2 exhaust pipes:
We take the enriched gas from the center and the top to go to the next
stage, and we take the depleted gas that has less uranium-235 from the
outside and the bottom to go to the previous stage.

wwOne centrifuge can’t do it all. We need centrifuges in parallel and


in stages. If we want to enrich uranium for a reactor, we need lots
of centrifuges in parallel with just a few stages. If we want to enrich
uranium for bombs, we need lots of stages and not as many in parallel.

wwDepleted uranium is a little less


radioactive than natural uranium,
Depleted uranium is very
because uranium-238 has a
dense (70% denser than
half-life of about 4.5 billion years
lead), so it is used for
and uranium-235 has a half-life
armor-piercing projectiles,
of about 0.7 billion years, which
counterweights, and
is 6 times shorter, so it’s 6 times
radiation shielding. more radioactive. We’ve reduced
the uranium-235 percentage from
0.7% to 0.3%, so we’ve reduced
the overall radioactivity of the
uranium by a few percent.

Lecture 15
| Splitting the Nucleus 161
Supplements

READINGS

Bodansky, Nuclear Energy, chaps. 5 and 6.


GlobalSecurity.org, “Uranium Enrichment Techniques.”
Lilley, Nuclear Physics, sections 10.1–10.3.
PhET Interactive Simulations, “Nuclear Fission.”
Rhodes, The Making of the Atomic Bomb.
US Nuclear Regulatory Commission, “Uranium Enrichment.”

QUESTIONS

1 If the half-life of uranium-235 was doubled to 1.4 billion years,


how would that affect the relative abundance of uranium-235 in
natural uranium?

2 What is the most common form of radiation from the daughter


nuclei from uranium fission?

3 Why are we so concerned with nonnuclear nations enriching


uranium to use in power plants?

ANSWERS

1 Doubling the half-life would increase the abundance of uranium-235 by


more than a factor of 10. If uranium-235 had twice the half-life (1.4 billion
years instead of 0.7), then there would have been only 3.5 half-lives
instead of 7 since the Earth was formed. After 7 half-lives, only 1/128 of
the original uranium-235 is left, which is less than 1%. However, after
only 3.5 half-lives, 10% of the original uranium-235 is left. Thus, instead
of being only 0.7% of natural uranium, if the half-life were twice as long,
uranium-235 would be 10% of natural uranium.

162 Nuclear Physics Explained


2 Uranium has about 1.5 neutrons for each proton. The daughter nuclei also
have about 1.5 neutrons for each proton. However, stable nuclei in that
mass range have only about 1/3 more neutrons than protons. Therefore, the
daughter nuclei have too many neutrons and need to convert neutrons to
protons. They do this by beta-minus decay.

3 We are concerned because the same technology used to enrich uranium


from its natural abundance of 0.7% to the 4% needed for power plants can
enrich the uranium further to the 90% needed for weapons. It takes much
less effort to enrich uranium from power plant grade to bomb grade than it
does to enrich it from natural abundance to power plant grade.

Lecture 15
| Splitting the Nucleus 163
16 NUCLEAR
WEAPONS
WERE NEVER
“ATOMIC” BOMBS

T
 e term “atomic bomb” was first used by
h
H. G. Wells in his 1914 book The World Set
Free. This launched an idea that is wildly
inaccurate. Conventional explosives use
chemical reactions and should be called
molecular bombs because they rearrange
molecules. Atomic bombs should be called
nuclear bombs because they rearrange
nuclei. But the Manhattan Project
scientists called their weapons “the gadget”
or “atomic bombs,” layers of misdirection to
conceal what they were doing. u

164
Nuclear Bombs and Chain Reactions
wwWe can make nuclear bombs from uranium-235. We want to use it
very highly enriched—preferably 90% enriched or better—because
uranium-238 absorbs some of the critical neutrons that we want to
go on to fission uranium-235 instead. Lower-enrichment bombs are
possible, but they need a lot more uranium.

wwWe can also make nuclear bombs from plutonium. Plutonium is a


reactor by-product from uranium-238. When uranium-235 absorbs
a neutron, most of the time it fissions. Uranium-238, instead, can
absorb neutrons. When uranium-238 absorbs a neutron, it becomes
uranium-239 and gives off a gamma ray. The uranium-239 then
beta-decays to neptunium-239, giving off an electron and a neutrino.
Neptunium-239 then beta-decays to plutonium-239, again giving off
an electron and a neutrino.

wwPlutonium-239 is a desired isotope, but it can absorb neutrons


and become plutonium-240, so you need to withdraw the fuel
before too much plutonium-239 can become plutonium-240. But
plutonium-240 fissions spontaneously, which could cause the bomb
to preignite and fizzle. So, weapons-grade plutonium has very little
plutonium-240; it’s almost all plutonium-239. Reactor-grade plutonium
has more plutonium-240 and relatively less plutonium-239.

Only 1 kilogram of chemical explosives can liberate 4 million


joules of energy. This gives us the definition of a kiloton:
1000 tons of chemical explosive, or 4 × 1012 joules.
Through the process of fission, 1 kilogram of uranium-235
releases about 1014 joules of energy, which is 30 million
times more energy than is released from 1 kilogram of TNT.

Lecture 16
| Nuclear Weapons Were Never “Atomic” Bombs 165
wwThe Chernobyl reactors—and similar reactors in the United States
in Hanford, Washington—were designed to remove and replace fuel
rods easily while the reactor is running. They make weapons-grade
plutonium by removing the fuel rods often, extracting the
plutonium-239 before it can become plutonium-240, and then putting
them back. Reactors like this can make 1000 kilograms of plutonium
per year from a gigawatt of electricity.

wwMost power reactors, called light-water reactors, are of a different


design and need to be shut down for refueling, so fuel rods are left
in for 1 to 3 years before being replaced. This makes reactor-grade
plutonium. It makes about 300 kilograms per year—enough for
about 30 bombs. They make less plutonium than the Chernobyl
reactors because a lot of the plutonium is fissioned in the power plant
to make more energy. Another way to make plutonium is to place
uranium-238 targets near a reactor core (and replace them often,
without disturbing the operation of the reactor).

wwHow much fissionable material is needed for a critical mass? We need


most of the neutrons to interact to fission the uranium or plutonium.
This means that we need a sphere of material whose radius is about the
same length as the neutron interaction length.

wwFast neutrons—the neutrons emitted by fission that are typically on the


order of millions of electron volts—have a cross section of about 1 barn
before they’re moderated and slowed down. This cross section is much
less than the thermal neutrons, which have been slowed down to room
temperature. The fast neutrons have a mean free path and interaction
length of about 17 centimeters in uranium and 12 centimeters in
plutonium. So, plutonium has a higher fission cross section.

wwWe can use a reflector placed outside of the uranium or the plutonium
to reflect some of the neutrons back and decrease the amount of
material that we need. For uranium, if there is no reflector, we need a
sphere radius of 9 centimeters, so the diameter is about 18 centimeters,

166 Nuclear Physics Explained


or 8 inches. If we have a reflector, then the radius of the sphere is
6 centimeters and the diameter is 12 centimeters, or 4 inches, so we
need between 13 and 25 kilograms of uranium. For plutonium, if there
is no reflector, then the sphere radius is only 5 centimeters, or 2 inches.
If we use a reflector, then the radius is 4 centimeters, which means that
we only need 5 or 10 kilograms of plutonium.

wwTo trigger the chain reaction, we jump-start it with a lot of neutrons.


We use an alpha emitter, such as americium, that gives off alpha
particles and a beryllium target, with a very thin wall between the 2 to
stop the alphas. When we want to start the chain reaction, we break the
wall. The alpha particles can hit the beryllium and knock out neutrons.

wwThere are a few ways to make a bomb. The simplest way is with a
gun-type bomb, where we shoot one subcritical mass of uranium onto
another subcritical mass to make a critical mass.

wwWith the Little Boy bomb, an 86-pound hollow cylinder of uranium


was shot at a 60-pound spike of uranium. The uranium was about 80%
enriched, was in a 6-foot-long barrel, and was hit at 300 meters per
second (about 600 miles an hour). It went critical at about 10 inches
away, or 0.3 meters. It needed to have very few spontaneous fissions
in the time it took those
2 pieces to move that last
0.3 meters—which, traveling
Little Boy
at 300 meters for 1 second,
was about a millisecond—to
avoid premature detonation
and fizzle. They didn’t test
this design; they were sure
that it would work. This
design would not work
with plutonium because
too much plutonium
spontaneously fissions.

Lecture 16
| Nuclear Weapons Were Never “Atomic” Bombs 167
Little Boy

168 Nuclear Physics Explained


wwThe other kind of bomb is an implosion bomb, where we compress fuel
using a shockwave that travels at 5000 meters per second. We need a
very symmetric shockwave, which means that we need to use specially
shaped high-explosive lenses and perfect timing. The compression gives
it a higher density, which decreases the neutron interaction length and
the critical mass that is needed. By compressing a subcritical mass of
plutonium or uranium, it becomes critical. We need less plutonium or
less uranium. But we need to make sure that the ignition starts when
it’s fully compressed and not an instant before so that there is very little
spontaneous fission.

wwThis was the design of the Fat Man bomb that was dropped in
Nagasaki with an energy release of 21,000 tons of TNT. This
implosion-type plutonium bomb was much more complicated than
the gun-type uranium bomb, and they needed to test it to make sure it
would work. This was the famous Trinity test, when the first nuclear
bomb was detonated.

Fat man

Lecture 16
| Nuclear Weapons Were Never “Atomic” Bombs 169
Fat Man

wwThe Little Boy, like the Fat Man, used a chain reaction to release as
much nuclear energy as possible. With a chain reaction, we need to
release all of the energy before the bomb material can disassemble.
We have to use fast neutrons to do the fissioning because they have
a smaller cross section; there is no time for the neutrons to bounce
around a moderator and slow down and thermalize. We want to
maximize the neutron multiplication factor (k) because the bigger it is,
fewer generations are needed to fission all the nuclei. Reactors want a k
of exactly 1 so that they keep running at constant power.

wwLet’s think about the reaction in


generations. In one generation, some Enrico Fermi, one of the
nuclei fission and release neutrons. In great physicists of the
the next generation, there are k times 20 th century, built the
more nuclear fissions. The time of a first atomic pile, which
generation is the neutron interaction had the first sustained
length divided by the velocity. nuclear chain reaction.

170 Nuclear Physics Explained


wwA 1-million-electron-volt neutron has a speed of 1.4 × 109, which
is about 3% of the speed of light. Its interaction length is about
10 centimeters, so the time it takes for it to interact is about
10−8 seconds, or 10 nanoseconds. That length of time was called a
shake by the Manhattan Project.

wwIt takes 80 generations to fission 1 kilogram of uranium. The time


required to do this is 80 shakes, or about 1 microsecond (1 millionth of
a second). Half of the nuclei fission in the last generation.

wwThe problem is that the warhead starts to disassemble and spread


out when about 1 in a million of the nuclei have fissioned. Enough
energy is released so that the fuel starts expanding. If it predetonates,
we get a fizzle yield of only about 1000 tons of TNT. We need to
keep the bomb together for another 20 shakes, which is a quarter of a
microsecond, to let the rest of the uranium fission.

Controlling the Fissile Material


wwBuilding bombs is the easy part; controlling the fissile material is the
difficult part. Uranium needs to be enriched, which is expensive and
difficult. But this is the path that Iran is taking with its thousands
upon thousands of centrifuges. Bomb-grade uranium (90%
uranium-235) can be denatured by mixing it with natural uranium.
We’ve reduced our Cold War stockpiles of highly enriched uranium
this way.

wwWith plutonium, a reactor needs to be operated and the plutonium has


to be chemically extracted from the very radioactive spent fuel. This
is the path that North Korea has followed. In some ways, this is easier
than enrichment, but you have to remove the fuel from the reactor
very often.

Lecture 16
| Nuclear Weapons Were Never “Atomic” Bombs 171
wwThis is done by reprocessing plants, which make plutonium much
more available. This is one of the reasons why the United States
stopped reprocessing fuel in the late 1970s. Plutonium can’t be
denatured, which means that Cold War stockpiles of plutonium need
to be guarded.

wwTo make a bigger bang, we use a fusion-boosted bomb, which involves


adding deuterium and tritium in the center of the bomb. If we can get
the deuterium and tritium to fuse, it will make helium-4 plus a neutron
and release 14 million electron volts. The extra neutrons boost the
fission yield, and we get more energy from the extra fissions (200 MeV
of energy) than we do from the fusion (14 MeV of energy).

wwDeuterium and tritium are gasses, which means they are not very
dense, so we typically add lithium deuteride, which is lithium hydride,
but instead of being made with hydrogen, it is made with heavy
hydrogen (deuterium), which is a solid. And the lithium deuteride
makes the tritium: When a neutron hits the lithium, it makes a
helium-4 and a tritium.

wwIf we want even more energy released, then we make a thermonuclear


bomb, where we start with a fusion-boosted fission bomb and then
add a careful layering of lithium deuteride, plutonium, and reflectors.
The fission bomb sets off the secondary, and the thermonuclear bomb
actually gets a lot of its energy from nuclear fusion.

wwHow do we miniaturize nuclear


warheads to fit in a missile? Fission or The biggest bomb ever
fusion-boosted warheads can be smaller set off—Tsar Bomba—
than thermonuclear ones. They can be was 50 million tons
made of uranium or plutonium, but of TNT, which is more
we need less plutonium for that critical powerful than all of
mass, so we can make a smaller weapon
the explosives used in
from plutonium. Then, we need to
World War II.

172 Nuclear Physics Explained


optimize the shockwave to minimize
A dirty bomb is not
the explosives needed to compress the
a nuclear explosion.
plutonium or uranium.
It uses conventional
explosives to disperse
wwA Trident II ballistic missile submarine
radioactive material,
carries 24 missiles. Each missile carries
up to eight 100- or 500-kiloton nuclear
spread the radiation
warheads that are inside their own
over a large area,
reentry vehicles so that they can survive
and contaminate
reentry from space. the ground.

Supplements

READINGS

Bodansky, Nuclear Energy, chap. 17.


Dyson, Project Orion.
Mahaffey, Atomic Awakening.
Rhodes, The Making of the Atomic Bomb.

QUESTIONS

1 Why is it easier to make a nuclear bomb from highly enriched


uranium-235 than from plutonium?

2 Is it possible to make a nuclear reactor that does not


produce plutonium?

Lecture 16
| Nuclear Weapons Were Never “Atomic” Bombs 173
ANSWERS

1 Uranium-235 has a much lower spontaneous fission rate. In other words,


many fewer uranium nuclei than plutonium nuclei fission on their own.
The spontaneous fission rate of uranium-235 is about 1000 times less than
that of plutonium-239. In the gun design used for the Little Boy uranium
bomb, it took 1 entire millisecond (100,000 shakes) between the time when
the 2 pieces of uranium in the bomb came close enough to form a critical
mass and the time the bomb was ignited. There was less than 1 chance in
1000 that a uranium nucleus would spontaneously fission during that time,
igniting the bomb prematurely and making it fizzle.

2 No, unless there is no uranium-238 in the reactor fuel. All the reactor fuel
used today is about 96% uranium-238 plus 4% uranium-235. Some of the
neutrons released by uranium-235 fission are absorbed by uranium-238 and
make plutonium-239. (Thorium reactors, which would not produce
plutonium, are discussed in lecture 19.)

174 Nuclear Physics Explained


HARNESSING
NUCLEAR CHAIN
17
REACTIONS

D
id you know that nuclear power plants
boil water to turn a turbine to move wires
near magnets to generate electricity,
just like coal-fired power plants do? The
uranium needed to fuel a nuclear power
plant for a year could fit under your
dining room table, but the coal needed to
generate the same amount of electricity
for a year would fill a 150-car coal train
every day for an entire
year. Nuclear fission At 5% enrichment, 20
provides reliable, tons of uranium—which
carbon-free electricity takes up about 1 cubic
to millions of homes and meter of space—is
needed to fuel a power
powers submarines and
plant that produces 1
aircraft carriers. u gigawatt of electricity
for 1 year.

175
Fission Reactors
wwFissioning uranium-235 releases energy. Fission rearranges the
nucleons in nuclei to release energy. (Chemical reactions, on the other
hand, rearrange electrons in atoms to release energy.) The energy
released is used to boil water to turn a turbine to turn a generator to
generate electricity.

wwWe want a critical mass of uranium so that the fission keeps on


going. When uranium fissions, it gives off about 2.5 neutrons. When
plutonium fissions, it gives off 2.9 neutrons. We want only one of those
neutrons to fission the next uranium or plutonium nucleus.

wwMost neutrons are prompt, but about 0.65% are emitted later from
radioactive decay. Those delayed neutrons are important for stable
reactor operation.

wwNeutrons come out in a broad energy range, from 0 to 8 million


electron volts. High-energy neutrons have a much lower cross section
to fission the uranium and are much more likely to be captured by
uranium-235, not fission it. So, the cross section depends on the
neutron energy. This means that the neutron interaction length
depends on the neutron energy, because the bigger the cross section, the
shorter the distance the neutron travels before it interacts.

176 Nuclear Physics Explained


wwSlower neutrons have a much bigger fission cross section and a much
shorter interaction length, so we need to slow the neutrons down, or
moderate them, through lots of elastic collisions with other nuclei.

wwWhich nuclei do we want to use? The closer in mass the 2 particles


are in a collision, the more they share the energy in a collision. We
want the neutrons to have elastic collisions and bounce off of light
nuclei—such as protons, deuterons, carbon nuclei—so we want to
moderate them.

wwAfter lots of collisions, the neutrons thermalize, which means that


they end up with the same average kinetic energy as the hydrogen,
deuterium, or carbon atoms. They are at room temperature, which
tells us that the average kinetic energy of each is about 0.025 of an
electron volt.

wwTo keep the chain reaction going, at least 1 neutron from each fission
has to first thermalize, which means dramatically decrease its energy;
not escape the reactor; interact in the fuel (not in the moderator, or in
the other elements); and fission the fuel, rather than being absorbed.

wwWe want the neutron multiplication factor (k) to equal 1 for stable
reactor operation. (We want only 1 neutron from each fission to go on
to fission other nuclei.) This factor depends on the fuel enrichment, the
moderator, and other material that might be in the reactor.

wwHow quickly do operators need to react when k gets bigger than 1? This
depends on the mean neutron lifetime between the emission of the
neutron and when that neutron goes on to fission the next uranium.
For prompt neutrons, that lifetime is 0.1 milliseconds. But 0.65% of
the neutrons come from delayed decay with a typical half-life of about
10 seconds. Delayed neutrons dramatically lengthen the multiplication
time, giving us lots of time to react. We just have to make sure that k
stays less than 1.006.

Lecture 17
| Harnessing Nuclear Chain Reactions 177
wwThe neutrons absorbed in the fuel can fission the uranium-235, be
absorbed by the uranium-235 without fissioning to make uranium-236,
or be absorbed by the uranium-238 to make uranium-239.
Uranium-239 beta-decays fairly quickly to neptunium-239, and the
neptunium-239 decays relatively quickly to plutonium-239, which is
fissionable (usable in reactors or bombs).

wwA typical light-water reactor will produce about 1/2 of a plutonium for
each uranium-235 that is fissioned. The plutonium also fissions in the
reactor, so it extends the fuel life of the reactor.

wwHow do we control the reaction rate? Typically, we insert or remove


materials that have large thermal-neutron capture cross sections (if we
can capture more neutrons, then fewer of them go on to fission more
uranium—or maybe we want to capture fewer neutrons). We can
use boron-10, with a cross section of 4000 borons and an abundance
of about 20% of boron, so we don’t have to enrich it. Or we can use
cadmium-113, with a cross section of 20,000 borons and a reasonable
abundance of 12%. To control the reaction rate, the control rods can
be inserted further or removed.

wwIf we want to start with more enriched fuel, we can put burnable
poisons, such as boron, in the fuel rods to absorb neutrons.
Boron-10 absorbs neutrons to become boron-11, which no longer
absorbs neutrons. Uranium-235 absorbs neutrons and fissions.

wwIf we start with more uranium-235 in the fuel rods, we’ll also put
in some boron-10. Over time, the uranium-235 level will decrease,
making the reactor less reactive. The boron-10 level will also decrease,
making the reactor more reactive. This keeps the total reactivity
approximately constant and lets the reactor operate more stably.

178 Nuclear Physics Explained


Human-Built Reactors
wwThere are generally 2 purposes for human-built reactors: to produce
electrical power and to produce electrical power and generate
plutonium. There are 2 main types of reactors: thermal-neutron
(slow-neutron) reactors, which use a moderator to slow the neutrons
down to thermal energies; and fast-neutron reactors, which don’t use a
moderator and are higher in energy.

wwA conventional reactor needs fuel—uranium-235, usually in the


form of uranium oxide, in pellets in fuel rods. A typical fuel rod is
about 1 centimeter in diameter and 4 meters long and weighs about
10 pounds. The reactor also needs a moderator, either light water,
heavy water (deuterium oxide), or carbon. It needs a coolant, which
transfers the energy from the hot fuel to the electrical turbine and
keeps the core from melting down. Typically, light water, heavy water,
or (if carbon is a moderator) a gas (such as carbon dioxide or helium) is
used. The reactor also needs control rods to control the reaction.

wwTo turn heat into electricity, we


use the heat to boil water to turn a
turbine to spin a generator. Then,
we need a cold sink to dump the
waste heat into. That cold sink can
be a river, a lake, or an ocean, or
we can use a cooling tower, which
cools by evaporating a lot of water
and has a shape that maximizes
airflow. The thermodynamic
efficiency is about 1/3, which
means that 2/3 of the energy has to
be dumped into a cold sink. The
same process of turning heat into
electricity is used for coal, natural
Cooling Towers
gas, and nuclear power.

Lecture 17
| Harnessing Nuclear Chain Reactions 179
Today’s Reactors
wwIn light-water (regular hydrogen, Light-water reactors were
not deuterium) reactors, the originally developed for the
water is both the moderator and US Navy under the guidance
the coolant. There are 2 general of Hyman Rickover.
types of light-water reactors. In
pressurized-water reactors, water
is under high pressure so that it doesn’t boil (higher pressure means
a higher boiling point), and the primary cooling water transfer its
heat to a secondary water system, which boils and turns a turbine. In
boiling-water reactors, the primary water boils and uses the steam to
turn a turbine, so it doesn’t need a secondary water system.

wwA pressurized heavy-water reactor, such as the CANDU plant in


Canada, uses a heavy-water moderator (deuterium oxide) and a
heavy-water primary coolant, so it can use unenriched uranium. The
heavy water transfers its heat to a light (regular) water secondary
system, which then boils and turns the turbine.

wwA water-cooled graphite-moderated reactor is similar to Fermi’s original


reactor in that it is graphite-moderated, but now it is water-cooled. It
can also run on natural uranium, so this is a Chernobyl-type reactor.
It’s built for power and for producing bomb-grade plutonium.

wwThe pressurized-water reactor—the most common reactor in use


today—uses pressurized water (at 150 atmospheres and about 300°
Celsius, or 600° Fahrenheit) as a coolant and a moderator in the
reactor vessel. It’s a closed loop, so the water doesn’t escape. There is
a reactor pressure vessel that is about 40 feet high, 15 feet in diameter,
and 8 inches thick. The primary water
system transfers its heat to a secondary
The first nuclear
water loop in the steam generator. It has a
reactors were placed
containment building with concrete and
on submarines.
steel to contain everything inside of it.

180 Nuclear Physics Explained


wwThe steam from the secondary water system turns a turbine, which
turns a generator. A condenser condenses the steam to make the
low-pressure side of the turbine. The cooling source for that condenser
can be an ocean, a lake, a river, or a cooling tower.

wwFor fuel, a pressurized-water reactor uses uranium oxide pellets in


4-meter-long, 1-centimeter-diameter zirconium alloy (Zircaloy) fuel
rods. There are about 50,000 fuel rods in the reactor core. The fuel
rods remain in the core for about 3 years. They are typically refueled or
replaced every year or 2. The plant has to be shut down for refueling.

wwWhat makes a pressurized-water reactor safe? First, there are multiple


levels of containment: Uranium oxide pellets retain most of the
radionuclides. The pellets don’t escape because they are enclosed
in zirconium alloy fuel rods, which are enclosed in a pressure vessel
and a closed primary coolant loop. And all of that is enclosed in a
containment vessel.

Lecture 17
| Harnessing Nuclear Chain Reactions 181
wwThere are active controls,
There are about 440 nuclear
in the form of controller
power plants around the globe,
rods that can move in and
which generated 11% of the
out. There are also passive
world’s electricity in 2014. There
controls: The moderator is
is no carbon dioxide emission.
the coolant, so if there is
any loss of coolant due to §§ In France, nuclear power is
boiling or to a leak, then used to generate 75% of
there is less moderator. their electricity.
Decreased moderation
§§ In Japan, nuclear power is used to
decreases the neutron
generate 30% of their electricity
multiplication factor and (but all of Japan’s plants were
stops the chain reaction. idled after Fukushima).

§§ In the United States, nuclear


power is used to generate about
20% of their electricity, using
Supplements about 100 reactors.

READINGS

Bodansky, Nuclear Energy, chaps. 6–8 and 14.


Cowan, “A Natural Fission Reactor.”
Lilley, Nuclear Physics, sections 10.5–10.6.
Mahaffey, Atomic Awakening.
Morrison, “Where Fiction Became Ancient Fact.”
“The Oklo Reactor,” re-actions.
World Nuclear Association, “Nuclear Power in the World Today.”

QUESTIONS

1 If a power reactor uses reactor fuel that is enriched to 4%


uranium-235, how could the reactor fission more than 4% of the
nuclei in the fuel to release more energy?

182 Nuclear Physics Explained


2 Do reactor control rods containing cadmium get used up?

3 Can a reactor be made using unenriched uranium?

4 Do all nuclear reactors use the iconic cooling towers? Are cooling
towers only used by nuclear reactors?

ANSWERS

1 Some of the uranium-238 in the reactor core will absorb neutrons and end
up (after 2 beta decays) as plutonium-239. Some of the plutonium-239 will
be fissioned in the reactor.

2 Yes. When cadmium-113, with a neutron absorption cross section


of 20,000 barns, absorbs a neutron, it becomes cadmium-114, with a
much lower neutron absorption cross section. As more and more of the
cadmium-113 nuclei absorb neutrons, the control rod becomes less and
less useful.

3 Yes, but it needs to use carbon or heavy water for the moderator (because
the hydrogen in regular water will absorb too many neutrons) and it needs
to be very large so that fewer neutrons will escape the reactor. The very first
nuclear pile used natural uranium.

4 No and no. Nuclear reactors use about 3 gigawatts of thermal power from
uranium fission to produce about 1 gigawatt of electrical power. This means
that they need to dump 2 gigawatts of thermal energy somewhere. Many
reactors dump the waste heat into rivers, lakes, or oceans. Coal- and natural
gas–fired power plants also only convert about 1/3 of their thermal energy
into electrical energy and also need to dump the other 2/3. Like nuclear
power plants, coal- and natural gas–fired power plants also use rivers, lakes,
oceans, or cooling towers.

