Atomic Physics For Everyone: Will Raven
Atomic Physics For Everyone: Will Raven
Atomic Physics
for Everyone
An Introduction to Atomic Physics,
Quantum Mechanics, and Precision
Spectroscopy with No College-Level
Prerequisites
Atomic Physics for Everyone
Will Raven
© The Editor(s) (if applicable) and The Author(s) 2025. This book is an open access publication.
Open Access This book is licensed under the terms of the Creative Commons Attribution 4.0 Inter-
national License (http://creativecommons.org/licenses/by/4.0/), which permits use, sharing, adaptation,
distribution and reproduction in any medium or format, as long as you give appropriate credit to the
original author(s) and the source, provide a link to the Creative Commons license and indicate if changes
were made.
The images or other third party material in this book are included in the book’s Creative Commons
license, unless indicated otherwise in a credit line to the material. If material is not included in the book’s
Creative Commons license and your intended use is not permitted by statutory regulation or exceeds the
permitted use, you will need to obtain permission directly from the copyright holder.
The use of general descriptive names, registered names, trademarks, service marks, etc. in this publication
does not imply, even in the absence of a specific statement, that such names are exempt from the relevant
protective laws and regulations and therefore free for general use.
The publisher, the authors and the editors are safe to assume that the advice and information in this book
are believed to be true and accurate at the date of publication. Neither the publisher nor the authors or
the editors give a warranty, expressed or implied, with respect to the material contained herein or for any
errors or omissions that may have been made. The publisher remains neutral with regard to jurisdictional
claims in published maps and institutional affiliations.
This Springer imprint is published by the registered company Springer Nature Switzerland AG
The registered company address is: Gewerbestrasse 11, 6330 Cham, Switzerland
I have many people to thank for helping me write this book. To properly give praise,
I need to provide a little background. I joined Smith College in 2013 and quickly
became part of the AEMES (Achieving Excellence in Mathematics, Engineering,
and Sciences) mentoring team, led by the amazing Dr. Valerie Joseph. This program
is designed to increase diversity in STEM fields through a variety of initiatives.
One of the key insights I gained was that providing students from underrepresented
and/or under-resourced backgrounds with early research experiences is one of the
best ways to keep them in STEM majors. Over time, the demand for research
experiences grew among students. To address both of these needs—fulfilling
the increasing demand for research experiences and providing early research
opportunities to students from under-resourced backgrounds—I developed, with the
encouragement and help of many people, a course-based research program. This
program teaches first-year college students (freshmen) atomic physics, giving them
time in the lab to conduct real science. The class has no college-level prerequisites;
we use no calculus, linear algebra, or assume any background in introductory
mechanics. However, we do use algebra and some trigonometry, specifically sines
and cosines. The class was included in the Broader Impacts section of a National
Science Foundation (NSF) grant that I was fortunate to be awarded: PHY-2110311.
People Who Helped Me Develop the Pedagogy for the Class I had casual
conversations about class development with a large number of people, but there
are a few who made substantial contributions including Prof. Maria-Teresa Herd
(Physics, Assumption University), Prof. Timothy Malacarne (Sociology and Data
Science, Nevada State University), and Prof. Patricia DiBartolo (Psychology and
Director of the Sherrerd Center for Teaching & Learning, Smith College). The Smith
College Office of the Provost and Dean of Faculty, led by Prof. Michael Thurston
(English Language and Literature, Smith College), were incredibly supportive and
provided personnel support to the physics department to free up my time to develop
and teach the class.
My original intent was to write this book as free PDFs for my website. It was
my NSF program manager Dr. John D. Gillaspy (Physics) who first encouraged me
to publish the book for broader distribution. But, if my goal was to attract students
from under-resourced backgrounds, it didn’t make much sense to have them pay
vii
viii Acknowledgments
for a book. With additional support from Smith College, we are able to publish this
book open access!
People Who Read the Book and Gave Me Lots of Feedback There are two
people in particular who read the book multiple times through all the iterations.
I owe Prof. Maria-Teresa Herd (Physics, Assumption University) and Prof. Doreen
Weinberger (Physics, Smith College) so, so much. They were so insightful about
approach and organization. This short paragraph seems incredibly insufficient for
what they have done for me.
I want to thank the numerous people who read a selected chapter (or chapters).
They are Prof. Michael Thurston (English Language and Literature, Smith College),
Prof. Gary Felder (Physics, Smith College), and Prof. Travis Norsen (Physics,
Smith College). And then there are all of the students who took this class with me.
They found countless typos and mistakes, and pointed out confusing sections. More
importantly, their questions and thoughtful exploration of the world of the super
small helped me organize, expand, and enhance the book.
I also want to directly thank a few people who helped me in the classroom. These
folks not only helped teach the class but also provided valuable feedback, sometimes
in real time. Three of these individuals were learning assistants in the lab sections
of the class the first time I taught it. They are Bárbara Cabrales, Molly Herzog, and
Chitose Maruko. Dr. Karl Ahrendsen (Physics), who was a postdoctoral scholar in
my lab, co-taught the class with me during the second iteration. He also gave an
incredible amount of feedback to this text.
A Brief Tour of the Book
This book is divided into two parts. The first part, titled Atom-Light Interactions,
explores spectroscopic techniques used to extract information about atoms. The
second part, titled Digging Deeper: Quantum Mechanics and Beyond, examines
the underlying physics of atomic structure, covering atomic notation, the principles
behind it, current scientific understanding, and ongoing mysteries.
Generally, Part I is for learners interested in how we acquire knowledge about
atoms, while Part II is aimed at those keen on understanding the fundamental
physics of atoms. Readers may skip Part I if they are solely interested in the physics
of atomic structure, but they should complete Chap. 1 before moving on to Part II.
The text also includes three appendices and a glossary that should serve as
valuable references for readers. Appendix A is a periodic table, Appendix B is a long
table that contains a lot of useful information about the atoms on the periodic table,
and Appendix C lists all of the quantum mechanic rules that need to be satisfied for
an electron to transition between two atomic states.
ix
Contents
xi
xii Contents
Index.............................................................................................. 261
Part I
Atom-Light Interactions
Introduction to Atoms and Light
1
Abstract
In this chapter, we explore the nature of light and atoms, focusing on their
dual nature as both particles and waves. We examine why atoms have discrete
energy levels and how only certain frequencies of light can excite electrons
within these atoms. Through this exploration, we will understand the relationship
between light’s frequency, wavelength, and photon energy. We also explore key
concepts such as wave interference and the historical experiments that shaped our
understanding of quantum mechanics. Most importantly, this chapter emphasizes
the scientific method, encouraging continual questioning of ideas, understanding,
theories, and results to uncover the fundamental nature of the universe.
Learning Goals
1 Words in the glossary are in bold type the first time we use them.
2 Inthe last half of the twentieth century, atomic physics combined with molecular physics and
optical physics to make what is now known as atomic, molecular, and optical physics. We use the
acronym AMO physics.
1.1 What Is Atomic Physics? 5
• Atoms are composed of three types of particles: protons, neutrons, and elec-
trons.
.◦ Protons and neutrons form the center of the atom, called the nucleus, and
charge. We call this a neutral atom. If we take an electron away from the
atom or give the atom an extra electron, it is now charged. We call this an ion.
• Molecules are made from multiple atoms bound together. They can be the same
atoms, like a nitrogen molecule consists of two nitrogen atoms, or different
atoms, like a water molecule that has two hydrogen atoms and one oxygen atom.
• Atoms can be in the gas phase, liquid phase, or a solid phase.3.,4 Most atomic
physicists study atoms in the gas phase to avoid the complexity of liquids and
solids. In liquids and solids, atoms are bonded to one another, which makes the
system more complicated. Our ideal system is a single atom all by itself away
from all outside interactions.
• The periodic table of elements is a common way to view all the elements we
know about. A copy of the periodic table can be found in Appendix A.
• Spectroscopists use light to interact with and learn about the world of the super
small.
This book is aims to teach quantum mechanics, atomic physics, and spectroscopy
without any advanced math. The most complicated math we use is algebra and
some trigonometry (sines and cosines). Spectroscopy is a subfield of atomic physics
that tries to learn about an atom through its interaction with light. Spectroscopy
can be thought of as both a subfield and a tool. As a subfield, it is using atom-
light interactions to try and understand the world of the super small. Additionally,
it serves as a tool. The goal of atomic physics is to understand how atoms work,
and spectroscopy is just a tool to accomplish this goal. The important thing is that
spectroscopists use light to interact with atoms with the end goal of understanding
the world of the super small. Scientists in lots of fields use spectroscopy, including
atomic physicists, nuclear physicists, chemists, geologists, atmospheric scientists,
and astronomers.
Atomic physicists try to reach this goal by starting with a simple system and
building up complexity over time. One of our simplest systems is a single electron
3 There are other phases of matter as well, including plasmas and Bose-Einstein condensates
(BEC), which were named after the Indian mathematician and physicist Satyendra Nath Bose and
the German born physicist Albert Einstein. A BEC is a really amazing state of matter that was first
postulated in 1924. It wasn’t until 1995 when American physicists Carl Wieman and Eric Cornell
used atomic physics techniques to create the world’s first BEC from rubidium atoms. Shortly after,
German physicist Wolfgang Ketterle made a BEC from sodium atoms. The three atomic physicists
won the 2001 Nobel Prize in Physics for this effort.
4 One of my readers suggested that adding nationalities would add a bit of fun and extra history (I
agreed!). Lots of sources list Einstein as ‘German born’ because he moved to Switzerland in 1895,
giving up his German citizenship. In 1901, he became a Swiss citizen.
6 1 Introduction to Atoms and Light
Fig. 1.1 A flowchart for trying to understand something. This is a version of the scientific method
showing the interplay between experiment and theory
5 To be fair, experiment does not always agree with the Standard Model. For example, the Standard
Model says that the universe should be nearly equal amounts of matter and antimatter (this is the
baryon asymmetry listed above), which would be bad since then all the matter and antimatter
would combine to destroy the universe. In other words, the Standard Model says that the universe
shouldn’t exist, which disagrees with experiment.
8 1 Introduction to Atoms and Light
We are not going to discuss the Standard Model right now. We are going to keep
things a bit simpler and discuss, conceptually, the atom. Later on in the book, we
will add to our conceptual model to make things more complete. So, what does
quantum mechanics say about the atom?
To answer that question, let’s first think about an electron that is orbiting around a
nucleus, as shown in Fig. 1.2. The electron has negative charge and the nucleus has
positive charge. According to electromagnetic theory, the electron should radiate
away energy. Imagine if you were flying a hand glider and you slowly lost energy.
In this scenario you would slowly drift down until you landed on the ground.
According to electromagnetic theory, the electron will also lose energy and should
orbit closer and closer to the nucleus until it collides with the nucleus. Experiment
shows this is not true. In fact, if it were true, we wouldn’t be here reading this book
because atoms wouldn’t exist and thus the universe wouldn’t exist,6 which disagrees
with experimental observation.
There were some amazing experiments conducted in the late 1800s and early
1900s that seemed to imply that the electron is not like the moon orbiting around
the earth. One of my favorite experiments that showed this behavior is called the
double-slit experiment. If we assume electrons are like little sticky balls and send
them through two small slits in a barrier, as shown in Fig. 1.3, we expect to have two
strips of balls stuck to the screen. However, that isn’t what is seen experimentally!
Fig. 1.2 A helpful conceptual, but incorrect, way of thinking about atoms. The model where the
electrons orbiting around the nucleus is called the Bohr model, named after Danish physicist Niels
Bohr
6 Well,
maybe the universe would exist. It would just exist without any matter, which wouldn’t be
much fun.
1.2 Conceptually Understanding the Atom 9
Fig. 1.3 A double-slit experiment where the electrons behave like little sticky balls. In panel A,
the electrons are moving towards a barrier that has two small slits in it. Behind the barrier with two
small slits is a screen with no holes. In panel B, the electrons have crashed into or passed through
the barrier to crash into the screen. Panel C shows a front view of the electrons that passed through
the barrier and crashed into the screen
Before we get to the real experimental results, we need to define some important
terms and explore the concept of wave interference. Below are three important
definitions about waves. Figure 1.4 is a visual representation of two of those
definitions: wavelength and frequency.
Fig. 1.4 A visual description of wavelength and frequency. Wavelength is the distance between
two “like” points on the wave, for example the distance between two adjacent peaks or two adjacent
troughs. If the wave is moving towards the right, frequency is how many peaks pass through the
dashed line every second. If both waves are moving with the same speed, the upper wave has a
higher frequency than the lower wave since more peaks pass the dashed line in 1 second
10 1 Introduction to Atoms and Light
Definitions
• Wavelength: The distance between any two “like” points on a wave, such
as two adjacent wave peaks. The variable we use for wavelength is the
lowercase Greek letter lambda, .λ. Since wavelength is a distance, we use
length units such as meter, centimeter, or nanometer.
• Frequency: The number of oscillations per second. We use the variable f
for frequency. The unit for frequency is 1/seconds, which is called a hertz,
named after the German physicist Heinrich Hertz. A hertz is shortened to
Hz. Some other units that we use for frequency include a megahertz (1 MHz
.= 1 × 10 Hz), a gigahertz (1 GHz .= 1 × 10 Hz), and a terahertz (1 THz
6 9
.= 1 × 10
12 Hz).
• Period: The time it takes for the wave to complete one full oscillation. We
use the variable T for period, and the unit is seconds. Frequency and period
are related by the formula .T = 1/f .
t 0s t 0.6 s t 0.9 s
Fig. 1.5 Two pulses that constructively interfere with each other: The thick gray line is what we
would actually see. The blue dashed line shows the pulse traveling to the right while the red dashed
line shows the pulse traveling to the left. When they pass through one another, they add to create a
larger pulse. At .t = 1.0 s, this is 100% constructive interference. At all other times, the pulses are
only partially constructively interfering
1.2 Conceptually Understanding the Atom 11
t 0. s t 0.6 s t 0.9 s
t 1. s t 1.4 s t 1.8 s
Fig. 1.6 Two pulses that destructively interfere with each other: The thick gray line is what we
would actually see. The blue dashed line shows the pulse traveling to the right while the red dashed
line shows the pulse traveling to the left. When they pass through one another, they perfectly cancel
each other out creating no disturbance on the actual rope for a brief period of time. At .t = 1.0 s,
this is 100% destructive interference. At all other times, the pulses are only partially destructively
interfering
Try This
Find a piece of thin rope or a slinky and a stopwatch. If you have a friend
nearby, have them hold one end of the rope or slinky. If not, tie or connect one
end of the rope or slinky to a door knob. Stand a distance apart so that there is
a bit of tension on the rope or slinky. Send a pulse down the rope or slinky and
watch what happens when the pulse reflects off your friend or the doorknob.
Next, start creating a sine wave motion with your hand, see Fig. 1.4. Try to
move your arm up and down so that the rope or slinky creates the shapes seen
in Fig. 1.7.
Fig. 1.7 The first three standing waves of a one dimensional rope or slinky. Each standing
wave is drawn with the same amplitude A. Interestingly, the frequency of a standing wave
is independent of the amplitude of the standing wave
(continued)
12 1 Introduction to Atoms and Light
You will find that your hand has to move up and down with a very specific
frequency to create these shapes, which are called standing waves. When
you produce standing waves, the wave that you are creating with your hand
is constructively interfering with the reflected wave. No other frequencies
produce perfect constructive interference. Using your stopwatch, find the time
for ten full oscillations (the time it takes for your hand to move up and down
ten times) for each of the first three standing waves. The period of the standing
wave is that time divided by 10. If you count twenty full oscillations, you
would divide the time by 20 to find the period. Next use the formula .f = 1/T
to find the frequency needed to produce that standing wave. The frequency of
the standing wave with one “loop” (the left picture) is called the fundamental
frequency. You should find that the second picture with two loops has twice
the fundamental frequency. The third picture with three loops should have a
frequency that is three times the fundamental frequency. In general, .fn = nf1 ,
where n is how many loops the standing wave has. Notice that the higher mode
standing waves (the waves with more loops) require more shaking energy!
This will be important later.
Back to the Double-Slit Experiment Ok, let’s run a different experiment. Instead
of little, sticky balls, let’s send a wave towards the screens, as shown in Fig. 1.8.
Panel A shows the wave traveling towards the slits. The vertical lines are supposed
to indicate the peaks of the wave. The troughs are halfway between each peak, so
you are visualizing the wavelength of the wave. Panel B shows the wave as it passes
through the slits. The straight wave turns into two arc waves, one coming from each
slit. The arc waves are basically half circles with the center of the circle at the slit.
I changed the color of one of the circular waves to better visualize the evolution of
each wave. The two circular waves interfere with each other on the screen. Places
Fig. 1.9 Experimental results of a double-slit experiment with electrons. Image Credit: Dr.
Tonomura and Belsazar from Wikimedia Commons CC BY-SA 3.0
where peaks of the waves overlap, indicated by the black lines between the slits
and the screen, are where the two circular waves constructively interfere. Half way
between the points of constructive interference are points of perfect destructive
interference. Panel C shows what we see on the screen. The bright parts are where
the two waves constructively interfere. The dark parts are where the two waves
destructively interfere. At the very center of the dark parts, the waves are 100%
destructively interfering.
Now, let’s do the experiment with electrons! Real experimental results can be
seen in Fig. 1.9. The electron behaves like a wave! There is actually a lot more
to the double-slit experiment, but we aren’t going to go into any more detail in
this book. You will learn much more about it if you take a Modern Physics class.
The really important thing we need to conceptually understand is that the electron
has wavelike behavior. This is incredibly important information for us because an
electron that is orbiting a nucleus is not like the moon orbiting the earth, which is
depicted in Fig. 1.2. Experiments show that electrons behave like waves.
This is a really hard concept to wrap our brains around, but experiments seem
to indicate this idea is correct. How does a wave “orbit” around a nucleus? As an
analogy, imagine a wave that wraps around upon itself in a circle, see Fig. 1.10. This
figure is just a conceptual example since there is no start or end to the wave. But
to explore this concept we are going to wrap the electron wave counter-clockwise
around the nucleus, and the wave is going to interfere with itself. If the electron wave
does not perfectly wrap back around so that two peaks don’t overlap perfectly, the
electron will destructively interfere with itself. If an electron destructively interferes
14 1 Introduction to Atoms and Light
Fig. 1.10 A conceptual exploration of an electron wave orbiting a nucleus that would result in
destructive interference. In (a), we imagine the electron wave starting at the top of the circle and
traveling counter-clockwise. In (b), the electron wave has made one full orbit, but notice that the
wave does not line up with its starting point. For (c), the wave continues and should be interfering
with itself, but we aren’t going to add the waves together quite yet. In (d), the wave continues
for 12 orbits. To explore destructive interference, (e) shows adding together this wave after 4 full
rotations. (f) shows destructive interference of this wave after 10 full rotations
with itself, there is no wave! If there is no wave, then there is no electron. Since
electrons exist, the electron wave must constructively interfere with itself.
A lot has happened, so let’s do a quick recap. We have learned that:
Pause here and see if you can come up with a conclusion. When you have reached a
conclusion, read on. My conclusion is below this fun fact:
Fun Fact
Many musical instruments make noise (hopefully pleasant noise!) because of
standing waves. For example, when a guitarist plucks a guitar string, which is
fixed at both ends, they are creating many standing waves (see Fig. 1.7) with
different amplitudes all at the same time. Each of those standing waves has a
frequency that is a multiple of the fundamental frequency. The amplitude of
each of those standing waves is what gives the guitar it’s distinctive sound. To
(continued)
1.2 Conceptually Understanding the Atom 15
play a different note, the guitarist changes the length of the string by pressing
on the string in a different spot, which in turn changes the fundamental
frequency.a The same is true for a piano. A piano key strikes a wire that
produces standing waves that result in a (hopefully!) pleasant noise. The
amplitudes of all the standing waves on a struck piano wire are different than
the amplitudes of a guitar string, which is why they sound different.
A note played on a trumpet or saxophone also produces standing waves,
but the standing waves are created in the air and are physically manifested as
places of high and low air density.
a
They could also pluck a different string that has a different mass density or change the
tension on the string. All of these changes will result in a different fundamental frequency.
Conclusion The electron must constructively interfere with itself. This idea is
conceptually shown in Fig. 1.11.
An electron can only “orbit” around the nucleus if it satisfies a standing wave
condition. Each standing wave has an energy associated with it. Just like how the
standing wave on a string with 2 loops has a shorter wavelength and requires more
shaking energy than the standing wave with 1 loop, the higher energy “states” of
an electron have shorter wavelengths. Thus the allowed energies of the electron are
“discrete”: they can only have the specific values corresponding to these standing
waves.
We call the discrete energies an electron orbiting a nucleus can have “energy
levels” or “energy states” and describe them in diagrams like Fig. 1.12. The lowest
energy state is called the ground state of the atom, or the ground state for short.
States with higher energy are called excited states. These energy level pictures
are sometimes called Grotrian diagrams, named after German astronomer and
astrophysicist Walter Grotrian. The SI unit for energy is a joule, which is named after
ve
wa
ron
el ect
nucleus
Fig. 1.11 Four conceptual examples of an electron wave orbiting a nucleus that would result in
constructive interference. Going from left to right is going from a lower energy state to higher
energy states
16 1 Introduction to Atoms and Light
the English physicist James Joule. A joule is shortened to .J.7 A joule is shorthand
for .kg m2 /s2 . For reference, the energy of a baseball moving at 50 mph is about
36 joules. The energy of an apple moving at 1 meter/second is about 0.05 joules.
The energy difference between the ground state and the lowest energy excited
state in the hydrogen atom is .0.000000000000000001634 J = 1.634 × 10−18 J.
That is a small number! A more common unit for energy in atomic physics is the
electronvolt (eV). The conversion is .1 eV = 1.602 × 10−19 J, so the energy of the
lowest excited state is .10.2 eV above the energy of the ground state.
One Final Thing The above description of an electron in an atom is a good starting
point to understanding how and why atoms behave the way they do. Think about this
chapter as making a first pass-through Fig. 1.1. The conceptual idea you just learned
is correct. However, the electron wave turns out to be more complicated than the
simple picture given above. For one thing, the electron wave is going to be in 3
dimensions. But, more importantly, the actual wave that describes the electron in
an atom, which is called a wavefunction, is a bit messier to deal with. All of the
concepts like wave interference and energy levels in the atom are still correct, but
the presentation of the electron wave has been simplified. You will refine these ideas
in future classes such as Modern Physics and Quantum Mechanics.
7 Notice the unit ‘joule’ is not capitalized. In general, a unit named after a person is not capitalized.
1.3 Photons and Spectroscopy 17
Suppose we want to measure the energy difference between the ground state and any
one of the excited states. How do we do it? Answer: Spectroscopy! We shine light
onto the atoms and look at what happens to the light. In modern spectroscopy, we
use laser light. Light, including laser light, is composed of tiny particles known as
photons.8 As an analogy, think about a stream of water. The water looks continuous,
but it is actually made up of tiny water molecules. The same thing is true with light.
Light is composed of tiny particles we call photons. There is, however, a really
important difference between water molecules and photons. The energy of a photon
is determined only by the frequency of the light wave. The more the light wave
oscillates up and down in one second, the larger the energy of an individual photon
that makes up that light wave. The energy of a water molecule, on the other hand, is
related to how fast the molecule is moving, rotating, and vibrating. We will explore
the connection between the frequency of the light wave and the energy of a photon
more in the next section.
8 German physicist Max Planck won the 1918 Nobel Prize in physics for discovery of the photon.
Albert Einstein won the 1921 Nobel Prize in physics for explaining the photoelectric effect, which
determined that a beam of light is made up of a bunch of photons. I may be biased, but Einstein’s
explanation of the photoelectric effect was the most important discovery of the twentieth century.
18 1 Introduction to Atoms and Light
Thought Experiment Time Imagine you have an electron behaving like a wave
orbiting around a nucleus. That electron wave has to constructively interfere with
itself, and we can think about that standing wave as having some amount of energy.
Let’s call this energy .E1 . If we wanted to add another loop to the standing wave,
we need to add energy to the system. Suppose an electron with 1 more standing
wave loop has energy .E2 . Remember that the electron needs to be a standing wave,
which has a specific energy. If it wasn’t, the electron would experience destructive
interference. We want to move an electron from the lower energy standing wave
to the higher energy standing wave. How much energy do we need to put into the
system? The answer is in the footnotes.10
This is the basic idea behind spectroscopy. If we want to excite an electron from
one energy level, which has some energy, to another energy level, which has a
different energy, we need to provide the system with the correct difference in energy.
Restating that in terms of an actual experiment: We send a laser through an
optically clear container, called a cell or vapor cell, filled with atoms. We will
smoothly scan the energy of the photons over time and monitor the transmission
of the laser through the cell. If the energy of the photons does not match the energy
difference between the ground state and an excited state, the laser light will pass
right through the atoms with no losses. If the energy of the photons matches that
energy difference, light will be lost from the laser and the transmission will decrease.
Experiments that match the above description are known as absorption spectroscopy
experiments. We, as experimentalists, simply monitor the transmission of the laser
through an atomic sample as we change the photon energy. When the amount of
transmitted light drops, we learn what energy is required to excite the atoms from
the ground state to an excited state. Believe it or not, that is, conceptually, all there
is to it. In the lab things are a bit more complicated, but this is the basic idea. We
will finish up this section with a new definition.
9 This important statement is super important, but it also isn’t 100% correct. We are going to start
with this statement to get at some important concepts. In the next few chapters, we are going to
discuss some more physics and then restate this important statement to something more correct.
It’s kind of like the flowchart from the beginning of the chapter. We start simple and build up
complexity.
10 We would need to add precisely .E − E of energy.
2 1
1.4 Math 19
Definitions
• Resonance: When an atom gets excited by a photon from one state to
another, we say the atom “goes through resonance.” This is similar to
playing the trumpet. When you blow correctly into a trumpet, a standing
wave is excited in the pipe to create a note. The same thing happens with an
atom. When you excite an atom with the right energy photon, the electron
goes from one standing wave mode to another. Atomic physicists use the
word “excitation” and “resonance” interchangeably.
1.4 Math
The speed of light, the wavelength of the laser light, and the frequency of the laser
light are all related using the formula:
c = f λ,
. (1.1)
where .c = 299, 792, 458 m/s ≈ 3 × 108 m/s is a constant of nature known as the
speed of light.11 The energy of a photon is given by the formula:
Eph = hf,
. (1.2)
11 This is the speed of light in a vacuum. It is exactly .299, 792, 458 m/s.
20 1 Introduction to Atoms and Light
between the two numbers requires multiplying by the correct combination of the
fundamental constants h and c. This energy unit is explored more in Problem 1.6.
We can now make a new definition:
Definitions
• Resonance frequency: When the frequency of the laser is just right to excite
the atom from the ground state to an excited state, we call that the resonance
frequency. We use the variable .fr to represent resonance frequency. Since
resonance frequency is a frequency, it uses units of Hz, MHz, GHz, or THz.
Common Misconception Power is not the same thing as energy. Power is the rate
at which energy is exiting the laser. Imagine two laser beams with two different
frequencies. The light from laser #1 has a higher frequency than the light from laser
#2. We now know that the photons that make up laser #1 have more energy than the
photons that make up laser #2. Power is how much energy is exiting the laser per
second. If the same number of photons per second are leaving both lasers, laser #1
has a larger power. In math form, the power of a laser is:
P = NEph ,
. (1.3)
where N is how many photons per second that leave the laser and .Eph = hf is the
energy of a single photon. The unit for power is a watt, named in honor of Scottish
chemist James Watt. A watt is shortened to .W and is equivalent to joules per second.
Imagine you have a 1 watt laser and a 10 watt laser that have photons with the same
energy. The 10 watt laser emits 10 times as many photons per second as a 1 watt
laser.
The frequency of light tells us how many times that light wave oscillates up and
down in 1 second. Polarization tells us the direction the light is oscillating. There are
three major groupings of light polarization: linear, circular, and elliptical. Linearly
polarized light is the most common light we use in spectroscopy, so we aren’t going
to talk about circularly or elliptically polarized light here. However, there are types
of spectroscopy experiments that do use circularly polarized light. If you are curious
about them, image search the phrases to find some neat animated gifs showing
light with different polarizations moving through space. For now, we will focus
on linearly polarized light.
1.5 Extra: Polarization 21
Fig. 1.13 Examples of oscillating electric fields propagating along the z-direction. The left picture
shows horizontal linear light while the right picture shows vertical linear light
Linear polarization is light that is oscillating up and down in a single plane. There
are two specific types of linearly polarized light: horizontal and vertical. Horizontal
light oscillates . . . horizontally with respect to some surface, and vertical light
oscillates . . . vertically to that surface, see Fig. 1.13 Remarkably, we can use these
two polarizations to describe any linearly polarized light. For example, suppose the
light was oscillating at a 45.◦ angle. We would describe the light as half horizontal
and half vertical.
In the lab, there are optical devices called half-waveplates, sometimes written as
.λ/2 plate or just .λ/2. A half-waveplate can rotate the linear polarization of light.
If you have horizontally polarized light, you can use a half-waveplate to change
the polarization so that it is 10% vertical and 90% horizontal, 50% vertical and
50% horizontal, 75% vertical and 25% horizontal, etc. You can even make the light
exiting the half-waveplate be completely vertical. That might seem like a neat trick,
but the real usefulness comes when we put a second optical device after the halfwave
plate called a polarizing beam splitter (PBS), see Fig. 1.14.
Fig. 1.14 Using a half-waveplate and a polarizing beam splitter, we can create two beams of light.
You can rotate the half-waveplate to control the ratio of light in each path. In normal spectroscopy
setups, you are looking down on the light and optics from above. In this orientation, p-polarized
light is horizontal, indicated by the blue arrows and s-polarized light is vertical, indicated by the
dotted circles
22 1 Introduction to Atoms and Light
By tradition, the light that bounces off the PBS is called s-polarized light while
the light that passes through is called p-polarized light.12 This is because the
PBS can technically be in any spatial orientation. For safety reasons, we almost
always keep the light in a horizontal plane. In this typical setup, s-polarized light
is vertically polarized light while p-polarized light is horizontally polarized light. If
you were to rotate the PBS so that the s-polarized light was going straight up (don’t
do this, it is an eye hazard!!), that s-polarized light is now horizontally polarized
while the p-polarized light is now vertically polarized. This is why we use “s” and
“p”.
In the lab, we use half-waveplates and polarizing beam splitters to split a single
laser beam into two beams and also control the power in each of them. If the
waveplate is oriented such that the light polarization leaving the waveplate is
completely vertical, all of the light would bounce off the PBS. If the polarization
of the light leaving the waveplate was at 45.◦ , there would be equal power in both
lasers leaving the PBS.
Don’t be afraid to ask questions. Don’t be worried about asking for clarification.
Don’t expect to remember or understand every single concept the first time you read
or hear about it. Mastery is not a short journey. Developing critical thinking skills is
not a 5 minute activity that you figure out after watching a 3 minute YouTube video.
Practice deep learning, keep a growth mindset, and, most importantly, have fun!
Problems
1.1 Go through the chapter and write down all of the fundamental constants (there
are two of them). Don’t forget units.
1.2 Go through the chapter and write down each equation. For each equation, write
a brief description about what the equation means.
12 Fun fact: the “s” stands for the German word “senkrecht”, which translates to perpendicular; the
“p” stands for the German word “parallel”, which translates to parallel (that isn’t a misprint, the
German word for parallel is parallel). I like to use “s” for skip and “p” for pass-through.
1.6 The Most Important Equation in All of Science 23
(a) Using two formulas from Problem 1.2, derive a formula for the energy of a
photon in terms of only fundamental constants and wavelength.
(b) Check to make sure the units are correct. Physicists always check units. Always.
Hint: 1 J = 1 kg m2 /s2
1.4
(a) What is the energy of a photon that is in a laser that has a frequency of 647.8 THz
(647.8 × 1012 Hz)? Express your answer in both joules (J) and electronvolts
(eV). Your final answer should have 4 significant figures.
(b) What is the wavelength of the light in nanometers?
1.6 In atomic physics, energy is often measured in the units of cm−1 . The method
of calculating energy in these units is to first find the wavelength of the laser in
centimeters, and then take the inverse. In equation form, this is
( ) 1
E cm−1 =
. . (1.4)
λ(cm)
Note that in the above formula, the units to use for each symbol are included in
parentheses; the formula does not say “λ multiplied by cm”, but rather “make sure
wavelength has units of cm before plugging into the formula.”
Using λ = 852.347 nm, find the energy in units of inverse centimeters. Your
answer should have 6 significant figures.
1.7 The equation in Problem 1.6 can be confusing. Energy has units of joules and
not inverse length! Why are atomic physicists comfortable with using energy in this
weird unit? There are multiple correct answers here.
1.8 It isn’t only atomic physicists who use energy in weird units. Astronomers,
nuclear physicists, and a good number of chemists also use energy in inverse cen-
timeters. Some chemists and condensed matter physicists prefer to use electronvolts
(eV). Virtually no one uses the SI unit (joules)! Any thoughts why?
24 1 Introduction to Atoms and Light
Open Access This chapter is licensed under the terms of the Creative Commons Attribution 4.0
International License (http://creativecommons.org/licenses/by/4.0/), which permits use, sharing,
adaptation, distribution and reproduction in any medium or format, as long as you give appropriate
credit to the original author(s) and the source, provide a link to the Creative Commons license and
indicate if changes were made.
The images or other third party material in this chapter are included in the chapter’s Creative
Commons license, unless indicated otherwise in a credit line to the material. If material is not
included in the chapter’s Creative Commons license and your intended use is not permitted by
statutory regulation or exceeds the permitted use, you will need to obtain permission directly from
the copyright holder.
“Natural Light”
2
Abstract
In this chapter, we explore how light from sources such as the sun or a lamp
can be dispersed into its spectral components. We discuss various dispersive
elements, including gratings and prisms, and the phenomena of refraction and
diffraction that allow these elements to spatially separate light. The chapter
explores blackbody radiation and the historical significance of the ultraviolet
catastrophe. Additionally, we introduce the concepts of absorption and emission
lines, which provide insights into atomic and molecular energy levels.
Learning Goals
• the concept of spectral components and how white light is composed of various
wavelengths of light.
• the function and importance of dispersive elements like prisms and gratings in
spectroscopy.
• the difference between refraction and diffraction.
• blackbody radiation and the ultraviolet catastrophe.
• the principles of emission and absorption lines in spectroscopy and their
significance in understanding atomic structure.
• the Stefan-Boltzmann law and Wien’s displacement law in the context of
blackbody radiation, and how to determine temperature from the blackbody
spectrum.
• how light collected from the sun gave us our first evidence of atomic energy
levels.
Definitions
• Spectral component: “White” light is made up of many different wave-
lengths of light. Even light from the sun, which looks yellow, is made up of
many different wavelengths of light. A single wavelength of light that makes
up a broader spectrum of light is called a spectral component.
• Dispersive element: Anything that spatially separates light into its spectral
components.
• Spectrometer: Any tool that allows us to separate out and measure the
amount of each spectral component.
• Refraction: the redirection of a wave as it passes from one medium to
another medium (like air to water).
• Diffraction: when waves bend around the corners of an obstacle.
Fig. 2.1 A dispersive prism takes white light, which enters the prism from the left, and makes a
rainbow. Each wavelength is refracted at a different angle. The white light source I used to make
this picture is a tungsten lamp
2.1 Breaking Light into a Spectrum 27
Fig. 2.2 (a) An example of white light, which is composed of many spectral components,
reflecting off a reflection grating. For illustrative purposes, d, which is usually very small, is greatly
enlarged. We also only show one order to keep the example a little cleaner. (b) An example of
white light diffracting through a transmission grating. Again, d is greatly enlarged and we only
show one order. (c) An example of blue light with .λ = 455 nm hitting perpendicular to the
transmission grating and being diffracted. In this example, we show all diffraction orders. The
angles of diffraction are calculated using Eq. 2.1
wavelength refracts, you can send in an unknown wavelength, measure the angle of
refraction, and use math to determine the wavelength of the light. You can also send
in light from, for example, a hydrogen lamp and see what wavelengths or spectral
components are in that light. Knowing how each wavelength refracts allows you to
determine what wavelengths make up the hydrogen lamp spectrum.
In spectroscopy, the most common dispersive element is an optical grating,
which works due to diffraction. Optical gratings commonly come in two types:
transmission and reflection, see Fig. 2.2. Either way, the optical grating is a
dispersive element that will spatially separate light into its spectral components
because each spectral component diffracts at a different angle.
To create the diffraction, both types of gratings have small structures separated
by a distance d. A reflection grating has a bunch of small tilted mirrors called rulings
while a transmission grating has a bunch of small slits. The only physical criterion
for a grating is that .d > λ. The size of d in Fig. 2.2 has been greatly enlarged for
visual purposes. While d must be larger than .λ, in practice we also make sure that d
is typically less than about .5λ to ensure suitable diffraction angles.
We will start with the equations that describe how spectral components are
diffracted through a transmission grating; we will find that each spectral component
can be diffracted at multiple, but well defined, angles. The math behind the next
equation is a little complicated, but the end result is really what we care about. Let’s
start by keeping things simple and assume the incoming light is perpendicular to
the grating, see Fig. 2.2c. The light diffracts through the grating according to the
equation:
d sin θm = mλ
. (2.1)
for incident light perpendicular to a transmission grating
28 2 “Natural Light”
In this equation, d is the distance between the slits, m is an integer known as the
diffraction order (the fact that a spectral component diffracts at multiple but well
defined angles is mathematically represented by m), .λ is the wavelength of a spectral
component of the light, and .θm is the angle of diffraction for order m. m can be a
negative integer, zero, or a positive integer. If you know d and m, you can then
measure .θm and use that to calculate .λ.
Example A transmission grating has .d = 1 μm. There are three laser pointers: a
blue laser pointer with .λb = 450 nm, a green laser pointer with .λg = 532 nm, and
a red laser pointer with .λr = 650 nm. All three lasers hit the grating perpendicular.
What angles do the three lasers diffract for m = +1, 0, .−1, and +2?
First, we use the diffraction equation to solve for .θm : .θm = sin−1 (mλ/d), where
−1
.sin is the inverse sine function. For .m = 0, we plug in numbers to find .θ0,b = 0◦ ,
◦ ◦
.θ0,g = 0 , .θ0,r = 0 . So, let’s not use the .m = 0 diffraction order since we can’t
z = ±/ L
d 2
. ( mλ ) −1 (2.2)
for incident light perpendicular to a transmission grating,
where L is the distance between the diffraction grating and the screen. z is
positive for positive m and negative for negative m. You have the opportunity
to derive this formula in problem 2.4. Notice that the diffraction order in the
denominator is squared, so the sign of m only determines the sign of z.
angle (i.e. not a complex number). In the end, it doesn’t matter which diffraction
order we pick. As long as we know the angle at which a spectral component
diffracts, we can use that information to determine the wavelength of an unknown
spectral component.
A reflection grating does a similar job as a transmission grating, but the equation
looks slightly different:
d sin θm = −mλ
. (2.3)
for incident light perpendicular to a reflection grating
All that is different is the minus sign on the righthand side. Finally, let’s suppose
the incoming light isn’t perpendicular to the grating but hits the grating at an
angle .θi with respect to the perpendicular (in optics, the perpendicular is called
the “normal”), see Fig. 2.2b. The formula describing the diffraction of a spectral
component is:
While this formula is more complicated, it is important to note that the concept is
far more important than the formula. The concept is simply this:
Fun History
• The first known exploration of a diffraction grating was by the Scottish math-
ematician and astronomer James Gregory in the mid 1600s. He observed the
diffraction of sunlight caused by light passing through a bird feather. The
individual feathers acted like the small slits of a transmission grating. The
German physicist Joseph von Fraunhofer made the first diffraction grating in
1814. He discovered something amazing, which is the topic of the next section.
• Diffracting light from a light bulb whose gas is a particular element like
hydrogen, nitrogen, or oxygen is the topic of Sect. 2.3. Classical physics had
no way of explaining the observed spectral components, and this mystery was
one of the puzzling experiments that led to quantum mechanics.
I pointed a commercial spectrometer at the sky, see Fig. 2.3. Collecting this data was
spectroscopy! A few important notes before we continue:
• This spectrometer has a dispersive element that was pre-calibrated, so the angle
the light hits the detector is automatically converted to a wavelength.
• Spectrometers are not uniformly sensitive. This means that the detector you are
using might be more sensitive to red light than blue light. Graphs like Fig. 2.3
are useful for identifying the wavelengths present in the light, but unless the
detector has been calibrated to correct for wavelength sensitivity, the shape may
Intensity (arbitrary units)
(nm)
0.8
700
0.6
600
0.4
0.2 500
0.0
400 500 600 700 800 900 1000 400
(nm)
Fig. 2.3 The spectral components of light from the sun after it passes through the atmosphere
2.2 Blackbody Radiation 31
be incorrect. In fact, the real spectrum of the sky goes to far longer wavelengths,
but this spectrometer is not sensitive to light at those frequencies.
• When this type of spectrum was first taken in the 1800s, those dips were a
mystery, and physicists love a good mystery. Whenever a physicist is presented
with data, the first thing they ask themselves is, “Why does the data look the
way it does?” If we understand the concepts, we should be able to produce a
mathematical model to describe (and predict) the data.
• For this data, there are, at least, two things that need to be thought through. The
first is the overall shape. The second is, what are those dips?
Historically, it was a fun journey to figure out the overall shape. Physics known
in the 1800s predicted a far different shape. In fact, when challenged with describing
the overall shape, physics failed hard. When physicists tried to predict the spectrum
in Fig. 2.3, the model was close for large wavelengths but very, very wrong for small
wavelengths. Starting at the larger wavelengths and moving to shorter wavelengths,
the math says the spectrum should get larger and larger and larger instead of turning
around and getting smaller, which would require infinite energy. The dramatic result
of the math is this: the universe doesn’t exist. Whenever the model says the universe
doesn’t exist, there is something we don’t understand.1 This result is now known
as the “ultraviolet catastrophe.” It is a dramatic name, but the predictions from the
mathematical model were also dramatic. We clearly needed a better model!
A Short Aside
Imagine you had a room with perfectly reflecting, parallel walls. Only light
with particular wavelengths that create standing waves (perfect constructive
interference like in Fig. 1.7) can exist in this room. Even though only certain
wavelengths are allowed in the room, there are still an infinite number of
them (you can always add one more loop to the standing wave). According
to classical thermodynamics, a well tested and very successful theory, each
standing wave has the same amount of energy. If you learned about heat
capacity in high school chemistry, this classical thermodynamics model
predicted an infinite heat capacity. Yikes!
According to the model, since each mode has the same energy, Fig. 2.3
should keep going up and up at small wavelengths. As a practical example,
this mathematical model says that if you stood in a closed room and lit a
match, the entire room and everything in it would burst into flames due to the
infinite energy density at short wavelengths. Ultraviolet catastrophe indeed!
In 1901, after a lot of thought and model development, German physicist Max
Planck eventually figured out that if light was composed of photons, then there
would be way fewer higher-energy photons than lower-energy ones, which solved
the problem. In 1905, German born physicist Albert Einstein built upon this idea
with the photoelectric effect. The mathematical result of this new model predicted
the shape in Fig. 2.3, which is modeled by the complicated formula:
2π hc2 1
M(λ, T ) =
. , (2.5)
λ5 e λ khcB T − 1
2 Boltzmann’s constant was named by Max Planck in honor of the Austrian physicist and
telling you how much light is being emitted from a blackbody at a given wavelength. If you look at
Fig. 2.3, there is more light being emitted at 500 nm than 700 nm, so the spectral radiant exitance
is larger at 500 nm. This is what we mean by “per unit wavelength.”
2.2 Blackbody Radiation 33
Fig. 2.4 Left: Blackbody spectrum for very hot objects. The surface of the sun is 5800 K. Right:
Your blackbody spectrum, assuming you are human (310 K is about 98 F). Notice the horizontal
axis is a very different scale
The Stefan–Boltzmann Law and Wien’s Displacement Law Imagine you have
a blackbody with surface area A. Besides the spectrum of the blackbody, you can
also measure the total power emitted by the object. The total power emitted is
proportional to the area under a spectral radiant exitance graph. After some math
(calculus), the total power emitted by a blackbody:
P = σ ϵAT 4 ,
.
2π 5kB4 (2.6)
σ = 15 c2 h3 = 5.67 × 10−8 W
m2 K 4
Several years before Max Planck solved the ultraviolet catastrophe, German
physicist Wilhelm Wien experimentally noticed that the wavelength of the
maximum spectral radiant exitance graph, see Fig. 2.4, was inversely proportional
to temperature. He discovered that a blackbody with a temperature of 6000 K had
a maximum around .λpeak = 484 nm, 5000 K had a maximum around 580 nm,
4000 K around 725 nm, etc. After studying the trend between temperature and peak
wavelength, he determined:
b
λpeak =
. , (2.7)
T
4 Named after the Carinthian Slovene physicist, mathematician, and poet Josef Stefan, who
empirically found the relationship, and Austrian physicist and philosopher Ludwig Boltzmann,
who derived the equation.
34 2 “Natural Light”
where .b = 2.898 × 106 nm K. This formula was later derived from the spectral
radiant exitance formula. If you have the calculus background to find the maximum
of a function and want to derive this formula, you should know that, unfortunately,
the resulting equation is a transcendental equation and thus does not have an
analytical solution. However, you can find a numerical solution.
What About Those Dips? Let’s return to the experimental data we took for the
spectrum of the sky, see Fig. 2.3. Blackbody radiation explains the overall shape,
but what about the dips? What do you think is making those dips? The answer is
below the fun little puzzle in Fig. 2.5.
There are two sources of the dips: atoms and molecules in our atmosphere and
the sun itself. For example, the dip near 750 nm is due to oxygen molecules in the
atmosphere. There are photons coming from the sun that have the perfect energy to
excite oxygen molecules. Because oxygen molecules absorb only photons with the
perfect energy, some light is lost at very specific wavelengths before it reaches our
spectrometer.
The second source for the dips is the sun itself. While the sun emits blackbody
radiation, that light passes through gas at the surface of the sun. That gas absorbs
light just like atoms and molecules in our atmosphere. If we wanted just the
composition of the sun, we should collect this data from space!
This leads to one method of performing spectroscopy with a spectrometer. You
take a light source, which can be a blackbody, a flashlight, or really any light source
you want, and send it into your spectrometer. Record this spectrum. Next, place
an atomic sample between the light source and the spectrometer and record this
spectrum. Comparing the two spectra tells you which wavelengths got absorbed.
Each wavelength that is lost from the light due to absorption is the same wavelength
needed to excite the atom from one energy level to another. So measuring these
absorption dips gives you information about the energy levels of the atoms.
2.3 Discharge Lamps 35
Cons:
– The resolution is limited. You need to get different wavelengths far enough apart
to distinguish between them. For example, let’s see how far apart 450.334 nm is
Fig. 2.6 Spectrum collected from a helium discharge lamp collected using a spectrometer with an
accuracy of 0.5 nm
36 2 “Natural Light”
from 450.335 nm using Eq. 2.2 with .L = 10 cm, .m = +1, and .d = 1000 nm:
z(λ) = ± / L
d 2
. ( mλ ) −1
z(450.335 nm) − z(450.334 nm) = 140 nm.
That is small! A possible solution is to make a bigger spectrometer, but then you
will have to worry about thermal drifts, how well you can uniformly space the
grating slits, and how well you can measure d. In short, it becomes really hard if
you want really high resolution.
– The data can be hard to interpret. Getting all the data at once means you are
getting all transitions from your atom or molecule all at once. A spectroscopist
has to figure out what line comes from what transition (pair of energy levels),
which is not an easy task!
For high precision spectroscopy, most spectroscopy groups use a laser to study a
single transition instead of collecting light from many transitions at once. As we
will find in Chap. 4, having atoms moving around (i.e., a hot gas) is a bad thing if
we want to measure something like the energy difference between two energy states
to really high precision. Clever physicists developed a neat technique to still use hot
atoms to measure properties to high precision, which is the topic of Chap. 5, but we
need multiple lasers to do so.
Problems
2.1 Go through the chapter and write down all of the new fundamental constants.
Don’t forget units.
2.2 Go through the chapter and write down each equation. For each equation, write
a brief description about what the equation means.
2.3 What is the frequency difference between light that has a wavelength of
450.334 nm and light that has a wavelength of 450.335 nm? Express your answer
in MHz. Your final answer should have 6 significant figures.
2.5 When introducing gratings we stated, “The only physical criterion for a grating
is that d > λ.” Looking at Eq. 2.2, what would happen if d < λ?
2.6 Figure 2.6 shows the spectrum of a helium discharge lamp collected using a
spectrometer. Below is a table of helium energy levels. Pick any three spectral lines
from the Figure (except for the feature at 389 nm) and determine which transition
produced each line. The lower level for each transition is listed in the third column.
2.3 Discharge Lamps 37
For example, the feature at 389 nm is due to an electron transitioning from level 8
to level 2:
2.7 You want to design your own spectrometer to measure light collected from
a lamp. You need to collect data for light with wavelengths between 1000 and
3000 nm. You have a number of different gratings that you can pick from, each with
a different d. What are a few examples of a bad choice for d? What is an example
of a good choice for d?
(e) Your solar panel is 20% efficient at converting sunlight into usable electric
power. What is the power output of your solar panel? Is it enough to power
a 10 watt LED light bulb?
(f) A refrigerator needs 200 watts of power to run. What area solar panel do you
need to run the refrigerator? If the solar panel was a square, what are the
dimensions of that square?
(a) Figure 2.7 is an amazing figure made by Robert A. Rohde. It shows the spectrum
from the sun collected using an amplitude corrected spectrometer. The vertical
axis is the spectral irradiance.5 Included in the graph is a good approximation
of what the sun would emit if it were a perfect blackbody. Using your favorite
graphing program, estimate the temperature of the surface of the sun by plotting
the spectral radiant exitance. Don’t worry about the vertical scale. What you are
most concerned about is getting the spectral radiant exitance to be a maximum
around 500 nm.
Fig. 2.7 The spectrum from the sun both before the light enters the atmosphere (yellow) and at
the surface of the earth (red). A perfect blackbody spectrum is shown by the black curve. Image
Credit: Robert A. Rohde, CC BY-SA 3.0 via Wikimedia Commons
5 Spectral radiant exitance is the radiant flux emittedby a surface per unit area per unit wavelength.
Spectral irradiance is the radiant flux received by a surface per unit area per unit wavelength. You
can think of spectral radiant exitance as what leaves the sun and spectral irradiance as what hits the
earth.
2.3 Discharge Lamps 39
(b) Use Wien’s displacement law to assess your answer. (In other words, does
Wien’s displacement law confirm the peak of your graph in part (a)?)
(a) Using your favorite numerical program, numerically find and plot the wave-
length of maximum spectral wavelength (λpeak ) for temperatures between 3000
and 6000 K in steps of 100 K.6
(b) Fit your data to b/T to find b. Be sure to use the speed of light accurate to, at
least, 5 digits.
(a) Find the transcendental equation that you would need to numerically solve to
find Wien’s displacement law.
(b) Convert the spectral radiant exitance from a function of wavelength to a
function of frequency. To do this, you need to use the formula. Mλ (λ, T )dλ =
−Mf (f, T )df , which guarantees that the same amount of total energy is in
a spectral interval dλ as in the corresponding interval df . The minus sign is
because decreasing wavelength increases frequency.
(c) Find the spectral radiant exitance as a function of wavenumber (inverse wave-
length).
Open Access This chapter is licensed under the terms of the Creative Commons Attribution 4.0
International License (http://creativecommons.org/licenses/by/4.0/), which permits use, sharing,
adaptation, distribution and reproduction in any medium or format, as long as you give appropriate
credit to the original author(s) and the source, provide a link to the Creative Commons license and
indicate if changes were made.
The images or other third party material in this chapter are included in the chapter’s Creative
Commons license, unless indicated otherwise in a credit line to the material. If material is not
included in the chapter’s Creative Commons license and your intended use is not permitted by
statutory regulation or exceeds the permitted use, you will need to obtain permission directly from
the copyright holder.
6 In python, you can use the code scipy.optimize.fmin from the scipy library. The code only
finds the minimum, so you would multiply the spectral radiant exitance formula by -1 first. In
Mathematica, the function is FindMaximum.
Atoms at Rest
3
Abstract
In this chapter, we consider the factors that lead to complexity in atomic lines. We
will learn that all spectral lines have a fundamental (natural) width and exist as
a spread of frequencies rather than a single frequency. The width of atomic lines
can also be affected by external factors such as pressure and laser power. This will
be related to what we observe when probing an atom with a laser. In addition, we
will learn about how this fundamental width is related to the lifetime of a state,
how many photons per second an atom can absorb from a laser, and the two major
types of experimental plots known as absorption plots and transmission plots.
Learning Goals
• the relationship between the natural linewidth of a transition and the lifetime of
a state.
• absorption plots and transmission plots.
• the scattering rate and its influence on the absorption and transmission plot.
• the saturation intensity of a transition.
• the saturation parameter.
• power broadening.
Definitions
• Resonance: When an atom gets excited by a photon from one state to
another, we say the atom “goes through resonance.” This is similar to
playing the trumpet. When you blow into a trumpet, you excite a standing
wave in the pipe to create a note. The same thing happens with an atom.
When you excite an atom, the electron goes from one standing wave mode
to another. Atomic physicists use the word “excitation” and “resonance”
interchangeably.
• Resonance frequency: When the frequency of the laser is just right to excite
the atom from the ground state to an excited state, we call that the resonance
frequency. We use the variable .fr for resonance frequency. Since resonance
frequency is a frequency, it uses units of Hz, MHz, GHz, or THz.
Now, we do spectroscopy. We start by sending a laser through the vapor cell and
detecting how much light makes it through, see Fig. 3.1. At the beginning of the
thought experiment, the photons in the laser do not have enough energy to excite
the atoms from the ground state to the excited state. In other words, the frequency
of the laser is lower than the resonance frequency. As such, we don’t expect the
3.1 A Thought Experiment 43
atoms to absorb photons from the laser beam. Next, we will smoothly increase the
frequency of the laser until the photons in the laser beam have more energy than the
energy difference between the ground and excited states. In going from too small
to too big, the laser frequency will, at some point, be just right so that the photons
have the right amount of energy to excite the atom, which we call resonance. In the
lab, smoothly changing the laser frequency over time is referred to as “scanning the
laser.”
A transmission plot is a plot of the fraction of photons that make it through
the vapor cell as a function of laser frequency. The question is, what does our
transmission plot look like for this thought experiment? Based on what we’ve
learned so far, a completely reasonable guess would be that the atoms completely
ignore the photons unless the photons have the perfect energy to excite the atom
from the ground state to the excited state. So, you might guess that our transmission
plot looks like the sharp dip in Fig. 3.2. The plot has a single, sharp dip that occurs
at the resonance frequency. If the laser has any other frequency the photons do
not have the correct energy to excite the atom. However, this is not quite right. A
more realistic transmission plot can be seen in Fig. 3.3a. The transmission plot does
indeed have a dip at the resonance frequency, but the dip has a width. This width is
Fig. 3.2 A completely reasonable, but incorrect, guess for the transmission plot. For a transmis-
sion plot, 100% means that no photons are absorbed by the atoms, 50% means that half of the
photons are absorbed, and 0% means that all of the photons are absorbed by the atoms
44 3 Atoms at Rest
Fig. 3.3 (a) A more accurate transmission plot for the thought experiment. (b) A picture of a laser
beam whose frequency matches a resonance frequency near 459 nm (blue light) passing through a
vapor cell of cesium atoms. As we scan the laser frequency from below the resonance frequency
to above the resonance frequency, we visually see no glowing cylinder, followed by a glowing
cylinder (resonance), followed by no glowing cylinder. On resonance, the transmission decreases
as shown in (a) because the atoms take photons from the laser and re-emit them in all directions,
and almost none of these re-emitted photons continue along their original path to the detector. Note
that the transmission plot in (a) is a theory plot assuming that the atoms are at rest. In the vapor
cell the atoms are not frozen in place. Chapter 4 will explore a transmission plot for atoms moving
around
called the natural linewidth of the transition and is represented by the lowercase
Greek letter gamma, .γ . The natural linewidth is a frequency, so it has units of
hertz. Spectroscopists often just say “linewidth” instead of “natural linewidth.” In
an experiment, we would see our atoms start to glow as the laser frequency passes
through resonance, see Fig. 3.3b.
Often times spectroscopists would rather look at an absorption plot instead of
a transmission plot, see Fig. 3.4. An absorption plot is the fraction of photons lost
as a function of laser frequency. It looks very similar to a transmission plot, but it
has a bump instead of a dip. Both an absorption plot and a transmission plot tell us
the same information: atoms absorb (and re-emit) photons from the laser around the
resonance frequency. If you add absorption to transmission, you should get 100%
for all laser frequencies.
Fig. 3.4 An illustrative example of an absorption plot (left) and a transmission plot (right). If you
add these two plots together, you would get 100% for all laser frequencies
3.1 A Thought Experiment 45
Absoption
half maximum of this feature
20%
0%
fr – 2 fr fr + 2
f (THz)
The natural linewidth is the full width at half the maximum (FWHM) of the
absorption bump, see Fig. 3.5. It is a property of the transition that we cannot change.
As an analogy, think about the charge or mass of an electron. The charge of the
electron is simply the charge of the electron, which is .1.602 × 10−19 coulombs. The
mass of the electron is simply the mass of the electron, which is .9.11 × 10−31 kg.
These are intrinsic properties of the electron that we cannot change. The natural
linewidth of a transition is inherent for that transition, and we cannot change it.
The shape of the bump (or dip in the transmission plot) is often referred to as
a spectral feature or spectral profile. For completeness, the width of the spectral
feature we measure in the lab is always larger than the natural linewidth because
of various “broadening” mechanisms. One of these broadening mechanisms is laser
power, which we will discuss in Sect. 3.5. If the laser power was the only broadening
mechanism, we would find that as the laser power gets smaller and smaller, the
width gets narrower and narrower until it reaches the natural linewidth. The natural
linewidth is the minimum possible FWHM of a spectral feature.
The mathematical shape of the spectral feature is a Lorentzian function. In the
absence of any broadening mechanism, the mathematical form is:
A
L(f ) =
. , (3.1)
4(f −fr )2
1+ γ2
where .γ is the natural linewidth1 and A is the maximum absorption, which describes
the amount of laser light lost when traveling through an atomic sample when the
laser light is perfectly on resonance. A is a number between 0 (no light absorbed)
and 1 (all light absorbed, which is also 100%). You may be familiar with a similarly
shaped Gaussian function. A Lorentzian function looks very similar, but it is slightly
different in shape.
1 Thisis for the low power limit. In Sect. 3.5, we will refine this formula slightly to include power
broadening. I just want to start basic to get the concepts first.
46 3 Atoms at Rest
Math Assessment
Equation 3.1 is the mathematical shape of a spectral feature. According to
Fig. 3.5 (and the definition of natural linewidth), the absorption at .f = fr ±
γ /2 should be half as large as the absorption when .f = fr . So, let’s check to
make sure the formula matches that statement. To do this, we will first plug in
.f = fr into Eq. 3.1 to make sure that .L(fr ) = A:
L(fr ) = A
2
1+ 4(fr −f r)
. γ2
= A
=A
1+ 02
γ
L(fr ± γ /2) = A
2
1+ 4(fr ±γ /2−f r)
γ2
. = A
2 = A
2 /4) = A
2 = A
1+1
1+ 4(±γ2/2) 1+ 4(γ 1+ γ 2
γ γ2 γ
= A
2
Yay!
Definitions
• Transmission plot: A plot of the percentage or fraction of photons that
passes through a vapor cell as a function of laser frequency.
• Absorption plot: A plot of the percentage or fraction of photons lost from
a laser after it passes through a vapor cell as a function of laser frequency. A
dip in a transmission plot is seen as a bump in the absorption plot.
• Natural linewidth: The minimum possible full width at half maximum
(FWHM) of a spectral feature. The natural linewidth is a property of a
transition, with each transition in an atom having a unique natural linewidth.
One Last Thing We can now update that important statement from Chap. 1, which
is also at the top of this section.
Many formulas that spectroscopists and atomic physicists use contain variables with
angular units. This is more of a mathematical convenience, but since they are used
so often, I wanted to introduce them as a topic. Suppose we had a simple sine wave
that described an oscillation in time with frequency f and amplitude A. We would
write that as:
A sin (2πf t)
. (3.2)
The presence of .2π in the argument of the sine function stems from the inherent
periodicity of a sine wave, which repeats itself every .2π radians. Consequently, each
time the argument of the sine function increases by .2π , the waveform completes one
full cycle. Thus, the frequency denotes the rate at which the sine wave repeats within
a time span of one second. This periodicity of .2π is where the “angular” part comes
in. Instead of always writing .2πf , we simplify things by using angular frequency,
48 3 Atoms at Rest
ω = 2πf . That curly-looking symbol is the lowercase Greek letter omega. That
.
A sin (ωt).
. (3.3)
𝚪 = 2π γ .
. (3.4)
I want to emphasize that frequency, and not angular frequency, is the unit that we
work with in the⎛lab.⎞To help us keep these two parameters separate,
⎛ ⎞we use the unit
radians/second . rad
s for angular frequency and inverse seconds .
1
s or hertz (Hz)
for frequency. If you are given the information that .𝚪 = 10.5 × 106 rad
s , then we
10.5×106 rad/s
calculate .γ = 2π = 1.67 × = 1.67 MHz for use in the lab. If we
106 1s
want to change the frequency of a laser by one linewidth, we change the frequency
by .1.67 MHz, not .10.5 MHz. Again, to avoid confusion, we don’t use the unit MHz
for the angular frequency. Instead, we make sure to use rad/s: .10.5 × 106 rad
s .
constant resembles a lowercase h with a horizontal line drawn across its stem. We
call this constant “h-bar”. We now have two completely equivalent ways of writing
the energy of a photon:
It doesn’t matter what equation you use, you will always calculate the same energy.
The natural linewidth is, remarkably, related to the lifetime of the excited state. To
understand the lifetime of an atomic state, first imagine that all of the atoms from the
above thought experiment are in the excited state. Next, we turn off the laser beam.
As time progresses, the atoms will each emit a photon to go back to the ground state
at a random time. Spectroscopists often use the word “decay” when talking about
an atom emitting a photon to go back to the ground state. This is a random process,
so some atoms decay back to the ground state quickly while others take their time.
Statistically, each atom has a probability of decaying out of the excited state that will
look like Fig. 3.6. At .t = 0, all the atoms are in the excited state. As time marches
on, atoms start to decay and the excited state fraction gets smaller. For the graph in
Fig. 3.6, about 63% of the atoms have decayed back to the ground state in 6.25 ns,
and about 80% of the atoms have decayed back to the ground state in 10 ns. The
mathematical form of the decay curve is:
e−t/τ ,
. (3.7)
where .τ (this is the Greek letter lowercase tau) is called the lifetime of the excited
state. This function is called an exponential decay. Like the natural linewidth of a
transition, .τ has a unique value for every transition in an atom. The lifetime of the
excited state used for the decay in the graph is .τ = 6.25 ns, which means that after
.6.25 ns about 63% of the atoms have decayed. There is nothing special about 63%;
the atoms have decayed leaving .e−2 = 0.135, or 13.5%, of the atoms in the excited
state.