Lecture 17
| Harnessing Nuclear Chain Reactions 183
18 NUCLEAR
ACCIDENTS AND
LESSONS LEARNED

N
 clear power plants are extraordinarily
u
safe most of the time, but there have
been a few dramatic accidents, including
Three Mile Island, Chernobyl, and
Fukushima. From these accidents, we
have learned to make sure that the
safety system is multilayered; accidents
are caused by multiple mistakes aligning
badly, so layers of a safety system should
not have a common failure point. We
also learned to make sure that the chain
reaction stops. In addition, we learned
to make sure that the cooling continues;
we need long-term passive cooling, which
requires a continuous supply of water. u

184
Three Mile Island Core meltdowns
are typically caused
wwThe biggest disaster that didn’t kill
by a loss-of-coolant
anyone was in March of 1979 at Three
accident. A chain
Mile Island, which was a 900-megawatt
reaction stops in a
electric pressurized-water reactor built
pressurized-water
by Babcock & Wilcox. It had only been
reactor due to a loss
running since 1978.
of moderator, but
residual power is
wwAt 4 o’clock in the morning, the
still generated—200
secondary coolant water flow was
interrupted. This was probably caused
megawatts of
by system maintenance. There was an
power immediately
emergency shutdown of the reactor, and after, dropping
the control rods were inserted. This to 16 megawatts
is called a scram. The chain reaction after 1 day. If there
(fission) was stopped, and the emergency is no coolant, the
backup pumps were started. temperature of
the core rises
wwThe problem was that the emergency dramatically,
backup pump valves were closed for and the fuel rods
maintenance. This was the key failure. in the reactor
The operators didn’t notice this. The core melt down,
primary coolant—the water—heated forming corium.
up and turned to steam. A relief valve
A core meltdown can
opened to vent the steam from the
ruin a multibillion-
primary coolant. The relief valve stuck
dollar reactor, but
open, but the indicator for the valve in
there is typically no
the control room showed that it was
explosion, no escape
closed, so the valve stayed open for
of core material,
2.5 hours. During this time, the coolant
and no danger to
boiled off.
the public unless a
lot of radioactivity
is released.

Lecture 18
| Nuclear Accidents and Lessons Learned 185
Three Mile Island

wwThere was a lot of confusion in the control room. There were


more than 1000 dials, gauges, and indicators; 600 alarm panels;
and hundreds of switches. A minor problem would cause alarms
and blinking lights, and there could be 50 alarms at a time. The
temperature of the fuel increased, and the fuel melted down.

186 Nuclear Physics Explained


wwIt was recommended that pregnant women and preschool children be
evacuated, and another 130,000 people self-evacuated. They worried
that there was a hydrogen bubble inside the reactor vessel that might
explode. Fortunately, there wasn’t enough oxygen in the reactor
vessel for the hydrogen to react with, so there couldn’t have been an
explosion. But there was a lot of public concern.

wwHow bad was Three Mile Island? The core melted down, but the
corium—the melted-down core materials—was contained in the
reactor pressure vessel. The containment worked, and there was very
little radiation release. The maximum dose to any of the workers at the
power plant was about 4 rem, which is less than the annual safety limit
of 5 rem. The maximum offsite dose was less than 0.1 rem, which is
less than the annual safety limit for the general public.

wwOnly gases escaped the core. Iodine-131 bonded to the concrete of the
containment building, so only about 15 curies out of 64 megacuries
were released. Xenon-133 was released, but that had almost no
biological effect; xenon is a noble gas, so it doesn’t bond to anything,
and it dispersed quickly.

wwThe maximum dose rate just outside the reactor vessel was about
1.2 rem per hour. The total collective dose would cause about 1 extra
cancer (assuming the very conservative linear no-threshold hypothesis).
We would expect about 300,000 natural cancer deaths out of 2 million
people in the area, so which 1 was actually due to Three Mile Island?

wwThere were no measurable adverse health effects, but this disaster


disrupted the lives of more than 100,000 people due to evacuations. It
cost about a billion dollars to clean up and wrecked a 5-billion-dollar
nuclear power plant.

Lecture 18
| Nuclear Accidents and Lessons Learned 187
Is nuclear power too much trouble?
Comparatively:
§§ Oil can spill and has geostrategic risks.

§§ Coal is dangerous to mine and creates air pollution


when burned.

§§ Hydropower dams rivers and floods scenic canyons,


and dams can burst with catastrophic effect.

§§ Wind is intermittent and expensive, but it is getting


cheaper. It already kills hundreds of thousands of
birds yearly.

§§ Solar is intermittent and expensive, but it is


getting cheaper. It uses a lot of land, with impact
on ecosystems.

Most nonnuclear deaths are one at a time and


do not make the news. Estimates of accidental
and air pollution–related deaths in 2007 found
that nuclear had been safer and cleaner than
oil, coal, biomass, or gas. But 99% of those
nonnuclear deaths are attributed to air pollution,
and nuclear power does not emit carbon dioxide.
How do you want to generate our terawatts of
electricity?

188 Nuclear Physics Explained


Chernobyl
wwThe worst reactor accident was on April 26, 1986 at Chernobyl, which
was not a Western light-water reactor. It was a graphite-moderated and
water-cooled reactor, so there was water in vertical metal pipes passing
through the core, and the water absorbed the heat and boiled.

wwIt was an RBMK-1000 reactor, generating 1000 megawatts of


electricity. There were 4 operating at the Chernobyl site, and 2 more
were under construction. Unit 4 had just been built in 1983, and units
1 and 3 continued operation until 1996 and 2000, respectively. They
operated on natural uranium because they didn’t need to enrich the
uranium with a graphite moderator. This is similar to Fermi’s original
reactor, which he built in Chicago by piling up graphite and uranium.

wwThis reactor was optimized for plutonium production for bombs, which
meant that the fuel rods could be swapped into and out of the reactor
while it was running. This is similar to the reactors in the United States
that were built to make plutonium. We want plutonium-239 with
as little other plutonium isotopes as possible, which means that we
want to remove the fuel rods and extract the plutonium after some
plutonium-239 has been made, but before it can absorb another
neutron and make plutonium-240.

wwChernobyl had some design issues. It was much bigger than the
submarine-influenced light-water reactor designs in the United States
and was too large for a convenient containment structure. The core was
46 feet in diameter and 23 feet high, and 12-foot-long fuel rods were
inserted vertically.

wwThe moderation was mostly done by carbon. The cooling water could
do a little moderation, but it mostly absorbed neutrons. If the amount
of water decreased, then there was a little less moderation, which
is negative feedback, and less neutron absorption, which is positive
feedback. Chernobyl had net positive feedback—which is very bad.

Lecture 18
| Nuclear Accidents and Lessons Learned 189
wwThere was also a control rod design defect.
The control rods were removed by lifting
them upward. To keep the water from
filling the control rod channel (and also
from absorbing neutrons), a graphite
displacer was attached. But there was
about 1.25 meters, or 4 feet, of water at
the bottom of the control rod channels.
Inserting the control rod initially replaced
the water with carbon, and that initially
makes things worse. In addition, the
rods moved much more slowly than
needed—18 seconds to insert.

wwThe safety test went wrong. They planned


to test the function of the pump during
loss of power, so they reduced the power,
but they overshot. The buildup of xenon (fuel poison) caused the
operators to remove 205 out of 211 control rods. They ignored the
unexpected conditions and continued the test.

wwThe test reduced steam flow to the turbine, which decreased the
water flow in the reactor, which meant more boiling, less water, and
more reactivity—positive feedback. This led to an increase of control
rods, which made things worse, increasing the reactivity to prompt
supercritical. This means that the timescale for the exponential
increase of the reactivity was very short. If there was a sharp increase in
power output, it might have hit 30 gigawatts. It’s not entirely clear what
happened next.

wwWithin about 20 seconds, there were 2 explosions. The first was a


steam explosion, which exposed the reactor fuel to air. The second was
either a chemical explosion between the reactor material and the air or
runaway reactor criticality, which is like a very-low-yield nuclear bomb.
These explosions started fires.

190 Nuclear Physics Explained


wwThe big tanks of water under the reactor turned to steam and exploded,
blowing the 500-ton reactor cap off. Then, there was a hydrogen
explosion and fire, and the core and radioactive debris—and the
fission products in the core—were widely dispersed. Most of these are
volatile nuclides.

wwFirefighters weren’t told what they were fighting. Water doesn’t put out
graphite fires. Eventually, they figured it out and dumped a lot of sand,
lead, and boron, etc., from helicopters.

wwNo announcement was made. About 2 days later, in Sweden and


Finland, radiation alarms started going off and the radiation alarm
at a Swedish nuclear power plant went off because radioactive
contamination was tracked in on a plant worker’s shoes.

wwThere were about 50 tons of vaporized uranium. That wasn’t a


problem; the uranium is not significantly radioactive. But there was a
plume of other material, including noble gases, such as krypton-85 and
xenon-133; 48 megacuries of iodine-131 (which can damage the
thyroid gland), with a half-life of 8 days; 1.5 megacuries of cesium-134,
with a 2-year half-life; 2.3 megacuries of cesium-137, with a 30-year
half-life; and 0.3 megacuries of strontium-90 (which, like radon, seeks
bone), also with a 30-year half-life. This airborne plume of material fell
to ground—which is why it’s called fallout—across Europe.

wwAbout 130 plant workers and firefighters came down with acute
radiation syndrome, and 30 of them died from the radiation exposure.
There was population exposure from ingestion and ground exposure.
The linear no-threshold hypothesis predicted about 9000 excess
cancer fatalities.

wwAbout 300,000 people were evacuated. There were several thousand


measurable extra thyroid cancers due to iodine-131, which amounted to
about 15 extra deaths from thyroid cancer.

Lecture 18
| Nuclear Accidents and Lessons Learned 191
wwSignificant areas of the Ukraine, Belarus, and western Russia were
contaminated with about 5 curies or more per square kilometer. For a
100-kilogram person, this equates to an external dose of 6 × 10−4 joules
per kilogram in a year, which is a dose of 60 millirem in a year. Our
background radiation dose is 600 millirem, so this contamination was
relatively small.

wwThis predicted a 70-year population exposure of 600,000


person-sieverts, or 60 million person-rem. In the first year, this
increased the average radiation dose in non-former-Soviet Union
Europe by about 8%. If we apply the linear no-threshold hypothesis,
this predicts 9000 extra cancer deaths over 70 years—out of
100 million natural cancers.

wwTwenty years after the accident, the United Nations Scientific


Committee on the Effects of Atomic Radiation said that except for
thyroid cancer, there was no evidence of a major public health impact
caused by radiation exposure. (Thyroid cancers are otherwise very rare,
so it’s much easier to identify extra thyroid cancers than extra other
kinds of cancer.)

wwToday, there’s a new, very expensive dome enclosing the reactor. They
didn’t think they needed it before the accident. You can tour the
contamination zone.

Fukushima
wwThe most recent nuclear accident was at the Fukushima Daiichi nuclear
power plant, which had 6 boiling-water reactors, 3 of which were shut
down at the time for refueling. Seawater cooling was used, which
involves pumping in seawater to help cool the power plant. There was a
19-foot seawall to protect against tsunamis.

192 Nuclear Physics Explained


Fukushima

wwThere were lots of safety cooling systems. But those systems all needed
electricity, supplied first by the power plant, second by power from
the grid, third by 2 diesel generators per reactor, and fourth by battery
backups that could run those plants for 8 hours.

wwOn March 9, 2011, there was a magnitude 7.2 earthquake. The reactor
scrammed—so fission stopped—and then restarted. There was no
problem. Just a few days later, on March 11, there was a magnitude
9.0 earthquake. The reactor scrammed, and fission stopped. Japan
moved 8 feet closer to the United States. More than 383,000 buildings
were destroyed by the earthquake. Only 41 minutes later, a 13-foot
tsunami hit, with no damage, but 8 minutes after that, a 49-foot
tsunami overwhelmed the seawall.

Lecture 18
| Nuclear Accidents and Lessons Learned 193
wwThe water intake structures collapsed. The pumps were blown down.
The electrical lines were demolished. The basement generators were
flooded. One above-ground generator survived for units 5 and 6. Units
3 and 4 were on battery power, with backup passive cooling for unit 3.
Units 1 and 2 had no power.

wwThe operators scavenged all the car batteries in the parking lot to
briefly power the controls, but unfortunately, there was no cooling for
unit 1 and only backup passive cooling for unit 2. They needed cooling
for the residual core power in reactors 1, 2, and 3. But roads were
destroyed, the interconnect cables they could have used to connect the
functioning generators with the nonfunctioning reactors were lost, and
the existing 4-inch cables weighed a ton each.

wwUnit 1 had no power. The water boiled away in a few hours. The fuel
assemblies heated up, sagged, and collapsed. The 1-inch-thick reactor
vessel burst. There was hydrogen gas in the containment building, and
they needed to open the vent stack valves—manually, in a radioactive
environment—to vent the hydrogen.

wwThey failed to reconnect power to the other units. The passive cooling
eventually failed as the last of the water boiled away. There were
hydrogen explosions from gas generated by the meltdowns, and the
cores of units 2 and 3 also melted down.

wwNo spent fuel was damaged. All the radiation came from damaged
cores exposed to steam. All the cores were contained in the pressure
vessels. The maximum exposure was about 60 rem to 2 operators
whose respirators didn’t fit properly over their glasses and didn’t seal.

wwThe radiation release was about 10% of Chernobyl. Radiation levels


about 50 kilometers from Fukushima were only about 1/10 of what you
get in an airplane. There was also lots of confusion and fear.

194 Nuclear Physics Explained


wwThe earthquake and tsunami damaged or destroyed about a
million buildings, killed 18,000 people, and displaced more than
300,000 people. The nuclear accident killed zero people directly,
caused 150,000 people to be evacuated from a 20-kilometer radius (and
there were a few hundred deaths attributed to the evacuations), wrecked
4 very expensive reactors, and should cause about 130 expected extra
cancer deaths (if we assume the linear no-threshold hypothesis). As a
result of Fukushima, Japan shut down all of its power plants to evaluate
their tsunami risk.

Other Reactor Incidents


There were 5 major accidents before 1970 in
test or experimental reactors. Together, they
caused 3 deaths and 1 significant radiation
release that might have resulted in about 300
expected thyroid cancers.
In Idaho Falls in 1961, a technician manually
and rapidly removed a control rod. This
increased the reactor power, led to a steam
explosion, and killed 3 people.
At Browns Ferry in 1975, which was a
commercial power reactor, there was a fire in
the electrical wiring. There was no damage to
the reactor, but there was a lack of redundant
safety features that has since been corrected.
We have little knowledge of Soviet/Eastern
Bloc accidents.

Lecture 18
| Nuclear Accidents and Lessons Learned 195
Supplements

READINGS

Bodansky, Nuclear Energy, chaps. 15.


Dworschak, “Is Radiation as Bad as We Thought?”
International Nuclear Safety Advisory Group, INSAG-7—The
Chernobyl Accident.
Mahaffey, Atomic Accidents.
National Research Council, Health Risks from Exposure to Low
Levels of Ionizing Radiation
——— , Nuclear Physics, p. 209–212.
United Nations Scientific Committee on the Effects of Atomic
Radiation, Sources and Effects of Ionizing Radiation.

QUESTIONS

1 What made the Chernobyl accident so much worse than either


Fukushima or Three Mile Island?

2 Why can’t we measure the extra 9000 cancer deaths expected in


Europe due to the Chernobyl accident?

3 How many things had to go wrong in series to cause the


Fukushima meltdown?

196 Nuclear Physics Explained


ANSWERS

1 In Fukushima and Three Mile Island, the nuclear reaction had been
stopped. The reactor cores melted down because of the residual heat
generated by radioactivity in the spent reactor fuel. The core meltdowns
were contained by the containment vessels. At Chernobyl, the nuclear
reaction had not been stopped and probably went out of control. Therefore,
much more energy was released. In addition, the Chernobyl reactor did not
have a containment vessel, so the explosion distributed the spent reactor fuel
and its by-products over a huge area.

2 This is because about 1/5 of the population will naturally die of cancer, and
this amounts to a total of about 50 to 100 million natural deaths from
cancer. The extra 9000 deaths attributed to Chernobyl are only about
1/5000 of the overall rate. This tiny effect is impossible to measure, because it
is less than the natural variation in the cancer death rate.

3 (a) The tidal wave had to be higher than the seawall, (b) the tidal wave had
to knock out the power lines connecting the reactor to the outside world, (c)
the tidal wave had to knock out the emergency generators, and (d) they had
to be unable to connect the generator powering units 5 and 6 to the other
units. The tidal wave overwhelmed the seawall at Fukushima II, but one
of the power lines and some of the generators and pumps survived, so they
could keep the reactor cores cool.

Lecture 18
| Nuclear Accidents and Lessons Learned 197
19 THE NUCLEAR
FUEL CYCLE
AND ADVANCED
REACTORS

N
 clear fission power plants are getting
u
better and better. The first-generation
power plants, the initial plants, are all
retired. The second-generation power
plants, built beginning decades ago, are
the ones that are now in service. They
are safer, more reliable (providing power
more than 98% of the time that they’re
expected to), and more fuel efficient.
The third-generation power plants are
starting to be built in the 21st century,
and the fourth-generation power plants
are being designed. The goal is to improve
safety, radioactive waste, proliferation,
and cost. u

198
Third-Generation Power Plants
wwThe third-generation power plants are new light-water reactors. They
are an evolution of existing design. They use thermal neutrons, which
means that the neutrons are moderated, or slowed down. They have
more passive cooling and active safety systems and are designed to
be much simpler, with less wiring and piping, which should make
them cheaper.

wwDifferent designs are approved to be built in China, Europe, and the


United States, and a few new plants are being built. They should be
safer. But we don’t know yet if they will be cheaper. They are going
to burn more of their fuel, which means that they will require less
fuel, reducing waste. They are good for proliferation because they use
low-enriched uranium and leave the fuel in the core longer, reducing
the bomb potential of the plutonium.

wwHow can we produce nuclear power more cheaply, more safely, with less
waste and less risk of nuclear proliferation?

§§ According to numbers from 2014, construction and finance costs


dominate a nuclear plant, at 80% of the costs. The fuel costs are very
small, about 10% of the costs. There is also a small surcharge for
waste disposal. The big thing to reduce is construction costs.

§§ For safety, we need to make sure that the reaction stops and that the
coolant keeps flowing once the reaction has stopped.

§§ We can reduce waste in a few ways: by improving the thermodynamic


efficiency (if we can run the reactor at a higher temperature, then
we need less fuel), using more of the fuel (burn up a higher fraction
of the fuel in the reactor), and processing the waste to reduce the
amount of waste. We can also transmute transuranics —elements

Lecture 19
| The Nuclear Fuel Cycle and Advanced Reactors 199
heavier than uranium, such It takes 3 gigawatts of
as neptunium, plutonium, energy released from
americium, and curium—or nuclear fuel to make 1
have a reactor that doesn’t make gigawatt of electric power.
transuranics and uses thorium.

§§ For proliferation, we want to make less plutonium-239. When we do


make it, we want to contaminate it with other plutonium isotopes so
that it’s much more difficult to use in a bomb.

Spent Fuel
wwNuclear waste comes from the nuclear fuel cycle. About 2/3 of the
nuclear fuel is used once through and then dumped. We put the fuel in
a reactor, fission the uranium, get the energy out, and remove the spent
fuel and dispose of it. In the United States, that’s all that is done with
the fuel.

wwFrance, Great Britain, etc., reprocesses the spent fuel. They remove
the uranium and plutonium from the spent fuel. This reduces the
volume of the waste tremendously, and they can reuse the uranium and
plutonium in other reactors.

wwWe can also breed more fuel and then reprocess and use it. We can use
uranium-238 to breed plutonium and then fission that plutonium in a
power plant. But uranium is a tiny fraction of the total cost of nuclear
power, so there is very little financial incentive to reprocess the fuel or
to breed more fuel.

wwHow do we use more of the fuel? We fission, or burn, more of the


uranium. We can either fission the uranium-235 or convert the
uranium-238 to plutonium-239 and fission the plutonium-239.

200 Nuclear Physics Explained


wwThe burnup rates have increased from about 2% in 1973 to 4.5%
to 5% today. We use more enriched fuel; it’s now about 4.5%
uranium-235. To make that more enriched fuel usable, add burnable
poisons to keep the reactivity constant. We leave the fuel in the reactor
longer, which means that we need less fuel. More uranium-235 is
consumed, and more plutonium is made and fissioned. We need to
refuel less often. This is good because power plants need to be shut
down to refuel and less time is wasted refueling.

wwA higher burnup rate increases the plutonium production, but the
fuel is spending more time in the reactor, which means that more
plutonium-239 absorbs neutrons to make plutonium-240 and
plutonium-242. The ratio of plutonium-240 to plutonium-239
increases. Plutonium-240 has a much higher spontaneous fission rate,
which means that it’s much more difficult to make bombs with it.

wwHow much radioactive waste has been made in the United States so far?
It’s split into categories. There are 20 million cubic meters of low-level
waste. This is 90% of the volume of all nuclear waste but only 1% of
the total radioactivity. Almost all of that is disposed in near surface, or
special landfills.

Decommissioning a nuclear power plant is paid


for by funds collected from the power plant or put
aside by the power plant during operation. The plant
material is much less radioactive than the spent
fuel. In fact, almost all of it is low-level waste. Many
power plants have been safely decommissioned.

Lecture 19
| The Nuclear Fuel Cycle and Advanced Reactors 201
wwHigh-level waste includes the spent fuel. There are 22,000 cubic meters
of high-level waste, which is about 1000 times less than the amount of
low-level waste by volume. About 1000 cubic meters per year is being
made. This is not that much for all of a country’s nuclear power plants.
High-level waste needs cooling. It’s about 3% of the volume of all the
waste we make and about 95% of the radioactivity.

wwThe spent fuel is very, very radioactive. There are some


very-short-half-life nuclides that are very radioactive but decay rapidly.
When the spent fuel is taken out of the power plant, there are about
7 gigacuries initially, which puts out about 40 megawatts of power, so it
needs lots of cooling. A year later, there is 100 times less radioactivity,
at 74 million curies, and 100 times less power, at 0.5 megawatts—so
it still needs lots of cooling, via water bath. After 10 years, it decreases
to about 14 million curies (about 5 times less), so we can take it out of
water baths and air-cool it.

wwWhat is left in the spent fuel? Strontium-90 and cesium-137,


which have 30-year half-lives, decay first. Plutonium-241 decays
to americium-241, which has a 400-year half-life. The long-term
radioactivity (more than 100 years) is due to the transuranics—such as
neptunium, plutonium, americium, and curium—and their daughter
nuclei. After 10,000 years, the radioactivity is comparable to the
natural radioactivity of the Earth.

wwThe spent fuel is initially very hot, both thermally and radioactively,
so the fuel assembly is moved remotely. The fuel assembly is put in a
water-filled cooling pool. It will be cool enough to reprocess the fuel
after a few years, and it will be cool enough to put in dry storage after
about 5 years. Dry storage involves steel casks with air cooling and
multiple levels of containment. We don’t have any long-term storage
solutions yet.

202 Nuclear Physics Explained


wwHow do we safely dispose of anything for millennia? We use multiple
barriers. We immobilize the waste in an insoluble matrix (borosilicate
glass or uranium oxide fuel pellets). Then, we want to put the waste in
a corrosion-resistant container, made of stainless steel. We isolate the
waste from people and the environment deep underground in a stable
rock formation. We fill it with an impermeable material to reduce
water transport.

wwAbout 1/3 of all spent fuel has been reprocessed. To reprocess it, we
separate the uranium and the plutonium from the rest. The uranium
and plutonium are much less radioactive, so the remaining radioactive
waste has much less mass (but the same radioactivity).

wwTo recycle this spent fuel, we remotely dissolve the radioactive fuel rods
in acid and extract the uranium and plutonium chemically. There is a
proliferation risk from extracting the plutonium.

wwWe can reuse the plutonium. We make mixed oxide fuel: 95%
uranium-238 and about 5% plutonium. Then, we can fuel our
power plants with 1/3 mixed oxide fuel and 2/3 low-enriched
uranium. The reason we don’t use more mixed oxide fuel is that the
plutonium-239 has a smaller delayed neutron fraction, so that makes it
more difficult to control as a reactor fuel.

Yucca Mountain in the United States was designated by


Congress as the sole US depository 30 years ago. It’s not
yet in operation because it’s very politically contentious.
The Waste Isolation Pilot Plant in New Mexico is 2000
feet below ground in a salt formation. It’s for bomb-related
transuranic waste only, which is a much smaller volume
of waste. In 2014, one of the drums of waste exploded,
releasing some radioactive material, but it reopened in 2016.

Lecture 19
| The Nuclear Fuel Cycle and Advanced Reactors 203
Advanced Fission Reactors
wwThere are 2 general types of advanced fission reactors: ones that use
fast neutrons and ones that use slow neutrons (moderated thermal
neutrons). Slow-neutron reactors use uranium-235 and a graphite or
water moderator to slow down the neutrons.

wwFast reactors use uranium-238 or a mix of uranium-235 and


plutonium-239. They don’t use a moderator; instead, they use a heavy
coolant. They use a mix of fissile and fertile fuel (the fissile part can
fission and the fertile fuel can absorb a neutron) and make more fuel.
Uranium-238 becomes plutonium-239, or thorium-232 becomes
uranium-233, or plutonium is burned to reduce stockpiles. There are
accelerator-driven subcritical reactors, and the advantage is that they
are always subcritical.

wwAn example of a new thermal-neutron reactor design is the


high-temperature gas-cooled reactor, which is helium-cooled and
graphite-moderated. Helium doesn’t absorb neutrons, which means
that if there is less coolant, then there is no positive feedback from
cooling water boiling, like Chernobyl had.

wwThis type of reactor runs in higher temperatures (1000° Celsius),


which means more thermal efficiency. If the temperature in the core
increases, then the uranium-238 atoms in the fuel move faster, which
means that the uranium-238 can absorb more neutrons, which is
negative feedback.

wwWith this type of reactor, there is also a smaller ratio of fuel to


moderator, so there is a greater thermal inertia and less temperature
increase if it loses cooling. The fuel is stable up to 2000° Celsius, but
air must be kept out because carbon can burn and cause a fire. This
reactor uses 8% enriched fuel. It makes 1-millimeter microspheres of
fuel encapsulated in layers of protective carbon. It should be safer, and

204 Nuclear Physics Explained


we don’t know if it will be cheaper. But there will be more thermal
efficiency, so it needs less fuel and produces less waste. There’s no
change in proliferation.

wwOne of the concepts for this high-temperature gas-cooled reactor is a


pebble-bed reactor, where there are 1500 of these microspheres in each
6-centimeter carbon pebble. It uses about half a million pebbles for a
reactor. The pebbles are continually inserted at the top and removed
from the bottom, so there is no need to shut down the power plant
to refuel.

wwThere are also fast reactors, where “fast” refers to the neutrons. These
use a mix of fissile and fertile fuel—plutonium or uranium-235 to
fission and uranium-238 and thorium-232 to make more fuel. These
2 isotopes of uranium and thorium are 99% abundant. We don’t need
to do any enrichment.
PEBBLE BED REACTOR SCHEME
wwThere is no moderation
that uses fast neutrons.
The plutonium has a
low-absorption cross
section for neutrons
but a high-fission cross
section (of 2 barns), so
it’s much more likely
to fission than to
absorb a neutron. The
coolant needs to be
a high-mass coolant,
such as sodium metal
or sodium chloride.

Lecture 19
| The Nuclear Fuel Cycle and Advanced Reactors 205
wwWe can reprocess the spent fuel. We can separate out the transuranics
and return the uranium and plutonium to the reactor in new fuel rods.
The safety of the fast-neutron reactor is due to negative temperature
feedback. Increased temperature means the core expands thermally,
which means less density and less reactivity.

wwWe don’t know yet if a fast-neutron reactor is cheaper, but it does


produce less waste because it breeds more fuel and burns the waste.
It could be run to just burn plutonium, reducing the plutonium
stockpiles, or it could make more plutonium and chemically separate it
for use in other reactors.

wwThere are different coolants for fast-neutron reactors. They can use
helium at high pressure and at high temperature, so there is high
thermal efficiency. We can use molten lead, the advantage of which
is that it’s low pressure and relatively inert. Or we could use sodium,
which is also low pressure and has a more advanced design.

wwWe can even dissolve the fuel in the coolant, which would be a
molten-salt reactor. We dissolve the fuel in the molten salt, and it
circulates through separated channels in graphite and is brought
together in the core, where it can fission. There is some moderation
because of the graphite. It’s self-breeding and uses fertile fuel, either
uranium-238 or thorium-232. It can achieve very high burnup, but
there are no prototypes yet.

wwAnother exciting possibility is a liquid fluoride-thorium reactor,


which would be a fast breeder. We would flow liquid thorium fluoride
through the core and mix it with lithium fluoride or beryllium fluoride
to get the melting point down to 360° Celsius. Fission occurs only in
the core, and the heat is transferred at 700° Celsius to a gas, such as
carbon dioxide or helium, which then drives turbines.

206 Nuclear Physics Explained


wwWe start with some low-enriched uranium and thorium to get the
reaction started. The uranium-235 fissions, and the thorium-232
absorbs neutrons to make fissionable uranium-233. It breeds new fuel,
so we don’t need any more low-enriched uranium after startup. We
continuously remove the waste and add new fuel to the liquid.

wwThe liquid fluoride-thorium reactor should be safer. It works at


atmospheric pressure, so we don’t need high-pressure systems. The
waste should be much less radioactive, with a much shorter half-life—
centuries, not millennia. There is a reduced proliferation risk because it
is much more difficult to make a uranium-233 bomb due to some very
radioactive thallium, which is the decay product of uranium-233. We
don’t know yet if it would be cheaper.

wwThen there is accelerator-driven fission. We take a subcritical reactor


and supply extra neutrons from the accelerator. We accelerate protons
to about 1.6 billion electron volts, hit a molten mercury or lead target,
and knock out lots of neutrons. We moderate the neutrons, and this
then makes the reactor go. In terms of safety, it’s always subcritical,
but—because we would need the accelerator—it would be more
complicated and more expensive.

Lecture 19
| The Nuclear Fuel Cycle and Advanced Reactors 207
wwTo reduce costs, we can make lots of reactors from the same design.
We can use small modular reactors, for which there are many projects
and designs. They should have reduced size and reduced complexity, be
built in factories so that they’re cheaper, and be transportable so that
there are more places that they can be used.

wwThere are many fascinating possible reactor technologies on the


horizon. There are many private companies involved in developing
improved light-water reactors, pebble-bed reactors, fast-neutron
reactors, thorium-fluoride (molten-salt) reactors. All of them should
be safer. Most designs are more complicated, such as breeders, or need
a lot of research and development, such as thorium fluoride. The
small modular reactors are most likely to expand usage and reduce
construction costs. There are many interesting possibilities, but unless
they are a lot cheaper to construct, they will never become the next
generation of nuclear power.

Supplements

READINGS

“Asgard’s Fire,” The Economist.


Bodansky, Nuclear Energy, chap. 16.
Fountain, “On Nuclear Waste, Finland Shows U.S. How It Can Be
Done.”
MIT Center for Advanced Nuclear Energy Systems Symposium,
Nuclear beyond LWRs.
Nutall, Nuclear Renaissance.
Temple, “Small Reactors Could Kick-Start the Stalled Nuclear
Sector.”
World Nuclear Association, “Advanced Nuclear Power Reactors.”
——— , “Storage and Disposal of Nuclear Waste.”

208 Nuclear Physics Explained


QUESTIONS

1 How much less radioactive is the spent fuel from thorium fission
than from uranium fission?

2 Why is negative feedback in a reactor so important?

ANSWERS

1 There is little difference in the radioactivity in the first 100 years, because
that is dominated by the radioactive decay of fission products, especially
cesium-137 and strontium-90. However, after 200 years, most of the
radioactivity comes from transuranics (plutonium isotopes and their decay
products). These are not present in spent fuel from thorium fission.

2 This is important for stable operation. Negative feedback means that


when the reactor rate increases for any reason, other properties will
change to decrease it, and when the reactor rate decreases for any reason,
other properties will change to increase it. For example, in a standard
boiling-water reactor, if the reactor rate increases, then more water will boil.
This will decrease the amount of water in the reactor vessel, which will
decrease the neutron moderation and therefore decrease the reaction rate.
The primary cause of the Chernobyl reactor accident was that the reactor
did not have enough negative feedback, so when the reactivity increased, the
only way to decrease it was to manually insert the control rods, which were
much too slow.

Lecture 19
| The Nuclear Fuel Cycle and Advanced Reactors 209
20 NUCLEAR FUSION:
OBSTACLES AND
ACHIEVEMENTS

N
 clear fusion means fusing hydrogen
u
to become helium, releasing energy. It
promises clean, carbon-free power with
no worries of catastrophes, nuclear
proliferation, nuclear waste, or running
out of fuel. Unfortunately, nuclear fusion
has been promised for 50 years. Some
say that fusion power is 50 years away,
has always been 50 years away, and
always will be 50 years away. u

210
Fusing Deuterium and Tritium
wwWith nuclear fusion, we’re bringing the energy supply of the cosmos
down to Earth. How can we tame it? We need to overcome the electric
repulsion so the 2 nuclei can meet and fuse. We want to fuse isotopes
with the fewest number of protons so that they have the lowest
electrical repulsion. We want to end up with helium-4 so that we can
get lots of energy.

wwWe could do what the Sun does and fuse proton to proton, but we
don’t have the billions of years that the Sun has to wait for a few
proton nuclei to collide. We could fuse a proton and deuterium, but
that has a very low probability and very little energy is released. We
could fuse 2 deuteriums to make helium-4 and a gamma ray; we get a
lot of energy that way, but there is a small cross section because it’s an
electromagnetic interaction, not a strong interaction. We could fuse
2 deuterons to make helium-3 plus a neutron, or tritium plus a proton,
but that’s relatively low energy.

wwOr we could fuse deuterium and tritium to make helium-4 plus a


neutron, which releases 18 million electron volts of energy. That looks
like the right one to try.

wwTritium is radioactive. What if instead we fuse deuterium and


helium-3 to make helium-4 plus a proton, releasing 18 million electron
volts of energy? This is great. Our fuel’s not radioactive. We don’t have
those pesky neutrons at the end—we have protons. The problem is
we’re using helium, with more electrical repulsion, so the reaction needs
temperatures that are 6 times higher. We’re probably not going to fuse
deuterium and helium-3 this century.

wwWe need to confine our fuel—so far, typically deuterium and tritium—
at high density and pressure for long enough for them to fuse. There
are 3 ways to confine the nuclei so that they will fuse: gravitational
confinement, inertial confinement, and magnetic confinement.

Lecture 20
| Nuclear Fusion: Obstacles and Achievements 211
wwWhy is fusion so attractive? There are no transuranics, such as
plutonium, and there is no long-term (thousand-year-plus half-life)
radioactive waste. There is unlimited fuel. We can get deuterium
from seawater by vacuum distillation, and we can make tritium
from lithium-6 by hitting it with a neutron. A neutron plus
lithium-6 becomes an alpha particle plus tritium.

wwAlso, fusion has 10 times more energy density than fission. Fusion is
fail-safe. The fuel is continuously supplied. If the confinement fails,
the reaction stops. There is no critical mass of uranium-235. There is
no stored energy, like there is with fission. With fusion, we don’t have
to worry about the residual decay heat for the loss-of-coolant accidents
that happened at Fukushima. The plasma is hot, but it has so little
mass that it has little stored energy.

wwThere are no proliferation concerns because there is no enriched


uranium. There is no plutonium. But we need tritium, which is
radioactive and biologically active, and it needs to be produced. Fusion
produces neutrons, and there is neutron irradiation, which damages
reactor materials and creates some radioactive material.

212 Nuclear Physics Explained


wwWe need very high temperatures and high densities for long enough
to make the nuclei fuse. The high temperature makes a plasma, which
is where the electrons and the nuclei move independently, such as in
fluorescent light bulbs and in candle flames.

wwWe need the density to be high enough so that nucleus-nucleus


collisions allow for fusion. We need high enough temperatures so
that nuclei tunnel through the electrical repulsion to fuse. This is the
same effect as tunneling for alpha decay. A small change in energy
makes a big change in the probability, so we need temperatures
of 10,000 to 20,000 electron volts, or 100 to 200 million Kelvin.
We also need confinement times that are long enough so that we get
enough fusion out.

Plasma Confinement and Heating


wwWith fusion, 2 nuclei at a temperature of about 20,000 electron
volts collide and release about 20 million electron volts. A
14.1-million-electron-volt neutron leaves the plasma, and we catch it in
a “blanket” to make heat to make electricity.

wwThe alpha particle has 3.5 million electron volts, and it deposits its
energy in the plasma. This helps keep the plasma hot. We get out
500 times more energy than we put in, but not all of the nuclei fuse.
The break-even point is at a temperature of 10,000 to 20,000 electron
volts, or 100 to 200 million Kelvin. The density times the time is
greater than 1020 nuclei per cubic meter times seconds, which would be
a confinement of 1 second at 3 atmospheres of pressure. At that point,
it would be about a million times less dense than air.

Lecture 20
| Nuclear Fusion: Obstacles and Achievements 213
wwThe Q value is the energy released divided by the energy needed to heat
the plasma. The break-even point is a Q value of 1, but 80% of the
energy is carried away by the neutron. It is used to make electricity, not
to heat the plasma. We need a Q value that is greater than 5 to keep the
plasma hot.

wwHow do you confine a hundred-million-degree plasma? We use a


magnetic bottle. Charged particles moving in a magnetic field feel a
force perpendicular to the field. There is no force that is parallel to
the field, which means that the particles spiral along the magnetic
field lines.

wwHow do you confine the particles in the third dimension (along the
field lines)? We make the magnetic field lines circular so there is
no end. This is called a tokamak, which is a Russian acronym for a
toroidal chamber with magnetic fields.

wwWe take a solenoid, which


is a wire wrapped around
a cylinder, and bend the
ends together to form a
circle. The coils make
a toroidal magnetic
field, or a doughnut.
The particles moving
radially outward
from the inside of the
doughnut spiral around
the magnetic field lines
and stay in the torus. But
the field is weaker farther from
the center, so the configuration is
not quite stable.

214 Nuclear Physics Explained


wwNext, we give the magnetic
field lines a twist. We pass
a current around the
torus to add a poloidal
field, and now the field
lines—instead of going
around in circles—trace
out helixes around the
axis of the torus. That
will confine the particles.

wwTo make a toroidal current


to get the poloidal field, we make a
transformer—which has a primary and a secondary—with N windings
around the primary and an iron core to induce a current in the plasma
that makes the current N times bigger going through the secondary.

Lecture 20
| Nuclear Fusion: Obstacles and Achievements 215
wwMost transformers use alternating current, so the primary current
varies continuously in a sine wave. That creates a secondary current
that also is a sine wave. But the transformer for a tokamak uses direct
current, where the primary current increases steadily up to a maximum,
thus creating a constant secondary current.

wwBut it can’t run continuously, because there is a maximum current. So,


we turn everything on, increase the primary current to the maximum,
and then turn everything off to reset and repeat. The timescale of this
is between minutes and tens of minutes. This could be a problem if we
want to use a tokamak for baseload power generation.

wwOnce we confine the plasma, how do we heat it up? The current


provides some power, but it’s not enough. We can use radio waves
to heat the plasma at the natural frequency of the electrons or of the
nuclei in the plasma. This is like a microwave oven heating water at
its resonant frequency. We can also aim neutral particle beams at the
plasma, which adds energy and deuterium or tritium to the plasma.

wwA fusion experiment needs a vacuum vessel for the plasma,


superconducting magnetic coils to make the magnetic field to contain
the plasma, a transformer to create the toroidal current, and external
radio-frequency and beam heaters for the plasma. We need to protect
the superconducting magnetic coils, which are really cold, from
the hundred-million-degree plasma as well as protect the delicate
electronics from the neutron radiation.

wwIf we want to make a fusion power plant—not just do a fusion


experiment—we want to generate electricity, so we need a blanket to
absorb the neutrons and the energy of the neutrons. The blanket has
to have high-temperature tolerance. Graphite is good, but it absorbs
tritium. We also have to worry about the blanket being eroded by the
plasma; if that happens, some of the material from the blanket can get
into the plasma and contaminate it.

216 Nuclear Physics Explained


wwNext, we need a coolant to extract the heat from the blanket. We
typically use water or helium gas. The coolant then boils the water to
turn a turbine to make electricity. We need lithium in the absorbing
blanket to absorb neutrons and generate more tritium. Neutron plus
lithium-6 gives us helium-4 plus tritium. We also need a diverter to
remove the helium-4 “ash” from the plasma. (Ash is the product of our
burning or fusion, and it gets in the way of more reactions.)

The layers of a fusion power plant


include the plasma with a diverter
to remove the helium-4, a thermal
blanket with lithium to make tritium
to be recycled as fuel, a neutron
shield, the vacuum vessel (contains
the plasma), the magnet coils and
magnet cryostat, and the biological
radiation shield.

Lecture 20
| Nuclear Fusion: Obstacles and Achievements 217
Cold Fusion
Stanley Pons and Martin Fleischmann announced in 1989 at a
press conference that they had discovered cold fusion. They didn’t
give any real details, but there was a scientific frenzy around the
world to reproduce the experiment. There were lots of anomalous
results or partial confirmations.

Pons and Fleischmann ran a current through 2 electrodes in an


electrolysis cell, which separates the water into hydrogen and
oxygen. They used palladium electrodes; palladium can absorb lots
of hydrogen. They used heavy water (deuterium) and claimed to
see fusion.

But what they actually saw was anomalous heat. They claimed
to see heat above and beyond what they put in. They saw a few
neutrons. How could this possibly work at room temperature?
They claimed it was due to the much higher hydrogen density in the
palladium electrodes. But the deuterium nuclei still need to tunnel
through the electrical repulsion barrier to fuse. They’re at a much
lower temperature and therefore have much less kinetic energy
and a much smaller fusion probability. The alpha-decay half-lives
change dramatically with kinetic energy.

In addition, deuterium-deuterium fusion should also release


radiation. Deuterium plus deuterium fuses to make helium-3 plus
a neutron. But they didn’t detect enough neutrons by a factor of
about 1 million. With the energy release they claimed, there should
have been a lot of radiation. There were 10 watts of fusion power,
so they should have gotten a few rem every second.

When they tried the experiment again with regular water


(protons), which can’t fuse to make helium, they got the same
results as they did with heavy water. That was a problem. If there
was fusion, they should have seen helium in the electrodes, but
they didn’t find any. There are still some enthusiasts looking for
cold fusion, but most physicists and chemists are not.

218 Nuclear Physics Explained


wwFor inertial confinement fusion,
we hit a pellet of frozen fuel from How close are we getting
all directions at the same time to to controlled fusion?
compress it. We hit it with lasers or
particle beams. Today, we mostly The maximum fusion
use lasers. power was produced by
the Joint European Torus
wwAt the National Ignition Facility (JET) tokamak in 1997
(started in 2009) in the United with a Q of 0.64 for
States and at Laser Mégajoule in about 10 seconds. There
France, they use lasers to implode has been tremendous
frozen deuterium-tritium pellets. progress since the
They heat the pellet surface, 1960s, but there is still a
creating a shockwave that implodes long way to go.
the fuel.

wwThere is also indirect drive, which involves hitting a cylinder called


a hohlraum with lasers, creating x-rays that compress the fuel pellet.
The expected density is about 1000 times the density of water, and the
pressure to do this is 300 billion times atmospheric pressure.

Other Techniques to Confine Plasma


wwMagnetized target fusion, which is a mix of inertial and magnetic
confinement, involves making a spinning liquid metal sphere with a
hole through the middle and shooting plasma rings in from the top and
bottom to make a target in the center. Then, a few hundred pistons
ram the surface hard enough to send a shockwave through the liquid
metal to compress the plasma target. This is being done by General
Fusion in Vancouver.

Lecture 20
| Nuclear Fusion: Obstacles and Achievements 219
wwA colliding beam fusion reactor involves making 2 plasma clouds,
accelerating them to fusion energies, colliding them, and then
heating them further with plasma beams. This is being done by Tri
Alpha Energy, which in 2015 achieved a stable plasma that lasted for
5 milliseconds (a really long time for a plasma).

wwInertial electrostatic confinement, or polywell fusion, is being


pursued by many groups. External electric fields by themselves can’t
contain charged particles, so they use magnetic fields, too. They trap
electrons and use the trapped electrons to attract ions. They have
achieved 10,000 volts, which is enough to accelerate the ions and
achieve low-rate fusion. It’s easy to make and confine plasmas, but
it’s more difficult to make some fusion, and it’s not clear whether it’s
possible to achieve enough fusion for breakeven.

Next Steps for Tokamaks


wwThere are some variants of tokamaks:

§§ Spherical tokamaks look like a cored apple rather than a ring


doughnut. The advantage is better magnetic pressure.

§§ A stellarator is a twisted tokamak. The advantage is there is no need


for a pulsed transformer to make a toroidal current. It can operate
continuously. But the disadvantage is that it has a more complicated
magnetic field and coils. This is being pursued in Germany with a
$1 billion prototype.

wwThe next step for tokamaks, though, is bigger ones. The International
Thermonuclear Experimental Reactor (ITER) Tokamak, which will
be huge and expensive (about $14 billion), is being built in France by a
collaboration of the European Union, the United States, Japan, Russia,
China, India, and Korea.

220 Nuclear Physics Explained


wwITER was started in 2013, and the first experiments are expected in
2025. Full deuterium-tritium fusion measurements are expected in
2035. The goal is to generate 500 megawatts for 1000 seconds with a Q
factor of 10 (which is 10 times more energy out than put in).

wwTechnical challenges of ITER include the fact that ash diverters are
needed to remove the alpha particles, and they have to work at extreme
temperatures with lots of neutron irradiation. They need to get the
blanket right to absorb the neutron energy. ITER needs to have high
reliability if it’s going to be a power plant, and it needs structural
integrity and material strength to survive the high temperatures,
neutron bombardment, and stresses from cycling magnetic fields.

wwThe plan is 10 years to build it, 10 more years of experiments building


up to deuterium-tritium fusion, 15 more years to build a demonstration
power plant, and another 15 years to commercialize it—or one of the
novel startups might succeed. If all goes well, controlled nuclear fusion
might actually be here in 50 years.

Supplements

READINGS

Clery, “Fusion’s Restless Pioneers.”


Hamilton, “A New Approach to Fusion.”
Lilley, Nuclear Physics, chap. 11.
Nutall, “Fusion as an Energy Source.”
——— , Nuclear Renaissance, chaps. 9–11.
Park, Voodoo Science.

Lecture 20
| Nuclear Fusion: Obstacles and Achievements 221
QUESTIONS

1 The difference in the binding energies of hydrogen and helium is


7 MeV per nucleon, and the difference in the binding energies of
uranium-235 and its decay products is only 1 MeV per nucleon. Why
does fissioning 1 uranium-235 nucleus yield almost 10 times more
energy than fusing hydrogen into 1 helium nucleus?

2 Neutron irradiation from fusion power plants is a concern. How


does this compare to neutron irradiation from existing fission
power plants?

3 Why are future fusion power plants (once they are made to work)
likely to be less reliable than fission power plants?

ANSWERS

1 Helium only has 4 nucleons, so fusing hydrogen into 1 helium nucleus only
releases 4 × 7 = 28 MeV total. Uranium has 235 nucleons, so fissioning
uranium-235 releases 235 × 1 ≈ 200 MeV. However, in terms of energy
released per kilogram of fuel (rather than per nucleus fissioned or fused),
fusion releases 7 times more energy.

2 Uranium fission releases 2 to 3 neutrons for each fission, or about 1 neutron


for each 100 MeV of energy released. Deuterium-tritium fusion releases
1 neutron for each fusion, or 1 neutron for each 28 MeV of energy released.
Thus, for the same power production, fusion will release about 4 times more
neutrons than fission. In addition, fusion plants will be more complicated,
with superconducting magnets and other apparatuses that might be more
susceptible to neutron damage.

222 Nuclear Physics Explained


3 It took the fission industry decades to figure out how to run their power
plants reliably. Fission power plants just boil water like coal-fired power
plants. Fusion power plants will be much more complicated than fission
power plants. Fusion power plants will have hundred-million-degree plasmas
just a few meters from ultracold superconducting magnets. They will have
complicated transformers, beam heaters, ash removers, etc. In addition, 80%
of the energy is carried away by the neutrons, so they will need to capture
the neutrons and convert that energy to heat. All of this adds tremendous
complexity that will make fusion power plants less reliable.

Lecture 20
| Nuclear Fusion: Obstacles and Achievements 223
21 KILLING CANCER
WITH ISOTOPES,
X-RAYS, PROTONS

T
 e field of radiation oncology uses radiation
h
to treat cancerous tumors. It employs
2 kinds of radiation: internal and external
sources. Internal sources use radioactive
materials that are encapsulated and placed
in and around the tumor. They are best for
well-defined tumors in a single location but
can also be attached to a molecule that
seeks out the tumor and is injected into the
body. This is the best method if the delivery
is selective enough. External sources use
beams of photons (x-rays or gamma rays),
protons, neutrons, or ions and are aimed at
the tumor from the outside. u

224
In 1916, a gynecologist named
Dr. Howard Kelly, who was
one of the founders of The
Johns Hopkins Hospital,
used radiation to treat
350 patients with advanced
deep-seated cancers. He
“milked” radon-222, which
is a gas that has a half-life
of 4 days, from radium.
There is alpha, beta,
and gamma radiation
from the decay chain. DR. HOWARD KELLY
1858–1943
He placed the radium
in tiny capsules called
seeds, and lots of these seeds were positioned in and
around the tumors. He cured 20% of his patients. This
was a tremendous success for patients who were
otherwise going to die. Today, our success rate for these
cancers is about 70%.

Internal Radiation
wwWhat are the ideal characteristics of an internal radioactive source? We
want the energy to be deposited in the tumor, not the healthy tissue. If
the isotope travels to the tumor (i.e., iodine-131 travels to the thyroid
gland), then we prefer that the radiation has the distance of microns, so
we use very-short-range radiation, such as alpha particles or low-energy
beta particles.

Lecture 21
| Killing Cancer with Isotopes, X-Rays, Protons 225
wwIf it’s an implanted source (i.e., iridium-192 to treat prostate cancer),
then the seeds or catheters where the iridium-192 is placed will be
spaced millimeters apart, so we want the radiation to travel millimeters
to cover all of the tumor in between the seeds or catheters. Therefore,
we want to use short-range, low-energy x-rays or gamma rays. If the
isotope is left in, as it is with radioactive seeds, then ideally the decay
time of the isotope should be matched to the treatment length.

wwHow can we get the radioactive material—the radiopharmaceutical—


to travel to the tumor? Iodine (element 53) travels to the thyroid as
sodium iodide. It’s used to treat thyroid cancer and hyperthyroidism.
Iodine-131 has a half-life of 8 days. The decay chain of iodine-131 is
mostly beta rays with 0.6 MeV maximum energy and 0.4 MeV gamma
rays. The range of the beta rays is about 2 millimeters.

wwThe beta rays irradiate the tumor. The gamma rays (photons) are have
a longer range and irradiate the whole body, giving the whole body a
dose of about 10% of what the thyroid gets. The radiation is not ideal,
but iodine targets the thyroid so well that we use it.

wwWe make the iodine in a nuclear reactor. We place tellurium,


which is 34% tellurium-130, in a nuclear reactor. A neutron hits
tellurium-130 and makes tellurium-131, which decays to iodine-131.
Most of the iodine-131 is gone from the body after 3 to 5 days. It is
excreted via the urine or decays.

wwIf we want to treat bone cancer metastasis, then we take advantage of


the fact that radium is a bone seeker. It’s in the same column of the
periodic table as calcium and strontium. Strontium-90 emits beta rays,
which have a longer range and are not as useful. Radium is heavier than
strontium and thus alpha-decays. The alphas have a range of between
2 and 10 cells, and they deposit much more energy in one place, so they
are much more lethal to cells. The half-life is 11 days.

226 Nuclear Physics Explained


wwTo make radium-223, we put radium-226 in a reactor, and a
neutron plus radium-226 makes radium-227, which beta-decays
to actinium-227 (with a half-life of 22 years), which decays to
thorium-227 (with a 19-day half-life), which then decays to
radium-223. We purify the actinium. The actinium decays and makes
more radium, and we extract the radium as needed. This is called an
actinium/radium generator.

wwIf we can’t get the radioactive material to transport itself to the tumor,
then we implant the material. This is called brachytherapy. We can
implant it in the body either temporarily via implanted tubes (catheters)
or permanently in seeds. We insert lots of seeds or catheters to get good
dose uniformity throughout the cancer. This allows us to concentrate
the dose in some tumors better than with external beams. The total
dose delivered to the tumor is typically between 30 and 150 gray, or
between 3000 and 15,000 rad.

wwThe dose rate depends on the treatment plan designed by the


oncologist for each patient. There are 2 dose rates: high-dose rate and
low-dose rate.

§§ High-dose rates (more than 12 grays an hour, or 1000 rad an hour)


are given as discrete treatments. We infuse the radioactive material
into the catheters, preferably remotely, and remove the radioactive
material after each relatively short treatment. Isotopes used are
iridium-192 or cobalt-60.

§§ Low-dose rates (less than 2 grays an hour, or 200 rad an hour) are
typically given with implanted seeds that are left in place. The seeds
are filled with iodine-125 or palladium-103.

wwWe want to implant isotopes with a medium half-life (days to weeks).


The shorter the half-life, there is more radioactivity, but we want the
half-life long enough so that we can handle it. We want the half-life
matched to the treatment length for permanent implants (seeds).

Lecture 21
| Killing Cancer with Isotopes, X-Rays, Protons 227
The first x-ray beam was applied
to Hodgkin’s disease by
Dr. Henry Kaplan at Stanford
in the late 1950s. Hodgkin’s
disease is a disease of
lymph nodes, typically in the
chest. The tumors are
deep but have well-
defined locations. The
cancer progresses
one lymph node at
a time, so it doesn’t
metastasize to
faraway locations in
the body. He cured
50% of early-stage
patients. Today, the
cure rate is more DR. HENRY KAPLAN
than 90%. 1918–1984

wwWe want to use x-rays or low-energy gamma rays so that the range of
the particle is about the same as the spacing of the seeds. Alpha and
beta particles have too short of a range to get uniform coverage.

wwWe don’t want the decay products to be gaseous because gas molecules
can escape and migrate, and they increase pressure because they take
up more space. We want this to be nontoxic, and we need to be able to
sterilize everything.