Amazingly, the natural linewidth and the lifetime are related! The formula
relating the two quantities is:
1 1
τ=
. = . (3.8)
𝚪 2π γ
As you can imagine, atomic physicists and spectroscopists often use the angular
frequency form of natural linewidth when thinking about lifetime.2 These two
quantities are inversely related to each other. This means:
• If the lifetime is small, the natural linewidth is large. This means that excited
states with a small lifetime have absorption and transmission plot spectral
features that are relatively wide.
• If the lifetime is large, the natural linewidth is small. This means that excited
states with a large lifetime have absorption and transmission plot spectral features
that are relatively narrow.
Definitions
• Waist of a laser: The half width of the intensity plot where the intensity
is 13.5% of the maximum intensity, see Fig. 3.7. We use the variable w to
represent waist.
• Intensity: The intensity of a laser beam is the power of the laser divided by
the cross sectional area of the laser beam, denoted as I = P /A.
If you shine a laser on the wall, you will see something that appears to be a circle.
However, it isn’t actually a circle. If we plotted the intensity of the laser, it would
look like Fig. 3.7a. The waist of a laser beam is defined in Fig. 3.7b. It is a bit
Fig. 3.7 (a) The intensity of a laser beam on the wall. (b) The intensity along any one of the
axes. The waist is the half width of the intensity plot when the intensity is 13.5% of the maximum
intensity
2A s
careful reader might notice the units for lifetime look like . rad . You can think of radians as
simply a placeholder to remind us to use radians (.2π) instead of degrees (.360◦ ) in our math when
calculating frequency. We do not keep the radian unit for lifetime. That unit for lifetime is seconds
(or .μs or ns).
3.4 The Scattering Rate and Saturation 51
confusing since most people think of the waist of a laser as the diameter, but it is
more similar to a radius. The intensity profile is described by a Gaussian function.3
The area of a laser beam is:
1
.A= π w2 (3.9)
2
This is, of course, for a perfect laser beam. In real life, the intensity profile is not a
perfect Gaussian function and the waist in the x direction and y directions could be
different. In the latter case, the cross sectional area of a laser is . 21 π wx wy .
How many photons an atom scatters depends upon a few different ideas: (1)
how far the laser frequency is from the resonance frequency relative to the natural
linewidth, (2) how long an atom spends in an excited state, and (3) how a transition
reacts to photons from a laser whose frequency matches the resonance frequency.
Let’s unpack all of these ideas.
(1) We now know that the shape of the absorption bump has the mathematical
form of a Lorenztian function. For spectroscopists, the natural linewidth really
matters. Suppose an atom has a natural linewidth of .γ = 100 MHz and the
laser is .25 MHz below the resonance frequency, see the left spectral profile in
Fig. 3.8. For this transition, the atom will “scatter” (i.e., absorb and re-emit)
many photons from the laser. The right spectral profile in Fig. 3.8 corresponds
to a transition that has a natural linewidth of .γ = 1 MHz. For this narrower
transition, the atom will not scatter many photons at all even though the laser
frequency is still set to .25 MHz below the resonance frequency.
To quantify how far the laser frequency is from the resonance frequency, we
define a new parameter called detuning. Detuning, which is represented by the
lowercase Greek letter delta .δ in normal frequency units and capital delta .Δ =
2π δ for angular frequency units, gives us this information. Mathematically,
detuning is .δ = f − fr , where f is the frequency of the laser. Notice that
if .f < fr , then .δ < 0. In the lab, we call this “red detuning,” which will
3A Gaussian function is very similar to a Lorentzian but it is slightly different. We will explore
Gaussian functions more in Chap. 4.4.
52 3 Atoms at Rest
Fig. 3.8 Two examples for how light from a laser interacts with two different transitions. The left
transition has a large natural linewidth while the right transition has a small natural linewidth. The
laser frequency (red dashed line) is 25 MHz below the resonance frequency for both transitions
make more sense after we discuss the Doppler effect in Chap. 4. Likewise, if
f > fr , then .δ > 0, which we call “blue detuning.” In our thought experiment,
.
the atoms will absorb the largest number of photons when .δ = 0 (the laser
frequency exactly matches the resonance frequency). In the absence of power
broadening (i.e., when the laser power is kept low), the atoms will absorb half
as many photons when .δ = γ /2 or .δ = −γ /2 compared to when .δ = 0.
As discussed above and displayed in Fig. 3.8, detuning is important, but
so is the natural linewidth. What we really want to know is how far the laser
frequency is from the resonance frequency relative to the natural linewidth. For
example, if we set the laser frequency such that .δ = −γ /2, the laser will lose
half as many photons compared to when .δ = 0. This is true for any transition.
The quantity that matters for the scattering rate is the ratio of detuning to
natural linewidth. So, we expect .δ/γ to show up in the relevant scattering rate
equation.
(2) The scattering rate tells us how many photons are absorbed (and re-emitted) by
an atom. From point (1), we know that detuning matters. The average time an
atom spends in the excited state also matters. If an atom, on average, spends
only 1 ns in the excited state, it decays very quickly, freeing itself up to be
excited again. Contrast that with an atom that spends, on average, 2 seconds in
the excited state. That atom will have a very small scattering rate. From Eq. 3.8,
we know that the average lifetime is related to the natural linewidth. A large
natural linewidth means a short lifetime. Therefore, a large natural linewidth
results in a large scattering rate.
(3) Let’s consider two transitions with different resonance frequencies. Transition
#1 has a resonance frequency .fr1 and transition #2 has a resonance frequency
.fr2 . We will assume that .fr1 > fr2 and that both transitions have the
same natural linewidth. We also have two lasers with the same intensity. The
frequency of laser #1 is set to .fr1 and is sent through a vapor cell with atoms
that have transition #1. The frequency of laser #2 is set to .fr2 and is sent
through a vapor cell with atoms that have transition #2. In other words, the
only difference between the two experiments is that transitions have different
3.4 The Scattering Rate and Saturation 53
resonance frequencies. Will both transitions scatter (i.e., absorb and re-emit) the
same number of photons per second, or will the scattering rates be different?
Surprisingly, the answer is that the scattering rates are different for the two
transitions! The reason is that the cross section (or “size”) of the photons is
different for the two transitions. Roughly, the cross section of a photon is .λ2 .
If a transition has a smaller resonance frequency, the laser light has a longer
wavelength and photons with a bigger cross section. That means the photons in
the laser are “big” and more likely to “hit” the atom and cause the transition.
If a photon has a small cross section (high frequency, short wavelength), it is
less likely to hit the atom and cause the transition. For the example above, the
photons from the laser with .f = fr1 have a smaller cross section than the
photons from the laser with .f = fr2 . If we want the same number of scattered
photons from the two transitions, the atom with .fr1 needs to be exposed to
more photons than the transition with .fr2 . In other words, to achieve the same
scattering rate for the two transitions, the transition with .fr1 requires a higher
intensity than the transition with .fr2 .
Summary
(1) Detuning relative to the natural linewidth matters. .δ = 0 should have the largest
scattering rate.
(2) The natural linewidth (or lifetime of an excited state) matters. Large .γ should
have a large scattering rate.
(3) Photon cross section matters. A large frequency, or short wavelength, has a
smaller scattering rate.
To quantify points (2) and (3), we introduce a parameter known as the saturation
intensity .Is . The saturation intensity contains all the information about a particular
transition that helps us understand how easily an atom interacts with photons in a
laser beam whose frequency matches the resonance frequency of a transition (we
will quantify this statement soon). Suppose we send a laser beam whose frequency
matches the resonance frequency for some transition through a sample of atoms. If
an atom absorbs a photon, it will spend some amount of time in the excited state
before decaying back to the ground state, where it is free to absorb another photon.
The saturation intensity is the laser intensity for an on-resonance laser (.f = fr )
such that 25% of the atoms are in the excited state at any given time. A transition
with a small saturation intensity means that we only need a small laser intensity
to have 25% of the atoms in the excited state. A transition with a large saturation
intensity means we need a large laser intensity to make that happen.
The ratio of the intensity of light to the saturation intensity is called the
saturation parameter .s = I /Is . We like to use s because, like .δ/γ , it means the
same thing for every transition. Saying .s = 1 means that we set the laser intensity
equal to the saturation intensity. Some transitions might have a high saturation
intensity, like a transition in the beryllium atom that has .Is = 885 mW/cm2 , while
54 3 Atoms at Rest
other transitions have a low saturation intensity, like a transition in the cesium atom
that has .Is = 0.40 mW/cm2 . For that transition in the beryllium atom, we would
need .885 mW of power for a laser with cross sectional area .A = 1 cm2 to have
.s = 1, which means 25% of the atoms are in the excited state. 885 mW is a lot of
laser power! For the transition in the cesium atom, we only need .0.40 mW of power
to have 25% of atoms in the excited state. The formula for the saturation intensity
is:4
2π 2 hcγ
Is =
. . (3.10)
3 λ3
Notice that the saturation intensity is proportional to .γ . This is point (2). Also notice
that there is a .λ3 in the denominator. One of those .λ terms is grouped with hc in the
numerator; .hc/λ represents the energy of the on-resonance photon. The remaining
2
.λ comes from the cross section of a photon. The saturation intensity is a property
The scattering rate, which is derived using quantum mechanics, is how many
photons per second an atom will absorb (and re-emit). The scattering rate, .r𝚪 (Δ, s)
using angular frequency variables and .rγ (δ, s) using normal frequency variables, is:
r𝚪 (Δ, s) = 𝚪 s
2 1+s+4( Δ )2
𝚪
. (3.11)
rγ (δ, s) = π γ s
.
1+s+4( γδ )2
The equation using normal frequency units is more practical, but you will rarely see
that formula written anywhere. Almost every atomic physics textbook will use the
scattering rate formula that uses angular frequency variables. Notice that the ratio
.δ/γ is in the denominator. If that number gets big, the scattering rate decreases. The
saturation parameter is in the numerator and the denominator, which you might not
have guessed. It makes sense for it to be in the numerator because if I increased the
4 Forcompleteness, this is the formula for doing spectroscopy with linearly polarized laser light.
Experiments with circular polarized light have a slightly different formula.
3.4 The Scattering Rate and Saturation 55
laser intensity there are more photons for the atom to interact with making it more
likely to absorb and emit a photon. But what about the extra s in the denominator?
The first thing to notice is that if the saturation parameter is small, than .1+s ≈ 1,
and s would only be in the numerator. The extra s in the denominator comes from
the fact an atom will always spend some amount of time in the excited state. If we
got rid of the s in the denominator, the number of photons scattered per second (i.e.,
the scattering rate) is linear with laser power. That means if we increase the laser
power by some factor, we increase the scattering rate by the same factor. However,
the atom spends, on average, time .τ = 1/(2π γ ) in the excited state. If the atom
is already in the excited state, it can’t absorb a photon. If all the atoms were in the
excited state, there are no atoms left in the ground state to absorb any photons.
This is important, so let’s explore it a little more. Suppose an atom always spends
a time .τ = 10 ns in the excited state before decaying back to the ground state.5 In
the most extreme case of super high laser intensity, an atom would be immediately
re-excited back to the excited state before having to wait another 10 ns to decay. In
this most extreme case, the atom can only absorb 1 photon every .10 ns. That means,
at most, an atom can absorb . 1 photon
10 ns
1 photon
= 10×10 −9 s = 10 photons every second.
8
Without the s in the denominator, the scattering rate would increase without bound
as the power increases. However, with the s in the denominator, the on-resonance
scattering rate will “saturate” at .π γ .6
Summary of Formulas
rγ (δ, s) = π γ s
2
1+s+ 4δ2
γ
I = P /A
.
A = 12 π w 2 (for a laser beam) (3.12)
s = IIs
δ = f − fr
2
Is = 2π3 hcγ
λ3
One Final Thing: The fraction of atoms in an excited state was one of the key
concepts we used to explore the scattering rate. So, it is no surprise that scattering
rate also tells us what fraction of atoms are in the excited state. The fraction of
atoms in the excited state is .r𝚪 (Δ, s)/ 𝚪 (use the angular frequency formulas for
calculating the excited state fraction).
5 Remember that an atom actually decays probabilistically with a characteristic time .τ . This is just
a thought experiment to understand the concept of saturation.
6 A careful reader might notice the maximum scattering rate is only half as big as our thought
experiment predicted. The idea of the thought experiment is correct, but we are ignoring an effect
known as coherent state transfer, which includes stimulated absorption and stimulated emission.
This is a massive, complicated topic, and one that is beyond the scope of this book. In fact, most
quantum mechanics classes don’t get to this idea until the very end of the semester, if they get to it
at all. Including this extra physics reduces the maximum scattering rate by a factor of 2.
56 3 Atoms at Rest
The scattering rate tells us how many photons per second an atom takes from the
laser. A large scattering rate must correspond to a larger amplitude spectral feature
in an absorption plot. In fact, atoms scattering photons is the only way to produce
a spectral feature. If that is the case, shouldn’t the scattering rate be a Lorentzian
function (Eq. 3.1) like our spectral features? It looks close, but there is that extra s
in the denominator of the scattering rate that is not in Eq. 3.1. While it might not
look like it, the scattering rate is a Lorentzian function. We just need to do a little
algebra to convert the scattering rate into a new form. The algebraic step is to factor
out .1 + s from the denominator:
s ⎛ πγ ⎞ s
.rγ (δ, s) = π γ = . (3.13)
4δ 2
1 + s + γ2 1 + s 1 + 24δ 2
γ (1+s)
√
We are going to define a new parameter .γs = γ 1 + s, known as the power
2 2
broadened linewidth. With this definition, we can write . γ 24δ
(1+s)
as . 4δ
γs2
. To help us
assess this formula, we will also switch the positions of s and .π γ in the numerator.
With those changes, we have:
⎛ s ⎞ πγ
rγ (δ, s) =
. . (3.14)
1 + s 1 + 4δ22
γs
1.5 × 107
1.0 × 107
r ( )
s=100
s=10
5.0 × 106 s=1
0
– 20 – 10 0 10 20
(MHz)
Fig. 3.9 The scattering rate for a transition with .γ = 5.22 MHz as a function of detuning for
different saturation parameters. The red dashed line is the maximum possible scattering rate
16.2 million photons per second .s = 100. For a typical transition, scattering
hundreds of thousands to millions of photons per second is not unusual. Notice that
when s is small, the scattering rate at .δ = −20 MHz is almost 0; mathematically, it
is about 270,000 photons/sec for .s = 1. However, the scattering rate for .s = 100
remains quite sizable.
For completeness, there are other factors that can broaden the width of a
transition including temperature (this is the topic of Chap. 4) and pressure. We do
not cover pressure broadening, also known as collisional broadening, in this book.
3.6 Example
A lot has happened in this chapter, but two of the important concepts are that (1) the
scattering rate tells us how many photons per second an atom absorbs (and emits)
from the laser and (2) the cumulative effect of all the atoms taking photons results
in the absorption profile. Let’s solidify these concepts with an example. Throughout
this example, we will also introduce some commonly used language in atomic
physics. A lot of the language will be explored in more detail as we work our way
through this book.
A particular type of barium atom7 called barium-135 has one ground state and
three closely spaced excited states, see Fig. 3.10. The spacing between the excited
states is called hyperfine splitting, which will be explored in detail in Chap. 9. The
reason there are three excited hyperfine states is because the nucleus has angular
momentum, a concept we will begin exploring in Chap. 7. If the nucleus did not
7 All barium atoms have 56 protons in the nucleus. Neutron number can vary from about 58–97!
We call a particular barium atom, for example an atom with 56 protons and 79 neutrons, an isotope
of barium. Isotopes are explored in Chap. 10.
58 3 Atoms at Rest
Fig. 3.10 A simplified Grotrian diagram for a transition in the atom known as barium-135. The
energy spacings are not to scale. The energy spacings between the excited hyperfine levels are
calculated from the results of Baird et al.[1]. The center of gravity frequency is extracted using
numbers from both Baird et al. [1] and Karlsson et al. [2]. The uncertainty in the center of gravity
frequency is about 30 MHz. [2] While that isn’t a terrible uncertainty, modern day spectroscopic
methods can do better!
have angular momentum, there would be only one excited state.8 Again, we will
learn the physics behind hyperfine splitting starting in Chap. 7. Before we do an
example, we need a few more definitions.
Definitions
• Center of gravity: The energy of a state if the nucleus had no angular
momentum.
• Hyperfine splitting: When the nucleus has angular momentum, the single
energy level at the center of gravity splits into multiple energy states. Each
of these states is going to shift in energy a small amount compared to the
center of gravity energy.
• Bra-ket notation: Atomic physicists sometimes use “bra-ket” notation
when working with states in an atom. A “ket” is a way to represent a
particular state and it looks like this: .|put state label here〈. A “bra”, which is
not used in this book but you will use it a lot if you take quantum mechanics,
looks like this: .〉put state label here|. Bra-ket notation is also referred to
as Dirac notation, named after the English mathematical and theoretical
physicist Paul Dirac.
8 There are atoms with ground state hyperfine splitting; barium-135 is just not one of those atoms.
3.6 Example 59
Note
Some learners may use only Part 1 of this book. I wanted to introduce you to
bra-ket notation so that you have seen it at least once before you take quantum
mechanics. We are going to use it in this example and in Problem 3.9, but it
will not be used again until Part 2. You are, of course, welcome to use bra-ket
notation if you want, but it is not necessary. You may also see bra-ket notation
if you take a linear algebra class.
| 〉
As seen in Fig. |
| 3.10,〉 we | are going
〉 to| label the
〉 ground state . gF =3/2 and the three
excited states .|eF =1/2 , .|eF =3/2 , and .|eF =5/2 . F is called a quantum number, and
it is always a positive integer, a half integer, or zero; quantum numbers are explored
starting in Chap. 6. For now, these quantum numbers are just being used to label
our states. Notice there is a center of gravity frequency, .fcog , that tells us the energy
difference between the center of gravity for the two states. This tells us that our
transitions are all around 541.4 THz, or 553.7 nm. Each hyperfine level is shifted
from their center of gravity state by a small amount. I want to emphasize that if a
nucleus has angular momentum, the center of gravity states do not exist in real life!
The hyperfine states are the actual states. However, we can learn a lot of physics
by determining how the hyperfine levels shift from the center of gravity, which is
explored in Chaps. 9 and 10.
With that nice long intro, let’s get into the actual example. We have four states:
one ground state and three excited states. We are going to assume that our atoms are
all at rest and send a laser beam| through
〉 the sample, see Fig. 3.1. All of the atoms
will start in the ground state .|gF =1/2 . An electron can be excited to a higher energy
level as long as the following rule, derived from quantum mechanics, is satisfied: A
transition can change F by .−1, 0, or 1. Memorize this rule! For completeness, there
is one exception to this rule. If the ground state has .F = 0, that electron cannot be
excited to an .F = 0 excited state.9
The Rule
. ΔF = −1, 0, + 1 (3.15)
F = 0 /→ F = 0
In this example, an electron in the barium ground state can be excited to any
of the excited states. However, if there were an excited state with quantum number
.F = 7/2, an electron in the ground state could not be excited to this state because
.ΔF = 2.
9 We only need this rule for the moment. A full list of the rules that need to be satisfied for an
electron to transition between two atomic states is given in Appendix C.
60 3 Atoms at Rest
Fig. 3.11 A simulated absorption plot for a transition in barium-135. In this example, we are
assuming that all the atoms are at rest and the experimental setup is shown in Fig. 3.1. The
horizontal axis is the laser frequency with respect to the center of gravity frequency
Problems
3.1 Go through the chapter and write down all the new fundamental constants.
Don’t forget units.
3.2 For each of the formulas in Eqs. 3.12 and 3.15, write a brief description of what
each equation means.
3.6 Example 61
3.3 In the lab, we measure laser power. However, we actually care about laser
intensity. Why?
3.5 A transition in the Europium atom has a natural linewidth of γ = 25.5 MHz.
The wavelength of light at the resonance frequency is λ = 466.188 nm. Calculate
the saturation intensity in units of mW/cm2 and mW/mm2 .
Hint: 1 W=1 J/s
3.6
(a) For the transition in Problem 3.5, calculate the on resonance scattering rate (δ =
0) for a saturation parameter of 0.1, 1, 5, 10, and 100.
(b) Find the excited state fraction for each of the above saturation parameters.
3.8 If a laser beam has a waist of w = 1 mm, what power should we set the laser in
order to get saturation parameters of 0.1, 1, 5, 10, and 100? Assume the saturation
intensity is Is = 1.2 mW/mm2 .
3.9 (Graphing Problem) Plot the scattering rate versus detuning for a transition
with γ = 10 MHz for different saturation parameters of 0.1, 1, 5, 10, and 100.
(a) Using the rule shown in Eq. 3.15, find the resonance frequency for all possible
transitions (there are five of them).
62 3 Atoms at Rest
Fig. 3.12 A simplified Grotrian diagram for a transition in the atom known as rubidium-80. The
energy spacings are not to scale. The energy spacings are taken from the work of Thibault et al. [3]
References
1. Baird, P.E.G., Brambley, R.J., Burnett, K., Stacey, D.N., Warrington, D.M., Woodgate, G.K.:
Optical isotope shifts and hyperfine structure in λ553.5 nm of barium, Proc. R. Soc. Lond.
A365567–365582 (1979). http://doi.org/10.1098/rspa.1979.0035
2. Karlsson, H., Litzén, U.: Revised Ba I and Ba II wavelengths and energy levels derived by
fourier transform spectroscopy. Phys. Scripta 60, 321 (1999). https://doi.org/10.1238/Physica.
Regular.060a00321
3. Thibault, C., Touchard, F., Büttgenbach, S., Klapisch, R., de Saint Simon, M., Duong, H.T.,
Jacquinot, P., Juncar, P., Liberman, S., Pillet, P., Pinard, J., Vialle, J.L., Pesnelle, A., Huber, G.:
Hyperfine structure and isotope shift of the D2 line of 76–98Rb and some of their isomers. Phys.
Rev. C 23, 2720 (1981). https://doi.org/10.1103/PhysRevC.23.2720
References 63
Open Access This chapter is licensed under the terms of the Creative Commons Attribution 4.0
International License (http://creativecommons.org/licenses/by/4.0/), which permits use, sharing,
adaptation, distribution and reproduction in any medium or format, as long as you give appropriate
credit to the original author(s) and the source, provide a link to the Creative Commons license and
indicate if changes were made.
The images or other third party material in this chapter are included in the chapter’s Creative
Commons license, unless indicated otherwise in a credit line to the material. If material is not
included in the chapter’s Creative Commons license and your intended use is not permitted by
statutory regulation or exceeds the permitted use, you will need to obtain permission directly from
the copyright holder.
Atoms in Motion
4
Abstract
In this chapter, we explore how motion can affect the perceived frequency of
waves, with significant implications for spectroscopy. We cover the broadening
and shifting of atomic line frequencies due to atomic movement and investigate
the roles of velocity and temperature in these phenomena. Key topics include
the Doppler effect, Doppler broadening, the Maxwell-Boltzmann velocity distri-
bution, and the Equipartition Theorem. Additionally, the chapter discusses the
application of the Doppler effect in astronomy.
Learning Goals
The Doppler effect is likely a phenomenon you have encountered before. When
an ambulance, police car, or racecar travels past you, the sound you hear changes
pitch. This happens because the motion of the vehicle compresses or extends the
Fig. 4.1 Left: An ambulance at rest emitting a sound wave from its siren. The wavelength λ0 and
frequency f0 of the sound wave is the same in all directions. Right: Now the ambulance is moving
to the right. The sound wave in front of the ambulance is compressed, which means the perceived
wavelength is smaller and the perceived frequency is larger (higher pitch). The sound wave behind
the ambulance is expanded, which means the perceived wavelength is larger and the perceived
frequency is smaller (lower pitch)
sound waves. Figure 4.1 shows the sound waves emitted by a stationary ambulance
(left) and a moving ambulance (right). Let’s focus on the stationary ambulance.
Imagine that you are standing in front of or behind the ambulance. The wavelength
of the sound wave that hits your ear is the same for both scenarios, so you would
hear the same pitch independent of where you are standing. Now, imagine the
ambulance is moving. If you were standing in front of the ambulance (OK, maybe
a bit to the side . . . we don’t want you to get hit, even in a thought experiment),
the wavelength of the sound wave that reaches your ear is shorter compared to
the stationary ambulance. If you were standing behind the moving ambulance, the
wavelength is longer compared to the stationary ambulance. The formula that relates
the frequency (pitch) that you hear to the wavelength should look really familiar.
It is .vs = f λ, where .vs is the speed of sound in air (replace .vs with c and you
have Eq. 1.1 from p. 19). The apparent shift in frequency due to an object moving
is known as the Doppler effect, named after Austrian physicist and mathematician
Christian Doppler. It is a very important concept in spectroscopy.
The Doppler effect occurs for any type of wave. Whether it is a sound wave, a
light wave, or a water wave created by a duck swimming in a pond, the relative
motion of the object with respect to the observer will change the wavelength, and
thus the frequency of the wave. The EMT driving the ambulance hears no change in
pitch because they are stationary with respect to the siren. If you yelled positive
4.2 Laser Frequency From an Atom’s Perspective 67
encouragement at the ambulance as it passed, the driver would hear your pitch
change as they passed by you. Likewise, you don’t hear your pitch change as the
ambulance passes by you. What is important here is that the Doppler effect is
something experienced by the observer because the source of the wave is moving
with respect to them.
Definition
• Doppler effect: An increase or decrease in the frequency of sound, light,
or other waves as the source and observer move toward or away from each
other.
Important Reminder
Frequency, energy, and wavelength are all the same quantity. Each of these
parameters is related to the other parameters only by constants.
Imagine a laser beam traveling to the right, as shown in Fig. 4.2. Also imagine
there are three atoms: atom 2 is traveling to the left, atom 1 is stationary, and atom
3 is traveling to the right. For this thought experiment, we will assume the speeds of
atom 2 and atom 3 are the same, just in opposite directions.
In this experiment, the atom is the observer because it is interacting with the
laser light and not producing it. To understand the Doppler effect, it is important
to recognize that each atom perceives itself as stationary. Atom 2 would claim that
68 4 Atoms in Motion
2 1 3
Laser
Fig. 4.2 A simple experimental to explore how motion of atoms impacts the interactions between
the atoms and laser light
atom 1 is moving to the right and that atom 3 is moving twice as fast as we (as the
scientists looking from the outside) would say atom 3 is moving. Both atom 1 and
us, as the observing scientists, will agree on the frequency of the laser. Because of
the Doppler effect, atoms 2 and 3 will disagree. To make this idea a little clearer,
let’s say that the laser frequency is .652.0000 × 1012 Hz = 652.0000 THz (terahertz)
and that this is the resonance frequency for the atom. Both the scientists and atom
1 will agree that the laser frequency is .652.0000 THz; atom 1 will absorb photons
from the laser beam. However, atom 2 and atom 3 will disagree with this claim since
atom 2 is moving towards the laser and atom 3 is moving away from the laser.
Answer Atom 2 is moving towards the laser source, so it will perceive the laser
frequency as higher than it actually is. Atom 2 will only absorb a photon from the
laser if it thinks the laser frequency matches the resonance frequency. Therefore, we,
in the observing frame, need to set the laser frequency smaller than the resonance
frequency .(652.0000 THz) so that the actual laser frequency plus the frequency shift
due to the Doppler effect results in the resonance frequency in the frame of atom 2.
In equation form, this is represented as:
4.2 Laser Frequency From an Atom’s Perspective 69
fatom2 = fL + ΔfD ,
. (4.1)
where .fatom2 is the laser frequency according to atom 2, .fL is the actual laser
frequency (i.e. the frequency measured in the laboratory/stationary frame), and .ΔfD
is perceived shift in frequency due to the Doppler effect. In this example, .ΔfD > 0
for atom 2, so if we want .fatom2 = fr , then we need to set the laser frequency
smaller than the resonance frequency such that .fr = fL + ΔfD .
Likewise, atom 3 is moving away from the laser source, so it will claim the laser
frequency is lower. As a result, the actual laser frequency will have to be higher than
the resonance frequency for atom 3 to absorb a photon.
• From atom 2’s reference frame, the frequency of the laser is just right to excite
atom 2 from the ground state to the excited state.
• From the laboratory reference frame, the frequency of the laser is too low.
• The Doppler effect tells us that because atom 2 is traveling towards the
laser beam, it will see a higher frequency than what we, as scientists in the
laboratory/stationary frame, measure.
Definitions
• Doppler shift: The shift in the frequency of a laser seen by an atom due to
the Doppler effect.
• .v‖ : The velocity component of an atom in the direction of the laser beam.
.v‖ is negative if the atom is traveling towards the laser and positive if it is
The frequency of laser light as seen by the atom is given by the formula:
⎛ ⎞
v‖
fatom = fL 1 −
. (4.2)
c
where .fatom is the frequency of the laser as seen by the atom, .fL is the frequency
of the laser measured in the lab (rest) frame, and .v‖ is the velocity component in
the direction of (parallel to) the laser beam. Using .c = fL λ, the right-hand side of
Eq. 4.2 can be written as:
v‖
fatom = fL −
. , (4.3)
λ
where .λ is the wavelength of light measured in the laboratory frame.
Important
v‖ is a weird variable in that the sign of .v‖ depends on whether the atom
.
is moving towards (.v‖ < 0) or away (.v‖ > 0) from the laser. Messing up
the sign of .v‖ is a very common mistake when using this formula. When
in doubt, just remember that an atom moving towards the laser beam sees a
higher frequency.
Comparing Eqs. 4.1 and 4.3, we find the formula for the Doppler shift:1
v‖
ΔfD = −
. . (4.4)
λ
Fig. 4.3 (a) Only the component of velocity in the direction of the laser beams results in a Doppler
shift. The first two atoms have different velocities, but the same component in the direction of the
laser, .v‖ . As such, they will experience the same Doppler shift. The third atom has a velocity
component in the opposite direction, so it will have a different Doppler shift. The last atom is
completely stationary. (b) All three of these atoms have no velocity component in the direction of
the laser, so they all have zero Doppler shift
A Bit More About Velocity Components The velocity component in the direction
of the laser beam is an important, but sometimes confusing, idea when you first
encounter it. So, let’s spend a bit more time thinking this idea through using Fig. 4.3.
In Fig. 4.3a, the first atom’s velocity is pointing directly towards the laser, so .v‖ < 0.
For this atom, there is no perpendicular component to the atom, .v⊥ = 0. If the laser
was traveling towards the left, .v‖ > 0 for this atom because the sign of .v‖ only
depends upon if the atom is moving towards or away from the laser beam.
The second atom has both a perpendicular component and a parallel component.
Only the parallel component causes the Doppler shift, and the parallel component
tells us the atom is moving towards the laser source, so .v‖ < 0. Notice the parallel
component for the two first two atoms are the same size and pointing in the same
direction. Therefore, they will have the same Doppler shift.
The third atom has a perpendicular component, which we don’t care about, and a
parallel component pointing away from the laser, so .v‖ > 0. This atom will absorb
photons with a different laser frequency than atoms 1 and 2. The last atom is not
moving at all. It has no perpendicular or parallel component: .v‖ = 0 and .v⊥ = 0.
Figure 4.3b shows three examples of atoms with .v‖ = 0. Each of these atoms will
absorb photons when .fL = fr .