228 Nuclear Physics Explained


External Radiation
wwIf it’s too difficult to implant seeds or put in catheters—perhaps
because the tumor is the wrong shape or size or too diffuse—then
we use beams of external radiation. We can use x-rays or gamma rays
(photons), protons, or other particles.

wwThe particles have to have enough energy so that they can reach
the tumor, and we need to minimize the dose to the surrounding
tissue. We spread the dose over time to let the healthy tissue recover
in between.

wwRapidly dividing cells are more sensitive to radiation, and that doesn’t
depend on the dose rate, but slowly dividing cells (such as skin cells and
normal body tissue) are less sensitive to radiation, and the sensitivity is
reduced at lower-dose rate as the cells recover.

wwWhen an x-ray or gamma-ray photon passes through the body, it


interacts discretely and transfers all of its energy to an electron, which
keeps moving and ionizes atoms along its path until it runs out of
energy. This means that the photon beam is most intense as it enters
the skin (and deposits more energy there) and then loses intensity as the
photons interact. The beam therefore deposits less energy the deeper
into the body it goes.

wwTo make x-rays or gamma rays, we use an electron linear accelerator


with between 6 and 20 million electron volts. The electrons hit a
tungsten target and make x-rays. All of the x-ray or gamma-ray energies
from zero to the electron energy are made. This is the same way that we
make lower-energy x-rays for imaging purposes.

wwWe choose the energy to match the depth of the tumor. The deeper
the tumor, the higher the energy and increased range of the x-rays or
gamma rays. We want to collimate the beam horizontally so that the

Lecture 21
| Killing Cancer with Isotopes, X-Rays, Protons 229
shape of the beam matches the shape of the tumor. One beam will
deposit most of its energy at the skin; 2 beams from different angles
that overlap at the tumor will give more dose to the tumor.

wwModern treatments aim the beam at the tumor from multiple


orientations to maximize the dose to the tumor and minimize the
dose to the surrounding tissue. They know where to aim because very
careful imaging is done before and, especially, during treatment. Image
guidance is used. Careful computations and simulations are done by a
medical physicist to maximize the dose to the tumor and minimize the
dose to the healthy tissue.

wwIn intensity-modulated radiotherapy, each of the individual beams is


shaped to best match the tumor size and shape. The high-dose regions
(where we want to deliver a high dose to the tumor) are much tighter,
or much better matched to the tumor, but the low-dose regions (where
we irradiate healthy tissue) are bigger.

wwA newer method, first approved for treatment of certain tumors


in 1988, is proton therapy. Protons and photons interact very
differently in matter.

wwTo attack cancer, we want to deliver a lot of energy to the cancer


cells and as little energy as possible to the healthy cells around them.
When a photon beam travels through the body, each photon only
interacts in one location, and the photon is either absorbed or not. The
number of photons (and dose rate) decreases exponentially as it travels
through body tissue. The highest dose is at the skin, but there is some
dose at all depths. A higher-energy beam penetrates further than a
low-energy one.

wwBecause protons are charged particles, as a proton beam travels through


the body, it interacts with the electrons, exciting them and depositing
energy as it goes. The energy it deposits increases the more slowly the
proton travels.

230 Nuclear Physics Explained


wwUnlike the photon beam, where
Neutrons have the same
the energy deposited is at a
mass as protons and can
maximum as it enters the body
also do a lot of damage
and then decreases slowly as it
to tissue, but neutrons
travels through the body, the
are much more difficult
energy deposited by a proton beam
is at a minimum as it enters the
to accelerate and control
body, increases dramatically as it
because they have no
gets to where the tumor is, and
charge. Neutrons are
then the protons stop because they much more difficult to aim
have run out of energy. because they penetrate
materials. Neutrons don’t
wwWith a proton beam, we can be slow and stop precisely at
much more precise in how we a given depth, like protons,
deposit energy in the body; the because they have no
proton beam deposits most of its charge. There have
energy in the tumor and no energy been attempts to treat
past it. Another advantage to a cancer with neutrons, but
proton beam, in comparison to a the neutrons were not
photon beam, is that we can have a effective enough to be
pencil-thin beam that only hits the worth the extra effort.
area of the tumor.

wwIf proton beams can deposit energy in the tumor so much more
precisely than photon (x-ray or gamma-ray) beams, then why doesn’t
everyone use them? Proton beams are much more difficult to make.
We need protons of about 200 million electron volts to penetrate the
body. We need to use a cyclotron to accelerate the protons. We ionize
the hydrogen atoms and inject them into the cyclotron. The protons
in a cyclotron spiral outward in a magnetic field, getting a small push
twice each orbit. When the protons reach 230 million electron volts
(about half the speed of light), they leave the cyclotron and are steered
by magnets to the different treatment rooms.

Lecture 21
| Killing Cancer with Isotopes, X-Rays, Protons 231
wwHow do we control the depth of the proton beam? We can make a
narrow beam of protons, scan it across the tumor, and change the
energy of the beam as it scans to match the tumor depth at each point.
The lower-energy protons don’t penetrate as far; they deposit their
energy closer to the surface. Alternatively, we can diffuse the beam and
then make a plastic “mask” to match the depth profile of the tumor.

wwProton therapy should be better than photon therapy, because there


is better control of the dose. More importantly, we now have precise
imaging to take advantage of that better control. But protons are much
more expensive than photons, so we need to measure the improvement
to see if it’s worth the cost.

wwThere is a major National Institutes of Health initiative to perform


randomized, controlled trials of proton versus x-ray therapy.

wwWhat’s next? With targeted radioimmunotherapy, we attach a


short-lifetime alpha emitter to a biologically active molecule that targets
the cancer (i.e., antibodies). With boron-neutron capture therapy, we
attach boron to a cancer-seeking molecule and irradiate the patient
with a thermal-neutron beam. The boron absorbs neutrons much
more than anything else in the body. Boron-10 plus a neutron gives us
lithium-7 plus an alpha particle, which kills the cancer.

wwRadioisotopes can kill cancer cells, but radiation is much more


common in medical diagnosis.

232 Nuclear Physics Explained


Supplements

READINGS

Jorgenson, Strange Glow, chap. 6.


Lilley, Nuclear Physics, chap. 9.
National Research Council, Nuclear Physics, p. 150–158.

QUESTIONS

1 Why are radiation doses for cancer treatment about 10 times greater
than the lethal radiation dose for people?

2 Why do patients often feel sick after receiving radiation treatment


for cancer?

3 Why do we make sure that implanted radioactive seeds do not use


any materials with gaseous daughter nuclei?

4 If a 1-inch-diameter tumor in a 10-inch-thick patient is irradiated


with x-rays, then what fraction of the radiation in the beam is
delivered to the tumor?

ANSWERS

1 The lethal radiation dose for people is a whole-body dose. The specified
radiation dose for cancer treatment is the dose to tumor. They try to
minimize the radiation dose to healthy tissue. An 800-rad whole-body dose
will not kill all the cells in the body, but it will kill enough cells so that the
body can no longer function. However, when applying radiation to a tumor,
doctors want to kill almost all the cancer cells in the tumor and therefore
apply doses of around 8000 rads.

Lecture 21
| Killing Cancer with Isotopes, X-Rays, Protons 233
2 Even though oncologists and medical physicists do their best to maximize
the radiation dose to the tumor and minimize the radiation dose to healthy
tissue, and even though fast-dividing cancer cells are more susceptible to the
effects of radiation than most normal tissue, some healthy tissue still receives
large amounts of radiation. This can produce some of the effects of acute
radiation syndrome discussed in lecture 5.

3 This is because as gas accumulated in the sealed seeds, it would dramatically


increase the pressure in the seeds and possibly make them burst, damaging
nearby tissue and releasing their contents into the body.

4 We can crudely approximate x-rays as delivering about the same amount of


energy at all locations in their path (in reality, they deposit more energy at
the skin and then progressively less as they go deeper). If the tumor is about
1 inch in diameter and the patient is about 10 inches thick, then about
1/10 of the x-ray energy will be delivered to the tumor and the other 9/10 (or
90%) will be delivered to healthy tissue. By using many different beams
from many different angles, the 90% of energy delivered to healthy tissue
will be spread out over as much healthy tissue as possible to reduce the dose
delivered to any one spot. However, 90% of the energy of the beam is still
deposited in healthy tissue.

234 Nuclear Physics Explained


MEDICAL IMAGING:
CT, PET, SPECT,
22
AND MRI

T
 e increasingly sophisticated use of
h
radiation has transformed diagnosis and
treatment in many fields of medicine with
mammograms, PET scans, CT scans,
bone-density tests, MRIs, etc. How do they
work? What do they reveal? When, if ever,
might the doses of radiation be more than
might be regarded as acceptable? u

235
Types of Medical Imaging
wwWe measure biological effect in rems or sieverts. The limit for radiation
workers is 5 rem per year, or 50 millisieverts. Twenty rem will give
someone an extra 1% chance of developing cancer in their lifetime,
according to the linear no-threshold hypothesis. Acute radiation
syndrome sets in at about 100 rem, or 1 sievert, with symptoms of
fatigue and vomiting.

wwRadiation from a single x-ray is between 0.1 and 10 millirem, or


between 0.001 and 0.1 millisievert. The medical technician leaves the
room when you get an x-ray, but they could stick around for hundreds
or thousands of those before hitting their annual dose limit.

wwIn reality, the average dose for medical workers is about 300 millirem
per year, or 3 millisieverts. They typically get more to their fingers
from injecting radioactive materials. This is much less than the
5000-millirem-a-year limit set by the US Department of Energy for
radiation workers.

wwNewer types of images are like multiple x-rays: A mammogram is


approximately equivalent to 5 chest x-rays, or 1 long airplane flight; a
CT scan of the head is like 20 chest x-rays; and a CT scan of the chest
is like 100 chest x-rays, or about 20 long airplane flights.

Looking inside the body is like choosing an


e-reader: You can either have external
illumination, which uses reflected
light, or radiation, to see the
difference in the reflectivity
of ink and paper; or internal
illumination, where the
screen glows with
different colors.

236 Nuclear Physics Explained


wwIn the case of using radiation, there A Harvard report puts
is external illumination—x-rays and a cardiac stress test
CT scans—which shine a beam of with nuclear imaging at
radiation on the body and measure 40 millisieverts, or 4
the transmitted radiation to see the rem, which is 80% of the
difference in absorption between recommended limit for
bone and tissue. This technique sees medical workers.
structure in the body.

wwThere is also internal illumination—PET and SPECT scans—where


radiation-emitting substances are injected or inhaled into the body,
which transports the molecules. We measure where the radiation
is emitted from, which tells us how the molecules are transmitted,
and that tells us about function. Here, the radiation dose might
be somewhat higher because the nuclide stays in the body after
the procedure.

wwThen there is external-ish illumination— magnetic resonance imaging


(MRI). This used to be called nuclear magnetic resonance, but no
ionizing radiation is involved, so the name was changed to not scare
people. What MRI sees is mostly structural.

X-Rays
wwX-rays use x-rays—specifically, they use 50,000- to 150,000-electron-
volt x-rays. They take a 2-dimensional picture, measure the differential
absorption, and really only see bones and not-bones. That’s because
calcium has a larger number of protons than most of the other
elements in the body, so it absorbs a lot more x-rays. We can also see
injected contrast agents, such as barium, in the body. Fortunately,
modern cameras are much more sensitive than they used to be, which
dramatically reduces the radiation dose needed.

Lecture 22
| Medical Imaging: CT, PET, SPECT, and MRI 237
wwTo see more, we can take lots of pictures and let the computer figure
it out. We rotate the x-ray emitter and the detectors around the body.
For each point in the body, we combine the intensities of all the x-rays
that pass through that point. The sum depends mostly on the x-ray
absorption of the tissue at that point.

wwThen, we use a computer to deconvolute the images to get the


absorption of the tissue at each point. We then add it all back together
to recreate a 2-dimensional slice of the body. We move the scanner
one step along the body and measure another slice. This gives us very
high spatial resolution, but we’re still just using x-rays, so we only see
structure, not function.

PET and SPECT Scans


wwTo image a specific body function, we attach a radioactive isotope to a
bioactive molecule. We then watch the uptake of the molecule to trace
its function. We track the molecule by the radiation emitted by that
radioactive isotope.

wwAlpha and beta particles don’t escape the body. Gamma rays can be
measured one at a time. A beta-plus particle (positron) travels a short
distance in the body, finds an electron, and annihilates to produce
2 back-to-back gamma rays each with exactly 511,000 electron volts.
Then, we measure the emitted gamma rays and make an image.

wwFor positrons, we want low-energy positrons so that they annihilate


quickly. For gamma rays, we want the energy high enough that it
can penetrate tissue and leave the body. (This is the opposite need of
radiation oncology.) We want the energy low enough that we can see it
clearly in a gamma camera, ideally 100,000 to 200,000 electron volts.
We want the half-life short enough for high activity that dies away
quickly but long enough to use.

238 Nuclear Physics Explained


wwTo detect gamma rays, we use a gamma camera, which uses scintillator
crystals and photomultiplier tubes—the “film” for the camera. But
there are no lenses to focus gamma rays onto the film to make an
image. Therefore, if it’s a PET scan, we have 2 back-to-back gamma
rays, and we get the direction by detecting both. If it’s a SPECT scan,
which uses 1 gamma ray, then we need a collimator in front of each
crystal to define the direction.

wwWe use a lot of crystals and a lot of photomultiplier tubes to cover


a large area. The larger the area of the camera, the less radiation is
needed. The smaller the crystals that we use, the better the resolution.
But the larger the area and the smaller the crystal, the more crystals
and photomultiplier tubes we need, which means that the detector
costs more.

Lecture 22
| Medical Imaging: CT, PET, SPECT, and MRI 239
wwIn PET scanning, the positron is emitted and travels a few
millimeters, finds an electron, annihilates with it, and produces
2 back-to-back 511,000-electron-volt gamma rays. We have a large ring
of detectors to detect both gamma rays at exactly the same time.

wwTo reconstruct the PET image, we apply the exact same computing
techniques as CT scanning, only using internal illumination instead of
external illumination. The typical resolution that we get is about 5 to
10 millimeters, or 1/4 of an inch.

wwIf there’s only 1 photon, then we don’t need as large a detector to detect
1 at a time. We use a collimator so that each pixel in the detector
only sees the “light” (the gamma rays) coming from 1 direction. A
collimator is a thick lead plate with 1 hole in it for each pixel. This
gives us a 10% resolution, so if we’re looking at something 1 foot away,
it gives us a resolution of 1 inch. If it’s closer, the resolution is better.

wwThe camera gives us a 2-dimensional image. To get a 3-dimensional


image, we take pictures from many different angles and combine
them. Or we can leave the camera where it is and tilt the collimator,
take new images, and then combine them. This is under development
at Jefferson Lab. We apply the exact same computing techniques
of CT scanning, just like we do for PET scans. SPECT stands for
single-photon emission computed tomography.

wwThe big advantage of a 3-dimensional image is that we can make


slices so that we don’t have to see the things in front of or behind
the important organ. We can view the organ in different slices at
different orientations.

wwTo get the isotope to go someplace interesting (a specific organ),


we choose the appropriate molecule and attached the isotope to it.
For example, radioactive carbon and fluorine have a lot of different
biomolecules. Glucose (sugar) is transported to metabolically active

240 Nuclear Physics Explained


regions, such as the brain or tumors. Technetium-99 can be attached to
a lot of different molecules with different uses; about 85% of SPECT
procedures are done with technetium-99.

wwWe look at the distribution of radiation that gives us information on


the metabolic uptake of the molecule containing the isotope as well
as structures in the body. For example, the intensity and imaging of
iodine radiation shows the metabolic activity of the thyroid as well as
its size and shape.

wwHow do we make radioactive isotopes for imaging? PET scans use


positron emitters, which are isotopes such as carbon-11, nitrogen-13,
oxygen-15, or fluorine-18. They decay by positron emission, which
means that they started with too many protons, so we hit a stable
nucleus with protons from a cyclotron.

wwThese isotopes typically have 2- to 20-minute half-lives. This means


that the cyclotron has to be in the hospital so that they don’t decay
before you can get them to the hospital. This is expensive. Also,
because the half-life is so short, we need to do the chemistry very
quickly to get the isotopes into those useful molecules. The problem is
that this is hot chemistry. These isotopes are radioactive, so we have to
do this by remote control.

wwWe also make gamma-ray emitters for SPECT, typically in generators


with a longer-lifetime isotope that decays to the shorter-lifetime
desirable isotope. We want the isotope that we inject to have a relatively
short lifetime to reduce the radiation to the patient after the procedure
is over. The generator makes it possible to use short-lifetime isotopes.

wwOne of the more popular positron emitters is fluorine-18, which is


incorporated into a glucose analog called FDG, which is taken up by
cells needing energy, such as the brain, tumors, and places where there
is inflammation, healing, or muscular activity.

Lecture 22
| Medical Imaging: CT, PET, SPECT, and MRI 241
wwThe fluorine-18 accumulates in the cells and is seen by the 2 gamma
rays it gives off when the positron it emits annihilates with an electron.
Using this fluorinated glucose analog lets us image where the brain
needs energy (where it’s working). Different areas of the brain light up
doing different tasks.

wwPET resolution is limited by detector size and positron propagation


(how far the positron goes before it annihilates into the photons that
we detect). Most of the time, we combine high-resolution structural
CT scans with lower-resolution (5- to 10-millimeter) functional PET
images to get enough information.

wwHow much radiation do patients receive from these scans? In PET and
SPECT scans, adults receive about 1 rem, which is equivalent to about
2 to 3 years of background radiation. It corresponds to an increased risk
of cancer of about 0.05%, assuming the linear no-threshold hypothesis.

wwPatients become slightly radioactive. With a PET scan, patients emit


about 1 millirem per hour at 1 meter away from them immediately
after scanning. This decreases very quickly because the half-life is
between 2 and 20 minutes.

MRIs
wwWe can image the body without using ionizing radiation with MRI,
which images the spins of the hydrogen nuclei (protons) in the body.
We use an electromagnet to apply a really large (several tesla) magnetic
field to make about 1 in a million of the protons in the body spin-align.

wwThe resonant frequency of the protons is proportional to the magnetic


field, so we apply a radio-frequency wave at the resonant frequency to
rotate the proton spins. The spins are no longer parallel to the magnetic

242 Nuclear Physics Explained


field. The protons spin-precess at the resonant frequency. The protons
emit radio waves at this resonant frequency as they precess, and we
detect these radio waves.

wwTo make the images, we apply a large magnetic field that points along
the patient. We add another parallel field that increases with height.
For example, the magnetic field is smaller at the patient’s feet and larger
at the head. The resonant frequency will now depend on the location
along the patient’s body. We apply a perpendicular radio-frequency
magnetic field at the resonant frequency of the desired slice of the body.
This rotates the proton spin around in the selected slice of the patient.
Then, we turn off the 2 extra magnetic fields.

wwNext, we turn on another magnet to make a magnetic field that varies


along another axis in the slice. The frequency of the radio waves
emitted by the precessing protons depends on the distance along that
axis. The magnetic field in selects a slice of the slice. We repeat this for
different angles and frequencies and deconvolute the data, like we do
with a CT scan.

wwMRIs image hydrogen density, which is different in different tissue. For


example, the hydrogen density in water is less than it is in fat. We can
also see different things, such as how quickly the MRI signal decreases.

wwFunctional MRI (fMRI) compares oxygenated (diamagnetic, like


normal tissue) and deoxygenated (paramagnetic, which is different)
blood. Hemoglobin, despite containing iron, is not ferromagnetic. The
more brain or tissue activity, the more oxygen needed for metabolism,
the more oxygenated blood, the less local magnetic variation, and the
longer the relaxation time.

wwFortunately, there is a growing panoply of nuclear physics–based tools


to look at structure and processes inside our bodies. So, what will
come next? There will be better MRI as well as smaller, faster, better
gamma cameras.

Lecture 22
| Medical Imaging: CT, PET, SPECT, and MRI 243
Supplements

READINGS

Christian and Waterstram-Rich, eds., Nuclear Medicine and


PET/CT.
Gould and Edmonds, “How MRI Works.”
Lilley, Nuclear Physics, chap. 9.
National Research Council, Nuclear Physics, p. 150–158.
Tilakaratna, “How Magnetic Resonance Imaging Works,
Explained Simply.”

QUESTIONS

1 Why does internal radiation (e.g., SPECT or PET scans) show


function rather than structure?

2 When would a doctor choose to perform a SPECT scan rather than a


PET scan (or vice versa)?

ANSWERS

1 This is because the radioactive materials are transported by the body


to different locations, and doctors identify those locations by where the
radiation is emitted from. Therefore, they are measuring the function of
the body in transporting different molecules containing the radioactive
isotopes to different locations. For example, technetium-99* is transported
to different locations depending on which biologically active molecule it is
attached to.

244 Nuclear Physics Explained


2 It depends on which body function needs to be imaged and which molecules
are available. Radioactive isotopes for PET scans have half-lives measured
in minutes or tens of minutes. PET scans are limited to molecules that can
be very rapidly synthesized that contain carbon, oxygen, or fluorine. In
practice, most PET scans are done with FDG, a glucose analog containing
radioactive fluorine. Because the half-life is so short, the radioactivity of the
patient decreases to almost zero within a few hours. SPECT scans can use a
wider range of radioactive isotopes and molecules. In practice, about 80% of
SPECT scans use technetium-99* in a remarkably wide variety of molecules.
Technetium-99* has a 6-hour half-life, so there is more time to perform
the necessary hot chemistry to incorporate the technetium-99* into the
appropriate molecule. On the other hand, it takes a few days for the patient’s
radioactivity to decrease to almost zero.

Lecture 22
| Medical Imaging: CT, PET, SPECT, and MRI 245
23 ISOTOPES AS
CLOCKS AND
FINGERPRINTS

I
s otopes with unstable nuclei are like clocks and
fingerprints, which we use to date and identify
many features of our world. We can use
carbon-14 analysis to date both archeological
findings and the vanilla in your kitchen. In the
first case, we measure the age of an unknown
object to learn more about it. In the second
case, we measure the “age” of a known
object to look for fraud. In this lecture, you will
discover the many things that can be learned
from measuring isotopes. u

246
How do we date objects using radioactive decay?
We choose an object that stopped replenishing its supply
of a certain element at a certain time, such as death
for animals and plants, chemical formation for rocks,
and separation from the atmosphere for water. These
radioactive isotopes come from either the uranium, thorium,
or potassium decay chains or from cosmic-ray interactions in
the atmosphere.
We choose an isotope with an appropriate half-life and
appropriate chemistry. If we want to date really old rocks,
then we use uranium, with a half-life of billions of years. If we
want to date groundwater, we use krypton-81, with a half-life
of 230,000 years. If we want to date organic material, we
use carbon-14, with a half-life of almost 5700 years.

Carbon Dating
wwCarbon is absorbed by all living organisms. Hydrogen, nitrogen, and
oxygen are also absorbed, but they don’t have any convenient isotopes.
The unstable isotope of hydrogen, tritium, has a half-life of 12 years,
and the unstable isotopes of nitrogen and oxygen have half-lives that
are less than 5 minutes.

wwCarbon-14 has a 5700-year half-life, which is well matched to


archeological times. We can date objects back to about 10 times the
half-life, or about 60,000 years. This technique was developed by
Willard Libby in the 1940s; he received the Nobel Prize in Chemistry
for this in 1960.

Lecture 23
| Isotopes as Clocks and Fingerprints 247
wwCarbon-14 is continuously created in the atmosphere by cosmic rays,
which are high-energy charged particles coming in from outer space,
interacting with the nuclei in our atmosphere, and, in this case,
transmuting nitrogen-14 to carbon-14. Carbon-14 is only 1 part per
trillion of the carbon in our atmosphere, so it’s not enough to give us
significant radioactivity, but it is enough to measure.

wwIf we want to date a sample of carbon, then we count the radiation


given off by the carbon-14 decay. To get 99% accuracy, we need to
count 10,000 decays. We need 4 grams of carbon, which is actually
a big sample of a precious, unique archeological artifact. That might
be an entire page from a rare manuscript or a very large swatch of
ancient cloth.

wwWe can use smaller samples, but we don’t wait for the atoms to decay.
We put a tiny sample in an accelerator mass spectrometer and
count the atoms. The accelerator mass spectrometer will separate the
carbon-12 from the carbon-13 from the carbon-14, and we can count
the relative numbers of atoms.

wwUsing this technique, we can use milligram or microgram samples—


samples that are 1000 or 1 million times smaller than an entire page
of an ancient manuscript. This technique has a sensitivity of 1 part
in 1015, or 1 part per quadrillion, and the spectrometers that we use are
relatively small and compact.

wwAnything organic (in the chemical sense of organic, which means it


contains carbon) can be dated—including the medieval manuscripts
in the library at the University of Seville, the Shroud of Turin (a length
of linen cloth bearing the alleged image of Jesus that some believed
was Jesus’s burial shroud but was dated to between 1260 and 1390, not
the time of Jesus), and the Artemidorus papyrus (which some people
thought was a 19th-century forgery but was dated to about 1900 to
2000 years before the present).

248 Nuclear Physics Explained


wwWhen we do carbon dating, we date based on the decay of carbon, but
then we have to calibrate it because the abundance of carbon-14 in the
atmosphere varies slightly over time. This is because of 2 effects. First,
the cosmic-ray rate in the atmosphere (which makes carbon-14) varies.
Second, volcanoes and modern industrial activity emit old carbon
dioxide (that has no carbon-14 left) into the atmosphere that dilute the
carbon-14 in the atmosphere.

wwTo calibrate the date, we measure samples of a known age—from other


dating methods, such as tree rings—to make the calibration curve.
We just need the calibration curve for 1 sample of a particular date
and then we know how to calibrate things at that date. We use the
calibration curve to convert the carbon-14 date to a real date.

wwEstablishing 3000-year-old dates, such as the Megiddo archeological


site in Israel, is particularly tricky due to wiggles in the calibration
curve at that time. The calibration curve gives a 200-year range of
reconstructed dates, which is a long time for a historical site.

Lecture 23
| Isotopes as Clocks and Fingerprints 249
wwHow do bomb tests affect carbon dating? The carbon-14 concentration
doubled from 1955 to 1963 due to bomb tests. This is called the bomb
peak. The concentration of carbon-14 then decreased as carbon was
exchanged between the atmosphere and the oceans. It’s still slightly
above normal.

wwWe can use this information to precisely date post-1955 objects, and
we can detect modern forgeries of pre-1955 objects. For example, a
painting by Fernand Léger was supposedly painted in 1913 or 1914 was
in the Peggy Guggenheim collection, but carbon dating showed that
the painting had 30% too much carbon-14, so it had to have been
made after 1955 and was a modern forgery.

250 Nuclear Physics Explained


By testing the carbon-14 levels
in sparkling wine, we can tell
the difference between natural
fermentation (with carbon-14)
and artificial carbonation (carbon
dioxide from fossil fuel burning,
which has no carbon-14).

wwWe can also use carbon dating to date expensive


wines; to date modern animal parts, such as
elephant tusks, to determine if they were harvested
before or after certain laws were enacted; and for
neurological research, such as to determine when
individual neurons were created and whether the brain
continues to make new neurons.

wwWhy should we carbon-date food? Real food products


come from plants and should have modern
carbon-14 abundances. Artificial products are
synthesized from chemical feedstocks derived
from fossil fuels and have no carbon-14.

wwBecause food products are not precious items and there are no limits
on sample size, we can burn the sample to reduce the carbon to carbon
dioxide. Then, we can convert the carbon dioxide to methane, include
the sample gas in a wire chamber, and count the carbon-14 decays.

wwAlternatively, we can convert the carbon dioxide to benzene, include


it in a liquid scintillation counter, and count the decays. These
detectors are just like the ones used at Jefferson Lab for nuclear physics
experiments. Modern material counts about 15 disintegrations per
minute for each gram of carbon.

Lecture 23
| Isotopes as Clocks and Fingerprints 251
Vanilla is an expensive
natural flavoring,
and it has the standard
ratio of carbon-14 that is
in the atmosphere. Vanillin, the
primary component of vanilla, can be
synthesized using chemical feedstocks,
typically derived from natural gas, that does
not contain carbon-14.
Early testing showed no carbon-14 in many vanillas that were
being sold as natural. Forgers now synthesize vanillin from
carbon-14 and add it to samples. In other words, they are
adding radiation (albeit a negligible amount) to our food to
make it look natural.

Other Types of Dating


wwCarbon-14 dating won’t work for inorganic or very old objects, so we
have to use other tools. For example, if we want to date really old rocks,
they have no carbon and are too old, so we typically use either uranium
or potassium-40. Both of these have very long half-lives; which one we
choose depends on the kind of rock that we want to date.

wwIf we want to date groundwater or deep ocean currents, then we could


use carbon, but the problem is that carbon is chemically active, so it
won’t stay put. Instead, we use radioactive isotopes of inert gases. There
are no useful isotopes of helium or neon; their half-lives are too short.
Instead, we use argon-39, which has a half-life of 269 years, for ocean
currents, and we use krypton-81, with a half-life of 230,000, years
for groundwater.

252 Nuclear Physics Explained


wwWe can do uranium-lead dating. Uranium-238 has a half-life of
4.5 billion years; there is a long decay chain that ends up with
lead-206. Uranium-235 has a half-life that is about 6 times shorter,
710 million years, and its long decay chain ends up at lead-207. The
ratio of how much lead is in the rock to how much uranium is in the
rock for those different isotopes gives us ages. But we don’t know the
initial concentration of uranium, so we use both the uranium-238 to
lead-206 ratio and the uranium-235 to lead-207 ratio to correct for loss
of lead due to leaching out of the mineral. We want to choose a mineral
to date that has uranium in it but no lead, such as zircon, which is used
to date rock formations up to the age of the Earth.

wwWe can do potassium-argon dating. Potassium-40 decays to


argon-40 with a half-life of 1.25 billion years. Argon-40 is stable (in
fact, that’s where the argon-40 in our atmosphere comes from). The
argon that results from potassium decay is trapped in the rock matrix.
The ratio of the argon-40 to the potassium-40 gives us the age of the
rock. It measures when the rock last solidified because when the rock is
molten, the argon can escape.

wwWe can also do krypton-81 dating of groundwater. Krypton-81 is


produced in the upper atmosphere by cosmic rays hitting stable
krypton isotopes. This is similar to how carbon-14 is produced. The
half-life of krypton is 230,000 years. Krypton is inert—it doesn’t
react—so it’s ideal for tracing water. But it’s difficult to measure. Like
carbon-14, the concentration is in parts per trillion, but it has very low
solubility. There are only about 2000 krypton atoms in a liter of water.

wwWe now use a more sensitive technique, called atom trace trap analysis,
which involves laser manipulation of individual atoms to count
them. We extract the krypton from several tons of groundwater. One
application of this technique is to date old ice from ice cores, which
helps us establish dates for global temperature records.

Lecture 23
| Isotopes as Clocks and Fingerprints 253
wwDeep ocean water circulation takes about 1000 years and is crucial
for global heat transport. Krypton has too long a half-life, so we use
argon-39 to measure the age of water samples. Just like krypton and
carbon-14, argon-39 is produced in the atmosphere by cosmic rays
interacting with stable argon-40 to make argon-39, which has a half-life
of 269 years. The concentration, unfortunately, is 1 part in 1015, or
1 part per quadrillion. The amount of argon-39 remaining tells us how
long that water was flowing at the bottom of the ocean.

wwIonizing radiation creates defects (electrons and holes) in mineral


crystal grains in ceramics or rock. These defects accumulate over time.
They can be erased by baking or exposure to sunlight. We measure
the time since the object was last baked (over 500 centigrade), which
applies to ceramics and glasses, or we measure the time since the object
was last exposed to sunlight, which applies to glasses, sediment layers,
and old streets.

wwHow do we measure this damage? Heating the material lets the


electrons and holes recombine and emit light, and we measure
the thermoluminescence. Or we can expose the sample to specific
wavelengths of light and measure the optically stimulated luminescence,
and the amount of light given off (the luminescence) is proportional to
the age of the sample. This is called luminescence dating.

wwThen, we determine the yearly radiation dose by burying standard


test pellets for a year and measuring the nearby radioactive isotopes.
By applying these techniques, we can study the age of quartz
strains sensitive up to about 300,000 years. There is a wide range of
applications of this, including dating bricks in the Saint James Church
in Torún, Poland, and dating strata at Megiddo.

254 Nuclear Physics Explained


wwUnstable isotopes let us date objects. What can we learn from stable
isotopes? There are 2 main different photosynthetic paths in plants
(called C3 and C4), giving 2 slightly different ratios of carbon-13 to
carbon-12. We measure the carbon-13 to carbon-12 ratio by burning
it to carbon dioxide, putting it in a mass spectrometer, and comparing
masses 44, 45, and 46. If it’s carbon-12 and oxygen-16, it has a mass of
44; if it’s carbon-13 and oxygen-16, it has a mass of 45.

wwUsing this information, we can look for corn syrup (path C4)
adulteration in products from plants that use path C3, such as honey,
maple syrup, and fruit juice. Corn syrup is much cheaper, so it’s a lot
cheaper to adulterate and replace some of these with corn syrup.

wwWhat other isotopes can we learn from? Heavy water (deuterium oxide
or hydrogen plus oxygen-18) evaporates slightly less than regular water.
The abundance of heavy water decreases as you go from the equator
north or from the equator south. Heavy water has gone through more
evaporation cycles and is colder. One way to use this information
is to measure the average global temperature from heavy-water
concentration in Greenland ice cores because the evaporation rate of
heavy water depends on temperature.

wwWe can use isotopic abundances both to determine the date of an


object and its source. These radioactive isotopes need to be either from
the uranium, thorium, or potassium decay chains in the Earth or made
in the atmosphere by cosmic rays, such as carbon, argon, or krypton.
The effects and uses of these isotopes are opportunities to explore—for
chemists, geologists, archeologists, biologists, and everyone else.

Lecture 23
| Isotopes as Clocks and Fingerprints 255
Supplements

READINGS

Arnold, “Cold War Bomb Testing Is Solving Biology’s Biggest


Mysteries.”
Collon, “Using Nuclear Techniques to Analyze Art.”
Lilley, Nuclear Physics, chap. 8.
Mackova, et al., eds., Nuclear Physics for Cultural Heritage.

QUESTIONS

1 Could someone have used carbon-14 dating to prove that the


1983 Hitler Diaries, which were sold for more than 9 million
deutsche marks, were forgeries?

2 How much radioactive material would a forger need to add to


artificial vanilla so that it has the same carbon-14 content as
natural vanilla?

3 Why can’t we use carbon-14 dissolved in water to date groundwater


or ocean water, rather than using argon or krypton?

256 Nuclear Physics Explained


ANSWERS

1 Probably. Analysis of the paper would probably have shown that it had too
high a content of carbon-14 and had been made after the 1950s bomb pulse.

2 Because the abundance of carbon-14 is only 1 part per trillion, 1 teaspoon


of vanilla (about 5 grams) would need at most 5 picograms (5 trillionths of a
gram) of carbon-14. This is about 5 × 1010 carbon-14 atoms. Carbon-14 has
a half-life of about 6000 years, or 2 × 1011 seconds. Therefore, one of the
5 × 1010 carbon-14 atoms would decay about once every 4 seconds. Thus,
the forgers would need to add 1/4 becquerels of carbon-14 to each teaspoon of
vanilla. This is much less than a banana equivalent dose.

3 This is because carbon is chemically and biologically active and therefore


will not stay put. The argon or krypton could only enter the water when it
was in contact with the air. Carbon could enter the water through chemical
reactions or biological transport (e.g., fish excreta).

Lecture 23
| Isotopes as Clocks and Fingerprints 257
24 VIEWING THE
WORLD WITH
RADIATION

M
 asuring radiation, either emitted by an object
e
or passing through it, gives us a new way to
look at the world. There are many options
with radiation. We can do passive scanning,
where we look for radioactive materials; we
measure gamma rays and maybe neutrons,
but we don’t measure alpha and beta particles
because they don’t escape to be detected.
We can also do active scanning, such as
measuring absorption or scattering in
materials with lots of protons using x-rays or
CT scans, exciting radioactivity using x-rays or
gamma rays, doing surface scans with x-ray
fluorescence, and scanning the volume of an
object using nuclear resonance fluorescence
or neutron activation analysis. Furthermore,
we can use radioactive tracers, just like we
do in nuclear medicine. u

258
Why would anyone want
Passive Scanning to put a nuclear physics
experiment in a 1-foot
wwPassive scanning means looking hole, miles underground,
for the radiation emitted by right behind a drill bit?
an object. For example, we The environment is
look for radioactive material in incredibly hostile, with high
scrap metal going to steel mills. temperature, high pressure,
Radioactive steel was made into and lots of vibration and
reinforcement bars for apartment abrasion. This is done to
buildings in Taiwan in the early study the surrounding rock
1980s. Those apartments are still to look for oil and gas.
slightly radioactive.

wwWe can detect nuclear tests. In 1945, the Kodak company received
lots of complaints that their x-ray film was fogged. Kodak was careful
to avoid radium contamination of the packaging for their film.
But they traced the fogging to packaging made in Indiana that was
contaminated with nuclear fallout from the Trinity test.

wwIn 2006, we detected radioactive xenon-133 two weeks after a North


Korean nuclear bomb test. The test was underground, but it’s very
difficult to keep noble gases like xenon from escaping the test site, even
underground. There is a worldwide network of monitoring sites that
look for isotopes like xenon to detect nuclear bomb tests.

wwWhat else can we learn about nuclear tests or explosions? If we can


sample the ground where the test took place, the distribution of
actinides (such as uranium and plutonium) tells us whether it was a
uranium bomb or a plutonium bomb. We can figure out the grade of
fuel if it was a plutonium bomb from the ratio of plutonium-240 to
plutonium-239. We can tell whether it was bomb grade or just
reactor grade from the ratio of uranium-235 to uranium-238. We
can see how enriched the uranium was. And we can determine the
age of the plutonium in the device from the ratio of other isotopes,
americium-241 to plutonium-241.

Lecture 24
| Viewing the World with Radiation 259
Cobalt-60 from a medical source contaminated a truck in
Mexico in 1983. When the truck was scrapped and recycled,
the cobalt contaminated 5000 tons of steel. This was
discovered when a load of that steel entered Los Alamos
National Laboratory and set off the radiation alarms.

wwWhat material was used? Did they use a deuterium-tritium boost?


To determine this, we don’t need to be on the ground; instead, we
can fly through the fallout plume and sample the gases. Then, we
take ratios of isotopes because lots of background and noise cancel
in ratios. We look at the ratio of iodine-130 to iodine-135 versus
cesium-136 to cesium-137. This tells us whether it was a uranium
bomb or a plutonium bomb and by how much the yield was boosted by
fusion boosting.

260 Nuclear Physics Explained


wwHow do we detect smuggling of radioactive materials for a bomb or
a dirty bomb? We have radiation portal monitors at places where
lots of material enters the United States. We usually have radiation
monitors for gamma rays and sometimes for neutrons. We use plastic
scintillators (which are relatively cheap detectors) for screening. If we
see radiation, then we measure the gamma-ray energies more precisely
with higher-resolution crystal scintillators, such as sodium iodide or
high-purity germanium.

wwWe can also use neutron detectors that use helium-3 and take
advantage of the reaction that if a neutron hits helium-3, it can make
helium-4 and give off a gamma ray. Medical patients and truckloads
of bananas set off alarms. Uranium and plutonium are not very
radioactive—they don’t emit many gamma rays—so they’re very
difficult to detect passively.

wwWe can also use natural radiation (cosmic-ray muons) to scan objects.
About 1 cosmic ray passes through your hand every second. These are
high-energy muons, of several billion electron volts, so they’re very
penetrating. We can put detectors below the objects to see if we have
more muons in one place than in another, and if that is the case, then
they pass through more mass in one place than another. This was used
to detect hidden chambers in the second pyramid of Giza.

wwAlternatively, we can put muon detectors above and below the object
to measure where muons change direction, or scatter. From this
information, we can make a density map of the object or, for example,
we could see where the high-atomic-number materials (uranium and
plutonium) are in trucks. One application of this has been to measure
the inner structure of the dome of the Santa Maria del Fiore in Florence
to look for possible iron reinforcing chains that helped hold it together.

Lecture 24
| Viewing the World with Radiation 261
Active Scanning
wwWe can learn much more if we can actively scan with a beam that we
control. For example, we can use x-rays to measure x-ray absorption
and measure the density profiles of materials (especially those with lots
of protons). One application is looking for weld defects when 2 pieces
of metal are welded together.

wwWe can go beyond density and measure what something is made


of by using a few different techniques. X-ray fluorescence involves
exposing a target to x-rays from 20 to 60 kilovolts and then looking
for lower-energy x-rays emitted by the target. Because x-rays have very
limited range, this technique only measures materials on the surface.
The different x-ray energies are characteristic of different elements.

262 Nuclear Physics Explained


wwIn 2011, 53 paintings sold in Germany that were allegedly by 20th
century masters were determined to be forgeries. This was $30 million
in art sales. One painting, allegedly by Max Ernst, sold for more than
$1 million. The authenticator swore that it was genuine Max Ernst and
didn’t need scientific testing. But x-ray fluorescence showed titanium,
which was not used in contemporary pigments, so it had to be a forgery.

wwAnother technique is ion beam analysis, which involves measuring the


same characteristic x-rays but exciting them with low-energy (of a few
million electron volts) ion or proton beams that hit the sample. This
uses a relatively small accelerator. Similar to x-ray florescence, this
technique is sensitive to the surface of the material, but the advantage
of this over x-ray florescence is that the beam energy allows us to select
the depth of analysis (on the order of micrometers).

wwFrom ion beam analysis, we can learn the composition of gilding in


late antiquity jewelry, the source of alkali in early medieval glass beads,
whether the death of Tycho Brahe was due to mercury poisoning (it
probably wasn’t), and the date of Galileo’s writings by analyzing the ink
he used.

wwCuriosity and Sojourner Mars rovers carried alpha-particle x-ray


spectrometers to measure the composition of the Martian soil. If
we increase the photon energy from x-rays to gamma rays, we can
excite the nuclei, not just atomic electrons. One big advantage of this
technique—called nuclear resonance florescence—is that gamma rays
can travel much greater distances.

wwWe use a 6- to 10-million-electron-volt linear accelerator to make


photons that range from zero up to 6 to 10 million electron volts. These
photos are usually called gamma rays but are sometimes sloppily called
x-rays. The photons at resonant energies excite narrow nuclear states.
The nucleus deexcites by emitting a very specific photon energy. We
measure the photon energies to determine the elemental composition of
the sample.

Lecture 24
| Viewing the World with Radiation 263
wwWe can combine nuclear
resonance florescence with other
nuclear techniques to scan cargo
A company called Passport
containers. We need to scan
Systems uses a 9-million-
quickly—1 or 2 minutes per
volt linear accelerator
container—and not use much
radiation because we don’t want
to scan a pencil beam
to harm the contents or any
of high-energy photons
possible stowaways. across a cargo container
and measure everything
wwAnother way to study the that can be measured—
composition of objects absorption of transmitted
nondestructively is through photons or x-rays and
neutron activation analysis, backscattered photons—
which involves bombarding a to give a 3-dimensional
sample with neutrons, typically density profile of the high-
either from a reactor, a spallation atomic-number material in
source, or a radioactive source. the container. Then, nuclear
The slower neutrons have a resonant florescence
higher interaction probability. is used to measure the
photons reemitted at lower
wwThe atom absorbs the neutron, energies, which allows the
now in an excited state, so it identification of specific
emits a prompt gamma ray to elements. A system like
deexcite. The gamma-ray energy this can detect alcohol
depends on the neutron energy from the ratio of carbon to
and the nuclear binding. Some oxygen; explosives from the
nuclei are now unstable, so they ratios of carbon, oxygen,
decay and emit delayed photons. and nitrogen; and special
nuclear materials from
density and from emitted
neutrons.

264 Nuclear Physics Explained


wwThe benefit of neutron activation analysis is that the neutrons penetrate
the entire volume of the sample, so we can measure the overall
composition of the object. With this technique, we’re not just looking
at the surface.

wwBecause all elements emit prompt gamma rays, we only see the
major components of the sample. The signal from trace elements
is overwhelmed. But only some elements emit delayed gamma
rays. They have different half-lives and characteristic gamma-ray
energies; therefore, this measurement is much more sensitive to lower
concentrations of elements.

Lecture 24
| Viewing the World with Radiation 265
wwSpecific gamma-ray energies
(spectral lines) correspond to
specific elements. Therefore, The Fermi Gamma-Ray
in a particular sample, we Space Telescope detected
might see the biggest peaks gamma-ray pulsars and
from manganese and sodium gamma-ray bursts with
and much smaller peaks from 10,000 times the power of
titanium and magnesium, a supernova.
meaning that there are just trace
amounts of those elements.

wwUsing neutron activation analysis, archeologists have discovered sources


for Mayan obsidian that was found at Chichén Itzá, which tells us
about neolithic trade patterns in Central America. Archeologists have
also measured the silver content of Roman coins with prompt neutron
activation analysis to see which coins were debased by which emperors.
They found that the silver content of the Roman coins decreased
particularly rapidly during times of political instability.

Radioactive Tracers
wwIn addition to active and passive scanning, we can use radioactive
tracers. This is just like what we do in nuclear medicine with
technetium or PET scanning.

wwOne application is radioactive tracers of different plant nutrients. We


can add the appropriate isotopes either to the plant fertilizer to see
how it takes up the fertilizer or to the atmosphere to see how it absorbs
carbon dioxide. Then, we can measure the radiation given off by the
tracers with PET or SPECT scanning.

266 Nuclear Physics Explained


wwFor example, how do plants react to higher carbon dioxide
concentrations? We can add carbon dioxide made with carbon-11 to
the atmosphere and measure carbon uptake as we vary the carbon
dioxide levels, or the nutrient levels, etc.

wwWe can measure radiation given off in the leaves or stems, but the roots
are underground. How do we measure the root system? A rhizobeta
detector, which is under development at Jefferson Lab, uses lots of
plastic scintillator balls connected by wavelength-shifting fibers. The
signals are measured with photomultiplier tubes, and as a result, we
know where the radiation was detected.

wwVery few photomultiplier


tubes are used, so we can
cover a large area relatively
inexpensively. A detector
buried around the roots
of plants can tell us, for
example, how plants absorb
nutrients by measuring
the electrons given off
by phosphorous-32 and
phosphorus-33 beta decay.
We can use the same
detector to potentially
measure radiation leaks
in groundwater from
nuclear sites.

Lecture 24
| Viewing the World with Radiation 267
Supplements

READINGS

Baras, “Exploding Stars Could Have Kick-Started Our Ancestor’s


Evolution.”
Blitz, “When Kodak Accidentally Discovered A-Bomb Testing.”
Collon, “Using Nuclear Techniques to Analyze Art.”
Glascock, “Overview of Neutron Activation Analysis.”
Lilley, Nuclear Physics, chap. 8.
Mackova, et al., eds., Nuclear Physics for Cultural Heritage.

QUESTIONS

1 What technique could we use to determine if a wedding ring was


solid gold or just gold plated?

2 How does the half-life of PET isotopes limit PET studies of carbon
uptake by plants?

3 Why don’t we use neutron activation analysis to scan cargo


containers for special nuclear material?

ANSWERS

1 We would need to use a technique that is sensitive to the entire volume of


the object, not just its surface. We could use neutron activation analysis and
detect the prompt gamma rays emitted by the material of the ring to see its
major components.

268 Nuclear Physics Explained


2 Carbon-11 has a half-life of 20 minutes, so PET studies can only look at the
immediate uptake of carbon from the atmosphere, rather than the transport
of carbon within the plant.

3 Cost, speed, and resolution. We can make a narrow, intense beam of gamma
rays for nuclear resonance fluorescence measurements with a relatively
inexpensive compact electron accelerator. We cannot make a narrow, intense
beam of neutrons. A wider beam would have much worse spatial resolution.

Lecture 24
| Viewing the World with Radiation 269
THE PERIODIC TABLE
OF THE ELEMENTS

Electron shells

270 Nuclear Physics Explained


The Periodic Table of the Elements 271
...CONTINUED

Proton shells

272 Nuclear Physics Explained


The Periodic Table of the Elements 273
THE CHART OF NUCLIDES
Visit https://www.nndc.bnl.gov/chart/
to see a comprehensive and interactive
chart of nuclides.

Magic Numbers:
2, 8, 20, 28, 50, 82, 126

274 Nuclear Physics Explained


The Chart of Nuclides 275
GLOSSARY

A
abundance: The fraction of an element comprised by a given isotope of that
element. For example, carbon-12 comprises 98.93% of naturally occurring
carbon, so its abundance is 98.93%. By contrast, the abundance of
uranium-235 is only 0.72%.

accelerating cavity: Uses metal shaped into resonant cavities that build
up electromagnetic fields (usually microwave or radio wave frequency)
whose polarity is reversed many times each second in time with the arrival
of charged particles so as to push and pull the charged particles along,
accelerating the particles almost to the speed of light. Often made from
superconducting material to reduce power used and heat dissipated in
the cavity.

accelerator: Device that uses an electric field to speed up charged particles


to high energies. This includes an electrostatic generator, such as a Van de
Graaff generator, which accelerates the charged particles through a large
constant (DC) voltage; and a radio-frequency accelerator, such as a linear
accelerator or a cyclotron, which accelerates groups of particles (pulses)
that are synchronized with the frequency of the radio waves.

accelerator mass spectrometer: A very sensitive mass spectrometer


that can measure the masses of individual atoms. Used for carbon dating
and other forms of radioactive dating because it only needs tiny samples of
the material.

276 Nuclear Physics Explained


acceptance: Used to describe the spectrometers that detect the particles
knocked out from reactions at accelerator labs. Large-acceptance
spectrometers can typically detect particles emitted at almost all angles
and momenta from the collision, while small acceptance spectrometers
can typically only detect particles emitted in a very small range of
angles and momenta. Jefferson Lab (Newport News, Virginia) has
large-acceptance spectrometers in Halls B and D and small acceptance
spectrometers in Halls A and C.

actinides: Elements from 89 (actinium) to 103 (lawrencium). See


also transuranics.

acute dose: A large dose of radiation received in a short period of time.

alpha decay: The radioactive decay of an unstable heavy nuclide where the
nuclide emits an alpha particle, reducing the number of neutrons and the
number of protons in the resulting nucleus by 2 each.

alpha particle (or alpha rays): The nucleus of a helium atom, containing
2 protons and 2 neutrons. Emitted by the radioactive decay of unstable
heavy nuclides.

atomic bomb: Colloquial term for a nuclear bomb or nuclear weapon.

B
background radiation: The radiation people receive from terrestrial, cosmic,
and medical sources.

becquerel (Bq): Metric system unit of radioactivity, corresponding to


1 nuclear decay per second. A nonmetric unit is the curie.

Glossary 277
beta decay: Radioactive decay of a nucleus that keeps the same atomic weight
but increases or decreases the number of protons by 1. Beta-minus (β −)
decay refers to the radioactive decay of a neutron-rich isotope, where
1 neutron decays into a proton, an electron, and an antineutrino; beta-plus
(β+) decay refers to the radioactive decay of a proton-rich isotope,
where 1 proton decays into a neutron, a positron (the antiparticle of the
electron), and a neutrino.

beta particle (or beta rays): A high-energy electron (“beta-minus”) or


positron (“beta-plus”) of keV to MeV), when emitted by the beta decay of
an unstable nuclide.

big bang: The expansion of the universe from its original high-temperature,
high-density state.

binding energy: Average energy per nucleon needed to completely remove


all the nucleons from a nucleus. Alternatively, it is the energy gained by
assembling the nucleons into the nucleus.

bomb peak: An excess of carbon-14 in the atmosphere due to above-ground


nuclear bomb tests during the 1950s and 1960s. Used to precisely
carbon-date modern samples.

brachytherapy: A type of radiation therapy for cancer treatment through


either permanently implanted seeds or temporary introduction of
radioactive material via implanted tubes (catheters).

bremsstrahlung: Meaning “braking radiation,” this radiation consists


of x-ray and gamma-ray photons emitted by high-energy electrons as
they pass through matter. Matter containing nuclei with large numbers
of protons, such as lead or tungsten, causes electrons to emit more
bremsstrahlung radiation than matter with lighter nuclei. This effect is
used by x-ray machines to produce x-rays.

278 Nuclear Physics Explained


burnup rate: The fraction of the fuel of a nuclear reactor that is fissioned—
typically about 5% today.

C
calorimeter, electromagnetic: A detector that measures the energy of
a particle, typically alternating layers of lead (or iron) and scintillator.
Electromagnetic calorimeters use bremsstrahlung radiation and pair
production in the lead to make an electromagnetic shower and measure
the total energy of the showering particles in the scintillators.

carbon dating: Method using the amount of radioactive carbon-14


(half-life = 5700 years) in an organic sample to determine when the plant
or animal it came from died.

centrifuge: A rapidly spinning device used to enrich uranium.

chain reaction: A reaction where the neutrons released from the fission of
one nucleus of uranium or plutonium go on to fission other nuclei. If the
neutrons from one fission go on to fission exactly one more fission, then
the chain reaction is critical and continues at the same rate.

charge density: Amount of electric charge per unit volume (or per unit area,
or per unit length).

chart of nuclides: Any chart or table arranging all the known nuclides by
number of protons versus number of neutrons, together with their decay
mode, half-life, and other properties. The nuclear equivalent of the
periodic table. See PAGE 274.

Cherenkov (Cerenkov) counter: A type of particle detector that measures


particle speed by detecting Cherenkov radiation (the flash of light)
given off by a particle as it traverses the material in the detector. There
are several variants, including a threshold Cherenkov counter, which

Glossary 279
just detects the flash of light, and a ring-imaging Cherenkov (RICH)
counter, which measures the opening angle of the emitted cone of light.
The threshold Cherenkov counter just determines if the particle’s speed
is faster than the speed of light in the material; the RICH measures the
speed from the opening angle of the cone.

Cherenkov radiation: Light emitted when a charged particle that passes


through a medium is traveling faster than the speed of light in that
medium. This electromagnetic boom is analogous to the sonic boom
produced when an object, such as an airplane or the tip of a whip, travels
faster than the speed of sound in air. The speed of light in vacuum is
always 2.9979 × 108 meters per second, but the speed of light in material
depends on the material. Light is up to about 0.1% slower in gasses and
about 25% slower in water.

Chernobyl (Ukraine): Site of the worst nuclear power plant accident. The
graphite-moderated, water-cooled reactor suffered an uncontrolled
nuclear reaction and fire that resulted in the dispersal of radioactive
material across very wide areas.