Important
In high precision spectroscopy, we want to extract information from atoms
that have .v‖ = 0. The rest of the atoms in our sample make our spectrum
(continued)
72 4 Atoms in Motion
less precise. How the Doppler shift changes an absorption plot is going to be
explored in the rest of this chapter. Chapter 5 introduces a clever experimental
trick known as saturated absorption spectroscopy. This technique allows us to
remove the issues brought about from the Doppler shifts that we are about to
discuss.
Fig. 4.4 A simulated transmission plot (left) and absorption plot (right) for the three atoms
4.4 The Maxwell-Boltzmann Velocity Distribution 73
Table 4.1 A table summarizing the Doppler shifts for the 3 atoms
Atom 2 Atom 1 Atom 3
Motion Towards laser Stationary Away from laser
The atom sees Higher frequency light Actual laser frequency Lower frequency light
Resonance happens At lower frequency At actual frequency At higher frequency
the resonance frequency. The spectral feature for atom 3 is at .δ = +150 MHz. This
important information is summarized in Table 4.1.
In a real vapor cell, atoms are moving with all sorts of different velocities.
Unlike energy levels in an atom, velocity is a continuous variable. In the next
two sections, we are going to discuss the distributions of velocities inside a real
vapor cell (Sect. 4.4) and use that information to develop what the transmission and
absorption plots look like for a vapor cell of atoms at a given temperature (Sect. 4.5).
2 Named after the Scottish mathematician James Clerk Maxwell and the Austrian physicist Ludwig
Boltzmann.
74 4 Atoms in Motion
where .kB = 1.38 × 10−23 J/K is the Boltzmann constant, T is the temperature of the
gas (the unit is kelvin), and m is the mass of an atom in the gas (the unit is kilogram).
The hotter the gas, the wider the velocity distribution, and thus the larger the average
speed of the atoms in the gas. The mass of an atom is in the denominator, so atoms
with larger masses have smaller average speeds compared to an equally hot gas of
smaller mass atoms.
How to Use Distributions This section isn’t really needed to understand the
velocity distribution. However, distributions are incredibly important in many areas
of science, so I wanted to spend a bit of time talking about how we use them. The
velocity distribution represents the fraction of atoms that fall within a particular
velocity range. Since an atom must have some velocity, the total area under the
curve in Fig. 4.5 is 1. This is equivalent to saying that each atom must have some
velocity between .+∞ and .−∞.
To use a distribution, you ask questions like, “What fraction of atoms have a
positive parallel velocity component?”, “What fraction of atoms have a parallel
velocity component between .−2 and 10 m/s?”, or more generally, “What fraction
of atoms have a parallel velocity component between .va and .vb ?”, where .va and .vb
are any velocities we choose.
The answer is the area under the distribution between .va and .vb . If we want to
know the total number of atoms from the sample that have velocity components
between those two values, we multiply that fraction by the total number of atoms in
the sample. For example, here are three plots with different choices of .va and .vb :
Suppose we have 5000 atoms in our sample. As shown in the first plot, there
are .5000 × 0.726 = 3630 atoms with parallel velocity components between .va =
−100 m/s and .vb = +200 m/s. As shown in the second plot, there are .5000 ×
0.061 = 305 atoms with parallel velocity components between .va = −200 m/s and
.vb = −150 m/s, and in the third plot there are .5000 × 0.347 = 1735 atoms with
Important Reminder
Due to the Doppler effect, atoms with different velocities will absorb photons
at different laser frequencies.
A Gaussian looks similar to the Lorentzian function that models the shape of a
spectral feature, but the two functions are different. Figure 4.6 shows a plot of both a
Gaussian function and a Lorentzian function with the same area and FWHM. Notice
the Lorentzian has larger “tails” and is more spread out compared to the Gaussian
function. Both functions are very common in physics and math.
Extra Math for Those Who Have Taken Statistics In statistics, Gaussian func-
v2
−
tions are written as .e 2σ 2
, where .σ is called the standard deviation. / For the
Maxwell-Boltzmann velocity distribution, the standard deviation is . kBmT . The
FWHM of a Gaussian function defined using the standard deviation is:
/
√ kB T
ΔvFWHM = 2 2 ln (2)σ ≈ 2.355σ = 2.355
. (4.7)
m
3 This is a distribution for 1 dimension since we are only interested in the velocity component for
a single direction. In the future, you might encounter a Maxwell-Boltzmann velocity distribution
that has a power of 3/2 instead of 1/2 on the expression in front of the Gaussian function. That
would be a velocity distribution for all components of velocity, not just the parallel component.
76 4 Atoms in Motion
Definitions
• Doppler broadening: The widening of a spectral feature due to a vapor cell
having a given temperature.
• Doppler profile: The name given to a spectral feature that is broadened
because of temperature.
• Doppler width: The FWHM of a Doppler profile.
In Sects. 4.2 and 4.3, we explored a transmission plot with only three atoms with
different velocities. What if we had one hundred thousand atoms? Figure 4.7 shows
1.00000001 1.00000001
1.00000000
1.00000000
Transmission (mW)
Transmission (mW)
0.99999999
0.99999999
0.99999998
0.99999998 0.99999997
0.99999996
0.99999997
0.99999995
0.99999996
0.99999994
0.99999995 0.99999993
– 1000 – 500 0 500 1000 – 1000 – 500 0 500 1000
(MHz) (MHz)
1.0000001 1.0000010
1.0000005
Transmission (mW)
Transmission (mW)
1.0000000 1.0000000
0.9999995
0.9999999 0.9999990
0.9999985
0.9999998 0.9999980
0.9999975
0.9999997 0.9999970
– 1000 – 500 0 500 1000 – 1000 – 500 0 500 1000
(MHz) (MHz)
1.000000 1.00000
0.999995 0.99995
Transmission (mW)
Transmission (mW)
0.999990 0.99990
0.999985 0.99985
0.999980 0.99980
0.999975 0.99975
0.999970 0.99970
– 1000 – 500 0 500 1000 – 1000 – 500 0 500 1000
(MHz) (MHz)
the results for a simulation of transmission plots as we add more and more atoms
to a vapor cell. For this simulation, I assumed that we had a two-level atom with
a mass of .m = 2.33 × 10−26 kg (this is the mass of a nitrogen atom), a vapor
cell temperature of .T = 400 K, an excitation wavelength of .λ = 940 nm (.f ≈
319 THz), and a natural linewidth of .γ = 5 MHz. I randomly picked a velocity
component using the Maxwell-Boltzmann velocity distribution for each atom I add
to the cell.
Note that each transmission plot has a different vertical scale. Individually, a
single atom isn’t going to absorb a large fraction of a laser’s photons. However, the
more atoms you have interacting with the light, the more spectral features you have
piling up on each other. Ultimately, you get a transmission plot that looks like it has
a single feature.
This feature, which is called a Doppler profile, is much wider than a spectral
feature from a single atom and is pretty close to the same shape as .1−Af (v‖ ), where
A is some constant and .f (v‖ ) is the Maxwell-Boltzmann velocity distribution.
Notice the center of the Doppler profile is still at the resonance frequency .δ = 0.
The width of a Doppler profile can be found from FWHM of the Maxwell-
Boltzmann velocity distribution, which is a velocity. We can convert this velocity
to a frequency using the Doppler shift formula. The FWHM of a spectral feature
broadened by temperature, which is called the Doppler width, is given by the
formula:
/
2.355 kB T
.ΔfFWHM = (4.8)
λ m
1 2
.K= mv . (4.9)
2
Since kinetic energy is a type of energy, the unit is a joule. Imagine you have 3 atoms
in your gas. We will assume that all the atoms have the same mass but different
speeds. The average kinetic energy of the atoms in the gas would be:
⎛ ⎞
1 1 2 1 2 1 2
. mv1 + mv2 + mv3 . (4.10)
3 2 2 2
78 4 Atoms in Motion
If we had N atoms in our gas, all with the same mass, the average kinetic energy
would be:
⎛ ⎞ ⎛ ⎞
1 1 2 1 2 1 2 1 v12 + v22 + ... + vN
2
. mv + mv + ... + mvN = m . (4.11)
N 2 1 2 2 2 2 N
That last term that is in parentheses is called the average squared speed. We denote
this last term as .〈v 2 〉. In fact, whenever you see the mathematical expression
between two√ angle brackets, .〈 〉, you are being asked to take the average of that
property. . 〈v 2 〉 has a special name which is called the root mean squared speed
.vrms , which you may have learned about in high school Chemistry. Putting this all
together, we find that the average kinetic energy of all the atoms in the gas is:
1
〈K〉 =
. m〈v 2 〉. (4.12)
2
Reading the Above Equation For a gas composed of atoms with the same mass,
the average kinetic energy of the atoms is proportional to the average squared speed.
The Equipartition Theorem tells us that the energy of a gas is equally distributed
among all “degrees of freedom.” Degrees of freedom indicate the number of ways
an atom or molecule can move. Atoms have three degrees of freedom because they
can move in three dimensions. Molecules have more degrees of freedom because
they can rotate and vibrate, so a molecule has more ways to distribute its energy
than an atom. The Equipartition Theorem tells us that each degree of freedom has
1
. kB T of energy. In this book, we are only working with atoms, but, in the future, if
2
you work with molecules, the following formulas will be slightly different.
Imagine you have a vapor cell of atoms at some temperature T . The average
kinetic energy of the atoms in the gas is:4
1 3
. m〈v 2 〉 = kB T . (4.13)
2 2
The 3 on the right hand side represents
⎛ ⎞ the three degrees of freedom of an atom
1
. k
2 B T for each degree of freedom . This formula is very useful because it directly
relates the temperature of a sample of atoms to a characteristic speed of the atoms,
specifically .〈v 2 〉.
4 You can derive this formula from the Maxwell-Boltzmann velocity distribution, but you will need
to use calculus.
4.7 Application to Astronomy: Light from the Stars 79
In our experiment, both we, the scientists, and the laser light source are stationary
while the atoms, which act like the observers of the laser light, are moving. The
same principles of the Doppler effect apply whether the light source is moving and
the observer is stationary, the light source is stationary and the observer is moving,
or if both are moving. All that matters is whether the source and observer are moving
towards each other or away from each other.
The Doppler effect is a powerful tool in astronomy. Suppose we are using a
telescope to collect light from a distant star that is mostly composed of hydrogen
gas. That star is emitting light with frequencies corresponding to the difference of
the hydrogen energy levels. We now know that if the star is moving towards us, the
frequency of light leaving that star will look to us to have a higher frequency than
what we would observe if we just had a hydrogen light bulb in our lab. In astronomy,
this phenomenon is called blue-shifted light because the light has a higher frequency
than we would expect if we measured the spectrum of hydrogen here on earth. If the
star is moving away from us, which is far more common in astronomy, the frequency
of light that is emitted from the star looks to be lower frequency compared to what
we would measure from a source here on earth. This is called red-shifted light.
In summary, if the star is moving towards us, we will see a spectrum that is
shifted to higher frequencies compared to what we measure in the lab (blue-shifted;
.v < 0), and we will see a shift to lower frequencies if the star was moving away
fem
fobs = 1+z
. (4.14)
z = vc ,
where .fobs is the Doppler shifted frequency measured on earth, .fem is the frequency
of the light emitted from the star, and v is the speed of the star or galaxy in the
direction of earth. Astronomers also use the parameter .z = vc to describe blue-shift
light (.z < 0 → v < 0; the star is moving towards the earth) and red-shifted light
(.z > 0 → v > 0; the star is moving away the earth). You will have the opportunity
to derive this formula in Problem 4.7.
Finally, astronomers like to use wavelength instead of frequency. Writing
Eq. 4.14 using wavelength and solving for z gives:
z = λλobs − 1,
. em (4.15)
the formula astronomers use
80 4 Atoms in Motion
Problems
4.1 For each of the following equations, write a brief description of what each
equation means.
4.2 Assess Eq. 4.8. The purpose of any assessment is to increase or decrease our
confidence in something. Assessments are challenging because we inherently want
our calculations to be correct! To combat this bias for assessing a formula, I find
it is easiest to write down all the parameters on the right hand side and then try to
forget the formula all together. Then you ask yourself the question, “If I increased
T , then the Doppler width should get because .” You need to decide
if “larger” or “smaller” goes into the first blank and explain, using a physics reason,
why that should happen in the second blank. Next, repeat that process for every
parameter. After I think through each parameter, I go check the formula to make
sure my statements match the formula.
If your statement does not match your formula, then either your formula is wrong
or your reasoning is wrong. Either way, you now have an opportunity to learn
something! But, more importantly, you will understand an equation more after you
assess it.
4.3 An atom at rest is excited from the ground state to an excited state by a photon
from a laser with frequency f = 315.11254 THz.
(a) Suppose the laser is positioned to send photons to the right, and an atom is
moving towards the laser with a velocity component of v‖ = −200 m/s (the
minus sign indicates the atom is moving towards from the laser), see atom 2
from Fig. 4.2 on p. 68. What frequency should the laser be for this atom to
absorb a photon?
(b) Now the laser is pointed to send photons to the left, so now the atom is moving
away from the laser source. What frequency should the laser be for this atom to
absorb a photon?
4.4 Explain qualitatively how the motion of an atom affects the energy (frequency)
of a photon it will absorb compared to an atom at rest. Specifically, describe the
difference in photon energy required for an atom moving towards the light source
versus an atom moving away from the light source.
4.5 A vapor cell has strontium-84 atoms. A strontium-84 atom has 38 protons and
46 neutrons (notice 38 + 46 = 84). The mass of a strontium-84 atom is 1.393 ×
10−25 kg.
4.7 Application to Astronomy: Light from the Stars 81
(a) If the temperature of the vapor cell is 350 K, what is the full width at half
maximum of the Maxwell-Boltzmann velocity distribution?
(b) There is a transition from the ground state to an excited state at 650.5032 THz.
There are no other energy levels nearby, so you can treat this transition as a
two-level atom. What is the Doppler width for this spectral feature? Give your
answer in MHz.
(c) Sketch the transmission plot of a laser beam as it passes through a vapor cell
held at 350 K. You can pick any amplitude you want for the Doppler feature.
Hint: You should be using your answer from part b) in this sketch.
4.7 (The Full Doppler Shift Formula: Moving Observers and Sources) In non-
relativistic physics, the formula for the Doppler shift for a light wave is:
⎛ ⎞
c ± vobs
fobs =
. fem , (4.16)
c ∓ vem
where fobs is the frequency measured by the observer and fem is the frequency
emitted by the source. vobs is the speed of the observer relative to some background
and is always a positive number (it is a speed). It is added to c in the numerator if
the observer is moving towards the source and subtracted if the observer is moving
away from the source. vem is the speed of the source with respect to that same
background and is also always a positive number (it is a speed). It is added to c in
the denominator if the source is moving away from the observer and subtracted if
the source is moving towards the observer.
(a) Come up with 4 different scenarios for the 4 different sign combinations. For
example, what is a scenario where the observer is moving away from the source
and the source is moving towards the observer? In this scenario, you would use
the formula:
⎛ ⎞
c − vobs
.fobs = fem , (4.17)
c − vem
(b) Check to make sure this formula agrees with Eq. 4.2.
(c) Check to make sure this formula agrees with Eq. 4.14.
4.8 Figure 4.8 is a picture of the spectrum from a distant galaxy that you can
download from the Sky Server database.5 The Sky Server ID for this galaxy is
5 Image and data is from the Sloan Digital Sky Survey. Funding for the Sloan Digital Sky Survey
(SDSS) has been provided by the Alfred P. Sloan Foundation, the Participating Institutions,
the National Aeronautics and Space Administration, the National Science Foundation, the U.S.
82 4 Atoms in Motion
Department of Energy, the Japanese Monbukagakusho, and the Max Planck Society. The SDSS
Web site is http://www.sdss.org/.
The SDSS is managed by the Astrophysical Research Consortium (ARC) for the Participating
Institutions. The Participating Institutions are The University of Chicago, Fermilab, the Institute
for Advanced Study, the Japan Participation Group, The Johns Hopkins University, Los Alamos
National Laboratory, the Max-Planck-Institute for Astronomy (MPIA), the Max-Planck-Institute
for Astrophysics (MPA), New Mexico State University, University of Pittsburgh, Princeton
University, the United States Naval Observatory, and the University of Washington.
4.7 Application to Astronomy: Light from the Stars 83
Use the data graphed in Fig. 4.8 to estimate the wavelengths of those
lines that astronomers measured here on earth and find the speed of galaxy
582102012537667624 relative to the earth.
Open Access This chapter is licensed under the terms of the Creative Commons Attribution 4.0
International License (http://creativecommons.org/licenses/by/4.0/), which permits use, sharing,
adaptation, distribution and reproduction in any medium or format, as long as you give appropriate
credit to the original author(s) and the source, provide a link to the Creative Commons license and
indicate if changes were made.
The images or other third party material in this chapter are included in the chapter’s Creative
Commons license, unless indicated otherwise in a credit line to the material. If material is not
included in the chapter’s Creative Commons license and your intended use is not permitted by
statutory regulation or exceeds the permitted use, you will need to obtain permission directly from
the copyright holder.
Saturated Absorption Spectroscopy
5
Abstract
Learning Goals
1.0 1.0
0.8 0.8
Transmission
Transmission
0.6 0.6
0.4 0.4
0.2 0.2
0.0 0.0
fA fr fB fA fr fB
f (MHz) f (MHz)
Fig. 5.1 An illustrative example showing the transmission plots for a two-level atom. On the left is
the Doppler profile we learned about in Chap. 4. On the right is the transmission plot for a saturated
absorption setup
sentence! It is really quite amazing. Suppose the atoms are at 400 K. We know that
if we use a single laser beam, we would expect to see a Doppler broadened spectrum
from these atoms that is Gaussian in shape. Saturated absorption spectroscopy uses
two laser beams, resulting in a small Lorentzian feature on top of the Gaussian
shape, as shown in Fig. 5.1. The small Lorentzian feature comes only from those
atoms that have zero speed in the direction of the laser.1
This is how we do it: we send two laser beams into a vapor cell from opposite
directions, see Fig. 5.2. The laser beam that starts on the left and moves to the
right has a small amount of power. We call this laser the probe beam. In saturated
absorption spectroscopy, we measure the transmission of the probe beam. The other
laser starts on the right and is moving to the left and has a large amount of power.
We will call this laser the pump beam. In most experimental setups, the probe beam
and the pump beam originate from the same laser. The laser can be split into two
paths using, for example, a .λ/2 plate and a polarizing beam splitter, see Sect. 1.5.
One would adjust the orientation of the .λ/2 plate so that the probe beam has less
power than the pump beam.
Fig. 5.2 Saturated absorption spectroscopy needs two laser beams: a probe beam and a pump
beam. Not pictured is the photodiode with which we monitor the transmission of the probe beam
1 Remember that an atom can be moving perpendicular to the laser beam, and it will experience
no Doppler effect. Only the velocity component parallel (towards or away) with the laser will
contribute to a Doppler shift.
5.1 Saturated Absorption Spectroscopy 87
Transmission
transition in europium-156 0.6
atoms with a temperature of
400 Kelvin 0.4
0.2
0.0
fA fr fB
f (MHz)
To explore how this technique works, we will use our simple two-level atom. If
the pump beam were not present, we know the transmission of the probe beam looks
like Fig. 5.3. This plot is calculated using the mass of a europium-156 atom and a
vapor cell at 400 Kelvin. The natural linewidth of the transition is about 25 MHz,
which is much smaller than the Doppler width of 320 MHz.
Now, let’s add in the pump beam. Our ultimate goal is to determine how the
transmission plot of the probe beam changes with the addition of the pump beam.
Let’s start by thinking about the transmission of a laser through the vapor cell when
the laser frequency is at .fA , see Fig. 5.1. Since both the pump and the probe beam
come from the same laser, they have the same frequency. The only difference is that
they are moving in opposite directions. We want to ask the question: Which atoms
interact with each laser beam?
As a reminder, the Doppler shift for an atom moving with velocity component .v‖
is given by the formula:
v‖ v‖ c
ΔfD = −
. = − fL → |v‖ | = ΔfD (5.1)
λ c fL
where .v‖ is negative if the atom is moving towards the laser source and positive if it
is moving away.
I find it useful to use numbers, so let’s say that the frequency of the laser is set
to .fA , and .fA is 200 MHz below .fr = 652.0000 THz. Since we have two laser
beams moving in different directions, we are going to use the magnitude of .v‖ and
the descriptors “to the left” and “to the right” for the following discussion.
Let’s start with the probe beam, which is moving to the right. Since the frequency
of the laser is below resonance, we know that atoms which interact with the probe
beam have to be moving to the left with a specific .v‖ so that, according to those
atoms, the Doppler effect shifts the laser frequency into resonance. That means the
atoms would have to be moving at:
3 × 108 m/s m
v‖ = (200 × 106 Hz)
. = 92 to the left (5.2)
652 × 10 Hz
12 s
Now let’s think about the pump beam, which is moving to the left. That means
the atoms that absorb light from the pump beam must be moving to the right. The
math is the same, but the direction of movement is opposite:
3 × 108 m/s m
v‖ = (200 × 106 Hz)
. = 92 to the right. (5.3)
652 × 1012 Hz s
Spend a few minutes on the above argument to make sure it all makes sense.
Here is the important take home message: When the frequency of the laser is at
.fA , different atoms interact with the probe beam and the pump beam. Both lasers
are losing photons, but they are losing photons to different atoms. Since we are
monitoring the probe beam transmission, the probe beam transmission is the same
whether the pump beam is on or off. Again, this is a very important concept so make
sure it makes sense before moving on.
Your turn! The laser frequency is now at frequency .fB , which we will assume
is 200 MHz higher than .fr = 652.0000 THz. What velocity does an atom need to
have to absorb light from the probe beam? From the pump beam? The answers are
in the footnotes.2
The conclusion for when .fL = fB is the same as when .fL = fA : When the
frequency of the laser is at .fB , the probe beam transmission is the same whether the
pump beam is on or off. The trick happens when the laser frequency is at .fr . The
Doppler shift is 0, so both the probe and the pump beams interact with the same
atoms. When the laser frequency is at .fA or .fB (or any frequency except .fr ), the
pump and the probe lasers interact with different atoms. When the laser frequency
is at .fr , the two laser beams compete for the same atoms.
To explore this more, let’s do a thought experiment. First, we either block or turn
off the pump beam so that there is only a probe beam. The probe beam frequency
is set to the resonance frequency, and we’ll assume it hits a single atom at rest.
Let’s say the probe beam has 10 photons that pass by the atom for every lifetime
of the excited state. From those 10 photons, the atom absorbs 1 photon reducing
the probe beam transmission to 9 photons; this is a 10% reduction in probe beam
transmission. Now we turn the pump beam back on. The pump beam has more
power than the probe beam. Let’s say the pump beam provides an additional 990
photons. The atom will randomly pick 1 photon from a possible 1000 photons (10
from the probe and 990 from the pump). Most likely the atom is going pick a photon
from the pump beam. Since all 10 photons make it through, the transmission of the
probe beam is larger when the pump beam is on. Every once in a while, the atom
will randomly absorb from the probe beam, decreasing its transmission percentage.
However, the transmission of the probe beam is, on average, larger when the pump
beam is present. If we increase the number of atoms in the vapor cell, each atom
with .v‖ = 0 will randomly absorb from either the pump beam or the probe beam.
2 Probe:92 m/s (to the right); Pump: 92 m/s (to the left); notice the directions are switched from
when the laser frequency was .fA .
5.1 Saturated Absorption Spectroscopy 89
The conclusion is: When the laser frequency matches the resonance frequency, the
transmission of the probe beam is larger when the pump beam is present.
What is really important is that this only happens for the atoms that have .v‖ = 0.
For any other velocity, the probe beam transmission is exactly the same whether the
pump beam is on or off. Let’s recap all of this in a table. To make things easier, we
are going to define “moving to the left” (towards the probe beam) as negative and
“moving to the right” as positive.
.fL Velocity of atoms needed to Velocity of atoms needed to How does the pump beam change
absorb from probe beam. absorb from pump beam. the transmission of the probe?
.fA .−92 m/s .+92 m/s It doesn’t
.fr 0 m/s 0 m/s Transmission increases
.fB +92 m/s .−92 m/s It doesn’t
Summary
If the laser frequency is not on resonance, the probe beam and pump beam
are interacting with different atoms. In other words, the probe beam is losing
photons to different atoms than the pump beam. On resonance, the two lasers
compete for the same atoms, which results in less photons being absorbed
from the probe beam.
A simulation of the transmission of the probe beam with the pump beam off (left
plot) and with the pump beam on (middle) is shown in Fig. 5.4. If we subtract the two
plots (right), we are left with a spectral feature with a full width half maximum equal
to the natural linewidth of the transition.3 This is the same plot as the absorption plot
from the thought experiment that we did in Sect. 3.1, which was an absorption plot
Transmission
Fig. 5.4 The transmission plots of the probe beam for a 2 level atom with just a probe beam (left),
the probe beam and a pump beam (middle), and the difference between the two transmission plots
(right)
3 Assuming that the width is not broadened from some other effect like power broadening.
90 5 Saturated Absorption Spectroscopy
for atoms at 0 Kelvin. We call this plot a saturated absorption plot. Neat trick,
huh!
5.2 Crossovers
Saturated absorption spectroscopy is super cool.4 It allows us to use hot gas and still
produce spectral features as if all of the atoms were frozen in place (absolute zero or
0 K). However, there is a trade off if there are multiple ground or excited states, and
that trade-off is additional fake spectral features in our spectrum called crossovers.
There are three types of crossovers: V crossovers, .Λ crossovers (the Greek letter
capital lambda, so we call them “Lambda crossovers”), and X crossovers. The
reason for the names should be clear as you examine the energy level diagrams for
each in Fig. 5.5. I want to point out that the labels, .F = 4, .F = 5, etc. have meaning
that we explore in Chaps. 7 and 8. Even though we haven’t connected those labels
to physics yet, I wanted to remind you about “The Rule” from Eq. 3.15:
The Rule
. ΔF = −1, 0, + 1 (5.4)
F = 0 /→ F = 0
Fig. 5.5 The three types of crossovers. A V crossover is due to the probe beam and pump beam
exciting atoms from the same ground state to two different excited states. For a V crossover, the
pump beam “steals” atoms from the probe beam. A Λ crossover is due to the probe beam and pump
beam exciting atoms from the two different ground states to the same excited state. A X crossover
is due to the probe beam and pump beam exciting atoms from two different ground states to two
different excited states. For both Λ and X crossovers, the pump beam “gives” atoms to the probe
beam
4 Yay puns!
5.2 Crossovers 91
Fig. 5.6 The experimental setup and energy levels to think about V crossovers
5.2.1 V Crossovers
Suppose you have an atom with one ground state that can be excited to two
different excited states, as shown in Fig. 5.6. The two excited states have resonance
frequencies .fr1 and .fr2 . Using the arguments explored in Sect. 5.1, you might
expect transmission plots shown in Fig. 5.7. In the “Pump off” plot, I also plotted
the individual Doppler profiles for both transitions (red and blue dashes). If you add
these two together you will get the black curve. With the pump beam on, you might
(correctly) expect to get Lorentzian features at laser frequencies .fr1 and .fr2 .
That prediction is close, but not quite correct. If you do the experiment, you will
find that you have an additional Lorentzian shaped spectral feature exactly halfway
between .fr1 and .fr2 . This extra feature is called a V crossover. Let’s explore why
this happens with an example. I always find it easier to use numbers, so let’s say that
.fr2 -.fr1 = 600 MHz. The feature occurs when the laser frequency is set to 300 MHz
above .fr1 and 300 MHz below .fr2 . We are also going to define an atom moving
right as positive velocity and an atom moving left as negative velocity.
Your turn: Having the laser frequency precisely between .fr1 and .fr2 , calculate
the velocity that an atom would need in order to absorb from the pump beam to
excited state #1, from the probe beam to excited state #1, from the pump beam to
excited state #2, and from the probe beam to excited state #2. Use .λ = 500 nm for
the math. Make sure you have an answer before moving on. Here is a crossword
puzzle to separate the question and answer.
Transmission
Fig. 5.7 A very reasonable, but incorrect guess for a saturated absorption spectrum for an atom
with one ground state and two excited states
92 5 Saturated Absorption Spectroscopy
First, notice that the pump and probe beams are exciting different atoms to
excited state #1. The two beams are also exciting different atoms to excited state
#2. Specifically, for an atom to be excited by the probe to excited state #1 it would
have to be moving away from the probe beam at +150 m/s. An atom would have to
be moving away from the pump beam at .−150 m/s to be excited by the pump beam
to excited state #1. The pump beam and probe beam are interacting with different
atoms; nothing new here.
Now notice that the pump beam is trying to excite atoms moving at +150 m/s
to excited state #2 while the probe beam is trying to excite those same atoms to
excited state #1. Those atoms that are moving at +150 m/s get to pick which laser
to absorb from! They can absorb from the probe beam and be excited to excited
state #1 or absorb from the pump beam and be excited to excited state #2. The
atom is more likely to absorb a photon from the pump beam leaving fewer atoms
for the probe to interact with. Even though the two lasers are trying to excite to
different states, the pump beam still “steals” atoms from the probe beam meaning
the transmission of the probe beam will increase at that frequency. Similarly, the
atoms moving at .−150 m/s also get to pick between the pump and the probe beam.
Therefore, there will be an additional spectra feature that comes from two velocities
of atoms (+150 m/s and .−150 m/s) when the frequency of the laser is precisely
between .fr1 and .fr2 . With this new information, the transmission plot of the probe
beam has three spectral features: two that correspond to the actual frequencies of
the transitions and a third exactly halfway between that we call a crossover peak,
see Fig. 5.8.
The two real transitions come from atoms that have no velocity components in
the direction of the laser beams. The crossover peak comes from atoms that are
moving. I think now is a good time to remind everyone that the amplitudes of the
peaks in the above graphs are completely made up. Because there are two sets of
5.2 Crossovers 93
Transmission
0.6 0.6 0.6
0.4 0.4 0.4
0.2 0.2 0.2
0.0 0.0 0.0
fr1 fr2 fr1 fr2 fr1 fr2
f (MHz) f (MHz) f (MHz)
Fig. 5.8 A more accurate simulation of a saturated absorption spectrum with one ground state and
two excited states. The amplitudes of the spectral features are made up; in a real experiment, all
three features will have different amplitudes
atoms contributing to the crossover peak (.v = +150 m/s and .v = −150 m/s), the
crossover peak often turns out to be larger than the actual transitions. Also, the
amplitude for resonance 1 will not be the same as the amplitude for resonance 2.
▶ Important Comment If the two transitions are separated such that the
Doppler profiles of each transition are separated, you will not have any
crossovers because there are no atoms moving with the correct speeds
to cause the crossover feature, see Fig. 5.9.
Transmission
Transmission
Fig. 5.9 A simulation of a saturated absorption spectrum with one ground state and two excited
states, but the two excited states are separated by a large energy. There are no atoms with the
correct velocity to create the V crossover. The amplitudes of the spectral features are made up; in
a real experiment, the features will have different amplitudes
If the vapor cell was heated to increase the Doppler width, the crossover peak
would return, see Fig. 5.10.
Transmission
Transmission
Fig. 5.10 Now the vapor cell is heated so there are a few atoms with the correct velocity needed
to create a V crossover
94 5 Saturated Absorption Spectroscopy
Summary
If there are (1) two excited states and one ground state and (2) the Doppler
profiles for the two individual transitions are overlapping one another, there
will be a V crossover directly between the two transitions.