China syndrome: Hypothetical accident scenario where the nuclear fuel in


a reactor core not only experiences a core meltdown, but also continues
melting downward through the pressure vessel, thereby escaping to release
huge amounts of radiation. It has not happened yet.

chronic dose: A dose of radiation received over an extended period of time.

CNO cycle: A process for fusing 4 protons into helium in stars, where a
nucleus of carbon, nitrogen, or oxygen (CNO) serves as a catalyst.
Less important than the PP chain in the Sun, but more important in
heavier stars.

collimator: Comblike device in front of a gamma camera that defines the


direction of the measured gamma rays.

280 Nuclear Physics Explained


computed tomography (CT): Three-dimensional x-ray image reconstructed
from many different beams of x-rays, requiring 100 to 1000 times as much
radiation exposure (roughly 1 to 20 millisieverts) as conventional x-rays.

control rods: Rods made of a material, such as cadmium or boron, with a


very large neutron absorption cross section that can be inserted into a
reactor to absorb neutrons and reduce the reaction rate.

core meltdown: See meltdown.

cosmic radiation: Radiation from outer space. Caused by high-energy


protons and alpha particles hitting the atmosphere and making a shower
of other particles. Most cosmic radiation is screened by the Earth’s
magnetic field and atmosphere. Almost all of the particles that reach the
ground are muons.

cosmic ray: One particle of cosmic radiation; when reaching Earth, typically
a muon.

critical: A chain reaction is critical if it continues at the same rate without


increasing or decreasing. See also subcritical and supercritical.

critical mass: The amount of a fissile material required to sustain a


chain reaction.

Crookes tube: An early experimental electric discharge tube that was


discovered to emit x-rays.

cross section: Used to describe the probability that a high-energy particle


will interact with a target nucleus or target particle. The total cross section
is the effective area of the target as “seen” by the incoming particle. For
example, the total cross section for a high-energy proton to interact
with a uranium nucleus is approximately the cross-sectional area of the
nucleus. The total cross section for a high-energy electron to interact
with the same uranium nucleus is much smaller because it interacts via

Glossary 281
the electromagnetic force, which is much weaker than the strong nuclear
force. There are also differential cross sections, which are the probabilities
or effective areas for an incident particle to interact with the target in
specific ways, whether scattering at a particular angle, with a particular
energy, or through a particular reaction mechanism.

CT scan: See computed tomography.

curie (Ci): Unit of radioactivity corresponding to 3.7 × 1010 nuclear decays per
second. 1 Ci = 3.7 × 1010 Bq.

curve of binding energy: The binding energy of nuclei plotted versus


nuclear mass. It increases rapidly for light nuclei from hydrogen to helium
(7 MeV per nucleon), continues to increase up to iron or nickel (8 MeV
per nucleon), and then decreases slowly to the very heavy elements (about
7 MeV per nucleon).

cyclotron: A type of accelerator invented by Ernest O. Lawrence in 1934 that


uses a large magnet to make the accelerated particles travel in increasing
circles and an alternating voltage to accelerate the particles as they pass
through the gap between the magnet Ds. Compare with synchrotron.

D
dating, luminescence: Dating crystalline materials using the light emitted
(luminescence) when heated (thermoluminescence) or illuminated
(optically stimulated luminescence). The total light emitted is
proportional to the age of the sample.

dating, radioisotope: Dating materials using either remaining abundances


of radioisotopes, such as carbon-14 or krypton-81, or the ratios of
isotopes. See also carbon dating.

282 Nuclear Physics Explained


deep inelastic scattering: An interaction where an incident electron or
muon scatters from a quark in a nucleon or in the nucleus.

delayed neutrons: Neutrons emitted from the radioactive decay of fission


fragments following nuclear fission. See also prompt neutrons.

depleted uranium: Uranium with less than the natural abundance of


uranium-235.

detector: A device that amplifies and detects the tiny amounts of energy
left behind by a particle as it traverses the material of the detector to
determine the arrival time, speed, energy, or position of the particle. Types
of detectors include calorimeters, Cherenkov counters, drift chambers,
photomultiplier tubes, scintillators, spectrometers, and wire chambers.

deuterium: Heavy hydrogen, with 1 proton and 1 neutron orbited by


1 electron.

deuteron: The nucleus of heavy hydrogen, with 1 proton and 1 neutron.

dipole magnet: A magnet that makes a constant magnetic field perpendicular


to the direction of the charged particles to bend the trajectory of the
charged particles. The larger the magnetic field and the smaller the
particle momentum, the more the trajectory bends. Used in accelerators
to steer the beam and used in spectrometers to measure the momentum of
each particle.

dirty bomb: Radioactive materials dispersed by conventional explosive. Aims


to contaminate large areas by spreading the radioactive material widely.

diverter: Removes helium-4 “ash” from a tokamak reactor.

Glossary 283
drift chamber: A type of wire chamber that measures the time it takes for
the electrons knocked loose by the passage of the high-energy particle
through the wire chamber to reach a wire and be detected. It uses the
timing information to determine the location of the charged particle
much more precisely.

dynode: Intermediate electrode in a photomultiplier tube.

E
elastic scattering: A type of reaction where neither the incident particle
nor the target is changed and kinetic energy is conserved. This contrasts
with inelastic scattering, where either the incident particle or the target
is changed, whether by changing to an excited state, by creating more
particles, by melding the incident particle and the target, or by knocking
particles out of the target. In inelastic scattering, kinetic energy is
typically transformed to exciting or altering the target nucleus. See also
quasielastic scattering.

electron: Elementary particle with mass 0.5 MeV/c 2 and spin 1/2. Electrons
orbit the nucleus to form atoms. Emitted at high energy in beta decay
(specifically in β − decay). The antiparticle of the positron. Used in
accelerator labs to study nuclei in high-energy collisions.

electron volt (eV): Unit defined by the energy gained by an electron passing
through a potential difference of 1 volt. This unit is used to describe
subatomic energies. Because of the principle of special relativity and its
corollary, E = mc 2, this is used for energy (eV), momentum (eV/c), and
mass (eV/c 2). In nuclear physics, it is much more common to use keV (kilo,
or thousand), MeV (mega, or million), GeV (giga, or billion), and TeV
(tera, or trillion).

element: Any nucleus with a specific number of protons (e.g., any nucleus
with 6 protons is the element carbon).

284 Nuclear Physics Explained


EMC effect: A measurement by the European Muon Collaboration that
compared the quark distributions inside nucleons in heavy nuclei with
those inside deuterium. It found that quarks in nucleons in heavy nuclei
move more slowly than quarks in free nucleons or in deuterium.

enriched uranium: Uranium with more than the natural abundance of


uranium-235. Low-enriched uranium has less than 20% uranium-235;
high-enriched uranium has more than 20% uranium-235.

excitation energy: Energy difference between an excited state of a nucleus or


nucleon and the ground state of that nucleus or nucleon. For example, the
first excited state of carbon-12 has additional energy of 4.4 MeV.

excited state: State of a system (atom, nucleus, or nucleon) that has more
energy than the ground state. The difference in energy between the
ground state and an excited state is the excitation energy. See also
metastable isotope.

F
fertile isotope: An isotope that becomes fissile when it absorbs a neutron.
Examples include thorium-232 and uranium-238.

fissile isotope: An isotope that will fission when it absorbs a neutron. Fissile
isotopes include uranium-233 and uranium-235 as well as plutonium-239.

fission: The process of splitting a nucleus into 2 smaller nuclei, plus 2 to


3 neutrons. Energy can be released by fissioning nuclei much heavier than
iron. See also reactor.

fission, accelerator-driven: A subcritical fission reactor where the extra


neutrons needed to maintain the chain reaction are supplied by a
particle accelerator.

Glossary 285
fission fragments: Smaller nuclei resulting from nuclear fission.

Fukushima: Site in Japan of a nuclear power plant containing 6 reactors


where a tsunami resulted in a loss-of-coolant accident and the meltdown
of 4 of the reactors.

fusion: Process of fusing 2 nuclei into 1 larger nucleus. Typically requires


high temperatures and confinement of light nuclei under high pressures
at high densities. Energy can be released by fusing light nuclei into nuclei
lighter than iron (26 protons). Gravitational confinement is used in stars;
inertial confinement involves hitting a fuel pellet hard from all directions,
usually with powerful laser beams, to compress the fuel; and magnetic
confinement, used in tokamaks, relies on magnetic fields to confine
the fuel.

G
gamma camera: Detector used to make an image with gamma rays. It
comprises a collimator, a scintillator, and PMTs. Used for PET and
SPECT scans.

gamma rays: High-energy photons (with more than 100 keV). Can be
emitted by the decay of nuclei from an excited state to a lower-energy
state (typically from 100 keV to a few MeV). Can also be produced in
accelerators at higher energies.

gaseous diffusion: Technique used to enrich uranium.

Geiger counter: Detector consisting of a metal tube containing a gas with


a fine wire running down the center of the tube. When a particle or
photon interacts in the gas, it knocks atomic electrons loose. As the atomic
electrons drift to the wire, they are amplified by the potential difference
(voltage) between the tube walls and the wire. This amplifies them into a

286 Nuclear Physics Explained


large enough spark that they can make an audible click. Modern Geiger
counters don’t emit the audible click but do emit an electronic chirp when
they detect the signal from a particle.

Geissler tube: Early vacuum tube, first attempted in 1857, that led to more
sophisticated vacuum tubes, including x-ray tubes.

generator, radioisotope: Device used to make radioisotopes for


medical imaging. It contains a longer-lived isotope that decays to the
shorter-lived desirable isotope. For example, a technetium-99 generator
contains molybdenum-99, which has a 66-hour half-life and decays to
technetium-99*, which has a 6-hour half-life.

gluon: The particle that carries the strong nuclear force between quarks.
Protons and neutrons are each comprised of quarks and gluons.

graphite: Soft form of carbon that is extremely resistant to heat. See reactor,
graphite.

gray (Gy): The metric system unit of absorbed radiation dose corresponding to
an absorbed energy of 1 J/kg. See also rad.

ground state: Lowest energy state of a nuclide or of an atom. Contrasts with


excited state and metastable isotope.

H
half-life: Time it takes for half the nuclei of a radioisotope to decay.

heavy water: Water (H2O) where the hydrogen atoms are replaced with
deuterium (an isotope of hydrogen with 1 proton and 1 neutron).

helium burning: Process where stars fuse helium into carbon or oxygen. This
occurs after exhausting the hydrogen in their cores.

Glossary 287
I
inelastic scattering: See elastic scattering.

ion beam analysis: Irradiating a sample with low-energy ions from an


accelerator and detecting the characteristic x-rays emitted by materials on
the surface of the sample. The frequency of the x-rays tells us the isotopes
on the surface of the sample.

isotone: Nuclei with the same number of neutrons and differing numbers
of protons (e.g., carbon-12 and nitrogen-13 are isotones of neutron
number 6).

isotope (a.k.a. nuclide): Variants of an element with different numbers


of neutrons (e.g., carbon-12 is an isotope of carbon with 6 neutrons,
and carbon-14 is an isotope of carbon with 8 neutrons). See also
metastable isotope.

J
jet: Spray of particles produced when a very-high-energy quark is knocked out
of a nucleon.

L
light water: Regular water (H2O) composed of the most common isotopes of
hydrogen and oxygen: hydrogen-1 and oxygen-16.

linear accelerator (linac): A particle accelerator, typically used to accelerate


electrons, where the elements are all in a straight line. The Stanford Linear
Accelerator Center (SLAC) was the highest-energy electron accelerator in
the world for many years. Jefferson Lab has 2 linear accelerators, connected
by recirculating arcs, so that the electrons can be accelerated up to 5 times.

288 Nuclear Physics Explained


liquid-drop model: Accurately characterizes the masses of most nuclide by
describing the nucleus as a charged liquid drop. See shell model.

loss-of-coolant accident (LOCA): An accident at a reactor where the


fission chain reaction stops, but the heat generated by the residual
radioactivity in the core causes the temperature to rise dramatically. The
Three Mile Island and Fukushima accidents were both caused by loss
of coolant.

M
magnet: Creates the magnetic field used to bend the trajectories of charged
particles. There are several common magnet configurations: dipole
magnets, quadrupole magnets, solenoid magnets, and toroidal magnets.

magnetic resonance imaging (MRI): A method of imaging that relies on


aligning the spins of certain nuclei to measure their density. Typically
uses huge magnets to align the spins of some of the hydrogen nuclei in the
human body to measure their density. Previously called nuclear magnetic
resonance imaging (NMRI).

mass spectrometer: Uses electric and magnetic fields to measure the


masses of atoms.

meltdown: What can happen when the fuel rods in the core of a nuclear
reactor suffer a loss-of-coolant accident . The fuel rods can heat up and
melt down into a puddle of metal at the bottom of the reactor vessel.

metastable isotope: An excited state of an isotope that has a somewhat


“long” half-life, indicated with an asterisk (*) or “m.” For example,
technetium-99* or technetium-99m refers to a metastable state of
technetium-99 that has a half-life of about 6 hours.

Glossary 289
mixed oxide (MOX) fuel: Reactor fuel that is about 95% uranium-238 and
5% plutonium.

moderate: To slow neutrons down in a nuclear reactor, by having them


collide elastically with a light-nuclei moderator, such as hydrogen,
deuterium, or carbon. Moderating (i.e., slowing) the speed of neutrons
increases the probability that they will fission a uranium-235 nucleus and
continue the chain reaction. See thermalize.

muon: A heavy cousin of the electron, with 200 times the mass. The most
common component of cosmic radiation at the Earth’s surface.

N
neutrino: A neutral, ultralight particle that only interacts via the weak
interaction. Emitted in beta decay. Neutrinos have been detected from the
PP chain fusion in the Sun as well as from supernovas.

neutron: One of 2 particles that comprise the nucleus of the atom. Has
zero charge, mass 939.6 MeV/c 2, and spin 1/2. Made of 1 up quark and
2 down quarks.

neutron activation analysis: A technique of bombarding a sample with


neutrons. The composition of the entire sample can be determined by
measuring the radiation given off by activated nuclei in the sample—i.e.,
nuclei that were excited by or absorbed a neutron and are now radioactive.

neutron star: A star composed almost entirely of neutrons that can pack the
entire mass of the Sun into a sphere of radius about 10 kilometers. The
densest object known (other than black holes).

nuclear fission: See fission.

nuclear fusion: See fusion.

290 Nuclear Physics Explained


nuclear mass: The total number of neutrons and protons in a nucleus; often
abbreviated “A”.

nuclear physics: The study of the nucleus of the atom and


related phenomena.

nuclear reactor: See reactor.

nuclear resonance fluorescence: Light given off by samples after


being irradiated with gamma rays. The material of the sample can be
determined from the frequency of the emitted light.

nuclear weapon (or nuclear bomb): A bomb where the energy comes from
the fission of uranium or plutonium. See also thermonuclear weapon.

nucleon: Any individual proton or neutron.

nucleosynthesis: Process of making nuclei.

nucleus (plural nuclei): The entire group of protons and neutrons at the
center of an atom.

nuclide: Nuclear physics term for an isotope (for example, carbon-14 and
carbon-12 are nuclides).

P
pair production: The process where a gamma ray with more than 1 MeV
converts to a positron and an electron as it passes through matter.

particle, elementary: Particle with zero size and no structure. Examples


thought to include electrons, positrons, photons, quarks, and gluons.

Glossary 291
Pauli exclusion principle: The principle that no 2 identical particles can
have exactly the same spin, momentum, and position at the same time.

PET: See positron-emission tomography.

photomultiplier tube (PMT): A vacuum tube used to convert tiny flashes


of light into detectable electronic signals. The photons from the flash of
light hit a photocathode and electrons are knocked out. The electrons are
then accelerated by a potential difference (voltage) to the first dynode,
where they knock out more electrons. The amplification process continues
through between 4 and 12 dynodes until the amplified electronic signal
can be read out from the anode. PMTs are very sensitive to light and
can multiply a small current many times over, such that some PMTs
can detect a single photoelectron knocked out from the photocathode.
Photomultipliers can be used to count flashes of light from a scintillator.
Sometimes replaced by solid-state silicon PMTs.

photon: The fundamental quantum of light and carrier of the electromagnetic


force. Light travels as a wave and interacts as a particle. The
highest-energy photons are gamma rays, followed by x-rays.

pion: Lightest particle composed of a quark and an antiquark; interacts via the
strong force. Produced in accelerator collisions and when cosmic rays hit
Earth’s atmosphere, decaying in nanoseconds via the weak nuclear force
to a muon and a muon-neutrino. In some models, pions “carry” the strong
nuclear force between neutrons and protons.

Planck’s constant: The constant that sets the scale for quantum mechanical
behavior: h ≈ 6 × 10 −34 J-s = 197 MeV-fm. For example, the quantum
uncertainty in a particle’s position times the uncertainty in its momentum
is related to Planck’s constant: ∆x∆p ≥ h/2π.

polarized electrons: Electron beams with a preferential spin orientation,


where one spin state is preferred over another. Analogous to
polarized light.

292 Nuclear Physics Explained


positron: Antiparticle of the electron, with the same mass and spin but
opposite charge. When a positron hits an electron, the 2 particles can
annihilate and convert all their mass into energy in the form of 2 photons.
Also called a beta-plus (β+) particle because it is emitted in β+ decay.

positron-emission tomography (PET): Medical or other imaging using a


radioactive nuclide that emits positrons. When a positron hits an electron
in the material, the 2 particles annihilate and emit 2 photons (gamma
rays) in opposite directions. Contrast with SPECT.

PP chain: The primary source of energy production in the Sun, which starts
with 2 protons fusing to form a deuterium nucleus. The final result is that
4 protons fuse to form a helium nucleus, releasing a lot of energy.

Project Orion: A US study during 1958 to 1963 considering whether to


propel spacecraft by dropping nuclear bombs out the back and using the
blast to propel them forward.

prompt neutrons: Neutrons emitted immediately when uranium or


plutonium fissions. See also delayed neutrons.

proton: One of 2 particles that comprise the nucleus of the atom. Has positive
charge, mass 938.3 MeV/c 2, and spin 1/2. Made of 2 up quarks and 1 down
quark. The number of protons is often abbreviated as “Z.”

proton therapy: Cancer treatment using proton beams instead of x-ray photon
beams. Originally thought to be favorable mostly for solid tumors that are
localized and isolated, but the addition of pencil-beam aim and intensity
modulation have also made larger and multisite tumors more treatable.

Q
quadrupole magnet: Focuses a beam of charged particles similarly to how a
lens focuses light. Used in accelerators and spectrometers.

Glossary 293
quantum chromodynamics (QCD): Theory for understanding the strong
force between quarks that is carried by gluons. See strong nuclear force.

quantum tunneling: Process where a particle can emerge on the other side of
a barrier without actually passing through the barrier. Important in alpha
decay and in the first step of the PP chain for solar fusion.

quark: Constituent of the proton and neutron. Up quarks have mass ≈


5 MeV/c 2, spin 1/2, and charge +2/3. Down quarks have mass ≈ 7 MeV/c 2,
spin 1/2, and charge −1/3. A nucleon is composed of 3 valence quarks and an
indefinite number of virtual quark-antiquark pairs (sea quarks) and gluons.

quasielastic scattering: Elastic scattering from a moving nucleon inside


a nucleus where the nucleon is ejected from the nucleus but the rest of
the nucleus is left relatively unperturbed. This is different from elastic
scattering from the nucleus as a whole, where the entire nucleus is
left unchanged.

Q value: Ratio of energy released by a fusion reaction to the input energy


needed to heat the plasma.

R
rad: A unit of absorbed radiation dose, corresponding to an absorbed energy of
0.01 J/kg. 100 rad = 1 Gy. See also gray.

radiation: The high-energy (keV to MeV) particles emitted by radioactive


materials. The most common forms of radioactivity are alpha, beta, and
gamma rays.

radiation detector: See detectors.

radiation portal monitor: Detectors placed at ports and airports to look for
radioactive material.

294 Nuclear Physics Explained


radiation therapy: Applying radiation to tumors to treat cancer. Treatment
can either use external beams (e.g., x-rays, gamma rays, or protons) or
internal radiation sources (e.g., brachytherapy).

radioactive tracer: A radioactive isotope introduced into a person, an


animal, or a plant. By measuring the emitted radioactivity, we can trace
how the isotope was absorbed and traveled through the organism. Used in
the PET scan and the SPECT scan.

radioactivity: The activity of radioisotopes or of radioactive material.


Measured in becquerels (disintegrations per second) or curies
(3.7 × 1010 disintegrations per second). Each disintegration results in the
emission of at least 1 high-energy particle.

radio-frequency cavity: See accelerating cavity.

radioisotope: Short for radioactive isotope (e.g., fluorine-18, which is used in


PET scans). Any isotope that emits radiation.

reactor: A device for sustaining a nuclear fission chain reaction, typically


using fissile isotopes of uranium and plutonium. Nuclear reactors can
also convert fertile isotopes into fissile isotopes. Can be either thermal
reactors (moderated with light water, heavy water, or graphite) or
unmoderated fast reactors (including molten-salt and thorium reactors).
Thermal, or fast, breeder reactors produce more fissile material
than they consume. For a reactor design that relies on nuclear fusion,
see tokamak .

reactor, breeder: A fission reactor that produces more fissile fuel than it
consumes, by converting fertile isotopes into fissile ones.

reactor, fast: A fission reactor that uses unmoderated, or fast, neutrons to


fission nuclei.

Glossary 295
reactor, graphite: A fission reactor design that uses a graphite (carbon)
moderator. The Chernobyl-style reactors common in the Soviet Union
were graphite-moderated. An advanced (fourth-generation) graphite
reactor is a high-temperature gas-cooled reactor, which uses helium
coolant. For an example, see reactor, pebble-bed.

reactor, heavy-water: A thermal reactor that uses heavy water to


moderate the neutrons. Does not need enriched uranium to operate.

reactor, light-water: A thermal reactor that uses regular water to moderate


the neutrons. The most common type of power reactor in the world, used
extensively on nuclear-powered naval vessels and in commercial power
plants. Includes both pressurized-water reactors (PWR) and boiling-water
reactors (BWR).

reactor, molten-salt: A fission reactor design where the fissile material is


dissolved in a molten salt, which flows through channels in graphite and is
brought together in the core, where it can fission. See also reactor, thorium.

reactor, pebble-bed: A possible graphite -moderated, high-temperature


gas-cooled reactor that would have its fuel encased in layers of
protective carbon.

reactor, thermal: Any reactor that uses a moderator to thermalize the


neutrons emitted by fission reactions, to increase their absorption cross
sections and the probability that they will fission a subsequent nucleus.

reactor, thorium: A type of molten-salt reactor where molten thorium


fluoride flows through the reactor core. Thorium is fertile, so when
it absorbs a neutron, it ends up as uranium-233, which is fissile. The
uranium-233 sustains the chain reaction, releasing energy and more
neutrons that create more uranium-233.

reactor vessel: The thick steel pressure vessel that typically encloses
light-water reactors.

296 Nuclear Physics Explained


rem: A unit of radiation dose equivalent. It is the dose in rads adjusted for
the relative biological effect of different particles. 100 rem = 1 Sv. See
also Sievert .

reprocessing: The processing of spent nuclear fuel to remove leftover


uranium and plutonium (about 95% of the spent fuel). The reprocessed
uranium and plutonium can be converted into mixed oxide fuel.

residual radioactivity: About 1/5 of the power liberated by nuclear fission


comes from the delayed radioactive decay of the fission fragments. The
power released decreases from about 20% immediately after the chain
reaction stops to about 2% a day later.

S
scintillant: Any material that luminesces when excited by radiation. These
can be inorganic crystals, such as sodium iodide, or organic plastics
or liquids.

scintillator: Detector of both charged and uncharged high-energy particles


that combines a scintillant , which emits light when excited by radiation,
and a photomultiplier tube, which converts the flash of light into a
measureable electronic signal. Inorganic scintillator detectors typically
have excellent energy resolution. Organic (plastic) scintillator detectors
typically have excellent time resolution (1 nanosecond or less).

seed: A small encapsulated radioactive source implanted in or around a tumor


used for cancer treatment. See radiation therapy and brachytherapy.

shell model: A model of the nucleus where the constituent protons and
neutrons orbit in s-shell, p-shell, d-shell, etc., orbitals. Adapted from a
similar model of the atom, where the orbiting particles are electrons.

Glossary 297
Sievert (Sv): The metric system unit of radiation dose equivalent. It is
the dose in grays adjusted for the relative biological effect of different
particles. See also rem.

single-photon emission computed tomography (SPECT):


Gamma-camera technique that images the source of single gamma rays
emitted directly from a radionuclide in a person’s body. Contrast with
indirect emission of gamma rays in PET.

solenoid magnet: A cylindrical magnet where the magnetic field is


parallel to the cylinder. Used in some spectrometers so that the incident
particle beam is parallel to the magnetic field and is therefore not
deflected. Particles emerging from collisions between the incident beam
and the target travel at an angle to the field and hence are deflected.
This makes low-energy particles travel in small, tight circles so that
they are not detected and allows measurement of the momentum of
higher-energy particles.

SPECT: See single-photon emission computed tomography.

spectrometer: A magnet plus detectors. Usually built around a magnet.


The detectors measure the trajectory of the charged particles. Physicists
can reconstruct the particle momenta by measuring how much their
trajectories bend in the magnetic field of the magnet. Physicists combine
position-sensitive detectors, such as wire chambers, with time-, energy-,
and velocity-sensitive detectors to determine the particle’s momentum and
identity (proton, electron, pion, deuteron, etc.).

spent nuclear fuel: The fuel after being removed from a nuclear reactor.
Typically contains less uranium-235, more plutonium and other
transuranics, and a lot of highly radioactive fission products.

spin: The intrinsic angular momentum of a particle. The electron has spin 1/2,
so it can be either spin up or spin down. When particles are polarized,
they are spin-aligned.

298 Nuclear Physics Explained


strong nuclear force: The force, explained by quantum chromodynamics,
that holds quarks together to form nucleons and that holds nucleons
together to form nuclei.

subcritical: A chain reaction where the neutrons released from the fission
of 1 nucleus of uranium or plutonium go on to fission less than 1 other
nucleus so that the fission rate decreases steadily.

superconductivity: When some materials are cooled to very low


temperatures, their resistance to electrical currents drops to zero. This
means that no energy is wasted and converted to heat by the electrical
resistance. This allows the use of much higher electrical currents to
achieve much greater electric and magnetic fields. This is often used in
spectrometer magnets and in accelerating cavities. Superconducting
radio-frequency accelerating cavities are referred to as SRF cavities.

supercritical: A chain reaction where the neutrons released from the fission
of 1 nucleus of uranium or plutonium go on to fission more than 1 other
nucleus so that the fission rate increases exponentially.

supernova: Explosion of a massive star. This occurs after it has finished


fusing lighter nuclei to heavier nuclei and now has an inert iron core. The
star collapses and then rebounds to blow off its outer layers and briefly
shine brighter than its entire galaxy. This is the source of many of the
heavy elements in the universe.

synchrotron: Circular accelerator in which the radius of the particles is kept


constant by synchronizing the increasing strength of the magnetic field
with the increasing energy of the accelerated particles. This is used at the
Fermi National Accelerator Lab and Brookhaven’s Relativistic Heavy Ion
Collider in the United States and at the Large Hadron Collider at the
European Organization for Nuclear Research (CERN) in Europe.

Glossary 299
synchrotron radiation: As particles approach relativistic speeds, they
leak increasing amounts of electromagnetic radiation (usually x-rays),
which has practical applications in spectroscopy and crystallography
but also places an upper limit on the energy attainable by a
synchrotron accelerator.

T
technetium-99: Isotope whose metastable form is the most commonly used
for SPECT scans.

temperature: The average kinetic energy of the particles in the substance.


A temperature of 1 Kelvin corresponds to an average kinetic energy per
particle of 10 −4 electron volts.

Tesla coil: A transformer that produces very-high-voltage, high-frequency


electrical discharges.

thermalize: To reach a temperature equilibrium by losing or gaining energy


in the form of heat. A thermal neutron has had many collisions with
nuclei to end up with the same average kinetic energy as the nuclei it
collides with. This average kinetic energy is described by the temperature.
The thermal energy at room temperature (70° Fahrenheit, 20° Celsius, or
300 Kelvin) is about 1/4 of an electron volt. Neutrons in a nuclear reactor
are sometimes moderated to thermalize them.

thermonuclear weapon (or thermonuclear bomb): A bomb where the


energy released comes from both nuclear fission of uranium or plutonium
as well as nuclear fusion of deuterium and tritium into helium. Releases
much more energy than an ordinary nuclear bomb.

thorium: See reactor, thorium.

300 Nuclear Physics Explained


Three Mile Island: The site of a nuclear accident that resulted in a core
meltdown of 1 of the 2 nuclear reactors.

tokamak: One approach for pursuing controlled nuclear fusion. It confines


the ultrahigh-temperature plasma using magnets in the shape of a
torus; the word “tokamak” comes from a Russian acronym for “toroidal
magnetic chamber.” Examples include the Joint European Torus (JET)
and Japan Torus-60.

toroidal magnet: Magnet where the direction of the magnetic field follows
the donut shape of a torus. Used in the large-acceptance spectrometer
(CLAS) in Hall B at Jefferson Lab. Also used in tokamak .

transuranics: Isotopes with more than 92 protons (uranium). See


also actinides.

tritium: A heavy radioactive isotope of hydrogen that has 1 proton and


2 neutrons. Used in thermonuclear weapons and in controlled
nuclear fusion.

V
valley of stability: The most stable nuclide of each nuclear mass on the
chart of nuclides. Nuclides with more protons or more neutrons will
either be less bound or less stable (or both) than the nuclide at the bottom
of the valley.

Van de Graaff generator: An early particle accelerator invented at


Princeton in the 1930s. It uses a constant high voltage created by the
mechanical transport of electrons to one terminal of the generator to
accelerate charged particles. Some are still used as low-energy (several
MeV) accelerators today.

Glossary 301
virtual particles: Particles that only exist momentarily, usually in the form
of particle-antiparticle pairs. They pop out of the vacuum for a brief
moment and then annihilate.

W
weak nuclear force: The force that causes beta decay. Much weaker than
the strong nuclear force.

wire chamber: A type of particle detector that measures the position of a


high-energy charged particle passing through it. It is filled with gas and
contains hundreds or thousands of very fine wires at high voltage. A
high-energy charged particle passing through the gas will knock out many
atomic electrons. In the electric field formed by the wires, the electrons
will drift to the nearest sense wire. Close to the sense wire, the electric
field will cause an electron avalanche, amplifying the electric signal so
that it can be detected. The position of the charged particle is given by
the location of the sense wires that carried the electric signals. See also
drift chamber.

X
x-rays: High-energy photons (typically from 1 to 100 keV). Can result
from nuclear decay. Used commonly in medical imaging for standard
2-dimensional x-ray imaging and for 3-dimensional CT scans.

x-ray fluorescence: Lower-frequency x-rays given off by samples after being


irradiated with x-rays. The material of the sample can be determined from
the frequencies of the emitted x-rays.

302 Nuclear Physics Explained


SELECT NUCLEAR PHYSICS
RESEARCH FACILITIES

BNL: Brookhaven National Laboratory, Long Island, New York

CERN: French acronym of European Organization for Nuclear Research,


including the LHC, near Geneva, Switzerland

DESY: Deutsches Elektronen-Synchrotron, Hamburg, Germany

Fermilab: Fermi National Accelerator Laboratory, near Chicago, Illinois

FAIR: Facility for Antiproton and Ion Research, Darmstadt, Germany

FRIB [“eff-rib”]: Facility for Rare Isotope Beams, East Lansing, Michigan

ITER: International Thermonuclear Experimental Reactor, Provence, southern


France

JLab: Thomas Jefferson National Accelerator Facility, Newport News,


Virginia

J-PARC: Japan Proton Accelerator Research Complex, Ibaraki, eastern Japan

LHC: Large Hadron Collider, at CERN, near Geneva, Switzerland

MAMI: Mainz Microtron, Mainz, Germany 

PSI: Paul Scherrer Institute, Villigen, northern Switzerland

Select Nuclear Physics Research Facilities 303


RHIC [“rick”]: Relativistic Heavy Ion Collider, at BNL, Long Island,
New York

SLAC: Stanford Linear Accelerator Center, California; later named SLAC


National Accelerator Laboratory, Menlo Park, California

304 Nuclear Physics Explained


TIMELINE

1895 . . . . . . . . . . . . Wilhelm Röntgen discovers x-rays, earning


the first Nobel Prize in Physics in 1901.

1896 . . . . . . . . . . . . Henri Becquerel accidentally discovers


uranium emissions on photographic plates.

1896 . . . . . . . . . . . . J. J. Thomson discovers the electron.

1898 . . . . . . . . . . . . Becquerel’s student Marie Curie coins the word


“radioactive” and discovers thorium (element
90); Marie and Pierre Curie search for other
radioactive elements, discovering polonium
(element 84) and radium (element 88).

1899 . . . . . . . . . . . . Frederick Soddy and Ernest Rutherford discover


and name half-life of radioactive elements.

1903 . . . . . . . . . . . . Marie Curie, PhD, completes thesis, “Research


on Radioactive Substances,” and shares Nobel
Prize with Pierre Curie and Henri Becquerel.

1903 . . . . . . . . . . . . Rutherford names alpha and beta radiation.

1903 . . . . . . . . . . . . William Henry Bragg shows that protons or


alpha particles deposit their peak energy (later
named the Bragg peak) just before they stop.

1904 . . . . . . . . . . . . Doctor/dentist William H. Rollins recommends


shielding to protect patients and doctors
from unnecessary x-ray exposure.

Timeline 305
1908 . . . . . . . . . . . . Rutherford demonstrates that alpha
“rays” are particles of helium nuclei.

1908–1911 . . . . . . . . Rutherford’s lab assistant Hans Geiger develops


a device for counting alpha particles.

1910 . . . . . . . . . . . . Rutherford demonstrates by reflection


of alpha particles that atoms are mostly
empty, with only a tiny “nucleus” that
contains all of the positive charge.

1911 . . . . . . . . . . . . Hungarian physicist Georg Charles von Hevesy


uses trace amounts of radioactive lead to
demonstrate that leftover meat at a hostel had
been reprocessed into 2 different dinner meals.

1913 . . . . . . . . . . . . First mention of “isotopes” in print (Soddy).

1914 . . . . . . . . . . . . First mention of “atomic bombs” (in


H. G. Wells’s novel The World Set Free).

1919 . . . . . . . . . . . . First artificial nuclear reaction, as alpha beam


turns nitrogen into oxygen (Rutherford).

1920 . . . . . . . . . . . . Arthur Eddington proposes nuclear fusion


model of the Sun, based on Francis Aston’s
precise measurement of atomic masses.

1920 . . . . . . . . . . . . Rutherford proposes existence of neutrons to explain


why atomic weight differs from atomic number.

1923 . . . . . . . . . . . . Radioactive tracer proposed (chemist von Hevesy).

1927 . . . . . . . . . . . . Radioactive tracers used to diagnose heart


disease (Dr. Hermann Blumgart).

306 Nuclear Physics Explained


1928 . . . . . . . . . . . . Geiger improves his counter, in collaboration
with student Walther Müller, to detect
many kinds of ionizing radiation.

1929 . . . . . . . . . . . . Ernest O. Lawrence invents the cyclotron to


accelerate nuclear particles to high speeds,
eventually creating hundreds of radioactive
isotopes (1939 Nobel Prize in Physics).

1929 . . . . . . . . . . . . Linear accelerator first developed to


accelerate protons using high voltage (Sir
John Cockcroft and Ernest T. S. Walton).

1931 . . . . . . . . . . . . Harold Urey discovers heavy hydrogen


(deuterium) after constructing a chart
of known and missing isotopes.

1932 . . . . . . . . . . . . Heavy water deliberately concentrated using


electrolysis by Urey and Edward Washburn.

1932 . . . . . . . . . . . . Existence of the neutron confirmed when


Rutherford student James Chadwick
aims alpha particles at beryllium; Walton
uses protons to split lithium atom.

1933 . . . . . . . . . . . . Rutherford declares, “Anyone who says that with the


means at present at our disposal and with our present
knowledge we can utilize atomic energy is talking
moonshine.” (New York Times, September 12)

1933 . . . . . . . . . . . . Discovery of large magnetic moment for the proton


by Otto Stern (1943 Nobel Prize in Physics) suggests
that proton must have an internal structure.

Timeline 307
1934 . . . . . . . . . . . . Enrico Fermi discovers neutron-induced
radioactivity by bombarding stable nuclei
with neutrons (Nobel Prize 1938).

1934 . . . . . . . . . . . . Tritium created when deuterium hit


with deuterium nuclei by Rutherford,
Mark Oliphant, and Paul Harteck.

1934 . . . . . . . . . . . . Pavel Cherenkov first observes particles emitting


blue radiation when traveling faster than the
speed of light in water (Nobel Prize 1958).

1934 . . . . . . . . . . . . Irène and Frédéric Joliot-Curie hit nuclei with


alpha particles to create new radioactive isotopes.

1935 . . . . . . . . . . . . Yukawa theory (named after Hideki Yukawa) of


the strong nuclear force describes it as the result
of nucleons exchanging mesons between them.

1938 . . . . . . . . . . . . First attempted neutron therapy for cancer, using the


Lawrence cyclotron by brother John Lawrence, but
long-term side effects proved worse than for x-rays.

1938 . . . . . . . . . . . . Lise Meitner and Otto Hahn discover and


name nuclear “fission” after Hahn and Fritz
Strassman use neutrons to split uranium and
identify barium in the fission products.

1939 . . . . . . . . . . . . Liquid-drop model of the nucleus as “surface


tension” plus electrical repulsion proposed to
explain fission and describe the curve of binding
energy (Niels Bohr and John Wheeler).

308 Nuclear Physics Explained


1939 . . . . . . . . . . . . Hans Bethe calculates a proton-proton
chain reaction that describes nuclear
fusion in stars (Nobel Prize 1967).

1942 . . . . . . . . . . . . Manhattan Project team led by Fermi


achieves first self-sustaining nuclear chain
reaction (December 2, University of Chicago),
using uranium and ultrapure graphite.

1942–1943 . . . . . . . . Norwegian saboteurs and American bombers


undermine Nazi collection of heavy water, intended
as a uranium moderator for an atomic bomb.

1945 . . . . . . . . . . . . Atomic bombs dropped on Japan; data


from 120,000 survivors exposed to gamma
radiation became a common baseline for
assessing risks of other radiation exposure.

1946 . . . . . . . . . . . . First proposal for proton therapy by Lawrence


student Robert R. Wilson, who went on
to design cyclotron at Harvard for proton
radiosurgery and radiation therapy.

1947–1975 . . . . . . . . US Atomic Energy Commission assumes


civilian control over nuclear energy.

1947 . . . . . . . . . . . . “Nuclide” proposed by Truman P. Kohman as a


more general term than isotope to indicate a nucleus
with a specific number of protons and neutrons.

1949 . . . . . . . . . . . . First Soviet nuclear weapon tested.

Timeline 309
1949–1950 . . . . . . . . J. Hans D. Jensen and Maria Goeppert Mayer
independently develop shell model of the
nucleus to explain the stability of “magic”
numbers of nucleons (Nobel Prize in 1963).

1950 . . . . . . . . . . . . First true radioisotope imaging system,


combines a Geiger counter with crystal
components from photomultiplier tube (Dr.
Benedict Cassen’s “scintiscanner”).

1951 . . . . . . . . . . . . Electricity first generated from nuclear


energy and demonstrates that a breeder
reactor could produce more energy than
it consumed (EBR-1 at Arco, Idaho).

1952 . . . . . . . . . . . . France tests its first nuclear weapon.

1952 . . . . . . . . . . . . The United Kingdom tests its first nuclear weapon.

1952 . . . . . . . . . . . . First explosive to incorporate nuclear fusion, Ivy


Mike (November 1, Enewetak atoll, Marshall
Islands), leading to the first synthesis of elements
99 and 100 (einsteinium and fermium).

1953 . . . . . . . . . . . . “Atoms for Peace” speech at the United Nations


by President Dwight D. Eisenhower kicks off
multiyear US program to support peaceful
uses of nuclear energy around the world.

1953–1954 . . . . . . . . First nuclear-powered submarine, USS Nautilus.

1954 . . . . . . . . . . . . First full-scale nuclear power plant


in the USSR (Obrinsk).

310 Nuclear Physics Explained


1957 . . . . . . . . . . . . First full-scale nuclear power plant in the
United States (Shippingport, PA).

1957 . . . . . . . . . . . . Passage of the Price-Anderson Act in United


States limits liability of nuclear plant utilities,
vendors, and suppliers in case of accident.

1957 . . . . . . . . . . . . First instance of proton therapy (Sweden).

1959–1989 . . . . . . . . World’s first nuclear-powered surface vessel, the


icebreaker Lenin (20,000 deadweight tons).

1958 . . . . . . . . . . . . Saveli Feinberg proposes idea for a


“breed-and-burn” reactor.

1958–1961 . . . . . . . . Gamma camera (scintigraphy)


invented by Hal Oscar Anger.

1960 . . . . . . . . . . . . First fully commercial pressurized-water reactor


(Yankee Rowe, MA) and first commercial
boiling-water reactor (Dresden, IL).

1961 . . . . . . . . . . . . First US instance of proton therapy (Harvard).

1961 . . . . . . . . . . . . Nobel Prize to Robert Hofstadter for the first


measurement of the size of the proton.

1964 . . . . . . . . . . . . The People’s Republic of China


tests its first nuclear weapon.

1965 . . . . . . . . . . . . First nuclear reactor in space


(43 days, 500 watts of power).

1967 . . . . . . . . . . . . Gamma “knife” use of gamma rays in


surgery pioneered by Lars Leksell.

Timeline 311
1968 . . . . . . . . . . . . Wire chamber (“multiwire proportional
chamber”) particle detector invented by Georges
Charpak (1992 Nobel Prize), making possible
the study of very rare nuclear reactions.

1968–1969 . . . . . . . . Evidence for quarks first observed from


deep inelastic scattering of electrons from
the proton, Stanford Linear Accelerator
(1990 Nobel Prize to Jerome Friedman,
Henry Kendall, and Richard Taylor).

1970 . . . . . . . . . . . . Treaty on the Non-Proliferation of Nuclear Weapons


enacted to limit the spread of nuclear weapons.

1971 . . . . . . . . . . . . Magnetic resonance imaging (MRI) shows


radio signal for hydrogen is more pronounced
from tumors than for healthy tissue, due to
extra water in tumors (Raymond Damadian).

1972–1973 . . . . . . . . Computed tomography/computerized axial


tomography (CT/CAT) of magnetic resonance
information by Godfrey Hounsfield and Allan
Cormack becomes the basis for computer-assisted
planning of radiation treatment.

1973 . . . . . . . . . . . . Arab oil embargo leads to orders for 41 US


nuclear power plants in a single year.

1974 . . . . . . . . . . . . French government undertakes to replace all


oil-generated electricity with nuclear power.

1974 . . . . . . . . . . . . India becomes the sixth country to make a public


test of a nuclear weapon (12 kilotons), increasing
concerns about proliferation of nuclear weapons.

312 Nuclear Physics Explained


1975 . . . . . . . . . . . . US Nuclear Regulatory Commission
established to regulate nuclear power.

1977 . . . . . . . . . . . . First medical MRI of a human being


(Damadian) and first prototype for rapid MRI
via echo-planar imaging (Peter Mansfield).

1977–1981 . . . . . . . . United States entirely bans commercial


reprocessing of spent nuclear fuel in an
effort to hinder weapons proliferation.

1979 . . . . . . . . . . . . Three Mile Island nuclear reactor


partial meltdown (March 28).

1983 . . . . . . . . . . . . Finland declares that utility companies


are responsible for disposal of waste
from their nuclear power plants.

1985 . . . . . . . . . . . . Soviet leader Mikhail Gorbachev meeting


with President Ronald Reagan proposes
international collaboration on nuclear fusion,
leading to beginnings of the ITER tokamak.

1986 . . . . . . . . . . . . First successful passive shutdown demonstration


for a nuclear reactor (the sodium-cooled ERB-II
in Idaho) without control system or a backup.

1986 . . . . . . . . . . . . Chernobyl nuclear reactor accident (April 26), the


most deadly and expensive ever, sends a large plume
of radiation from Ukraine through northern Europe.

1987 . . . . . . . . . . . . Voters in Italy (65% turnout) place restrictions


on nuclear power via first national referendum;
all 4 Italian nuclear power plants close by 1990.

Timeline 313
1987 . . . . . . . . . . . . US Congress votes to declare Yucca Mountain,
Nevada, the sole site to be studied for a future
government-run repository for US nuclear
waste, despite the lack of nuclear power plants
in Nevada; by contrast, in Finland the same
year, 5 sites are selected for further study that
are close to power plants run by the 2 utilities
responsible for their own nuclear waste disposal.

1989 . . . . . . . . . . . . Claims about achieving nuclear fusion at


room temperature (“cold fusion”) in an
electrochemical cell by Stanley Pons and
Martin Fleischmann spark an unsuccessful
worldwide attempt to replicate their results.

1991 . . . . . . . . . . . . First gantry-based proton therapy


(Loma Linda, CA).

1993 . . . . . . . . . . . . Functional MRI (fMRI) introduced.

1993–2013 . . . . . . . Megatons to Megawatts Program converts


more than 20,000 Soviet-era ICBM
warheads into low-enriched uranium, which
is sold to US nuclear power plants.

1995 . . . . . . . . . . . . Jefferson Lab/CEBAF (Continuous Electron


Beam Accelerator Facility) begins experiments
with a 4-billion-electron-volt beam.

1996 . . . . . . . . . . . . Comprehensive Nuclear-Test-Ban Treaty


creates system of seismic and radionuclide
monitoring stations to distinguish nuclear
explosions from earthquakes.

314 Nuclear Physics Explained


1997 . . . . . . . . . . . . Joint European Torus (JET) tokamak fusion
reactor achieves 16 megawatts of gross power.

1998 . . . . . . . . . . . . Pakistan becomes the seventh country to


successfully test a nuclear weapon.

2000 . . . . . . . . . . . . The Relativistic Heavy Ion Collider


(RHIC) begins operation at Brookhaven
National Laboratory in New York.

2003 . . . . . . . . . . . . Northeast blackout in the United States and


Ontario caused 17 nuclear reactors to enter
“safe mode” shutdown (August 14).

2004 . . . . . . . . . . . . Construction begins on world’s first


long-term repository for high-level
radioactive waste (Olkiluoto, Finland).

2005 . . . . . . . . . . . . Lithuania begins shutting down its


Chernobyl-style reactors.

2005 . . . . . . . . . . . . US Energy Policy Act creates incentives to support


construction of new nuclear reactors; a few states
later allow utility companies to charge customers
for the construction of nuclear power plants.

2006 . . . . . . . . . . . . North Korea conducts its first nuclear


test, 4 years after exiting the Treaty on the
Non-Proliferation of Nuclear Weapons, with
subsequent tests in 2009, 2013, 2016, and 2017.

Timeline 315
2011 . . . . . . . . . . . . Fukushima core meltdown disaster due
to earthquake and tsunami that cause
failure of cooling systems for 4 nuclear
reactors. All of Japan’s 54 nuclear reactors
shut down over the following year.

2011 . . . . . . . . . . . . German government shuts down 8 reactors


built before 1985 and promises to close all
remaining reactors over the next decade.

2015 . . . . . . . . . . . . The first of Japan’s 43 operable nuclear reactors


is turned back on (Sendai, Kagoshima).

2016 . . . . . . . . . . . . World’s largest stellarator-type fusion reactor


(Wendelstein 7-X) begins operation to demonstrate
possibility of nonpulsed, continuous fusion.

2017 . . . . . . . . . . . . Jefferson Lab/CEBAF (Continuous Electron Beam


Accelerator Facility) completes upgrade from
6 billion electron volts to 12 billion electron volts.

2017 . . . . . . . . . . . . Westinghouse, the world’s oldest and largest supplier


of commercial reactor technology (owned by
Toshiba since 2006) files for bankruptcy in March.

2017 . . . . . . . . . . . . NuScale Power (Oregon) submits first US design


application for a small modular reactor.

316 Nuclear Physics Explained


NOBEL LAUREATES
Honored for Their Work in Nuclear Physics

1901 . . . . . . . . . . . . Wilhelm Röntgen for discovery of x-rays;


the first Nobel Prize in Physics

1903 . . . . . . . . . . . . (shared) Henri Becquerel for discovering


spontaneous radioactivity, and Pierre and Marie
Curie for their research into radiation (Physics)

1911 . . . . . . . . . . . . Marie Curie for discovering radium


and polonium (Chemistry)

1917 . . . . . . . . . . . . Charles Barkla for his discovery of the characteristic


radiation from different elements (Physics)

1921 . . . . . . . . . . . . Frederick Soddy for his investigations into the


origin and nature of isotopes (Chemistry)

1922 . . . . . . . . . . . . Francis Aston for discovering using a mass


spectrometer of isotopes in a large number
of nonradioactive elements (Chemistry)

1924 . . . . . . . . . . . . Karl Siegbahn for his discoveries in the


field of x-ray spectroscopy (Physics)

1934 . . . . . . . . . . . . Harold Urey for discovering heavy


hydrogen (Chemistry)

1935 . . . . . . . . . . . . James Chadwick for the discovery


of the neutron (Physics)

Nobel Laureates 317


1935 . . . . . . . . . . . . Frédéric and Irène Joliot-Curie for their synthesis
of new radioactive elements (Chemistry)

1936 . . . . . . . . . . . . (shared) Carl Anderson for the discovery


of the proton and Victor Hess for the
discovery of cosmic radiation (Physics)

1938 . . . . . . . . . . . . Enrico Fermi for his discovery of new radioactive


elements produced by neutron irradiation and for his
work on slow neutron nuclear reactions (Physics)

1939 . . . . . . . . . . . . Ernest O. Lawrence for inventing


the cyclotron (Physics)

1943 . . . . . . . . . . . . Otto Stern for discovering the magnetic


moment of the proton (Physics)

1943 . . . . . . . . . . . . Georg Charles von Hevesy for using


isotopes as tracers in the study of
chemical processes (Chemistry)

1944 . . . . . . . . . . . . Otto Hahn for discovering nuclear


fission (Chemistry)

1944 . . . . . . . . . . . . I. I. Rabi for his resonance method of recording


the magnetic properties of atomic nuclei (Physics)

1948 . . . . . . . . . . . . P. M. S. Blackett for his development


of the cloud chamber (Physics)

1949 . . . . . . . . . . . . Hideki Yukawa for his prediction of the


existence of mesons based on his theoretical
work on nuclear forces (Physics)

318 Nuclear Physics Explained


1950 . . . . . . . . . . . . Cecil Powell for his development of the photographic
method of studying nuclear reactions (Physics)

1951 . . . . . . . . . . . . Sir John Cockcroft and Ernest T. S. Walton for their


pioneer work on the transmutation of atomic nuclei
by artificially accelerated atomic particles (Physics)

1951 . . . . . . . . . . . . Edwin McMillan and Glenn Seaborg for


discoveries of transuranic elements (Chemistry)

1952 . . . . . . . . . . . . Felix Bloch and Edward Purcell “for their


development of new methods for nuclear
magnetic precision measurements and discoveries
in connection therewith” (Physics)

1958 . . . . . . . . . . . . Pavel Cherenkov, Il’ja Frank, and Igor Tamm


for their discovery that particles traveling
faster than the speed of light in water emit
electromagnetic radiation (Physics)

1960 . . . . . . . . . . . . Donald Glaser for his invention of


the bubble chamber (Physics)

1960 . . . . . . . . . . . . Willard Libby for his discovery and


use of carbon dating (Chemistry)

1961 . . . . . . . . . . . . (shared) Robert Hofstadter for using electron


scattering to study the structure of the proton and
Rudolf Mossbauer for his discovery of coherent
resonant absorption of gamma rays (Physics)

1963 . . . . . . . . . . . . (shared) J. Hans D. Jensen and Maria


Goeppert Mayer for developing the shell
model of the nucleus to explain the stability
of “magic” numbers of nucleons

Nobel Laureates 319


1967 . . . . . . . . . . . . Hans Bethe “for his contributions to the theory
of nuclear reactions, especially his discoveries
concerning the energy production in stars” (Physics)

1975 . . . . . . . . . . . . Aage Bohr, Ben Mottelson, and Leo


Rainwater for their discoveries in the
structure of the nucleus (Physics)

1979 . . . . . . . . . . . . Allan Cormack and Godfrey Hounsfield for the


development of computer assisted tomography
(CT scans) (Medicine or Physiology)

1990 . . . . . . . . . . . . Jerome Friedman, Henry Kendall, and


Richard Taylor for deep inelastic scattering of
electrons on proton and bound neutrons and
for their discovery of the quark (Physics)

1992 . . . . . . . . . . . . Georges Charpak for wire chamber


(“multiwire proportional chamber”) particle
detector invented, making possible study
of very rare nuclear reactions (Physics)

1995 . . . . . . . . . . . . (shared) Frederick Reines for the


detection of the neutrino (Physics)

2002 . . . . . . . . . . . . (shared) Raymond Davis and Masatoshi Koshiba


for the detection of cosmic neutrinos (Physics)

2003 . . . . . . . . . . . . Paul C. Lauterbur and Peter Mansfield for MRI


as a diagnostic tool (Physiology or Medicine)

320 Nuclear Physics Explained


BIBLIOGRAPHY

The following is a list of the best overall books to read, after which the
alphabetized bibliography begins:

Mackintosh, Ray, Jim Al-Khalili, Bjorn Jonson, and Teresa Pena. Nucleus: A Trip
into the Heart of Matter. 2nd ed. Baltimore, MD: The Johns Hopkins University
Press, 2011. A well-written, nonmathematical, brief overview of nuclear physics.
Lots of pictures and no equations. A good first book to read.

Lilley, John. Nuclear Physics: Principles and Applications. Chichester, England:


John Wiley and Sons, 2001. A well-written, comprehensive textbook with
some emphasis on applications aimed at undergraduate physics majors.
Much more comprehensive, in depth, and mathematically sophisticated than
Mackintosh’s Nucleus.

National Research Council. Nuclear Physics: Exploring the Heart of Matter.


Washington, DC: The National Academies Press, 2013. Available at https://
www.nap.edu/catalog/13438/nuclear-physics-exploring-the-heart-of-matter. An
up-to-date view of the scientific rationale and objectives of nuclear physics, with
an emphasis on current research and future directions. Written for policy makers
and other nonscientists. Nonmathematical. Complements the textbooks.