The speed of atoms needed to create a crossover feature can be derived from the
formula for the Doppler shift. The speed needed is:
|fr1 − fr2 |
|v‖ | = λ
. . (5.5)
2
If there aren’t any atoms in the vapor cell with that speed, there won’t be a
crossover feature. As with many things in experimental science, there are trade-
offs to saturated absorption spectroscopy. While saturated absorption spectroscopy
gives us really narrow spectroscopy features, it gives us more features to deal with.
Fortunately, we know precisely where those crossovers will be.
We can add a third excited state to the system.5 Let’s call the resonant frequencies
.fr1 , .fr2 , and .fr3 . We will get 3 crossovers features calculated using the same logic
as above. One crossover will be directly between .fr1 and .fr2 (i.e., (.fr1 +.fr2 )/2),
one between .fr1 and .fr3 (i.e., (.fr1 +.fr3 )/2), and one between .fr2 and .fr3 (i.e.,
(.fr2 +.fr3 )/2). In total, saturated absorption spectroscopy on an atom with one ground
state and three excited states will have 6 spectral features: 3 real features and 3
crossovers.
Λ crossovers and X crossovers, see Fig. 5.5, both occur when the pump beam excites
an atom with a particular velocity, such that, upon decay to a different ground state,
that atom has the correct velocity to be excited by the probe beam. As a reminder,
V crossovers occur because an atom gets to pick between absorbing a photon from
the pump beam and the probe beam. When an atom picks the pump beam over the
probe beam, the transmission of the probe beam increases resulting in a bump on the
transmission or absorption plot. Λ crossovers and X crossovers both occur because
the pump beam puts more atoms in the probe beam’s path. As such, the transmission
of the probe beam decreases resulting in a dip on the transmission or absorption plot.
Like V crossovers, this feature occurs when the laser frequency is precisely between
the two resonance frequencies:
5 We can’t add more than that for a single ground state due to “The Rule”.
5.2 Crossovers 95
fr1 + fr2
fcross =
. (5.6)
2
As before, the vapor cell needs atoms with a speed
|fr1 − fr2 |
|v‖ | = λ
. . (5.7)
2
to create these features.
There is one important difference for X crossovers. For V crossovers and Λ
crossovers, the pump and the probe beam are interchangeable. Consider an atom
that has two ground states and three exited states, see Fig. 5.11. We pick the ground
states to have labels F = 1 and F = 2 and the excited states to have labels F ' = 1,
F ' = 2, and F ' = 3.6 As a reminder, “The Rule” is that an atom can be excited
as long as ΔF = 1, 0, or -1 with the exception F = 0 /→ F = 0. Suppose we
have a Λ crossover that comes from the two transitions F = 1 → F ' = 2 and
F = 2 → F ' = 2. To add numbers, let’s say the vapor cell needs atoms with speed
v‖ = +35 m/s or −35 m/s to produce this crossover. It doesn’t matter if the pump
beam is exciting the first transition or the second. If the pump beam is exciting the
first transition, it is “pumping” atoms with v‖ = +35 m/s from the F = 1 ground
state to the F = 2 ground state via the F ' = 2 excited state. The probe beam is then
exciting those extra atoms on the F = 2 → F ' = 2 transition. If the pump beam is
exciting the second transition, it is “pumping” atoms with v‖ = −35 m/s from the
F = 2 ground state into the F = 1 ground state via the F ' = 2 excited state. The
probe beam is then exciting those extra atoms on the F = 1 → F ' = 2 transition.
The important thing to notice here is that the excited state of both transitions can
decay into either ground state. We say that there are two “velocity classes” of atoms
that are contributing to that crossover feature: +35 m/s and −35 m/s.
6I added primes to the excited states to help us distinguish between the ground states and the
excited states.
96 5 Saturated Absorption Spectroscopy
For X crossovers, there are some situations where the two transitions cannot be
interchanged. Consider the two transitions: F = 1 → F ' = 2 and F = 2 → F ' =
3. If the pump beam is exciting the first transition, it is “pumping” atoms with, say,
v‖ = +25 m/s from the F = 1 ground state into the F = 2 ground state via the
F ' = 2 excited state. The probe beam is then exciting those extra atoms on the
F = 2 → F ' = 3 transition. However, if the pump beam is exciting the second
transition, which would be the atoms with v‖ = −25 m/s, F ' = 3 cannot decay into
the F = 1 ground state. So, the probe beam transmission, which is exciting atoms
on the F = 1 → F ' = 2 transition, is not changed. This crossover only has one
“velocity class” that contributes to the crossover, so it tends to be smaller than a
crossover with two velocity classes.
We now have the basic building blocks to interpret a spectrum from an atom with
as many ground and excited states that we want. If our atom has energy levels as
shown in Fig. 5.11, we will have multiple real transitions and multiple crossovers.
For the real transitions, an atom in the F = 1 ground state can be excited to the
F ' = 1 or F ' = 2 excited states. An atom in the F = 2 ground state can be excited
to the F ' = 1, F ' = 2, or F ' = 3 excited states. Each of these five transitions will
have a Doppler profile that has a Doppler width associated with it.
A saturated absorption plot will have those five spectral features as well as
crossovers. For a crossover to occur, the vapor cell has to have atoms with the
velocity needed to create that crossover. To conclude this section, let’s recap the
three types of crossovers and list the possible crossovers for the atom with the energy
states shown in Fig. 5.11:
(1) V crossovers: If there are two excited states that are excited from the same
ground state and the Doppler profiles from the individual transitions are
overlapping, we will have a crossover whose frequency is directly between
the two transitions. Using Fig. 5.11 as an example, V crossovers occur due to
interference between:
Transition #1 Transition #2
F → F' F → F'
1→1 1→2
2→1 2→2
2→2 2→3
2→1 2→3
the two transitions. Using Fig. 5.11 as an example, Λ crossovers occur due to
interference between:
Transition #1 Transition #2
F → F' F → F'
1→1 2→1
1→2 2→2
Cesium-133, which has 55 protons and 78 neutrons, is one of the most studied atoms
on the periodic table. Figure 5.12 shows a simplified energy level diagram for a
transition that uses 455.6 nm light. The lower state, which has the label .6s 2 S1/2
(don’t worry about what that means right now, we will talk about the physical
meaning behind the labeling starting in Chap. 7), has two closely spaced ground
states with labels .F = 3 and .F = 4 (we will give meaning to these labels in Chaps. 8
and 9). The separation of these two states is just over 9 GHz. In energy units, that
would be .hf = (6.626 × 10−34 Js)(9.192 × 109 Hz) = 6.091 × 10−24 J = 38 μeV.
98 5 Saturated Absorption Spectroscopy
Fun Fact
This energy separation is how we define 1 second! Imagine you had a
pendulum that made exactly 9,192,631,770 oscillations in 1 second. Replace
that pendulum with a cesium atom and you have the official definition of a
second.
The excited state studied here has four levels. These four levels are far closer
together than the two ground state levels. To easily see all of the levels in the figure,
the energy spacing scale is different for the ground state and excited state; the energy
separation of the two ground states is over 100 times bigger than the excited state
separations. Below are 4 questions to work through. Answer the first two questions
together before answering the second two questions.
Question #1 What speed does an atom have to have to create a .Λ crossover between
the two ground states and the .F ' = 3 excited state?
Question #2 Using Eq. 4.5, what temperature would the cesium vapor cell be such
that the FWHM of the Maxwell Boltzmann distribution was half of the velocity for
Question #1? The mass of cesium-133 is .m = 2.207 × 10−25 kg. The answers are
below this fun anagram puzzle.
|v |
#1: .|v‖ | = λ |fr1 −f
2
r2 |
= (455.6 × 10−9 m) 9,192,631,770
2
Hz
= 2094 m/s → 2‖ =
1047 m/s. ⎛ ⎞2 ⎛ ⎞2
/
2.207×10−25 kg
#2: .vFWHM = 2.355 kBmT → T = v2.355 FWHM m
kB = 1047 m/s
2.355 1.38×10−23 J/K
=
3161 K.
5.3 Example with Cesium-133 99
Anagram Fun
Rearrange the letters in “cesium” to make a new 6 letter word.
How many 5 letter words can you create from the word “cesium”? (I found
one, but an online anagram solver found two!)
How many 4 letter words can you create from the word “cesium”?
This is really hot! For reference, room temperature is about 300 K. In short, we
don’t have to worry about .Λ crossovers (or X crossovers).
Question #3 But what about V crossovers? What velocity does an atom have to
have to create a V crossover between the .F = 4 ground state and the .F ' = 4 and
'
.F = 5 excited states?
For V crossovers, .|v‖ | is well within the full width half maximum of the
Maxwell-Boltzmann velocity distribution. So, we are definitely going to have
V crossovers. However, the .|v‖ | needed for .Λ crossovers and X crossovers is
well outside the distribution, so we won’t see any .Λ crossovers or X crossovers.
Figure 5.13 is a saturated absorption plot between the .F = 4 ground state and the
' ' '
.F = 3, .F = 4, and .F = 5 excited states. The three labeled peaks are the real
transitions. Notice there are additional Lorentzian features exactly halfway between
1.0
0.8 F'=5
Spectrum
0.6 F'=4
0.4
0.2 F'=3
0.0
– 50 0 50 100
f (MHz) + 657,932,340 MHz
Fig. 5.13 Experimental data taken by my research group showing a saturated absorption plot
from the .F = 4 ground state of cesium-133 to the .F ' = 3, .F ' = 4, and .F ' = 5 excited states,
see reference [1]. There are six spectral features. Three of them are real transitions and three are V
crossovers
= λ |fr1 −fr2 |
6
7 #3: .|v | −9 m) 8.29×10 Hz = 18.9 m/s.
‖ 2 / = (455.6 × 10 / 2
−23 J/K)(300 K)
#4: .vFWHM = 2.355 kBmT = 2.355 (1.38×10 2.207×10−25 kg
= 322 m/s. There are definitely atoms
in the vapor cell to make this crossover!
100 5 Saturated Absorption Spectroscopy
any two real transitions. Also notice that all of the amplitudes are different. The
crossover between .F ' = 4 and .F ' = 5 is really big while the real transition from
the .F = 4 ground state to the .F ' = 3 excited state turns out to be really small.
The peak directly to the right of .F ' = 4 is the V crossover between .F ' = 3 and
.F
' = 5. Even if this plot wasn’t labeled, we can still figure out which features
are the real transitions and which are the crossovers. We just look for the features
directly between two other features to find the crossovers. Also, the peaks at the
smallest and largest frequency values have to be real transitions; a crossover has to
be between two real transitions.
A lot of laser spectroscopy is done from the ground state to an excited state.
However, laser spectroscopy can also be performed between two excited states.
Figure 5.14a shows a simplified Grotrian diagram for a transition in neutral atomic
oxygen-16 (8 protons and 8 neutrons). The lower state, which we give the label
.J = 2, can be excited to three different excited states, which we give the labels
' ' '
.J = 1, .J = 2, and .J = 3. Like in the previous examples, just consider these
labels for now. The Rule for these transitions are the same as before, we just replace
F with J : .ΔJ = −1, 0, or +1 with the exception that .J = 0 /→ J ' = 0.8 We will
explore what the labels actually mean in Chaps. 7, 8, and 9. Oxygen, in its natural
form, is a molecule composed of two oxygen atoms. A discharge (basically think
Fig. 5.14 (a) A simplified energy level diagram for a spectroscopic study in atomic oxygen-16
near 926 nm. The spectrum is taken between two excited states, which I call the lower state and
the upper state. (b) Experimental data of a saturated absorption spectroscopy spectrum from the
J = 2 lower state of oxygen-16 to the J ' = 1, J ' = 2, and J ' = 3 upper states. There are only
five spectral features because the vapor cell wasn’t hot enough for the V crossover created by the
J = 2 → J ' = 1 and J = 2 → J ' = 3 transitions; the Doppler profiles for these two transitions
did not overlap
8 A full list of the rules that need to be satisfied for an electron to transition between two atomic
about a “neon tube” filled with oxygen molecules) can be used to both dissociate the
molecule into neutral atomic oxygen as well as excite the electrons into a variety of
excited states. Most of the atoms are not in the .J = 2 lower state, but there are
enough for us to do spectroscopy. It should be noted that the .J = 2 lower state also
has a lifetime of about 27 ns, so the discharge needs to continually repopulate the
lower state for us to do spectroscopy. Discharges are also typically hotter than room
temperature.
Next, take a look at Fig. 5.14b. There is a very visible V crossover created by the
large .J = 2 → J ' = 3 (real) transition and medium sized .J = 2 → J ' = 2 (real)
transition. Those transitions are about 3500 MHz apart, but the Doppler profiles of
these individual transitions are large enough to create a crossover. This crossover is
labeled as .J = 2 → J ' = 3/2.9 The V crossover with the label .J = 2 → J ' = 2/1
is created by the medium .J = 2 → J ' = 2 and the small .J = 2 → J ' = 1
transitions. It is also quite visible, although not as big as the .J = 2 → J ' = 3/2
V crossover. Those two transitions are about 3200 MHz apart. We did not see a V
crossover created by the .J = 2 → J ' = 3 and .J = 2 → J ' = 1 transitions,
which would have the label .J = 2 → J ' = 3/1. Those two transitions are about
6500 MHz apart. Because they are so far apart, the individual Doppler profiles don’t
overlap resulting in no crossover feature.
9 Just to be clear, that 3/2 is not the fraction equivalent to 1.5. It is meant to convey “a V crossover
where the two excited states are .J ' = 3 and .J ' = 2.”
102 5 Saturated Absorption Spectroscopy
Fig. 5.15 Left: The 6 hyperfine energy levels for the ground state of europium-151. The 6
hyperfine energy levels for a particular excited state that is about 642.9 THz (466.3 nm) above
the ground state. The numbers listed for .Δf are with respect to the center of gravity
spectral features overlap with each other. Figure 5.17 shows experimental results
for a saturated absorption plot collected by my research group on this transition in
europium-151. Look how complicated the spectrum is! Although it is a complicated
plot, there are spectral features that we can try to attribute to each transition or
crossover. Our job, as experimentalists, is to extract as much information as we can
from these plots.
1.0
Transmission
0.9
0.8
0.7
Fig. 5.16 A simulation of a transmission plot with a probe beam (no pump beam) traveling
through a vapor cell of europium-151 atoms held at 400 K. There are 15 transitions in total. The
Doppler profile for each transition is shown in blue-dashed lines. Some of the amplitudes are quite
small and not really visible by eye in this plot. The sum of all the individual Doppler profiles is the
black curve, which is what we would measure in the lab. The red vertical lines indicate the center
of each transition. As you can see, no single spectral feature can be resolved
5.6 Extra: Crossover-Free Spectroscopy 103
1.0
0.5
Signal
0.0
–0.5
Crossovers can be problematic because they introduce additional features into the
spectrum. Many times, those crossover features overlap each other or overlap the
features from real transitions. So, it isn’t too surprising that spectroscopists devel-
oped methods of getting sub-Doppler features without crossovers. The simplest idea
is to use an atomic beam, see Fig. 5.18.
An atomic beam is created by taking a sample, placing it in a vacuum-compatible
oven,10 and heating the oven. The oven has a small hole to allow the atoms to escape.
After the oven, metal pieces called collimators are typically used to block any atoms
diverging at large angles. The ideal spectroscopy experiment would have an atomic
Fig. 5.18 A sketch of how experimentalists use an oven and collimators to make a collimated
atomic beam. Since the atoms are not moving vertically, if we sent a laser perpendicular to the
atomic beam, then .v‖ = 0
10 Assuming the atoms are a solid at room temperature. For gaseous molecules, the oven is replaced
beam with zero divergence, resulting in a column of atoms exiting the oven. In
practice, there will always be some divergence of the atomic beam.
The laser beam intersects perpendicular to the atomic beam. In this experimental
design, there are no atoms moving towards or away from the laser so there are
no Doppler shifts and there are no crossovers. This type of setup does have a few
drawbacks. The first is that you really need to make sure the laser is perpendicular
to the atomic beam. If there is a small angle, there will be no atoms moving
perpendicular to the laser. And, you can’t really tell if there is a non-zero angle
either. You still have atoms absorbing from the laser, but they will all be absorbing
at the Doppler shifted frequency. So, the spectrum looks the same, but the resonant
frequency is off. A common technique to address this issue is to perform saturated
absorption spectroscopy on the atomic beam. The other issue you have to deal with
is that the atomic beam is never perfectly collimated. Often times, the atomic beam
will be diverging more in one direction than the other. That will cause an asymmetry
in the spectral signal, even when using saturated absorption spectroscopy.
Another clever method for doing spectroscopy is to have two laser beams that
are traveling in the same direction, see Fig. 5.19. Unlike typical saturated absorption
spectroscopy, the two laser beams have independent frequency control. In saturated
absorption spectroscopy, the pump and probe beams come from the same laser, so
changing the frequency of the laser changes the frequency of both the pump and
the probe beams. In this setup, the frequency of laser #1 is going to be fixed to
a transition, and the frequency of laser #2 is scanned. The transmission of laser
#1 is what we monitor. Laser #2 also has more laser power (a higher saturation
parameter).
One obstacle for this experimental setup is that you need two lasers, which can
be expensive. The other is that laser #1 has to be at the resonance frequency for one
of the transitions. If the frequency of laser #1 does not perfectly match a resonant
frequency, the frequency scale of your spectrum will be off.
Fig. 5.19 A comparison between a saturated absorption spectroscopy experimental setup and a
crossover-free setup. The crossover-free setup requires two separate lasers
5.6 Extra: Crossover-Free Spectroscopy 105
(a) Considering only the atoms moving with .v‖ = 0, explain why scanning the
frequency of Laser #2 across the .F = 5 → F ' = 5 transition results in a 0
Kelvin spectral feature on the transmission plot for Laser #1. Will you also get
a spectral feature when Laser #2 scans across the .F = 5 → F ' = 4 transition?
(b) When Laser #2 scans through the .F = 6 → F ' = 5 transition, Laser #2 will
excite atoms with .v‖ = 0 from the .F = 6 ground state to the .F ' = 5 excited
state. Even though Laser #1 is not resonant with that transition, there will be a
spectral feature at that frequency on the transmission plot. Why?
106 5 Saturated Absorption Spectroscopy
(c) Now consider atoms that are moving at a speed .v‖ ≈ 344 m/s towards Laser #1
(and also towards Laser #2). The atoms are moving with the perfect speed to be
excited by Laser #1 on the .F = 5 → F ' = 4 transition. Laser #2 now has its
frequency scanned. How do these atoms affect the transmission plot for Laser
#1?
(d) Next, consider atoms that are moving at a speed .v‖ ≈ 56 m/s towards Laser #1
(and also towards Laser #2). These atoms are moving with the perfect speed to
be excited by Laser #1 on the .F = 6 → F ' = 5 transition. How do these atoms
affect the transmission plot as the frequency of Laser #2 is scanned?
(e) After considering parts (a) through (d), how many features will our transmission
plot have?
Problems
5.1 For each of the following equations, write a brief description of what each
equation means.
5.2 To create a particular crossover, a vapor cell needs atoms that have a speed
given by Eq. 5.7. Derive this formula using the Doppler shift equation.
5.3 Consider a transition in an atom with three hyperfine ground states and two
hyperfine excited states. The ground states have labels F = 2, F = 3 and F = 4
and the excited states have labels F ' = 3 and F ' = 4.
5.4 Figure 5.21 shows energy levels for a transition in rubidium-87. The center of
gravity for the ground and excited states are shown on the far left. The ground state
labeled F = 1 has an energy of −4271.676 MHz with respect to the center of gravity
and the ground state labeled F = 2 has an energy of +2563.005 MHz with respect to
the center of gravity. The two ground states are separated by 6834.682 MHz, which
is much larger than the width of the Doppler profile for these transitions, which is
about 510 MHz, at 300 K. This means there will be no Λ or X crossovers. However,
all of the excited states are separated by frequencies smaller than the Doppler width,
which means the saturated absorption spectrum will have V crossovers. The natural
linewidth for this transition is 6 MHz.
5.6 Extra: Crossover-Free Spectroscopy 107
(a) Make a saturated absorption plot using the above energy levels assuming atoms
are only in the F = 1 ground state. Remember to use The Rule: ΔF = −1, 0,
or +1 with the exception F = 0 /→ F ' = 0. As always, don’t worry about the
amplitudes of the spectral features. The horizontal axis should be with respect
to center of gravity of the excited state. On your plot, label which features are
real transitions and which are crossovers.
(b) Make a saturated absorption plot assuming atoms are only in the F = 2 ground
state.
(c) The plots in part (a) and part (b) are separated by about 6830 MHz, see Fig. 5.22.
The 0 on the horizontal axis in Fig. 5.22 is with respect to center of gravity
frequency. Let’s assume the rubidium atoms are really hot. So hot that the
Doppler profiles from the two ground states are overlapping, which means we
will have more crossovers. Where on the above graph would the crossover be
due to the two transitions F = 1 → F ' = 1 and F = 2 → F ' = 1?
(d) In the scenario outlined in part (c), why would there be no crossovers due to the
two ground states and the F = 0 excited state?
0.4
0.2
0.0
–3000 –2000 –1000 0 1000 2000 3000 4000
f (MHz) + fcog
Fig. 5.22 A simulation of a saturated absorption plot (i.e., a pump on - pump off plot) scanning
across all possible transitions. The 0 on the horizontal axis is the center of gravity frequency
384.2 THz. The amplitudes for the spectral features are all set to be the same. In a real experiment,
the amplitudes will all be different
108 5 Saturated Absorption Spectroscopy
5.5 A transition in sodium-23 has a Grotrian diagram that is very similar to the
transition studied in Problem 5.4 for rubidium-87. The Grotrian diagram for the
sodium transitions studied in this problem are shown in Fig. 5.23: The difference
is that the energy levels are much closer together. The natural linewidth for this
transition is 10 MHz.
(a) Make a saturated absorption plot assuming atoms are only in the F = 1 ground
state. Assume we have some power broadening so that the width of the spectral
features is 12 MHz. As always, don’t worry about the amplitude of the spectral
features.
(b) Reflect on your spectrum.
(c) Now assume you collected the spectrum but you used a crossover-free experi-
mental setup. What does your spectrum look like now?
References
1. Williams, W.D., Herd, M.T., Hawkins W.B.: Spectroscopic study of the 7p1/2 and 7p3/2 states in
Cesium-133. Laser Phys. Lett. 15(9), 095702 (2018). https://doi.org/10.1088/1612-202X/aac97
2. Maruko, C., Cölmek, N., Herd, M.T., Ahrendsen, K., Cabrales, B., Cannon, G., Davis, E.,
Guo, X., Karani, T., Wallace, A., Wisnauckas, K., Williams, W.D.: Spectroscopic study of the
4f7 6s2 8 S◦7/2 − 4f7 (8 S◦ ) 6 s6p(1 P◦ ) 8 P5/2,7/2 transitions in neutral europium-151 and europium-
153: absolute frequency and hyperfine structure. J. Opt. Soc. Am. B. 41, 1217–1223 (2024).
https://doi.org/10.1364/JOSAB.521181
References 109
Open Access This chapter is licensed under the terms of the Creative Commons Attribution 4.0
International License (http://creativecommons.org/licenses/by/4.0/), which permits use, sharing,
adaptation, distribution and reproduction in any medium or format, as long as you give appropriate
credit to the original author(s) and the source, provide a link to the Creative Commons license and
indicate if changes were made.
The images or other third party material in this chapter are included in the chapter’s Creative
Commons license, unless indicated otherwise in a credit line to the material. If material is not
included in the chapter’s Creative Commons license and your intended use is not permitted by
statutory regulation or exceeds the permitted use, you will need to obtain permission directly from
the copyright holder.
Part II
Digging Deeper: Quantum Mechanics and
Beyond
Quantum Mechanics vs. Classical Physics
6
Abstract
Learning Goals
I’m often asked some variation of the question, “What is the difference between
quantum mechanics and classical physics?” This is a great question! After teaching
quantum mechanics many times and having numerous conversations with other
atomic and nuclear physicists, I believe the best way to introduce the difference
to new learners is to first understand two concepts: states and superposition. By
grasping these two ideas, we can begin to explore why quantum mechanics is
essential for explaining the world of the super small.
A note from Will: Many concepts discussed in this chapter are based on the
orthodox or Copenhagen interpretation of Quantum Mechanics. Although
it is the most popular interpretation, other interpretations exist, such as
the many-worlds interpretation, de Broglie-Bohm pilot wave theory, and
various collapse theories including Ghirardi-Rimini-Weber (GRW) theory and
Continuous Spontaneous Localization (CSL). This remains an active area of
exploration for many physicists and philosophers.
In addition, there are a few places where I simplify things a bit to make sure
we focus on the main concepts. Finding the balance between 100% correct
and still keeping things accessible was a challenge. For example, there are
places where I will write a sum of functions where it should technically be an
integral, but since calculus is not required for this book, I purposefully chose
to leave them as sums to help us explore the concepts. For equations that
should be integrals, I include a footnote to indicate that. For those who have
not taken calculus, the ideas we discuss are still correct! We are just going to
simplify the math a bit.
The other concept I simplify a bit is concerning compatible observables
and sharing states. However, after discussions with my physics pedagogy
friends, I decided to keep that language instead of exploring the delicate topic
of projections and subspaces. For the advanced readers, keep this in mind. For
new learners to quantum mechanics, this chapter is for you! If in the future
you decide to learn more advanced quantum mechanics (and I hope you do!),
please come back to this chapter to explore the subtleties of “commutation”
and “subspaces.”
Definitions
. Momentum: A property of an object that is moving. For a classical object
like a baseball, the formula for momentum is p = mv, where p is the
object’s momentum, m is the object’s mass, and v is the object’s velocity.
An object’s momentum will change if something acts on the object from the
outside. The larger an object’s momentum, the harder it is to stop. The unit
of momentum is kg · m/s.
. Observable: Something we can measure experimentally. Observables
include, but are not limited to, energy, position, momentum, and angular
momentum (angular momentum is the topic of Chaps. 7 and 8).
6.1 What Is a State? 115
Physical systems, such as atoms, can exist in various conditions. They can be excited
or relaxed, located in different places, moving or at rest, and so on. We describe these
different conditions as “states,” and use a mathematical “state function” to specify
the exact state of a system at a given time.
As we talked about and explored in Part 1 of this book, atoms have discrete
energy states.
these discrete energy states in energy level diagrams throughout Part 1 of this book.
Figure 6.1 is a copy of Fig. 5.12 as a helpful reminder. This is so important that I
want to say it again: An electron in an atom can have energy .E1 , .E2 , .E3 , etc. We
will never measure the energy to be somewhere between .E1 and .E2 .
Everything we can measure, which we call an observable, is represented by a
state. The state of an observable is described by a mathematical function. Examples
of things we can measure include energy, position, and momentum. As such, there
are energy states, position states, and momentum states. Every energy we can
measure is associated with some energy state. For example, suppose a system is
in an energy state described by the state function .ψ1 . If we measure the energy of
that system, we will find the system has energy .E1 . We will never measure any other
value for energy. Often those states are discrete (like the energy of an electron in an
atom) and sometimes they are continuous (like the position of an electron hanging
Fig. 6.2 The first three standing waves of a one dimensional rope or slinky. For this example, the
rope is fixed at .x = −L and .x = +L
wavefunction encapsulates the behavior and properties of the system at that energy
level. In quantum mechanics, the equation we use to calculate the wavefunctions
and energies is called the Schrödinger equation. The Schrödinger equation is a
partial differential equation and solving it for a system provides not just the spatial
form of the wavefunction, but also how they evolve in time. If you take a quantum
mechanics class, you will spend a good amount of time solving the Schrödinger
equation for different physical scenarios.
Returning to our example of the vibrating quantum mechanical string, we can
now describe the wavefunctions for specific energy states. The wavefunction .ψ1
corresponds to the state with one loop of energy,
( ) denoted by .E1 . Its functional form,
see Fig. 6.2, is described by .ψ1 = A1 cos π2Lx , where .A1 represents the amplitude
( .ψ)2 represents the state with two
and 2L denotes the length of the string. Similarly,
loops and has a functional form of .ψ2 = A2 sin πLx .
Summary So Far
If a quantum mechanical system has a particular energy, we say the system
is in an energy state. Mathematically, that energy state is described by the
wavefunction .ψn . If a quantum mechanical system is in an energy state
described by the wavefunction .ψn and we measure its energy, we will measure
the system’s energy to be .En . We will never measure anything else.
For the example in this section, n represents how many loops of energy the system
has, so .ψ3 is the third picture in Fig. 6.2. In bra-ket notation, see Sect. 3.6, we
would write either .|ψn 〈 or .|n〈. There are other quantities (obersvables) that we can
measure, and each thing we measure can change the shape of the wavefunction.
Figure 6.3 shows an example of three position states. If we measure the position
of a quantum mechanical particle, we would find the particle somewhere between
the start and end of the string. Mathematically, we label those position states .ψ
with a subscript, like .ψx . Notice that the position wavefunction is zero everywhere
Fig. 6.3 Examples of 3 different position states. Each position state has a specific position
associated with it
118 6 Quantum Mechanics vs. Classical Physics
except for where the particle was measured. For the first example, the particle was
measured to have a position at .x = 0. For the second example, the particle was
measured to have a position at .x = −0.4 L.
Super Important
For this quantum mechanical system, the energy states are not position states.
They look completely different.
2 Aswe work through the next few chapters, we will add more parameters to describe the state.
Eventually, the state will look something like .|n 𝓁 s j mj 〈. The 3 in this example is the n.
6.1 What Is a State? 119
Stepping Back
Take a look at one of the energy states shown in Fig. 6.2 and ask the question,
“Where is the particle?” The best you can answer is, “The wavefunction is
spread out from .x = −L to .x = L. It is not at any single location.” The
energy has a single value, but the exact position is unknown.
Far less intuitive is if we repeated this thought experiment with position
states. Take a look at one of the position states shown in Fig. 6.3 and ask the
question, “What is the energy of the particle?” Our intuition tells us there is an
answer to this question. However, in quantum mechanics, a particle only has
a specific energy if it is in an energy state. A position state is not an energy
state. The best you can answer is, “The energy is spread out. It does not have
a single value.”
Take as much time as you need to try and understand that last paragraph.
That last paragraph is very counterintuitive to our everyday experience. In
classical physics, if a car is traveling down the road, I can tell you its position
and energy. In quantum mechanics, we cannot know both. If the energy is
well defined, it is in an energy state. If the position is well defined, it is in a
position state.
There are two really important things to take away from this section:
Definitions
. Incompatible observables: If two observables cannot be precisely mea-
sured at the same time, they are called incompatible. Measuring one
observable changes the system, making subsequent measurements of the
other observable unpredictable.
. Compatible observables: If two observables can be precisely measured at
the same time, they are called compatible. Measuring one observable does
not change the system in a way that affects the measurement of the other
observable. Remeasuring either observable will return the same value as
initially measured.
We have learned that for quantum mechanical systems, energy and position are
incompatible observables. If the system is in a position state, it is not in an energy
state and vice versa. If we measure the position of the system, it is now in a position
state. Since that is not an energy state, we cannot predict the exact outcome when
we measure the system’s energy.
If there are two observables that share the same set of states, we say that they
are compatible observables. For the system in Sect. 6.1, there are, unfortunately, no
compatible observables, but we will be able to explore compatible observables using
systems in Chap. 7. So, for now, let’s just suppose there is some other observable
that is compatible with energy. Let’s call that observable A, described by the
wavefunction .ψA,n with possible measured values .An . If we measure the energy
of a quantum mechanical system, the quantum mechanical system will now be in an
energy state. Let’s say we measured the energy of the system to be .E3 , which is the
third picture with three loops in Fig. 6.2. Now we measure A and obtain .A3 . If we
went back and measured energy, we will find that the system is still in the state with
three loops with energy .E3 . Next, we measure A. If energy and observable A are
compatible, we will measure .A3 . This is because two compatible observables share
a set of states. In other words, if the energy state has the wavefunction with three
loops (.ψ3 ), and we measure A, we will find .A = A3 and that the wavefunction is
unchanged. It still has three loops.
6.3 Superposition of States 121
As you see, states are extremely important in quantum mechanics. Every observable
has a set of associated states. Each energy state corresponds to a specific energy. If
we measure the system’s energy, it will be in one of those states with the associated
energy. Similarly, measuring position, momentum, or any other observable places
the system in one of the states for that observable. Some observables share a set of
states (they are compatible, and subsequent measurements do not alter the previous
values), while others have different states (they are incompatible, and knowing one
means we don’t know the other).
Definitions
. Superposition: A concept in math and physics that says a function can be
constructed as the sum of two or more other functions.
. Basis set: All of the states for a particular observable for a particular
system. All of the possible energy states (for example, 1 loop, 2 loops, 3
loops, 4 loops, etc. for the vibrating quantum mechanical string shown in
Fig. 6.2) make up the energy basis set for that system. Similarly, all possible
position states (Fig. 6.3 shows a few position states for the vibrating quantum
mechanical string) make up the position basis set.
Fig. 6.4 (a) The state for our thought experiment. (b) The amplitude of each energy state needed
to construct the state. (c) The probability that, upon measurement of energy, we find the system in
a particular energy state
where .Ai is the amplitude of .ψi that is mathematically determined to reconstruct the
desired state .ψ. The set of all states .ψi is called the energy basis set. This equation
represents the concept of superposition. We are constructing a specific state from
a superposition of states from the energy basis set. The math required to calculate
the amplitudes .Ai is complicated and requires integrals, so I am just going to show
you the results of the math in Fig. 6.4b. In practical terms, we start with the energy
state with 1 loop (left picture in Fig. 6.2) and set the amplitude to 0.44. Next, we
take the energy state with 2 loops (middle picture in Fig. 6.2), set the amplitude
to 0.14, and add it to the energy state with 1 loop that had an amplitude of 0.44.
We then take the energy state with 3 loops, set the amplitude to .−0.38, and add it
to the first two energy states. This process repeats, and the more energy states we
include, the closer we get to the actual state. This is illustrated in Fig. 6.5. Adding
the first two energy states together (top row) with the appropriate amplitude doesn’t
resemble the desired state, but after adding the first 30 energy states (third row),
the resultant graph starts to look like the desired state. A superposition of the first
6.3 Superposition of States 123
Fig. 6.5 An illustrative example of adding more and more energy states together to build the
desired state. The amplitude of the energy state with 1 loop is 0.44, the amplitude of the energy
state with 2 loops is 0.14, 3 loops has an amplitude of .−0.38, etc. For the bottom row, the energy
state with 200 loops looks flat to our eye, but it has a very small amplitude of 0.019
200 energy states (bottom row) results in a close reconstruction. Remarkably, loops
can construct this state! Even more astonishingly, you can create any state you want
(as long as it is a well-defined single-valued function) from the energy basis set. In
other words, you can construct any state from a superposition of energy states.
Since we can construct a state out of energy states, we say that the state is a
superposition of states from the energy basis set. Similarly, we could construct an
energy state from a superposition of states from the position basis set. All we need
is some mathematical method to determine the amplitudes of each of the position
states so that when we add them all together we get the desired energy state.
In quantum mechanics, this is a very general idea. We can always construct a
single state from any basis set using a superposition of states from another basis
set. For example, we can construct a momentum state from a superposition of states
from the position basis set, a position state from a superposition of states from the
momentum basis set, a momentum state from a superposition of states from the
energy basis set, and so on. This is a neat math trick, but is it useful? It is perhaps
the most useful math trick in all of quantum mechanics.
Suppose we are in the state given by Fig. 6.4a. Now we want to measure the
energy of the system. The amplitudes given in Fig. 6.4b can be used to calculate the
probability that, upon measurement of energy, we will find the system in a particular
energy state. All we need to do is square that amplitude. For example, the amplitude
of the first energy state (1 loop) is 0.44. The probability that, upon measurement of
energy, we find the system with one loop of energy is .0.442 = 0.19, or 19%. The
amplitude of the third energy state (3 loops) is .−0.38. The probability that, upon
measurement of energy, we find the system with 3 loops of energy is .(−0.38)2 =
124 6 Quantum Mechanics vs. Classical Physics
0.15, or 15%. The probabilities are given in Fig. 6.4c. This rule is universal. If the
system is in a particular energy state and we want to measure position, there is an
associated probability for where we would find that quantum mechanical particle.
Those probabilities are found by first writing the energy state as a superposition
of all the position states from the position basis set with the correct amplitudes.
Squaring those amplitudes3 will tell you the probability that, upon measurement of
position, a quantum mechanical particle will be found at a particular position.4
The above discussion is for two incompatible observables. The two incompatible
observables have distinct basis sets, one for each observable. Measuring one of
the observables puts the system in a single state from that observable’s basis set.
However, that single state can be constructed from a superposition of states from
the other basis set. But, what if we had two compatible observables? Let’s recap
some of the important concepts:
Interestingly, we can use superposition to answer this question. Suppose energy and
our made up observable A are compatible observables. We measure the energy of
the system and find the system has energy .E3 . The system is now in the energy state
.ψ3 , which is also the same as .ψA,3 . Let’s write that energy state as a superposition
For the system we are exploring, building the energy state from a superposition of
A states is really easy! .ψ3 and .ψA,3 are literally the same sate. The probability that,
upon measurement of A, we find the system with .A1 (the value associated with
.ψA,1 ) is 0%. The probability that, upon measurement of A, we find the system with
value .A3 (the value associated with .ψA,3 ) is 100%. The system is already in that
state! Constructing the energy state from the A basis set is easy because there is one
state from the A basis set that perfectly matches the energy state.
3 For completeness, the amplitudes could be complex numbers, which is something beyond the
scope of this book. In the future, if you see an amplitude that is complex, you calculate the modulus
squared of the amplitude.
4 For completeness (again ☺), position and momentum are, mathematically, a bit harder to deal
with since their measured values are continuous and not discrete like energy. The idea we explored
is the same, but instead of the summation in Eq. 6.1 we would have an integral. As such, we would
state something like, “There is a 25% probability we would find the particle between .x = 0.100 nm
and .x = 0.102 nm”. This is a minor detail, but one I wanted to include a footnote for those who
have taken some more advanced math.
6.3 Superposition of States 125
Fig. 6.6 Constructing the position state from the middle picture in Figure 6.3. We need over 400
energy states to build this one state! Notice the position state in blue is noticeably taller than the
reconstructed states even when we use the first 200 energy states. The position state is still a bit
taller even with 400 energy states!
Now that we have explored the concept of superposition, let’s build a more
realistic position state, similar to the top right graph in Fig. 6.3, from the energy
states. Figure 6.6 shows the mathematical results needed to construct that position
state from the energy basis set. As you can see, we needed at least twice as many
energy states to get a good approximation for a position state. Despite needing
many more energy states to accurately reconstruct a single position state, the result
remains the same: if the quantum mechanical system were in that position state and
we measured its energy, we could get many different answers with probabilities
given by the squared amplitudes.
(continued)
126 6 Quantum Mechanics vs. Classical Physics
If the two observables have different basis sets, knowing the value of one
observable does not inform you of the value of the other. These observables
are incompatible. Measuring one observable places the system in a super-
position of states for the other observable where none of the amplitudes
are 1. Therefore, the state of the second observable after measurement is
probabilistic.
States and superposition of states are core concepts in quantum mechanics. Let’s
summarize the entire chapter using two generalized observables .α and .β.
. Both .α and .β have states associated with them, denoted as .ψα,1 , ψα,2 , ... and
.ψβ,1 , ψβ,2 , ....
. The set of states that are needed to describe all possible values for .α is called the
basis set of .α. The set of states that are needed to describe all possible values for
.β is called the basis set of .β.
. Each state has a measurable value associated with it. If the system is in state .ψα,7
and we measure .α, we will get the value associated with that state, .α7 . We will
never measure, for example, .α5 .
. If .α and .β are incompatible, they have different basis sets. If they are compatible,
they share a basis set.
. A state for .α can be written as a superposition of states for .β, and vise-versa. For
example,
or
. The square of the amplitudes in the above equations tell us the probability, upon
measurement, we will find the system in that state with the value associated with
that state. For example, if the system was in state .ψβ,137 with value .β137 and we
measured .α, the probability the system is now in a particular state of .α, say state
.ψα,3 with value .α3 , is the amplitude squared, .A
2 .
α,3
. Suppose we measured .β and got .β137 . Next, let’s say we measure .α and get .α7 .
If .α and .β are incompatible and we then re-measure .β, it is highly likely we will
get something other than .β137 .
6.4 The Energy Basis Set for a Quantum Harmonic Oscillator 127
The quantum harmonic oscillator is one of the first quantum systems learners
encounter in a quantum mechanics class, and it is a very practical system to study.
It can be used to model many physical systems including molecular vibrations,
see Fig. 6.7. Consider two atoms connected by chemical bonds to form a single
molecule. A classical model for this system is to have the two atoms connected
by a spring. If the two atoms are not moving and the spring is not stretched or
compressed, the system would just sit there at rest and not vibrate. However, if
the spring is stretched slightly, it tries to pull the atoms back closer together. As
the atoms move closer, the spring passes its equilibrium length and compresses.
Once compressed, the spring tries to push the atoms apart until it stretches past
equilibrium again, and the process repeats. This is an oscillation. In a quantum
mechanical version, this system will have discrete energy levels, meaning it will
have specific energy states. If the spring from this classical analogy were behaving
quantum mechanically, the spring would only oscillate at specific frequencies.
We can use the Schrödinger equation to find the energy states and their energies
for the quantum harmonic oscillator. The energy states with the four smallest
energies for the quantum harmonic oscillator are shown in Fig. 6.8. Notice the
similarities with the standing waves on a string example in Sect. 6.1. Although
the shape of the states look different, the lowest energy state has 1 loop, the
second lowest energy state has 2 loops, etc. If we measure the vibrational energy
of a molecule modeled by the quantum harmonic oscillator, we would find the
system in one of the energy states such as those shown in Fig. 6.8. Each state
has a defined, discrete energy. All of the same bullet points that summarized
Sect. 6.3 are still true! Observables that are compatible with energy, for example
the frequency of vibration, have the same basis set, allowing us to measure all
Fig. 6.7 Left: Two atoms connected by a spring. This is a model for two atoms connected by
molecular bonds. Right: The separation of the atoms as a function of time
128 6 Quantum Mechanics vs. Classical Physics
Fig. 6.8 The energy states with the four smallest energies for the quantum harmonic oscillator.
Each of these energy states also corresponds to a specific energy or frequency of vibration of the
atoms in the molecule
Reminder Definition
. Reduced Planck’s constant (h-bar): h̄ = h
2π = 1.054 × 10−34 kg · m2 /s;
where .ψp,i are the states making up the momentum basis set. Similarly, we could
write the state as a superposition of states from the position basis set:
∞
⎲
ψ = Ax,1 ψx,1 + Ax,2 ψx,2 + Ax,3 ψx,3 + ... =
. Ax,i ψx,i
i=1
where .ψx,i are the states making up the position basis set. If the system was in the
state shown in Fig. 6.4a, and we measured momentum, we would get a range of
possible values. More specifically, we have a .A2p,1 chance of measuring .p1 , a .A2p,2
chance of measuring .p2 , and so on. Equivalently, we would have some average value
of momentum. For example, if we had a .100% chance of measuring .p4 , the range (or
spread) of possible measurements for momentum is 0 while the average value would
be .p4 . Let’s call the range .Δp and the average value we measure .〉p〈. Similarly,
if we instead had measured position, we would get a range of possible position
measurements, which we will call .Δx, and an average position measurement, which
we will call .〉x〈.
If two observables A and B are compatible, there is a scenario where
.(ΔA)(ΔB) = 0. That scenario would be when the system is in a state of A,
which also happens to be a state of B since they are compatible. In that scenario,
there is no range of measurements for either A or B: .ΔA = 0 and .ΔB = 0.
However, if the system was not in a state of A and B, there would be a range of
possible measurements for both observables. Therefore,
(ΔA)(ΔB) ≥ 0
. (6.5)
The uncertainty principle for two compatible observables.
This equation covers all possible scenarios for two compatible observables.
Now let’s explore incompatible observables, for example momentum and posi-
tion. Since momentum and position are incompatible, then there is no situation
where .(Δx)(Δp) = 0. If we did the math to determine the relationship, we
would find that there is a minimum possible value to this product. For these two
observables, the minimum possible value is
h̄
(Δx)(Δp) ≥
. (6.6)
2
5 Forthe advanced reader, the next two equations should technically be integrals since momentum
and position have a continuous range of possible values, unlike energy that has discrete values.
130 6 Quantum Mechanics vs. Classical Physics
No matter what state you consider, there will always be a range of possible
measurements for at least one of the observables. You might be asking yourself
the question, “But wait, if the system was in a position state, wouldn’t .Δx = 0 and
thus .(Δx)(Δp) = 0 × (Δp) = 0?” If the system was in a position state, the range
of possible measurements for momentum would be .∞. So, we would have .(0)(∞),
which is undefined. However, if we were to carefully take the limit as .Δx → 0 and
.Δp → ∞, the product is still greater than .h̄/2.
For every pair of incompatible observables, the product of the ranges is greater
than some minimum value. The result is easier to read for some pairs of incom-
patible observables than others. For example, Eq. 6.6 is fairly straightforward: the
range of possible measurements of position multiplied by the range of possible
measurements of momentum is always greater than or equal to the constant
.h̄/2. Other times, the uncertainty principle is harder to interpret. The uncertainty
h̄
.(Δx)(ΔE) ≥ |〉p〈| (6.7)
2m
This equation is a bit harder to read. I would interpret it as follows: for a given
state, the product of the spread of possible position measurements and the spread
of possible energy measurements will always be greater than a constant times the
magnitude of the average value of possible momentum measurements. This is a
formula that you must be careful with. There are scenarios where .|〉p〈| = 0, but that
does not mean that there is a scenario where position and energy are compatible.
This just says that the product of their ranges must be greater than or equal to 0.
This chapter solidifies some concepts we explored in the first part of this book while
adding some more. Some observables are compatible (i.e., they share a basis set and
subsequent measurements do not disrupt the values of the observables) and some
observables are incompatible (i.e., they do not share a basis set and measuring one
observable puts the system in a superposition of states for the other observable). The
next questions to think about are
A reflection from Will: This was a really hard chapter for me to write!a There
are so many fun things to explore in quantum mechanics that I had a difficult
time narrowing it down to just a few important concepts. So, please take this
chapter as an initial introduction. If you choose to go on and learn more about
quantum mechanics (and I hope you do!), you can use this chapter as a base
to build more understanding. There are many fascinating things to explore
like quantum entanglement, Bell’s Inequality, Ehrenfest’s theorem, and the
no-cloning theorem, just to name a few. The uncertainty principle also has
much more to explore and learn. As with all things worth learning in life,
a
And I learned a lot by doing so.
Problems
6.2 Suppose a quantum system has two incompatible observables. We now know
two important principles:
Explain.
(b) Now suppose there is a quantum system with two compatible observables.
Do the two principles listed in the paragraph above still hold true for two
compatible observables? Why or why not?
132 6 Quantum Mechanics vs. Classical Physics
6.3 For the quantum harmonic oscillator, are position and energy compatible
observables? Explain.
6.4 Suppose we had a quantum mechanical system in the energy state represented
by the second state in Fig. 6.2 (the energy state with two loops).
(a) Draw that energy state on a piece of paper. Reconstruct that energy state by
drawing position states (Fig. 6.3) so that when you add them together you get
that energy state.
(b) Now you want to measure position. Where are you most likely to find the
quantum mechanical particle?
(c) Let’s put the system back into that energy state. What is the probability, upon
measurement of position, that the particle is found exactly in the middle? Why?
6.7 Just for fun, here is a crossword with some terms from this chapter. For each
word, write a clue.
Open Access This chapter is licensed under the terms of the Creative Commons Attribution 4.0
International License (http://creativecommons.org/licenses/by/4.0/), which permits use, sharing,
adaptation, distribution and reproduction in any medium or format, as long as you give appropriate
credit to the original author(s) and the source, provide a link to the Creative Commons license and
indicate if changes were made.
The images or other third party material in this chapter are included in the chapter’s Creative
Commons license, unless indicated otherwise in a credit line to the material. If material is not
included in the chapter’s Creative Commons license and your intended use is not permitted by
statutory regulation or exceeds the permitted use, you will need to obtain permission directly from
the copyright holder.
Angular Momentum
7
Abstract
Learning Goals
• that angular momentum is a vector that has both magnitude and direction.
• that quantum mechanics restricts angular momentum to certain discrete magni-
tudes and directions.
• that there are multiple angular momentum vectors in an atom, and understanding
how these vectors add together impacts the energy of a particular state.
• that quantum numbers are key tools for understanding angular momentum in
quantum mechanics.
• the concept of compatible and incompatible observables in the context of angular
momentum measurements.
7.1 Definitions
Note There are a lot of new definitions in this section. We will talk about and try
to understand all of them, but don’t worry too much about memorizing everything
right now. We will use some of the definitions continually throughout the rest of the
book. In future discussions, you should come back to these definitions as needed.
We are all learners, and learning takes repetition.☺
Momentum This is the same definition from Chap. 6: a property of an object that is
moving. For a classical object like a baseball, the formula for momentum is p = mv,
where p is the object’s momentum, m is the object’s mass, and v is the object’s
velocity. An object’s momentum will change if something acts on the object from
the outside. The larger an object’s momentum, the harder it is to stop. The unit of
momentum is kg m/s.
Force Below are two definitions of force. The first is the most common definition.
However, it is only true if the object’s mass is not changing, which is a very common
scenario. The second is the real definition of force. The unit of force is kg m/s2 ,
which we also call a newton (N ).
we tend to not write the unit kg m2 /s2 for either energy or torque. We use joules
(J ) for energy and newton meters (Nm) for torque. Torque will be important in this
chapter.
do some math, but the only other parameters in the equation should be fundamental
constants. So, being told 𝓁 = 2 is the same thing
√ as being told that the magnitude of
the electron’s orbital angular momentum is 6h̄ = 2.582 × 10−34 Js. We will soon
find that we need many quantum numbers to describe a particular state in an atom.
Electrons, protons, and neutrons all have angular momentum. A classical analogy
is thinking about the moon orbiting the earth.1 The moon has orbital angular
momentum because it is orbiting the earth. The moon is also spinning on its own
axis, so it also has “spin” angular momentum. So, the moon has two types of
angular momentum: orbital and spin. Likewise, an electron in an atom can have
orbital angular momentum and spin, which is sometimes called intrinsic angular
momentum. Interestingly, an electron always has intrinsic angular momentum, but
it does not always have orbital angular momentum. One of the craziest things about
quantum mechanics is that we don’t have a great analogy to think about the orbital
angular momentum or spin of the electron. The electron is not orbiting around
the nucleus or spinning on its axis like the moon, but it has the same properties
as if it were. The electron, as far as we can tell, has no size! So, how can it be
spinning? Imagine how confusing that must have been when physicists were first
trying to understand the electron.2 It has all the properties one would expect for a
ball spinning on its axis, but it is not a ball and it is not spinning! Understanding
electron spin is still a wonderful mystery.
Since there are two types of angular momentum, we could ask the question,
“What is the total electronic angular momentum of the electron in an atom?” The
total electronic angular momentum is not a simple sum of orbital angular momentum
and spin. In other words, you cannot just add the two angular momenta together like
.2 + 3 = 5. Angular momentum is something called a vector, which is something
with magnitude (i.e., size or length) and direction. To get the total electronic angular
momentum you have to add the orbital angular momentum and spin together in their
vector forms. As a classical analogy, imagine you walk 30 m due north, see Fig. 7.1.
That is a vector because it has a magnitude (30 m) and a direction (due north). Next,
you walk 40 m due west. Once again, that is a vector. If you were to add the two
vectors together, you would be asking the question, “If I restarted my journey but
1 Remember that in quantum mechanics, an electron is not like the moon orbiting the earth but
more like a wave that is surrounding the nucleus. We are just using the moon and earth as an
analogy to introduce angular momentum.
2 There are, at least, two things in physics that really boggle my mind. The first is spin. The second
is something called the fine structure constant, which you should Google if you want your mind
blown!
7.3 Orbital Angular Momentum of a Single Electron 139
Let’s get back to talking about angular momentum in a quantum system. An electron
in an atom has a total angular momentum. That total angular momentum of the
electron comes from the vector addition of its orbital angular momentum and
its spin. There are quantum numbers associated with all three of these angular
momentum vectors (orbital, spin, and total). The rest of this chapter is going to
be focused on understanding angular momentum in a quantum system. We will find
that both the size and direction of an angular momentum vector are quantized (i.e.,
can√only have certain values like the size of an orbital angular momentum vector
is . 𝓁(𝓁 + 1)h̄ where .𝓁 is a positive integer or zero). We will start by exploring the
orbital angular momentum of a single electron in an atom and then build up the
complexity by exploring (and adding) more and more angular momentum vectors
to the system.
The magnitude of the orbital angular momentum vector for a single electron in an
atom is represented by the quantum number .𝓁. The direction (or orientation) is also
quantized and is represented by the quantum number .m𝓁 . Figure 7.2a) shows an
example of the three possible orientations of an orbital angular momentum vector
represented by .𝓁 = 1. Each of these vectors has the same magnitude (they all
have .𝓁 = 1), but they all point in different directions (they all have a different
140 7 Angular Momentum
Fig. 7.2 Two popular ways to think about angular momentum states for a quantum mechanical
system, like an electron in an atom. (a) A vector representation of the three possible vector
orientations for an electron with orbital angular momentum quantum number 𝓁 = 1. (b) The
same information using cones instead of a vector and a dotted circle
m𝓁 quantum number). While .𝓁 tells us the magnitude of the vector, .m𝓁 tells us
.
how much of the vector points along the z-axis; this is called the z-component or
z-projection. Figure 7.2a) has three values of .m𝓁 : .−1, 0, and +1. The dashed blue
circles are meant to indicate that the angular momentum vector points anywhere on
that circle. Because the angular momentum vector can point anywhere on the circle,
some people like to think of angular momentum as a cone, see Fig. 7.2b).3 These
cones are also referred to as angular momentum states, which we will make more
sense as we work our way through Chaps. 7 and 8.
Orientation of the angular momentum vector, height of the cone, or com-
ponent/projection along the z-axis are all valid ways to think about .m𝓁 . After
solving the math for quantum mechanics, we find the component of orbital angular
momentum along the z-axis to be .m𝓁 h̄. For the upside-down cone in Fig. 7.2b), the
orbital angular momentum vector is represented by the two quantum√ numbers .𝓁√ =1
and .m𝓁 = −1, or the size of the orbital angular momentum is . 1(1 + 1)h̄ = 2h̄
and the amount of orbital angular momentum along the z-axis (or height of the
cone) is .−h̄. The flat cone has the same size (.𝓁 = 1), but it has no component
along the z-axis (.m𝓁 = 0). Finally the upright cone has a component along the z-
axis represented by .m𝓁 = +1. Since .m𝓁 is the amount pointing along the z-axis,
that number is also restricted by the value of .𝓁. For example, if .𝓁 = 1, .m𝓁 cannot
be 2. If it was, the amount pointing along the z-axis would be larger than the total
magnitude!
3 I,
personally, like using cones. But, as long as you understand the concepts, it doesn’t matter
which version you use.
7.3 Orbital Angular Momentum of a Single Electron 141
Important
In the following paragraphs, all of the states .(ψ) are for angular momentum.
I am going to use .ψz to represent the z-component of the angular momentum
state, .ψy for the y-component, and .ψx for the x-component.
Suppose our system is in a state .𝓁 = 1, m𝓁 = 1. This is the upright cone in Fig. 7.2b).
Since the x- and z- components are incompatible, we don’t know what we will
measure along the x-direction. We know there will be three possible outcomes for
142 7 Angular Momentum
the amount of angular momentum along the x-axis (.+h̄, 0, or .−h̄), but we don’t
know which value we will find until we measure it. This concept is what is being
represented by the blue circles or the cones.
Mathematically, we can also think about this using the principle of superposition.
Each set of three cones forms a basis set. One basis set is the three cones representing
orbital angular momentum along z and other is the three cones along x. A state from
one basis set can be constructed from a superposition of the other basis set. For
example,
Each state represents a cone along a particular axis. We are certainly allowed to
measure any component of angular momentum we want. Upon measurement, we
will be in one of those states. However, since they are all incompatible observables,
measuring one of the components puts the system in a superposition of the states for
the other obervables. Interestingly, we have two separate ways to construct the state
.ψz (𝓁 = 1, m𝓁 = 1):
In a way, this makes sense. If the quantum mechanical system is in the state .ψz (𝓁 =
1, m𝓁 = 1) and we measure the x-component of angular momentum, we will get
.m𝓁,x = 1 with probability .c
x,1 , .m𝓁,x = 0 with probability .cx,0 , or .m𝓁,x = −1
2 2
2
with probability .cx,−1 . On the other hand, we could have decided to measure the
y-component at which point we would need to construct the state from the y-basis
set to predict what we would measure if we measured the y-component. I should
point out that we can calculate those amplitudes
√ using quantum mechanics. For this
particular system, .cx,1 = 1/2, .cx,0 = −1/ 2, and .cx,−1 = 1/2.
7.3 Orbital Angular Momentum of a Single Electron 143
Ok, so all three components are incompatible with each other, but what about
compatibility with the magnitude of the angular momentum vector? Interestingly,
the magnitude of the angular momentum vector is a compatible observable with all
three components of angular momentum! At first, this might seem a little weird;
how can all the components be incompatible with each other yet each component is
compatible with the magnitude? But, in a way, this also makes physical sense. The
length of the vector is the same whether the cones point along the x-direction, the
y-direction, or the z-direction. Measuring a component of angular momentum does
not change the length, but it does change the orientation. For
√ this system, measuring
the magnitude of angular momentum will always return . 2h̄ (or .𝓁 = 1). It doesn’t
matter if we are in a state for the x-component, y-component, or z-component,
.𝓁 = 1.
The rest of the chapter will be spent making the system more complex and
complete. To do this, we will need to pick one component of angular momentum to
represent the system. Since the x-, y-, and z- components are just rotations of each
other, it doesn’t really matter which one we use to represent direction. By tradition,
we pick the z-component.
In the end, the above discussion can be summarized as follows: for .𝓁 = 1, there
are three possible projections for an axis and they all look the same (.+h̄, 0, and .−h̄).
So, we need to pick some axis to represent the idea that there are three orientations
for the orbital angular momentum vector.
A lot has happened in this section, so let’s summarize:
Summary
The
√ magnitude of an orbital angular momentum vector is given by the formula
. 𝓁(𝓁 + 1h̄. The amount of the vector pointing along the z-axis (or the height of
the cone) is .m𝓁 h̄. Notice that the formulas contain only quantum numbers and the
reduced Planck’s constant. If you hear someone say, “Angular momentum comes in
units of .h̄”, this is what they mean. Being told that a quantum mechanical system
has quantum numbers .𝓁 = 4 and .m𝓁 = −2 tells us all we need √ to know about√ the
orbital angular momentum of the system: the magnitude is . 4(4 + 1)h̄ = 20h̄
and the z-component is .−2h̄ (the cone is pointed in the negative z-direction with a
height of .2h̄).
There are also mathematical restrictions on the quantum numbers themselves.
.𝓁 can be either zero (no orbital angular momentum) or a positive integer (it has
orbital angular momentum). The component along the z-axis must be smaller than
the magnitude of the orbital angular momentum, so .m𝓁 is restricted to be between
-.𝓁 and +.𝓁 in integer steps.
√
The form of . 𝓁(𝓁 + 1)h̄ and .m𝓁 h̄ are the same for other types of angular
momentum. All we have to do is replace the quantum numbers representing orbital
angular momentum with the quantum numbers that represent the type of angular
momentum we are interested in. In the next few sections, we are going to expand
our discussion to talk about electron spin, the total electronic angular momentum
(orbital + spin), the total orbital angular momentum of multiple electrons in an
atom, and more. Each of the above examples of angular momentum will have two
quantum numbers associated with it, one for magnitude and one for the z-component
(cone height). For example, electron
√ spin will have quantum numbers s and .ms . The
formulas for electron spin are . s(s + 1h̄ and .ms h̄. While the quantum numbers
for orbital angular momentum are integers, we will find that the other types of
angular momentum quantum numbers can be either integers or half-integers. But
they all use the same two formulas and tell us the same information, magnitude and
z-component (cone height).
Imagine you had 2 electrons in your atom. Each of those electrons has orbital
angular momentum. We want to ask the question, “What is the total orbital angular
momentum of the system?” To answer that question, we need to add together two
quantum mechanical vectors. As an example, suppose we have an electron with the
magnitude of the orbital angular momentum vector represented by .𝓁A = 1 (Fig. 7.2
with 3 possible orientations) and a second electron with .𝓁B = 3 (Fig. 7.3 with 7
possible orientations). We want to add these two vectors together as in Fig. 7.1, but
we have to be thoughtful for two big reasons: (1) the vectors are really more like
“vector cones” and (2) there might be incompatible observables we have to think
about.
While the math is surprisingly complicated, the result is quite nice:
where L is the quantum number for the composite system. For each value of L, there
are .2L + 1 possible projections represented by the quantum number .mL . This might
be, understandably, confusing. So, let’s unpack this statement using our example of
.𝓁A = 1 and .𝓁B = 3.
find all possible values of L, we make a list that starts with the smallest value and
continually add 1 until we reach the largest value. In this case L can equal 2, 3, or 4.
To put it another way, there are three different orientations for .𝓁A and seven
different orientations for .𝓁B . Depending on the orientations, adding those two
angular momentum vectors results in different possible magnitudes. Working
through all possible combinations results in three possible magnitudes. For each
possible magnitude, there are different orientations (cone heights). In this example,
the individual orbital angular momentum vectors √ can add together
√ to make .L = 2 (5
possible cones with the same length√ vector . 2(2 + √ 1)h̄ = 6h̄), .L = 3 (7 possible
cones with the same length vector√. 3(3 + 1)h̄ =√ 12h̄), and .L = 4 (9 possible
cones with the same length vector . 4(4 + 1)h̄ = 20h̄).
After all the math, there are .5+7+9 = 21 possible cones. This makes sense since
the original vectors had 3 and 7 possible orientations, and .3 × 7 = 21. However, I
want to caution you in that the new cones are not constructed simply by taking one
cone chosen from .𝓁A = 1 and one cone from .𝓁B = 3 and adding them together. This
is because the z-components of the individual electrons are incompatible with the z-
component of the composite system. So, the new cones, which we will call the basis
set for the composite system, are a superposition of all the cones from the individual
electrons, which we call the individual electron basis set. The complicated math,
which we skip in this book, tells us the amplitude or amount that we need from each
of the individual electron basis set to construct the new composite system basis set.
We will explore this more in Sect. 7.5.1.
Another example: Suppose .𝓁A = 3 and .𝓁B = 5. The largest value for L is
.3 + 5 = 8, and the smallest value is .|3 − 5| = 2. Therefore, L can have values of 2,
For just these two electrons, we have a bunch of quantum numbers floating around.