Rhodes, Richard. The Making of the Atomic Bomb. New York: Simon and
Schuster, 1986. Superbly readable account of the Manhattan Project and the
nuclear physics leading up to it. Complements the textbooks.

Jorgenson, Timothy J. Strange Glow: The Story of Radiation. Princeton, NJ:


Princeton University Press, 2016. Superbly readable account telling the history
of our knowledge of radiation with delightful stories and anecdotes about the
scientists who made the discoveries. Complements the textbooks.

Bibliography 321
Mahaffey, James. Atomic Awakening. New York: Pegasus Books, 2009. Up-to-
date and thorough discussion of the history and future of nuclear power. Well
written and sprinkled with humor (only partially macabre). Excellent inside look
at what really happened by a nuclear engineer. Excellent companion to lectures
15–20.

Agency for Toxic Substances and Disease Registry. “Radon Toxicity.” https://
www.atsdr.cdc.gov/csem/csem.asp?csem=8&po=0. Continuing Medical
Education module on the sources and dangers of radon.

Akpan, Nsikan. “This Supernova Blast Was So Close, It Littered the Ocean
Floor with Radioactive Dust.” PBS NewsHour, Science Wednesday, April 6,
2016. http://www.pbs.org/newshour/updates/supernova-radioactive-blast-litter-
ocean-dust/. Discusses finding radioactive Iron-60 in ocean cores and what it
means about the history of nearby supernovas about 13 million years ago.

American Association of Variable Star Observers. “Supernova 1987A.” https://


www.aavso.org/vsots_sn1987a. Good description of the discovery and
subsequent behavior of supernova 1987A and supernovas in general. Not much
detail on the neutrino detection.

Arnold, Carrie. “Cold War Bomb Testing Is Solving Biology’s Biggest Mysteries.”
NOVA Next, December 11, 2013. http://www.pbs.org/wgbh/nova/next/body/
bomb-pulse/. Discussion of the uses of bomb-pulse carbon dating for brain
research, forensic analysis of deaths, etc.

“Asgard’s Fire.” The Economist, April 12, 2014. https://www.economist.com/


news/science-and-technology/21600656-thorium-element-named-after-
norse-god-thunder-may-soon-contribute. Discussion of the prospects for
thorium reactors.

322 Nuclear Physics Explained


Bahcall, John N., and Raymond David Jr. “Solar Neutrinos: A Scientific Puzzle.”
Science 191 (1976): 264–267. Good description of the solar neutrino problem as
of 1976.

Baras, Colin. “Exploding Stars Could Have Kick-Started Our Ancestor’s


Evolution.” New Scientist, October 9, 2017. https://www.newscientist.
com/article/2149763-exploding-stars-could-have-kick-started-our-
ancestors-evolution/?cmpid=NLC|NSNS|2017-1210-GLOBAL&utm_
medium=NLC&utm_source=NSNS. Describes the hypothesis that cosmic rays
from a “nearby” star could have caused lightning that sparked forest fires that
transformed east African forest to savannah.

Blitz, Matt. “When Kodak Accidentally Discovered A-Bomb Testing.” Popular


Mechanics, June 20, 2016. http://www.popularmechanics.com/science/energy/
a21382/how-kodak-accidentally-discovered-radioactive-fallout/. The first
(inadvertent) remote monitoring of a nuclear test.

Bodansky, David. Nuclear Energy: Principles, Practices, and Prospects. 2nd ed.
New York: Springer and AIP Press, 2004. Superb coverage of radiation, nuclear
reactions, and nuclear energy. Some equations, most can be glossed over.

Bolus, Norman. “NCRP Report 160 and What It Means for Medical Imaging
and Nuclear Medicine.” Journal of Nuclear Medicine Technology 41, no. 4 (2013):
255–261. A more accessible summary of NCRP Report 160.

Borel, Brooke. “Making New Elements.” Popular Science, May 2013. https://
www.popsci.com/science/article/2013-04/making-new-elements. Readable
description of how new elements are made.

Bushberg, Jerrold T. “Radiation Exposure and Contamination.” Merck Manual


Professional Version. http://www.merckmanuals.com/professional/injuries-
poisoning/radiation-exposure-and-contamination/radiation-exposure-and-
contamination. Thorough discussion of the sources and effects of radiation
contamination. The authoritative medical source.

Bibliography 323
Capellaro, Paola. “Course 22.02: Introduction to Applied Nuclear Physics.”
MIT Open Course Ware course notes. https://ocw.mit.edu/courses/nuclear-
engineering/22-02-introduction-to-applied-nuclear-physics-spring-2012/. Aimed
at MIT undergraduates, so somewhat more advanced than this course. Very clear
notes. Excellent figures. The first chapter is very accessible.

Christian, Paul, and Kristen Waterstram-Rich, eds. Nuclear Medicine and


PET/CT. Maryland Heights, MO: Mosby Elsevier, 2007. One of the standard
textbooks for nuclear medicine technicians. More detail than you probably want.
Excellent pictures.

Clery, Daniel. “Fusion’s Restless Pioneers.” Science 345 (July 25, 2014): 370–375.
Overview of the privately funded fusion efforts.

Collins, John. “Parton Distribution Functions.” Scholarpedia 7, no. 7


(2012): 10851. http://www.scholarpedia.org/article/Parton_distribution_
functions_%28definition%29. Describes the mathematical formalism behind
the parton (quark and gluon) distribution functions. Intended for scientists, but
available to everyone.

Collon, Phillippe. “Using Nuclear Techniques to Analyze Art.” Presentation


at the Linda Hall Library. https://vimeo.com/66107935. An excellent 1-hour
lecture on using selected nuclear physics techniques to help attribute art works,
determine their composition, and even detect forgeries.

Cowan, George A. “A Natural Fission Reactor.” Scientific American 235, no. 1


(July 1976): 36–47. Best-known article about the Oklo natural fission reactor.

Cullings, Harry M. “Impact on the Japanese Atomic Bomb Survivors of


Radiation Received from the Bombs.” Health Physics 106 (2014): 281–293.
Summarizes the final results of the Life Span Study. Authoritative.

324 Nuclear Physics Explained


Deutch, John, et al. The Future of Nuclear Power. Cambridge: Massachusetts
Institute of Technology, 2003. Available at http://web.mit.edu/nuclearpower/.
Addresses the interrelated set of technical, economic, environmental, and
political issues related to deploying new nuclear facilities.

Deutch, John, et al. Update of the MIT 2003 Future of Nuclear Power.
Cambridge: Massachusetts Institute of Technology, 2009. Available at http://
web.mit.edu/nuclearpower/pdf/nuclearpower-update2009.pdf. An MIT study on
the future of nuclear power.

Drake, Nadia. “In a First, Gravitational Waves Linked to Neutron Star Crash.”
National Geographic, October 16, 2017. https://news.nationalgeographic.
com/2017/10/gravitational-waves-discovered-neutron-stars-pictures-science/?no-
cache. Description of the first detected neutron-star collision and of the evidence
for the copious production of gold and other heavy elements in the collision.

Dworschak, Manfred. “Is Radiation as Bad as We Thought?” Spiegel Online,


April 26, 2016. http://www.spiegel.de/international/world/chernobyl-hints-
radiation-may-be-less-dangerous-than-thought-a-1088744.html. Newspaper
discussion of the dangers of radiation on the 30th anniversary of Chernobyl.

Dyson, George. Project Orion: The True Story of the Atomic Spaceship. New York:
Henry Holt and Company, 2002. A detailed description of US efforts to use
nuclear bombs to propel a spaceship.

Ellis, John, and David Schram. “Could a Nearby Supernova Explosion Have
Caused a Mass Extinction?” CERN-TH.6805/93, 1993. https://arxiv.org/abs/
hep-ph/9303206. Examines the probability of nearby supernovas and their
effect on life on Earth, typically through destruction of the ozone layer and the
resulting excess of ultraviolet radiation reaching the surface of the Earth.

Engelmann, W. “Radium Emanation Therapy.” Lancet 1 (May 3, 1913):


1225–1258. The first paper describing radiation therapy.

Bibliography 325
Feltesse, Joel. “Introduction to Parton Distribution Functions.” Scholarpedia 5,
no. 11 (2012): 10160. http://www.scholarpedia.org/article/Introduction_to_
Parton_Distribution_Functions. Accessible description of parton distribution
functions and how they are measured. Intended for scientists, but available
to everyone.

Fields, Brian. “When Stars Attack! In Search of Killer Supernova Explosions.”


Public lecture presented at University of Kansas, April 18, 2016. https://
www.youtube.com/watch?v=ND6u4Rnq0g0. An entertaining lecture by a
top scientist.

Fountain, Henry. “On Nuclear Waste, Finland Shows U.S. How It Can Be
Done.” New York Times, June 9, 2017. https://www.nytimes.com/2017/06/09/
science/nuclear-reactor-waste-finland.html?mcubz=1. A news report on the
progress of the Finnish waste repository.

FRIB Users Organization. FRIB: Opening New Frontiers in Nuclear Science.


August 2012. https://www.frib.msu.edu/_files/pdfs/frib_opening_new_
frontiers_in_nuclear_science.pdf. Describes the physics motivation for building
the Facility for Rare Isotope Beams (FRIB). Very accessible. Aimed at readers in
the US Department of Energy and the US Congress.

Friedman, Jerome. “Deep Inelastic Scattering: Comparisons with the Quark


Model.” 1990 Nobel Prize Lecture. Reviews of Modern Physics 63 (1991): 615.
https://www.nobelprize.org/nobel_prizes/physics/laureates/1990/friedman-
lecture.html. Describes the experiments performed to receive the Nobel Prize.
Provides fascinating historical and technical detail.

“Fusion.” Nobelprize.org. Nobel Media AB 2014. http://www.nobelprize.org/


nobel_prizes/themes/physics/fusion/sun_1.html. Nontechnical discussion of the
age of the Sun.

326 Nuclear Physics Explained


Geiger, Hans, and Ernest Marsden. “On a Diffuse Reflection of the α-Particles.”
Proceedings of the Royal Society 82 (A): 495–500. Archived at https://web.archive.
org/web/20080102232956/http://dbhs.wvusd.k12.ca.us/webdocs/Chem-
History/GM-1909.html. The classic paper describing the Rutherford scattering
experiment. Written clearly.

Glascock, Michael. “Overview of Neutron Activation Analysis.” Archaeometry


Laboratory, University of Missouri Research Reactor. http://archaeometry.
missouri.edu/naa_overview.html. Excellent detailed description of neutron
activation analysis.

GlobalSecurity.org. “Uranium Enrichment Techniques.” http://www.


globalsecurity.org/wmd/intro/u-enrichment.htm. A thorough description of all
the different uranium enrichment techniques. Five pages per month are free.

Gould, Todd A., and Molly Edmonds. “How MRI Works.” http://science.
howstuffworks.com/mri.htm. Excellent article on how MRI works.

Grupen, Claus. Particle Detectors. Cambridge: Cambridge University Press,


2008. A description of modern particle detectors and spectrometers, aimed at
working scientists.

Hamilton, Tyler. “A New Approach to Fusion.” Technology Review, July 31,


2009. https://www.technologyreview.com/s/414559/a-new-approach-to-fusion/.
Describes the magnetized target fusion research of General Fusion.

Health Physics Society. “Radiation Exposure during Commercial Airline


Flights.” http://hps.org/publicinformation/ate/faqs/commercialflights.html.
Summarizes the published research on radiation doses received by people who
fly, including passengers and crew.

Henley, Ernest, and Alejandro Garcia. Subatomic Physics. 3rd ed. Singapore:
World Scientific, 2007. An advanced undergraduate textbook with excellent
descriptions of many phenomena. One of the standard texts. Very accessible in
places, even for the nonmathematical reader.

Bibliography 327
Hewitt, Paul. Conceptual Physics. 12th ed. New York: Pearson Press, 2014. An
excellent conceptual-level discussion of radiation, radioactivity, health effects,
fission, and fusion (chapters 33 and 34). Older editions are equally good and
much less expensive.

International Nuclear Safety Advisory Group. INSAG-7—The Chernobyl


Accident: Updating of INSAG-1. Vienna: International Atomic Energy Agency,
1992. http://www-pub.iaea.org/MTCD/publications/PDF/Pub913e_web.pdf.
The official international investigation report of the Chernobyl disaster.

Jefferson Lab. “Jefferson Lab Virtual Tour.” https://www.jlab.org/virtual-tour.

Kendal, Henry. “Deep Inelastic Scattering: Experiments on the Proton and the
Observation of Scaling.” 1990 Nobel Prize Lecture. Reviews of Modern Physics
63 (1991): 597. https://www.nobelprize.org/nobel_prizes/physics/laureates/1990/
kendall-lecture.html. Describes the experiments performed to receive the Nobel
Prize. Provides fascinating historical and technical detail.

Krane, Kenneth. Introductory Nuclear Physics. 3rd ed. New York: John Wiley and
Sons, 1987. One of the standard nuclear physics textbooks used in undergraduate
courses for physics majors. Comprehensive, but beginning to be dated. Assumes
a working knowledge of calculus.

LeBlanc, Francis. An Introduction to Stellar Astrophysics. Chichester, England:


John Wiley & Sons, 2010. Excellent description of stellar astrophysics at the
sophomore/junior level. The chapter on nucleosynthesis and stellar evolution is
very accessible and thorough.

Leo, W. R. Techniques for Nuclear and Particle Physics Experiments. Berlin:


Springer-Verlag, 1994. Thorough introduction to all the different kinds
of particle detectors and how they work. Intended for people who will use
the detectors.

328 Nuclear Physics Explained


Mackova, Anna, et al., eds. Nuclear Physics for Cultural Heritage. Nuclear Physics
Division of the European Physical Society. http://www.edp-open.org/images/
stories/books/fulldl/Nuclear-physics-for-cultural-heritage.pdf. A thorough
overview of the different nuclear physics techniques used to characterize
objects in the fields of history, art history, and archaeology. Also includes
many examples.

Mahaffey, James. Atomic Accidents. New York: Pegasus Books, 2014. Up-to-date
and thorough discussion of the causes and results of nuclear meltdowns and
disasters, from the first misadventures with radiation to Fukushima. Well written
and sprinkled with humor (only partially macabre). Excellent inside look at what
really happened by a nuclear engineer.

Markandya, A., and P. Wilkinson. “Electricity Generation and Health.” The


Lancet 370 (2007): 979–990. doi.org/10.1016/S0140-6736(07)61253-7. An
evaluation of the health risks of different energy-generation technologies.
(Caveat: Estimating the number of deaths from small doses spread over large
populations is always extremely uncertain.)

MIT Center for Advanced Nuclear Energy Systems Symposium. Nuclear beyond
LWRs: A Celebration of Neil Todreas’ Career and Passion for Advanced Reactors.
Cambridge, MA, November 1–2, 2016. https://energy.mit.edu/wp-content/
uploads/2016/11/CANES-Nuclear-Beyond-LWRs-2016.pdf. A collection of talks
on different advanced nuclear energy topics. Represents the state of the art as of
late 2016.

Morrison, Phillip. “Where Fiction Became Ancient Fact.” Scientific


American 278, no. 6 (June 1, 1998): 99–101. Good description of the Oklo
natural fission reactor.

Moskowitz, Clara. “Ancient Roman Metal Used for Physics Experiments


Ignites Scientific Feud.” Scientific American 18 (December 2013). https://
www.scientificamerican.com/article/ancient-roman-lead-physics-archaeology-
controversy/. Description of why physicists prefer Roman-era lead to today’s lead.

Bibliography 329
National Council on Radiation Protection and Measurements. Report No. 160:
Ionizing Radiation Exposure of the Population of the United States. 2009. Detailed
information on the radiation exposure of the US population.

National Research Council. Health Risks from Exposure to Low Levels of Ionizing
Radiation: BEIR VII Phase 2. Washington, DC: The National Academies Press,
2006. Available at https://doi.org/10.17226/11340. The definitive National
Academy of Sciences study of the effects of low levels of ionizing radiation.

Nutall, W. J. “Fusion as an Energy Source: Challenges and Opportunities.”


Institute of Physics Report, September 2008. http://iop.org/publications/
iop/2008/file_38224.pdf. Very accessible, thorough discussion of the prospects
for fusion power.

——— . Nuclear Renaissance: Technologies and Policies for the Future of Nuclear
Power. Bristol, UK: Institute of Physics Publishing, 2005. Good description of
modern nuclear technologies, along with practical and regulatory concerns.

Oak Ridge Associated Universities. “Shoe-Fitting Fluoroscope.” http://www.


orau.org/PTP/collection/shoefittingfluor/shoe.htm. Description of the nature
and history of shoe fluoroscopy, including many pictures of fluoroscopes and
related items.

Park, Robert. Voodoo Science: The Road from Foolishness to Fraud. Oxford:
Oxford University Press, 2000. An entertaining and well-researched exploration
of scientific error, bad science, and pseudoscience, focusing on cold fusion.

PhET Interactive Simulations. “Nuclear Fission.” University of Colorado


Boulder. https://phet.colorado.edu/en/simulation/nuclear-fission. Great
simulation of how one uranium-235 nucleus fissions, how a chain reaction
develops among many uranium-235 and uranium-238 nuclei, and how a
chain reaction proceeds in a reactor. This was developed by one of the top
physics education research groups in the country, founded by Nobel Laureate
Carl Weiman.

330 Nuclear Physics Explained


Ritchie, Hannah. “It Goes Completely against What Most Believe, But out of
All Major Energy Sources, Nuclear Is the Safest.” Our World in Data (blog).
July 24, 2017. https://ourworldindata.org/what-is-the-safest-form-of-energy/. A
nontechnical description of electricity generation and health using data from
Markandya and Wilkinson’s Lancet article.

Rosen, S. Peter. “Ray Davis: Indefatigable Neutrino Pioneer.” CERN Courier,


September 7, 2006. http://cerncourier.com/cws/article/cern/29699. Obituary of
Ray Davis, the founder of neutrino astrophysics and the first person to measure
neutrinos from the Sun.

Rutherford, Ernest. “The Scattering of α and β Particles by Matter and the


Structure of the Atom.” Philosophical Magazine 6, vol. 21 (May 1911). Archived
at http://www.lawebdefisica.com/arts/structureatom.pdf. The classic paper
interpreting the results of Geiger and Marsden’s measurements in terms of the
nuclear model.

Taylor, Richard E. “Deep Inelastic Scattering: The Early Years.” 1990 Nobel
Prize Lecture. Reviews of Modern Physics 63 (1991): 573. https://www.nobelprize.
org/nobel_prizes/physics/laureates/1990/taylor-lecture.html. Describes the
experiments performed to receive the Nobel Prize. Provides fascinating historical
and technical detail.

Temple, James. “Small Reactors Could Kick-Start the Stalled Nuclear Sector.”
Technology Review, July 17, 2017. https://www.technologyreview.com/s/608271/
small-reactors-could-kick-start-the-stalled-nuclear-sector/. Discussion of the
prospects for small modular fission reactors.

The KamLAND Collaboration. “Partial Radiogenic Heat Model for Earth


Revealed by Geoneutrino Measurements.” Nature Geoscience 4 (2011): 647–651.
Technical article describing the measurement of geoneutrinos and its relation
to the energy released in the Earth by radioactive decay of uranium-238 and
thorium-232.

Bibliography 331
“The Oklo Reactor: A Nuclear Detective Story.” re-actions 9, September 1993.
http://www.ans.org/pi/np/oklo/docs/reactions.pdf. Description of how scientists
figured out what happened at the Oklo reactor.

Thomson, Sir William (Lord Kelvin). “On the Age of the Sun’s Heat.”
Macmillan’s Magazine 5 (March 5, 1862): 388–393. From reprint in Popular
Lectures and Addresses, 2nd ed., 1: 356­–376. Available at http://zapatopi.net/
kelvin/papers/on_the_age_of_the_suns_heat.html. Kelvin’s original article on
the age of the Sun. A fascinating look into the thought processes of one of the
19th century’s greatest scientists. Somewhat archaic language and terminology.

Tilakaratna, Prasanna. “How Magnetic Resonance Imaging Works, Explained


Simply.” https://www.howequipmentworks.com/mri_basics/. Good step-by-step
description of MRI, with lots of images.

United Nations Scientific Committee on the Effects of Atomic Radiation.


Sources and Effects of Ionizing Radiation. Volume I: Sources and Volume II:
Effects. Vienna: United Nations, 2011. Available at http://www.unscear.org/
docs/reports/2008/11-80076_Report_2008_Annex_D.pdf. Summarized in The
Chernobyl Accident: UNSCEAR’s Assessments of the Radiation Effects, http://www.
unscear.org/unscear/en/chernobyl.html. Full description of the health effects of
the Chernobyl accident. The brief, readable summary gives a clear description of
the health effects.

US Department of Health and Human Services. “REMM Radiation Emergency


Medical Management.” https://www.remm.nlm.gov/index.html. Useful
information from the US government.

US Nuclear Regulatory Commission. “Radiation Protection.” https://www.nrc.


gov/about-nrc/radiation.html. Brief but thorough description on the sources and
effects of radiation from the official US government website. It includes a link to
the Personal Annual Radiation Dose Calculator.

——— . “Uranium Enrichment.” https://www.nrc.gov/materials/fuel-cycle-fac/


ur-enrichment.html. A good description of the main enrichment techniques.

332 Nuclear Physics Explained


White, Martin. “Big Bang Nucleosynthesis.” http://w.astro.berkeley.edu/
~mwhite/darkmatter/bbn.html. Good description of big bang nucleosynthesis
from an expert in the field.

——— . “Origin of the Light Elements.” http://w.astro.berkeley.edu/~mwhite/


darkmatter/bbndetails.html. Good description of big bang nucleosynthesis from
an expert in the field.

Wilczek, Frank. “The Origin of Mass.” MIT Physics Annual 2003. 24–35.
Available at http://frankwilczek.com/Wilczek_Easy_Pieces/342_Origin_
of_Mass.pdf. Good description of quantum chromodynamics, confinement,
asymptotic freedom, and the origin of mass.

Wilson, Mark. “Electron-Scattering Experiments Resolve Short-Range


Correlations among Nucleons.” Physics Today, July 2008. http://physicstoday.
scitation.org/doi/full/10.1063/1.2962997. Description of the latest research on
nucleon behavior in nuclei.

Wiringa, R. B., and Steven C. Pieper. “Evolution of Nuclear Spectra with


Nuclear Forces.” Physical Review Letters 89, no. 18 (2002): 182501-1–4.
State-of-the-art calculation of light nuclei using different nucleon-nucleon
interactions to investigate exactly why there are no stable mass 5 or 8 nuclei. Very
technical article.

Wong, Samuel S. M. Introductory Nuclear Physics. Englewood Cliffs, NJ: Prentice


Hall, 1990. A good advanced-undergraduate level nuclear physics textbook.

World Nuclear Association. “Advanced Nuclear Power Reactors.” http://www.


world-nuclear.org/information-library/nuclear-fuel-cycle/nuclear-power-reactors/
advanced-nuclear-power-reactors.aspx. A review of Generation III nuclear power
plants that covers their general properties as well as the specific designs.

Bibliography 333
——— . “Nuclear Power in the World Today.” http://world-nuclear.org/
information-library/current-and-future-generation/nuclear-power-in-the-world-
today.aspx. Survey of the current status of nuclear power generation. Updated
periodically. Pronuclear. One of many informative articles on the World Nuclear
Association’s website.

——— . “Storage and Disposal of Nuclear Waste.” http://www.world-nuclear.


org/information-library/nuclear-fuel-cycle/nuclear-wastes/storage-and-
disposal-of-radioactive-wastes.aspx. A review of radioactive waste storage
options. Pronuclear. One of many informative articles on the World Nuclear
Association website.

334 Nuclear Physics Explained


IMAGE CREDITS

6 . . . . . Library of Congress, Prints and Photographs Division, LC-DIG-ggbain-03392.


7 . . . . . Library of Congress, Prints and Photographs Division, LC-DIG-ggbain-07682.
20 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . The Teaching Company Collection.
21 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . © abadonian/iStock/Thinkstock.
25 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . © webarama/iStock/Thinkstock.
27 . . . . . . . . . . . . . . . . . . . . . . . . . . . . © chengyuzheng/iStock/Thinkstock.
36 . . . . . . . . . . . . . . . . . . . . . . . . . . . . © Photoman195/iStock/Thinkstock.
39 . . . . . . . . . . . . . . . . . . . . . . . © monkeybusinessimages/iStock/Thinkstock.
51 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . © diego_cervo/iStock/Thinkstock.
52 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . © Rawpixel/iStock/Thinkstock.
62 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . NARA.
70, 78, 80 . . . . . . . . . . . . . . . . . . . . . . . . . The Teaching Company Collection.
81 . . . . . . . . . . . . Boston Museum of Science/Wikimedia Commons/CC BY-SA 3.0.
83 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . The Teaching Company Collection.
84 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . © 2014–2017 CERN.
90–93 . . . . . . . . . . . . . . . . . . . . . . . . . . . The Teaching Company Collection.
96 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Argonne National Laboratory.
99 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . The Teaching Company Collection.
102 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Jlab.
110, 114 . . . . . . . . . . . . . . . . . . . . . . . . . . The Teaching Company Collection.
116 . . . . . . . . . . . . . . . . . . . . . . Thomas Jefferson National Accelerator Facility.
122–125, 127 . . . . . . . . . . . . . . . . . . . . . . . . The Teaching Company Collection.
133 . . . . . . . . . Library of Congress, Prints and Photographs Division, LC-B2-6358-11.
143 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . The Teaching Company Collection.
161 . . . . . . . . . . . . . . . . . . . . . Fastfission/Wikimedia Commons/Public Domain.
167 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . NARA.
168 . . . . . . . . . . . . . . . . . . . . . Fastfission/Wikimedia Commons/Public Domain.
169 . . . . . . . . . . . . . . . . . . . . . . . . Ausis/Wikimedia Commons/CC BY-SA 2.5.
170 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . NARA.
176 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . The Teaching Company Collection.
179 . . . . . . . . . . . . . . . . . . . . . . . . . . . . © DanielPrudek/iStock/Thinkstock.
186 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . © Dobresum/iStock/Thinkstock.
190. . . . . . . . . . . . . . . . . . . . . . . . . . . . . The Teaching Company Collection.
193 . . . . . . . . . . . . . . . . . . . Tokyo Electric Power Co., TEPCO/flickr/CC by 2.0.
205 . . . . . . . . . . . . . . . . . . . Picoterawatt/Wikimedia Commons/Public Domain.

Image Credits 335


214, 215, 217 . . . . . . . . . . . . . . . . . . . . . . . . The Teaching Company Collection.
225 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . National Library of Medicine.
228 . . . . . . . National Cancer Institute/General Motors Cancer Research Foundation.
236 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . © urfinguss/iStock/Thinkstock.
239, 249 . . . . . . . . . . . . . . . . . . . . . . . . . . The Teaching Company Collection.
250 . . . . . . . . . . . . . . . . . . Hokanomono/Wikimedia Commons/Public Domain.
251 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . © karandaev/iStock/Thinkstock.
252 . . . . . . . . . . . . . . . . . . . . . . . . . . . . © kolesnikovserg/iStock/Thinkstock.
260, 262, 265 . . . . . . . . . . . . . . . . . . . . . . . The Teaching Company Collection.
267 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Jlab.

336 Nuclear Physics Explained

You might also like