Each quantum number corresponds to something we can measure. For these two
electrons, we have six quantum numbers: .𝓁A , .m𝓁,A , .𝓁B , .m𝓁,B , L, and .mL . Which
of these represent compatible observables? Is there something we can measure that
will disrupt the value of another observable? As learners of quantum mechanics, this
is important to keep track of.
All of the magnitudes are compatible with everything. If we know .𝓁A = 1,
.𝓁B = 3, and .L = 2, we can measure any of the observables represented by the six
C = (A + B), (A + B − 1), . . . , |A − B|
. (7.5)
For each value of C, .mC = −C, . . . , C in integer steps. The compatibility of
the observables is as follows:
An electron, whether it is in an atom or free, always has spin. The quantum numbers
for spin are s and .ms for the magnitude and orientation (cone height), respectively.
For an electron, .s = 1/2 always, so the possible orientations are always .ms = −1/2
or .ms = +1/2. From tradition, we call .ms = +1/2 “spin up” and .ms = −1/2 “spin
down”.
If the electron is in an atom, it can also have orbital angular momentum. We can
add orbital angular momentum and spin together and ask the question, “What is the
total angular momentum of that electron?” This new total angular momentum of
the electron is represented by the quantum numbers j and .mj . Using our new rule
(Eq. 7.5), the possible magnitudes of the total are represented by .j = l+s, . . . , |l−s|
in integer steps. Thus, if .𝓁 = 1 and .s = 1/2, j can be .3/2 or .1/2. As before,
.m𝓁 can be any value between .−𝓁 and .+𝓁 in integer steps, .ms can be any value
between .−s and .+s in integer steps, and .mj can be any value between .−j and .+j
in integer steps. Knowing the quantum numbers j , .𝓁, and s tells us the magnitude of
the angular momentum vectors. Knowing .mj , .m𝓁 , and .ms tells us the orientation of
each of their respective angular momentum vectors. However, .mj is incompatible
with .ms and .m𝓁 .
Here is a table summarizing all of the current types of angular momenta for a
single electron in an atom (QN means quantum number):
√ / /
When .j = 7/2, the magnitude is . j (j + 1)h̄ = 72 92 h̄ = 63 h̄ ≈ 3.97h̄.
√ / /4
When .j = 5/2, the magnitude is . j (j + 1)h̄ = 52 72 h̄ = 35
4 h̄ ≈ 2.96h̄.
We now have all the building blocks we need to add together as many angular
momentum vectors as we need. We first add two together to come up with the
new possible quantum numbers and projections for this composite system. Next,
we treat the composite system as its own individual angular momentum and add
a third angular momentum. Then keep going until you have accounted for all
electrons in your atom. The magnitudes of each angular momenta will be compatible
with everything. The z-component of any summed angular momentum will be
incompatible with the two individual angular momenta.
In practice, we don’t go through all this effort. As we will see in Chap. 8, there
are shortcuts we can take and databases that tell us all we need to know. For atoms
with more than 1 electron, we do not keep track of subscripts (i.e., electron 1,
electron 2, etc.), we simply ask the question “how much orbital, spin, and total
electronic angular momentum do all of the electrons have?” The formulas and ideas
are identical, but we use capital letters when we ask how much angular momentum
the electrons as a whole have:
All of the magnitudes are compatible with everything. However, .mJ is incompatible
with .mL and .mS .
The Nucleus The nucleus can also have angular momentum. Traditionally, this is
called nuclear spin, but that is a bit of a misnomer. The angular momentum of the
nucleus comes from the spin of the protons and neutrons as well as any orbital
angular momentum of those nucleons. So, nuclear spin should really be called
total nuclear angular momentum but no one uses that phrase. Nuclear spin has the
quantum numbers I and .mI . Nuclear spin .(I ) can be added to the total electronic
angular momentum .(J ) find the total angular momentum of the atom. The total
angular momentum is represented by the quantum numbers F and .mF , where F is
found from J and I using Eq. 7.5:
F = (J + I ), (J + I − 1), . . . , |J − I |
. (7.6)
The Photon The photon has intrinsic angular momentum as well. Photon spin
is the quantum mechanical description of light polarization, and it has a quantum
number of 1. In Chap. 1, we talked about how light can be linearly polarized,
circularly polarized, or elliptically polarized. That statement is for a laser beam,
which is composed of many photons. A single photon is circularly polarized. To
help visualize this, imagine a photon traveling straight towards you. The photon
would be rotating either clockwise or counterclockwise. To be clear, a photon is not
actually rotating just like the electron is not actually spinning like a top. However,
both the electron and the photon behave and interact with other particles as if they
are spinning or rotating. If it is spinning counterclockwise, the height of the cone is
.+h̄ (the z-component to photon spin with a quantum number of +1). If the photon
Every quantum number in every table in Sect. 7.6 is something we can measure.
Consider the hydrogen atom, which has a single electron. For now, we will ignore
the fact that the nucleus of a hydrogen atom has angular momentum (this is the topic
of Chap. 9 so we will revisit compatible and incompatible observables again in that
chapter). Here is a list of things we can measure: .𝓁, .m𝓁 , s, .ms , j , and .mj .
Technically, we could also measure L, .mL , S, .mS , J , and .mJ , but, since the
hydrogen atom has only a single electron, measuring, for example, the total orbital
angular momentum of all the electrons is the same as measuring the orbital angular
momentum of the single electron in the system.
We want to ask the question, which of these observables is compatible with
energy? In other words, if the hydrogen atom was in an energy state, what other
properties could we measure and still leave the electron in that same energy state?
The answer is:
That means if the system was in an energy state and we measured the z-component
of the electron’s orbital angular momentum, the system is now in a superposition
4 Because the photon has no mass, the math, interestingly, forbids a z-component of 0.
152 7 Angular Momentum
Fig. 7.4 An example of adding together a spin cone and a orbital angular momentum cone to get
the total electronic angular momentum cone. Notice that the z-component of both the spin and
orbital angular momentum cones is not the same value at all points of the cone tops. This means
the z-component does not have a specific, discrete value, which means that both the z-component
of spin and the z-component of orbital angular momentum are incompatible with the z-component
of j
of energy states. But what makes .m𝓁 and .ms incompatible with .mj and energy?
While it isn’t obvious why .m𝓁 and .ms are incompatible with the energy state,
we can visualize why they are incompatible with .mj . Figure 7.4 visualizes the
incompatibility. This figure shows the addition of two angular momentum cones.
Since .mj is compatible with the energy state, we will define the z-axis to be along
the cone for j . I want to point out a few things. First, notice that the cones for both
.𝓁 and s are tilted with respect to the z-axis. This means that the height of those
two cones (i.e., the result of a measurement of .m𝓁 or .ms ) are no longer perfect
projections onto the z-axis. In other words, if we measured the z-component of spin
with respect to this z-axis, we will be in a superposition of the .mj basis set. Compare
this to the measurement of the z-component of j . The cone for j has a rim that is
constant along the z-axis, so we will always measure the same value of .mj . Since
.mj is compatible with the energy state, measuring .m𝓁 or .ms would also put the
electron shells in more detail in Chap. 8. If the electron in the hydrogen atom is
in an energy state represented by one of these kets, we can measure any of those
observables and still be in that same energy state. If we measured .m𝓁 or .ms , the
system would be in a superposition of energy states.
7.7 A Bit More on Compatible and Incompatible Observables 153
Example #1 How many states will there be for the electron in a hydrogen atom
that has .n = 2, .𝓁 = 1, and .s = 1/2?
To start this problem, let’s first calculate what values of j are possible. j can
range from .𝓁 + s to .|𝓁 − s| in integer steps, so j can be .1/2 or .3/2. If .j = 1/2,
then .mj can be .1/2 or .−1/2. If .j = 3/2, then .mj can be 3/2, 1/2, -1/2, or -3/2.
Therefore, there are 6 energy states with .n = 2, .𝓁 = 1, and .s = 1/2. Using
the notation .|n 𝓁 s j mj 〉, these states are: .|2 1 12 21 21 〉, .|2 1 12 21 - 12 〉, .|2 1 12 23 23 〉,
.|2 1
2 2 2 〉, .|2 1 2 2 - 2 〉, and .|2 1 2 2 - 2 〉.
1 3 1 1 3 1 1 3 3
The above math tells us that there is an energy state in the hydrogen atom that
is represented by, for example, the quantum numbers .|2 1 12 23 - 32 〉. If the hydrogen
atom√ is in this state, √ it has a defined magnitude of orbital angular momentum (in this
case . 𝓁(𝓁 + 1h̄ = 1(1 + 1)h̄ ≈ 1.41h̄), a defined magnitude for spin, a defined
magnitude for the total electronic angular momentum, and a defined z-component
for the total electronic angular momentum. That state does not have a well defined
z-component of orbital angular momentum nor a well defined z-component of spin.
We can measure any observable compatible with energy and the system will stay
in the .|2 1 12 23 - 32 〉 state. However, if we measure .m𝓁 or .ms , the system will be in a
superposition of energy states.
(a) What will we measure for the electron’s orbital angular momentum, the
electron’s spin, the electron’s total angular momentum, and the electron’s z-
component of the electron’s total angular momentum?
(b) If we measured the z-component of the electron’s orbital angular momentum,
what might we get?
Solution
√
/ + 1)h̄ ≈ 1.41h̄ with 100% probability
(a) .𝓁 = 1 : We will measure . 1(1
s = 1/2 : We will measure .
.
1 1
(2 + 1)h̄ ≈ 0.87h̄ with 100% probability
/2
j = 3/2 : We will measure .
.
2 ( 2 + 1)h̄ ≈ 1.94h̄ with 100%
3 3
probability
.mj = −1/2 : We will measure .- 12 h̄ with 100% probability
(b) We would measure either .−h̄, 0, or .h̄ with different probabilities. We would
need advanced math to calculate those probabilities, so I will just emphasize
that we will not measure a definite value for the z-component of the electron’s
orbital angular momentum. After the measurement, the system will no longer
be in the state .|2 1 12 23 - 12 〉, but a superposition of energy states. If we now went
back and measured j , we might get .j = 1/2 or .j = 3/2 with hard to calculate
probabilities.
Problems
(a) What are the magnitudes and projection along the z-axis for 𝓁?
(b) What are the magnitudes and projection along the z-axis for s?
(c) What are the possible magnitudes and projections along the z-axis for j ?
7.3 An atom has two electrons. One electron has 𝓁 = 1 and the other has 𝓁 = 0.
Hint: There will be a different set of J values for each combination of L and S. For
example, if your answer to part (a) was L = 2 or L = 1 (I hope you didn’t get these
numbers, because they aren’t correct) and your answer to part (b) was S = 0 or
S = 1, then there would be 4 sets of answers to part (c). Word your answer similar
to:
“For L = 2 and S = 0, J can be . . . ”
“For L = 2 and S = 1, J can be . . . ”
7.7 A Bit More on Compatible and Incompatible Observables 155
7.4 An atom has two electrons. One electron has 𝓁 = 1 and the other has 𝓁 = 2.
7.5 The atom in Problem 3 has a nuclear spin of I = 3. Find all possible values of
F.
Hint: There will be a different set of F values for each J value.
Open Access This chapter is licensed under the terms of the Creative Commons Attribution 4.0
International License (http://creativecommons.org/licenses/by/4.0/), which permits use, sharing,
adaptation, distribution and reproduction in any medium or format, as long as you give appropriate
credit to the original author(s) and the source, provide a link to the Creative Commons license and
indicate if changes were made.
The images or other third party material in this chapter are included in the chapter’s Creative
Commons license, unless indicated otherwise in a credit line to the material. If material is not
included in the chapter’s Creative Commons license and your intended use is not permitted by
statutory regulation or exceeds the permitted use, you will need to obtain permission directly from
the copyright holder.
Electronic Structure and Atomic Notation
8
Abstract
Learning Goals
• the Coulomb interaction and its role in determining the energy levels within an
atom.
• describe how electrons fill shells and subshells according to quantum mechanical
principles.
• the concept of angular momentum in the context of atomic physics and how it
affects atomic energy states.
• the influence of Coulomb interactions, electron shell filling, and angular momen-
tum on the energy of atomic states.
• interpret and describe electronic configurations and term symbols to understand
the angular momentum of individual electrons and the overall atom.
• explain the difference between fine structure splitting and hyperfine structure
splitting.
• distinguish between fermions and bosons and explain their significance in atomic
and nuclear physics.
Figure 8.1 shows the energy levels for hydrogen, helium, lithium, and europium.
As a reminder, hydrogen has 1 electron, helium has 2, lithium has 3, and europium
has 63. Notice how different the energy levels are! Hydrogen’s first excited state is
over 80,000 .cm−1 above the ground state, while helium’s first excited state is around
200 000
150 000
E (cm–1 )
100 000
50 000
0
Hydrogen Helium Lithium Europium
Fig. 8.1 The energy levels of hydrogen, helium, lithium, and europium. The red line at the top of
each element is the ionization threshold, which is the energy required to remove an electron from
the atom
8.1 Energy Level Spacings 159
E (cm–1 )
40 230
40 220
40 210 Europium
160,000 .cm−1 above the ground state!1 The hydrogen states seem to get closer and
closer to each other as the energy increases, but helium seems to “clump” a bit
more. And look at Europium; there seems to be a big gap between the ground state
before a really dense set of energy levels, a gap, and then even more! The red line
on each element is the energy needed to rip an electron from the atom. We call this
the ionization threshold.
Even though the energy levels are so densely packed, they are still discrete.
The energy levels looking like a solid band is just an artifact of making a picture
with many, many discrete energy levels (the europium diagram has 500 levels in
it). Figure 8.2 is a zoom in of the Europium energy levels from 40,210 .cm−1 to
40,250 .cm−1 , which is right in the middle of one of the dense patches. As you can
see, these 7 levels are very close together, but still discrete. Plotting all 500 levels
together really highlights the groupings.
Guiding Question
Why are the energy levels so different for each element?
Give it some thought and then read on.
1 Remember that the ground states of hydrogen, helium, lithium, and europium do not have 0
energy. Each element’s ground state has an energy like the lowest mode of a standing wave has
energy. The actual ground state energy of hydrogen is very different than that of helium or lithium.
But, we always do spectroscopy with respect to the element’s ground state energy. Figure 8.1 is
a nice learning tool, but it can be misleading by giving the impression that, for example, the first
excited state of lithium has a similar energy to the first excited state of europium. They do not!!
The energy difference compared to their ground state is similar. A subtle, but important difference.
☺
160 8 Electronic Structure and Atomic Notation
Definitions
• Coulomb force: The force between two charged particles, described by F =
ke qr1 q2 2 , where F is the force, ke = 8.987 × 109 N · m2 /C2 is Coulomb’s
constant, q1 and q2 are the charges, and r is the distance between them. It
represents the attraction or repulsion between the charged particles.
• Coulomb interaction: The interaction between two charged particles due
to their electric charges. It is governed by the Coulomb force and is
fundamental in understanding the behavior of charged particles.
We use the phrase Coulomb interaction when we are thinking about general
interactions between charged particles, such as how these interactions can affect
energy levels. We use the phrase Coulomb force when we are specifically
thinking about the force one charged particle exerts on the other.
There are many interactions that happen inside the atom. Each of these interactions
affects where the discrete energy levels end up. If we, as physicists, understand all
of these interactions we should be able to calculate properties of the atom like the
energy of a particular state, the natural linewidth of each transition, and what would
happen if we put the atoms in a particular situation such as inside an electric field, a
magnetic field, or if we put a bunch of these atoms together so strongly compressed
under gravity that the gravitational energy is converted into heat and eventually the
atoms fuse together.2 While there are many interactions happening in the atom,
the two biggest contributions that determine the energy of an atomic state are the
Coulomb interaction and how electrons fill up “shells.”
The Coulomb interaction is how we describe the interaction between charged
particles. For the hydrogen atom, there is only 1 Coulomb interaction. That
interaction is between the single electron, which has negative charge, and the single
proton, which has positive charge. In fact, this interaction is so simple we can exactly
solve the Schrödinger equation to find the effect this interaction has on the energy
levels. The helium atom has two electrons and two protons. The two protons are
inside the nucleus and, while there is technically a Coulomb interaction between the
two protons, that interaction is overwhelmed by the strong nuclear force, which is a
topic in Chap. 10. Therefore, we can think of the nucleus as a single particle with a
charge of .+2. The two electrons on the other hand are not confined tightly together
like the two protons in the nucleus. For the helium atom, there are effectively
3 Coulomb interactions: one between the two electrons, one between one of the
electrons and the nucleus, and one between the other electron and the nucleus. The
increased number of interactions causes most of the dramatic difference between
the energy levels of hydrogen and helium.
Lithium has 3 electrons and europium has 63. Like helium, we can think about
the nucleus of lithium as 1 particle with a charge of +3. Even though europium
has 63 protons in the nucleus, we can still think of the nucleus as 1 particle with
a charge of +63.3 Now we have a lot of interactions. Each electron has a Coulomb
interaction with every other electron and each electron has a Coulomb interaction
with the nucleus. One of the homework problems for this chapter will have you
count the number of Coulomb interactions for the first few atoms in the periodic
table.
The second major contribution comes from something called electron shells,
which are made up of electron subshells. The first shell is called 1, and it contains
1 subshell labelled 1s, see Fig. 8.3. The second shell is 2 and has two shells. The
third shell is 3 and has 3 subshells and so on. In Fig. 8.3, we stop showing all
possible subshells on shell 5. There is a 5g subshell, 6f, 6g, 6h, etc. that is not
shown. In quantum mechanics and atomic physics, the shell number .(1, 2, 3, . . . )
3 The neutrons in the nucleus also matter, but not as much as the proton charge. The effect the
neutrons have on the energy states is a topic in Chap. 10.
162 8 Electronic Structure and Atomic Notation
is a quantum number called the principal quantum number, represented by the letter
n. This is the same quantum number n explored in Chap. 6.
The electron subshells come about because an electron in an atom can have
orbital angular momentum and spin. The letter accompanying the principal quantum
number is the orbital angular momentum quantum number for the electron in that
subshell. The subshells use the historical letter designations discussed in Chap. 7.
As a reminder, if an electron has the label s, it has .𝓁 = 0. If it has a label p, it has
.𝓁 = 1, and so on. For example, 3p tells us that .n = 3 and .𝓁 = 1. From solving the
Schrödinger equation, we find that n is the upper limit on possible values for .𝓁. As
a reminder, .𝓁 is a zero or a positive integer. From the math, we find .𝓁max = n − 1.
For example, if .n = 3, then .𝓁 = 0, 1, or 2. That means the .n = 3 shell has three
subshells: 3s, 3p, and 3d. There is no 3f subshell since .𝓁 = n, which is not allowed.
An s-subshell can hold 2 electrons in total, one will have spin up .(ms = +1/2)
and the other spin down .(ms = −1/2). A p-subshell can hold 6 electrons in total: 3
spin up and 3 spin down. A d-subshell can hold 10 electrons (5 up and 5 down), an
f-subshell can hold 14, a g-subshell can hold 18, etc. The subshells fill in a specific
order roughly shown in Fig. 8.3.
The first subshell to fill is the 1s subshell, which can hold 2 electrons. After the 1s
subshell is filled, electrons start to fill the 2s subshell, which also holds 2 electrons.
Once the 2s subshell is filled, electrons start to fill the 2p subshell, which holds 6
electrons. Next is 3s followed by 3p, 4s, 3d, 4p, etc.
Why Can a p-Shell Hold 6 Electrons? An electron with the label p means that
.𝓁 = 1. There are three possible orientations for this angular momentum vector:
.m𝓁 = −1, 0, and 1. In addition, the electron can be either spin up or spin down.
That means there are 6 orientations for an electron to have .𝓁 = 1. Using the notation
(.m𝓁 , .ms ), the possible orientations for a p-subshell are:
⎛ ⎞ ⎛ ⎞ ⎛ ⎞ ⎛ ⎞ ⎛ ⎞ ⎛ ⎞
. 1, + 12 1, − 12 0, + 12 0, − 21 −1, + 12 −1, − 12
An electron with the label d means that .𝓁 = 2. There are 5 possible orientations
for this vector (.m𝓁 = -2,-1,0,1,and 2). Including spin up and spin down for each .m𝓁
gives 10 possible orientations. Using the notation (.m𝓁 , .ms ), the possible orientations
for a d-subshell are:
⎛ ⎞ ⎛ ⎞ ⎛ ⎞ ⎛ ⎞ ⎛ ⎞
2, + 12 1, + 21 0, + 12 −1, + 12 −2, + 12
. ⎛ ⎞ ⎛ ⎞ ⎛ ⎞ ⎛ ⎞ ⎛ ⎞
2, − 12 1, − 21 0, − 12 −1, − 12 −2, − 12
Figure 8.4 are Bohr model pictures of hydrogen, helium, and lithium in their
lowest energy state (the ground state). Remember that electrons are actually more
like waves, but I like to draw them as little balls to explore this concept. Hydrogen
has 1 electron so the 1s subshell is half filled. Helium has 2 electrons that fully fill
8.2 The Coulomb Interaction and Electron Shells 163
Fig. 8.4 A Bohr model picture of hydrogen, helium, and lithium. Electrons will first fill the 1s
subshell. Hydrogen half fills the 1s subshell, while Helium fills the 1s subshell completely. Lithium
has 3 electrons that completely fill the 1s subshell leaving a single electron in the 2s subshell
100 000
75 000
E (cm–1 )
50 000
25 000
0
Hydrogen Lithium Sodium Potassium Rubidium Cesium Francium
Fig. 8.5 The energy levels for the first column of the periodic table. Each of these elements has
fully closed shells except for a single electron in an s-shell
164 8 Electronic Structure and Atomic Notation
have 1 electron sitting outside a closed shell. Therefore, the energy levels of these
atoms are very similar to hydrogen.
We write the electron configuration with superscripts that tell us how many
electrons are in a subshell; if there is no superscript present, there is only 1 electron
in that subshell. For example, the electronic configuration for the ground state of
helium is .1s2 ; there are 2 electrons in the 1s subshell. If we excite the atom so that
1 electron moves to the 3d subshell, the electronic configuration is 1s3d; 1 electron
is in the 1s subshell and 1 is in the 3d subshell.
As we move through the periodic table, the atoms gain more and more electrons
filling up shells according to Fig. 8.3. I should note that while the chart does a really
good job, particularly for lighter atoms, it messes up sometimes for heavier atoms.
For example, the chart predicts that the ground state of silver (47 electrons) should
have a filled 5s subshell and 9 electrons in the 4d subshell:
1s2 2s2 2p6 3s2 3p6 3d10 4s2 4p6 5s2 4d9
.
However, the actual electronic configuration has a completely filled 4d subshell and
a single electron in the 5s subshell:
As physicists, we ask, “Why?” Why would silver decide that filling the 4d subshell is
better than following that chart and keeping the 5s subshell filled? The short answer
is that nature tends to try and get to a configuration with the lowest overall energy.
As an analogy, think about a ball in a bowl. The ball will want to hang out at the
bottom of the bowl, which is the spot of lowest energy. Because there are so many
electrons in silver and so many interactions, the electron configuration that fills the
4d shell has less energy than if the 5s shell was filled.
Quick quiz:
Considering the silver atom, how many electrons are in the 3d subshell? The
answer is in the footnotes.4
4 10
8.3 Term Symbols 165
Here is the electron configuration for the ground state of europium, which has 63
electrons:
1s2 2s2 2p6 3s2 3p6 3d10 4s2 4p6 4d10 5s2 5p6 6s2 4f7
.
From this electronic configuration we can see which shells are filled. After the 6s
subshell fills up, the 4f subshell starts to fill until we run out of electrons. There are
7 electrons in the unfilled 4f subshell. Combined with the Coulomb interactions, the
end result is a pretty complicated energy level diagram, see Figs. 8.1 and 8.2. This
is why I’m glad to be an experimentalist. I don’t have to try to calculate these levels,
just measure them ☺.
Summary
The dominant interactions in atoms that determine the energy level locations
are the Coulomb interaction and the way electrons fill shells.
Reminder Definition
• Torque: A measure of how effective a force is in causing an object to
rotate. Applying torque changes the angular momentum of an object. There
is energy associated with that torque.
We spent almost all of Chap. 7 studying angular momentum and emphasizing the
orientation of the vectors. The theory of electricity and magnetism reveals an
important fact: because electrons and the nucleus have charge, and the electrons
have angular momentum, everything in the atom exerts a torque on everything else.
A Brief Aside
The Coulomb interaction is used to describe the interaction between charged
particles, resulting in an electric field between the charges. This approach
ignores the motion associated with their angular momentum. When we
include angular momentum, it is more appropriate to use the electromagnetic
interaction because moving charges create magnetic fields. Therefore, the
interaction between moving charged particles involves both electric and
magnetic forces, which is what the electromagnetic force describes.
166 8 Electronic Structure and Atomic Notation
Definitions
• Electromagnetic force: A fundamental force encompassing both electric
and magnetic forces, including the Coulomb force and magnetic forces due
to moving charges.
• Electromagnetic interaction: The interaction between charged particles
due to both electric and magnetic forces, including interactions involving
stationary charges (electrostatics) and moving charges.
The magnetic forces from the moving charges cause the internal torques. The
magnitude of this internal torque depends on the orientation and size of the angular
momentum vectors. In short, both the size and the orientation of the angular
momentum vectors impact the energy of a state. That is so important, I’m going
to give the sentence its own red box.
Important
Both the size and the orientation of the angular momentum vectors impact the
energy of a state.
where S is the quantum number representing the total spin of all the electrons, L
is the quantum number for the total orbital angular momentum of all the electrons,
and J is the quantum number for the total angular momentum of all of the electrons.
For example, suppose an energy level has a term symbol .3 D2 . That tells us that the
spin of the electrons has quantum number .S = 1 .(2S + 1 = 3), the orbital angular
momentum of the electrons has quantum number .L = 2 (represented by the letter
D), and the total electronic angular momentum of the electrons has quantum number
.J = 2. We can use those numbers to calculate the magnitude of angular momentum.
8.3 Term Symbols 167
Table 8.1 The first 6 states of atomic oxygen. The last column is the frequency of a laser needed
if we were to try to excite the atom from the ground state to that state
Electron configuration Term symbol Energy .(cm−1 ) f (Hz)
2 2 4
.1s 2s 2p
3
. P2 0
2 2 4
.1s 2s 2p
3
. P1 158.265 .4.74 × 1012
.6.80 × 10
2 2
.1s 2s 2p
4 3
. P0 226.977 12
.4.76 × 10
2 2
.1s 2s 2p
4 1
. D2 15,867.862 14
.1.01 × 10
2 2 4
.1s 2s 2p
1
. S0 33,792.583 15
.2.21 × 10
2 2 3
.1s 2s 2p 3s
5
. S2 73,768.200 15
1. The term symbol is important not just because it tells us three quantum
numbers for the electrons in an atom as a whole, but also because those
three quantum numbers affect the energy of a state.
2. The observables represented by S, L, and J are all compatible with energy.
The shift of an energy level due to electrons having angular momentum is called
fine structure splitting.5 The nucleus can also have angular momentum (often called
nuclear spin), which is the topic of Chap. 9. The shift in the energy of a state due to
the nucleus having angular momentum is called hyperfine structure splitting.6
Definitions
• Fine structure splitting: The shift in the energy of a state due to the
electrons having orbital angular momentum and spin.
• Hyperfine structure splitting: The shift in the energy of a state due to the
nucleus having angular momentum.
Table 8.1 shows the six lowest energy states of atomic oxygen (8 electrons, 8
protons, and 8 neutrons). This particular type of oxygen is known as oxygen-16.
Oxygen-16 has no nuclear angular momentum, so the nucleus does not play a role
in the following discussion. Notice that the first five energy states all have the same
electronic configuration but different term symbols. The next few paragraphs are all
about adding lots and lots of angular momentum vectors together. We are going to
start with the ground state and then work our way through the table.
J = (L + S), . . . , |L − S|
.
in integer steps
The ground state of oxygen has the designation .1s2 2s2 2p4 .3 P2 . This designation tells
us that both the 1s and 2s subshells are completely filled, and we can essentially
ignore them. The 2p subshell is partially filled with 4 electrons (the 2p subshell
can hold up to 6 electrons). From the electronic configuration, we know the orbital
angular momentum of each electron. The 4 electrons in the 2p subshell each have
.𝓁 = 1, or each electron in this subshell has orbital angular momentum with size
√ √
. 𝓁(𝓁 + 1)h̄ = 2h̄.
Each of the electrons in the 2p subshell have the same magnitude of orbital
angular momentum (.𝓁 = 1), but have different orientations. If we were to add all
four of those vectors together, we will get a new vector that represents the orbital
angular momentum of all four electrons as a composite system. For the ground state,
the orbital angular momentum of the composite system has a magnitude represented
by quantum number .L = 1 (P in the term symbol).
Those same four electrons also have individual spin vectors that add up to .S = 1
.(2S + 1 = 3). Again, that spin vector represents the total spin of all four electrons
as a composite system. Since the orbital angular momentum vector for all four
electrons has a size and orientation and the spin vector for all four electrons has
a size and orientation, we can ask the question, “What are the possible magnitudes
for the total electronic angular momentum for the composite system?” The answer
is .J = 2, 1, or 0. The orientations that produce a vector whose magnitude is
represented by .J = 2 has the lowest energy, so that is the ground state.
8.3 Term Symbols 169
Now let’s move on to the next energy level. This level has an energy that is
.158.265 cm−1 larger than the ground state. The only difference in the designation
for the ground state and this state is that J in the term symbol (.J = 2 for the ground
state and .J = 1 for this state). The orbital angular momentum for all four electrons
and the spin for all four electrons still add together, but the result has a different
total electronic angular momentum for the composite system. The orientation of all
the electrons to produce this new composite vector has a different internal torque,
so that orientation has a different energy than the .J = 2 ground state.
The first 5 energy levels of atomic oxygen have the same electron configuration:
2 2 4
.1s 2s 2p . To determine how the angular momenta of the last 4 electrons combine,
we look at the term symbols. As you can see from the table, there are multiple ways
this can happen, each resulting in a different energy level. In a hypothetical world
without the internal torque, all 5 of these levels would be degenerate. However,
due to the internal interactions, these levels split into the 5 distinct levels observed
experimentally. This phenomenon is known as fine structure splitting.7
While the nucleus of oxygen-16 has no angular momentum, a different nucleus
might. That nuclear angular momentum, whose magnitude is represented by the
quantum number I , exerts an additional internal torque that, once again, shifts the
energy of the states. This additional internal torque “breaks” the degeneracy of the
fine structure states. For example, oxygen-17 has nuclear spin .I = 5/2 (6 possible
orientations). The ground state has a total electronic angular momentum of .J = 2
(5 possible orientations). Adding together vector J and vector I produces a new
angular momentum vector whose magnitude is represented by the quantum number
F with possible values that range from .J + I to .|J − I | in integer steps, or in
this case .F = 9/2, 7/2, 5/2, 3/2, and 1/2. So the ground state of oxygen-17 further
splits into six hyperfine levels, each with a different energy. From the “important
foreshadowing statement,” the orientation of a particular F vector will not change
the energy. If .F = 3/2, there are four orientations represented by .mF = 3/2, 1/2,
.−1/2 and .−1. All four of these states are degenerate. We will explore hyperfine
structure and how the nucleus affects our energy levels further in Chap. 9.
7 Forcompleteness, Einstein’s theory of special relativity, which deals with the physics of fast-
moving objects, also shifts the energy of a state and is included as part of fine structure splitting,
but we will not explore that here. If this book has inspired you to pursue further studies in quantum
mechanics ☺, you will learn about perturbation theory and the effects of electron velocity.
170 8 Electronic Structure and Atomic Notation
Common Question
If .J = 2, there are 5 states with different orientations that have the same
energy. Can we do anything to “split” those final orientations so they have
different energies? Yes! The internal torque doesn’t do it, but we can apply
a torque from the outside to “break the degeneracy”. Applying an external
torque using an external magnetic field is called the Zeeman Effect, named
after the Dutch physicist Pieter Zeeman. Applying an external torque by
applying an electric field is called the Stark Effect, named after the German
physicist Johannes Stark. Both of these effects “break” the degeneracy of
those states.
Summary
The orientation and size of angular momentum impact the energy of a
state. For an atom with no nuclear spin, a state can be described by the
individual electron configurations, which tells us the angular momentum
of each individual electron, and the term symbol, which tells us about the
orientation of the angular momentum vectors of the electrons as a whole.
You may have learned about orbitals in high school chemistry. Orbitals are visual
representations of different energy states for an electron in a hydrogen atom.
Figures 8.6 and 8.7 show some orbital pictures for different energy states in
hydrogen. These are analogous to the shaking energy modes we studied in Chap. 1
(Fig. 1.7) and Chap. 6 (Fig. 6.2). There are 3 numbers on each plot given in bra-ket
notation. The first number is the principal quantum number (also called the shell
number). The second is the orbital angular momentum quantum number, and the
last is the projection of the orbital angular momentum quantum number along the
z-axis (orientation). For example, .|4 3 -1〉 means .n = 4, .𝓁 = 3, and .m𝓁 = −1.
You can think of all of these orbitals as different standing waves of the electron.
The “loops” are a bit harder to see in 3-dimensional space, so in Fig. 8.6 we plot
a few different orbitals in 3D and in Fig. 8.7 we plot a few cross sectional views.
The reason they are so hard to visualize is that we only have 3 dimensions to view
4 dimensions of information.
8.4 Connecting Angular Momentum to Orbitals 171
Fig. 8.6 3D illustrations of different electron orbitals in a hydrogen atom. The darker the shade
the larger the amplitude of the standing wave
For example, if you look at Fig. 6.2 you’ll see that we need two dimensions to
see the energy state for 1 dimension of shaking energy. The vertical axis shows the
amplitude of the energy state while the horizontal shows position in 1 dimension.
An example of a 2 dimensional standing wave would be a drum head vibrating. To
visualize a 2 dimensional standing wave, we need a 3 dimensional plot: 1 dimension
for the amplitude and 2 for the position dimensions. We run into a problem for
Fig. 8.7 A cross sectional view of different electron orbitals in the Hydrogen atom. The closer the
color is to white, the larger the amplitude of the standing wave
172 8 Electronic Structure and Atomic Notation
In atomic and nuclear physics, particles are categorized into two major groups:
fermions and bosons. Fermions are particles with half-integer spin quantum
numbers (e.g., 1/2, 3/2, etc.). Examples of fermions are electrons, protons, and
neutrons, all of which have a spin quantum number of 1/2. Bosons are particles
with integer spin quantum numbers (e.g., 0, 1, etc.). So far, we have encountered
only one boson, the photon, which has a spin quantum number of 1.
This is important because two identical fermions cannot “hang out” in the same
space while two identical bosons can. This is the essence of the Pauli Exclusion
Principle: two fermions cannot simultaneously occupy the same quantum state; that
is, no two fermions can have the same set of quantum numbers within a quantum
system. As an example, consider the two electrons, both of which are fermions, in
the atomic ground state of a helium atom, which has the electronic configuration .1s2 .
Let’s write out the quantum numbers for the two electrons using bra-ket notation
with the quantum numbers .|n 𝓁 m𝓁 s ms . The two electrons have quantum numbers:
First Electron: |1 0 0 1
2 2〉
1
.
Second Electron: |1 0 0 1
2 - 12 〉
Notice that the two electrons have a different set of quantum numbers. If they
did have the same quantum numbers, the math from quantum mechanics shows
that the two electron wavefunctions would cancel each other out.8 This is a very
bad thing. If the individual electron wavefunctions added together to produce no
wavefunction, neither electron would exist. Therefore, the two electrons must have
different quantum numbers. This is what is meant by the phrase, “two identical
fermions cannot hang out in the same space”. They must have a different set of
quantum numbers.
8 More specifically, the wavefunction for the two-electron system is “antisymmetric” and would be
The Pauli Exclusion Principle is part of why electron shells and subshells
exist. Every electron in the atom must have a different set of quantum numbers.
Interestingly, bosons don’t have this problem. Two bosons can have the same set
of quantum numbers and not destructively interfere each other out of existence. So,
two identical bosons can hang out in the same space. In fact, the two bosons can
constructively interfere with each other.
There can also be composite fermions and bosons. For example, helium-4 is a
system that acts as a composite boson. Helium-4 contains 2 protons, 2 neutrons, and
2 electrons. All of these particles are fermions, but they can pair up to behave like
bosons. Other examples of composite bosons include Cooper pairs9 (important for
superconductors) and Bose–Einstein condensates10 . Helium-3, which has 2 protons,
2 electrons, and 1 neutron, is a composite fermion.
Problems
8.1 As discussed in the chapter, hydrogen has 1 Coulomb interaction and helium
has 3. Assuming the nucleus is one big particle with a positive charge, how many
Coulomb interactions do lithium, beryllium, and boron have?
Challenge: What is the general formula to calculate the number of Coulomb
interactions for an atom with N electrons, assuming the nucleus is one big particle
with positive charge?
8.2 The electron configuration for hydrogen is 1s. There is 1 electron in the first
s subshell. The electron configuration for helium is 1s2 . That means there are 2
electrons in the first s subshell, which completely fills the shell. Lithium has 3
electrons, so the electron configuration is 1s2 2s. Notice since the first shell is filled,
we begin to fill the second shell. Write out the electron configurations for
9 Also known as Bardeen–Cooper–Schrieffer pairs named after American physicists John Bardeen,
Note: The electron configurations are something that you can easily find on the
internet. You can also find the answers in Appendix B. Don’t search for the answer
before you try yourself first.
8.3 For each of the following term symbols, what is the magnitude of the total
electronic spin, the total electronic orbital angular momentum, and the total
electronic angular momentum?
(a) 3P
2
(b) 3P
0
(c) 1S
0
(d) 1F
2
8.4 Appendix B has a list of all of the elements with their ground state electronic
configurations. Look through the list and find all the atoms that will have an energy
level structure similar to hydrogen (i.e. all the shells are filled except for a single
electron in the last s subshell). All of these elements will have energy level diagrams
that look similar to hydrogen.
8.6 In Sect. 8.3 when exploring oxygen, we had the sentence, “This designation
tells us that both the 1s and 2s subshells are completely filled, and we can essentially
ignore them.” Let’s explore this sentence a bit more.
100 000
E (cm–1 )
50 000
0
Lithium Beryllium Ion
8.5 Fermions and Bosons 175
(a) Write the term symbol for the two electrons that fill the 1s subshell. Explain
the reasoning you used to determine the term symbol. The answer (without
explanation) is in the footnotes.11
(b) Write the term symbol for the two electrons that fill the 2s subshell. Explain
your reasoning.
(c) Write the term symbol for the four electrons with the electronic configuration
1s2 2s2 . This is the same as the ground state of a beryllium atom, which has 4
electrons. Explain your reasoning.
(d) Write the term symbol for the ground state of neon. Neon has 10 electrons, so
the 2p subshell is completely filled. Explain your reasoning.
(e) Write the term symbol for the ground state of lithium, which has 3 electrons.
Explain your reasoning.
(f) Write the term symbol for the ground state of cesium, which has all subshells
filled except for the last electron that is in the 6s subshell. Explain your
reasoning.
(g) Boron has all subshells filled except for the last electron that is in the 2p subshell.
For that last electron, there are two possible values for J . What are they?
(h) For part (g), the term symbol with the lower value of J corresponds to the lower
energy and thus the ground state. Write the term symbol for the ground state of
boron.
8.7
(a) Make a sketch of a saturated absorption spectroscopy spectrum where the lower
state has J = 0 and the upper state has J = 1. The transition frequency is
fr = 652, 000, 000 MHz and the natural linewidth is 1 MHz.
Next, we put the atoms in an external magnetic field.(The magnetic
) field shifts
the energy levels according to the formula ΔE = 1 GHz T h m J B ext , where
Bext is the external magnetic field and the unit T stands for Tesla (the unit for
magnetic field).
(b) We place the atoms in a uniform magnetic field with Bext = 0.1 T. Make a sketch
of a saturated absorption spectroscopy spectrum where the lower state has J = 0
and the upper state has J = 1. Don’t worry about the relative amplitudes.
(c) Optional Challenge: (Next, we )place the atoms in a magnetic field described by
the function Bext = 0.001 cm T
z. Make a sketch of the three J = 1 states as a
function of z. Use frequency units instead of Joules for the vertical axis (this is
equivalent to using a laser to excite the atoms).
11 1 S
0.
176 8 Electronic Structure and Atomic Notation
Open Access This chapter is licensed under the terms of the Creative Commons Attribution 4.0
International License (http://creativecommons.org/licenses/by/4.0/), which permits use, sharing,
adaptation, distribution and reproduction in any medium or format, as long as you give appropriate
credit to the original author(s) and the source, provide a link to the Creative Commons license and
indicate if changes were made.
The images or other third party material in this chapter are included in the chapter’s Creative
Commons license, unless indicated otherwise in a credit line to the material. If material is not
included in the chapter’s Creative Commons license and your intended use is not permitted by
statutory regulation or exceeds the permitted use, you will need to obtain permission directly from
the copyright holder.
Hyperfine Structure
9
Abstract
Learning Goals
• why hyperfine structure exists and identify the conditions under which hyperfine
splitting occurs.
• the difference between fine structure and hyperfine structure.
• how to interpret spectroscopic data to extract hyperfine constants.
The nucleus contains protons and neutrons, each of which has intrinsic angular
momentum. In addition to their intrinsic spins, nucleons (protons and neutrons) can
also have orbital angular momentum due to their motion within the nucleus. When
discussing the angular momentum of the nucleus as a whole, we primarily refer
to the combination of the intrinsic spins of the nucleons and their orbital angular
momentum. This combined angular momentum should be called the “total nuclear
angular momentum,” but it is commonly referred to as “nuclear spin.” Despite the
terminology, neither the nucleus nor the protons and neutrons in the nucleus are
literally spinning or orbiting in the classical sense. Nuclei with both an even number
of protons and an even number of neutrons tend to have zero nuclear spin because
the individual spins pair up (one spin-up and one spin-down) and cancel each other
out. Nuclei with an odd number of protons and/or neutrons generally have a non-
zero nuclear spin.
When an atom has a non-zero nuclear spin, the nucleus interacts with the
magnetic fields produced by the electrons. More specifically, because the nucleus
has both charge and angular momentum (I ), and the electrons have both charge and
angular momentum (J ), there is an additional internal torque between the nucleus
and the electrons. This interaction leads to the splitting of atomic energy levels into
closely spaced sub-levels known as hyperfine levels, analogous to the interaction
between an electron’s spin and its orbital angular momentum, which results in fine
structure splitting. I want to point out that atomic physicists use hyperfine levels
and hyperfine states interchangeably. We tend to use hyperfine levels when thinking
about energy splittings and hyperfine states when using quantum numbers, but the
two terms mean the same thing.
Table 9.1 is a copy of a table from Chap. 7 that summarizes all of the angular
momentum vectors and angular momentum quantum numbers for the system of
electrons and the atom as a whole. Let’s look at some examples. Oxygen has three
stable isotopes: oxygen-16 ( 99.76% of all oxygen on earth is oxygen-16), oxygen-
17 (.∼0.04%), and oxygen-18 (.∼0.20%). Isotopes are elements with the same
number of protons but different numbers of neutrons. Oxygen-16 has 8 electrons,
8 protons, and 8 neutrons. Oxygen-17 has 8 electrons, 8 protons, and 9 neutrons.
Oxygen-18 has 8 electrons, 8 protons, and 10 neutrons. Since each isotope has a
different number of neutrons, the transition frequencies are slightly different. This
small shift, called an isotope shift, will be discussed in Chap. 10.
Table 9.1 This table summaries all of the angular momentum quantum numbers
Type QN Rule Formula
√
Orbital L Zero or positive integer Magnitude: . L(L + 1)h̄
.mL .−L to .+L in integer steps Cone height: .mL h̄
√
Spin S Zero, positive integer, or half-integer Magnitude: . S(S + 1)h̄
.mS .−S to .+S in integer steps Cone height: .mS h̄
√
Total electronic J Zero, positive integer, or half-integer Magnitude: . J (J + 1)h̄
.mJ .−J to .+J in integer steps Cone height: .mJ h̄
√
Nuclear spin I Zero, positive integer, or half-integer Magnitude: . I (I + 1)h̄
.mI .−I to .+I in integer steps Cone height: .mI h̄
√
Total atomic F Zero, positive integer, or half-integer Magnitude: . F (F + 1)h̄
.mF .−F to .+F in integer steps Cone height: .mF h̄
9.1 Hyperfine Structure 179
Important Reminders
• We can represent the energy difference between two states using energy
units, wavelength units, or frequency units. For hyperfine structure, fre-
quency units are the most convenient unit because the energy spacing
between hyperfine levels is small.
• When adding two angular momentum vectors together, both the magnitude
and orientation of the individual angular momentum vectors will affect the
magnitude and orientation of the resulting vector. Each possible magnitude
of the composite angular momentum vector will have an effect on the
energy of a state.
• Center of gravity: The energy of a state if there was no nuclear spin.
The nuclear spin quantum number for oxygen-17 is .I = 5/2. The nuclear spin
quantum number for both oxygen-16 and oxygen-18 is .I = 0. Therefore oxygen-17
will have hyperfine structure while the other two isotopes do not. All isotopes of
oxygen have a ground state term symbol of .3 P2 , or .S = 1 (.2S + 1 = 3), .L = 1,
and .J = 2. However, the ground state of oxygen-17 looks different compared to
the other two isotopes, see Fig. 9.1. If oxygen-17 had no nuclear spin, it would have
a single level precisely at 0. Nuclear spin “splits” this single level into 5 hyperfine
levels. For example, the level labeled .F = 7/2 has a slightly smaller energy than
the center of gravity while the .F = 5/2 state has a higher energy. Both oxygen-16
and oxygen-18 have no nuclear spin, so they have a single ground state that would
be labeled 0 energy and have no F quantum number designation.
Adding together nuclear spin, represented by the quantum number I , and the total
electronic angular momentum, represented by the quantum number J , results in a
new angular momentum vector that we call the total angular momentum of the atom,
represented by the quantum number F . Hyperfine structure generally has a much
smaller energy splitting compared to fine structure. Like every angular momentum
vector we have encountered, the F vector can also point in different orientations. For
example, an .F = 3/2 state has four possible orientations described by the quantum
180 9 Hyperfine Structure
numbers .mF =3/2, 1/2, .−1/2, and .−3/2. All four of those orientations have the
same energy.
Compatibility with Energy We often care about what observables are compatible
with energy. At the end of Chap. 7, we used the ket .|n 𝓁 s j mj 〉 to represent the
states of the hydrogen atom (one electron). All the different quantum numbers in
this ket represent observables that are compatible with energy.
Important Reminder
The observables represented by .m𝓁 and .ms are not compatible energy.
Let’s update our ket with information we have learned from this chapter. When
we add together nuclear spin and total electronic angular momentum, the cones
representing those angular momenta are tilted with respect to the total atomic
angular momentum (F ). Just like when we added orbital and electron spin, the cone
heights of I and J , which are represented by the quantum numbers .mI and .mJ , are
no longer compatible with energy. Therefore, our new ket is .|n 𝓁 s j I F mF 〉.
The nucleus of a hydrogen atom has a nuclear spin represented by .I = 1/2. The
atomic ground state of hydrogen has the electronic configuration .1s, or .s = 1/2,
.𝓁 = 0, and .j = 1/2. Because the nucleus has spin, there will be two hyperfine
|1 0 1
2
1
2
1
2 1 1〉
|1 0 1
2
1
2
1
2 1 0〉
. (9.2)
|1 0 1
2
1
2
1
2 1 -1〉
|1 0 1
2
1
2
1
2 0 0〉
The three hyperfine levels with .F = 1 are degenerate and have the same energy.
The hyperfine level with .F = 0 has a different energy.
So, the ground state of hydrogen has two hyperfine levels represented by the
quantum numbers .F = 1 and .F = 0. For something a bit more complicated, let’s
look at the ground state of europium. Europium has 7 electrons in its last, unfilled
subshell. All isotopes of europium have a ground state term symbol .8 S7/2 , or .S =
7/2 (.2S + 1 = 8), .L = 0, and .J = 7/2. If europium had no nuclear spin, that would
9.2 Math 181
1s2 2s2 2p6 3s2 3p6 3d10 4s2 4p6 4d10 5s2 5p6 6s2 4f7 8 S7/2
. (9.3)
to have 0 energy. However, both stable isotopes of europium have nuclear spin:
europium-151 (63 electrons, 63 protons, 88 neutrons, and .I = 5/2) and europium-
153 (63 electrons, 63 protons, 90 neutrons, and .I = 5/2). Other isotopes of
europium will have different nuclear spin. For example, europium-152, which is
radioactive with a half-life of 13.5 years, has a nuclear spin quantum number of
.I = 3. Each of these europium isotopes have hyperfine levels.
Summary
If an isotope has no nuclear spin, there would be 1 ground state. If it does
have nuclear spin, there are multiple ground states. This is because the angular
momentum from the nucleus exerts a torque on the electrons, which results in
a small splitting and shift of the state’s energy.
There is a single exception to the above summary that we will discuss more in
Sect. 9.2. If the state has no total electronic angular momentum (.J = 0), there is
still only a single state; the single state does not split into multiple hyperfine levels.
However, that single state will still have an F quantum number as a label.
9.2 Math
We can find the possible values of F by using the largest to smallest in integer steps
rule. The rule is:
Example 1 Hydrogen-1 has a nuclear spin quantum number of .I = 1/2 and the
ground state has a total electronic angular momentum quantum number of .J = 1/2
(also .j = 1/2 since hydrogen has a single electron). Therefore, the possible F
values for the ground state range from .1/2 + 1/2 = 1 to .|1/2 − 1/2| = 0 in integer
steps. Thus, there are 2 hyperfine levels for the ground state of hydrogen-1.
Example 2 Oxygen-17 has a nuclear spin quantum number of .I = 5/2 and the
ground state has a total electronic angular momentum quantum number of .J = 2.
Therefore, the possible F values for the ground state range from .5/2 + 2 = 9/2 to
.|5/2 − 2| = 1/2 in integer steps. Thus, there are 5 hyperfine levels for the ground
182 9 Hyperfine Structure
state of oxygen-17. Those values, which are shown in Fig. 9.1, are .F = 9/2, 7/2,
5/2, 3/2, and 1/2.
Quick Quiz
Fig. 9.2 A fun maze to separate the quiz from the answers
Notice that beryllium-9 still has a single ground state even though the nucleus has
spin. This is because .(I + J ) = |I − J |, so the range of possible F values is 3/2
to 3/2. This always happens when .J = 0. However, the nucleus still has angular
momentum, so we include .F = 3/2 in the description of the state.
9.2 Math 183
Using quantum mechanics, we can derive the energy splitting of a hyperfine level
with respect to the center of gravity:
3
K(K+1)−2I (I +1)J (J +1)
ΔE = 21 KA + 2 2I (2I −1)2J (2J −1) B
.
K = F (F + 1) − I (I + 1) − J (J + 1) (9.5)
A = 0 unless both I > 0 and J > 0
B = 0 unless both I > 1/2 and J > 1/2
where A is called the magnetic dipole hyperfine constant and B is called the
electric quadrupole hyperfine constant. Notice that everything else in the above
equation apart from A and B is a quantum number. Theorists can calculate the
hyperfine constants A and B while experimentalists measure them. When we
perform spectroscopy on an atom with hyperfine structure, we can measure the
energy spacing between all of the hyperfine levels and use the above equation to
back out experimental values of A and B. Some hyperfine constants were measured
many years ago while others have yet to be measured. For example, the ground
state hyperfine constants for europium-151 and europium-153 were measured for
the first-time way back in 1960 by P.G.H. Sandars and G.K. Woodgate and published
in the journal Proceedings of the Royal Society A.1 Sandars and Woodgate found
that the magnetic dipole hyperfine constant and the electric quadrupole hyperfine
constant for the ground state of europium-151 is .A = −20.0523 ± 0.0002 MHz
and .B = −0.7012 ± 0.0035 MHz. To convert those numbers into energy, just plug
the hyperfine constants into the above equation and multiply the result by Planck’s
constant, h. If someone already measured those numbers, we can use those as a
starting point for our fitting algorithms. If not, we have to determine them ourselves.
For every state in an atom, the electrons have different quantum numbers. States
with higher n tend to be farther from the nucleus while the angular momentum
quantum numbers represent different orbitals. Therefore, for an atom with nuclear
spin, every state in that atom will have different hyperfine constants resulting in
a different hyperfine splitting. Even the same state in two different isotopes that
happen to have the same nuclear spin will have different hyperfine constants because
the nuclei of the two isotopes are slightly different.
Summary
The “splitting” of the center of gravity into hyperfine levels is described by
Eq. 9.5. The magnetic dipole hyperfine constant A is zero unless both .I > 0
and .J > 0. The magnetic quadrupole hyperfine constant B is zero unless both
.I > 1/2 and .J > 1/2.
1 How cool of a journal name is that?! See reference [2] for the full citation.
184 9 Hyperfine Structure
One Final Thing The magnetic dipole term (the term with A) in Eq. 9.5 tends
to be larger than the electric quadrupole term (the term with B). For example,
the .F = 6 ground hyperfine state of europium-151 has a splitting .ΔE =
(−175.458 MHz)+(−0.175 MHz) = −175.633 MHz. The first term in parentheses
is from the magnetic dipole term and the second is from electric quadrupole term.
There are additional terms to Eq. 9.5, but they are very small compared to the electric
quadrupole term. The next term in the formula is the magnetic octupole term,
which contains quantum numbers and the magnetic octupole hyperfine constant C.
This term is generally unnecessary unless you have exceptionally good data. The
magnetic octupole constant is zero unless both .I > 1 and .J > 1.
Suppose we have an atom with a nuclear spin quantum number of .I = 3/2. To help
distinguish between the lower and upper states, we will use primes on the quantum
numbers for the excited states. In this example, the lower state has a total angular
momentum quantum number of .J = 1/2 and the upper state has .J ' = 1/2. Our goal
is to write an equation for the transition frequency between two hyperfine levels. We
first need to find the possible values for F , which can range from .3/2 + 1/2 = 2 to
'
.|3/2 − 1/2| = 1 in integer steps, giving .F = 1 and .F = 2. Since .J = 1/2 as well,
' ' '
the possible values for .F are .F = 1 and .F = 2, see Fig. 9.3. In this hypothetical
example, the .F = 1 hyperfine level has a smaller energy than the .F = 2 hyperfine
level while the order is reversed in the excited state; the ordering of the quantum
number all depends upon the interaction with the nucleus.
Next, we want to find the transition frequency from the .F = 2 hyperfine level
to the .F ' = 1 hyperfine level. Using Eq. 9.5, we can calculate the hyperfine energy
splitting. We will use “LS” for lower state and “US” for upper state. Note that since
'
.J = 1/2 and .J = 1/2, both .BLS = 0 and .BUS = 0. Evaluating Eq. 9.5 with the
given quantum numbers, we find .ΔELS,F =2 = 34 ALS and .ΔEUS,F ' =1 = − 54 AUS .
According to Fig. 9.3, the .F = 2 state has more energy than the center of gravity for
the lower level, meaning .ΔELS,F =2 > 0. Since .ΔELS,F =2 = 34 ALS , we also learn
that .ALS > 0. For the upper state, the hyperfine energy splitting is also positive.
For this state, the formula is .ΔEUS,F ' =1 = − 54 AUS , which implies that .AUS < 0 in
order to make .ΔEUS,F ' =1 > 0.
Quick Quiz
Suppose an electron in our made up atom is in the .F = 2 lower hyperfine
level. What is the formula to calculate the transition frequency from the .F = 2
lower hyperfine level to the .F ' = 1 upper hyperfine level? Your answer should
contain the center of gravity frequency .fcog and the two magnetic dipole
hyperfine constants .ALS and .AUS . The answer is in the footnotes.2
The general formula to calculate the transition frequency between two hyperfine
levels is
For many learners, the signs of the shifts can be confusing, so let’s explore the signs
with our toy example. Let’s start with the minus sign in front of .ΔELS (ALS , BLS ).
According to Fig. 9.3, the .F = 2 hyperfine state reduces the transition frequency
compared to .fcog . Since .ΔELS (ALS , BLS ) > 0 for that state, the minus sign makes
sense. What about if we wanted the transition frequency from the .F = 1 lower
hyperfine level? For that state, .ΔELS (ALS , BLS ) < 0, so that minus sign turns
into a plus sign, increasing the frequency needed compared to .fcog . This is a
subtle but important point, so take some time to convince yourself the transition
frequency from any lower hyperfine level to any upper hyperfine level is given
by Eq. 9.6. Using this same logic, convince yourself that the plus sign in front of
.ΔEUS (AUS , BUS ) makes sense.
1.0
0.8 F'=5
Spectrum 0.6 F'=4
0.4
0.2 F'=3
0.0
– 50 0 50 100
f (MHz) + 657,932,340 MHz
Fig. 9.4 Experimental spectroscopic results of the 6s 2 S1/2 F = 4 → 7p 2 P3/2 transition in
neutral cesium-133. This is an experimental result from my research group, see reference [1]. A
simplified Grotrian diagram for the transition can be found in Fig. 5.12
the Doppler width for transitions from either state. Therefore, we will not have any
Λ crossovers despite having two lower states.
.
4A full list of the rules that need to be satisfied for an electron to transition between two atomic
states is given in Appendix C.
5 See Reference [4] for the full citation.
9.4 Example with Cesium-133 187
Let’s apply Eq. 9.6 using numbers from cesium-133. The spectrum in Fig. 9.4 is
from the .J = 1/2, .F = 4 ground state. The excited state has .J ' = 3/2 and .F ' =3,
4, and 5.6 Let’s find .ΔE for all of the features.
1. .F = 4 → F ' = 3
2. .F = 4 → F ' = 4
3. .F = 4 → F ' = 5
Let’s calculate the shift from the center of gravity for the .7p 2 P3/2 state for each
real transition. Since these are shifts for the excited state, we will use the primed
values in Eq. 9.5. We will also need Eq. 9.6 to find the transition frequency between
two hyperfine levels. Since the hyperfine splitting for the ground state is so large,
we will perform our spectroscopy with respect to the .F = 4 ground hyperfine level.
For ease of reference, here are those equations:
3
K(K+1)−2I (I +1)J ' (J ' +1)
ΔE = 12 KAUS + 2 2I (2I −1)2J ' (2J ' −1) BUS
K = F ' (F ' + 1) − I (I + 1) − J ' (J ' + 1)
. (9.7)
where .fF =4→cog = fcog − ΔELS,F=4 is the frequency of light needed to go from the
F = 4 ground hyperfine level to the center of gravity of the .7p 2 P3/2 state. I also
.
replaced the quantum numbers with primed values and the hyperfine constants with
“US” subscripts.
Let’s do the math for the .F = 4 → F ' = 3 real transition. First, let’s find K:
Now we can find the hyperfine splitting for the .F ' = 3 state from the center of
gravity for the .7p 2 P3/2 state.
6 Thereis also an .F ' = 2 excited hyperfine level, but we cannot excite an atom from the .F = 4
ground hyperfine level to this state.
188 9 Hyperfine Structure
Table 9.2 The energy shifts for the hyperfine levels with respect to the center of gravity for the
3/2 state in cesium-133. All of the real transitions in this example are from the .6s S1/2 F =
.7p
2P 2
.F = 4 → F' = 4 .4(4 + 1) − 72 ( 72 + 1) − 32 ( 32 + 1) = + 12 .+
1
4 AUS − 13
28 BUS
.F = 4 → F' = 5 .5(5 + 1) − 72 ( 27 + 1) − 32 ( 32 + 1) = + 21
2 .+
21
4 AUS + 14 BUS
3
K(K+1)−2I (I +1)J ' (J ' +1)
ΔE = 12 KAUS + 2 2I (2I −1)2J ' (2J ' −1) B
⎛ ⎞⎛⎛ ⎞ ⎞ ⎛US⎞⎛ ⎞⎛ ⎞⎛ ⎞
⎛ ⎞ 3 15
− 2 +1 −2 72 72 +1 32 23 +1
15
2 − 2
= 2 − 2 AUS +
1 15 ⎛
7
⎞⎛ ⎛
7
⎞ ⎞ ⎛
3
⎞⎛ ⎛
3
⎞ ⎞ BUS
2 2 −1 2 2 −1
⎛ ⎞⎛ ⎞ 2⎛ ⎞⎛ 2 ⎞⎛ ⎞ 2 2
− 45
4 − 13
2 −7 2 2 52
9 3
4 AUS +
= − 15 BUS
⎛ ⎞ ⎛ 7(6)(3)(2)
⎞
. 585
8 − 945
8
= − 15
4 AUS + ⎛ ⎞
252 BUS
− 360
8
= − 15
4 AUS + ⎛ 252 ⎞ BUS
−45
= − 4 AUS
15
+ 252 BUS
= − 15
4 AUS − 5
28 BUS
Table 9.2 shows the results of the math. The hyperfine constants for the .7p 2 P3/2
state are kept as unknowns that we find while fitting the data. If there weren’t
any crossovers, we would be done. Our fitting function would be the sum of three
Lorentzian functions7 whose centers are taken from the above table:
C1 C2
g(f ) = ( )2 + ( )2
f −(fF =4→cog − 15 5
4 AUS − 28 BUS ) f −(fF =4→cog + 41 AUS − 28
13 B )
US
1+ 1+
γ12 γ22
. C3 (9.8)
+ ( )2 .
f −(fF =4→cog + 21 1
4 AUS + 4 BUS )
1+
γ32
where .fF =4→cog is the frequency of light needed to go from the .F = 4 ground
hyperfine level to the center of gravity of the .7p 2 P3/2 state. To avoid accidentally
mixing up the magnetic dipole hyperfine constant of the .7p 2 P3/2 state and the
amplitude of the Lorentzian functions, I changed the variable for amplitude to C.
7 There are many subtleties that we are glossing over here. Spectra sometimes have an offset, a
sloped offset, or a Gaussian pedestal for various reasons. When you fit data, sometimes you have
to add something to your fit function. However, you should always have a reason for why you are
adding something to your fit function.
9.5 Optional: Amplitudes 189
The same procedure can be done for the other two crossovers. In the end, the fit
function is a sum of 6 Lorentzian functions. From the fit, we find .AUS , .BUS , and
.fF =4→cog , all with uncertainty. You will get to practice this in a homework problem.
where the primed quantum numbers are for the excited state. That last symbol
is called the Wigner 6-j symbol (it is squared in the above equation). Also, the
second element of the second row is the number 1 (lots of folks accidentally
read that as the nuclear spin quantum number I ). In Mathematica, you would
type SixJSymbol{J,I,F},{Fp,1,Jp}.2 (Mathematica doesn’t allow J’ or F’ as variable
names, so I replaced them with Jp and Fp). There are also online calculators
you can find by searching the internet for “Wigner 6-j symbol calculator”. Notice
this formula has no units. We use this formula to take ratios to find relative
amplitudes.
190 9 Hyperfine Structure
1.0
0.8 F'=5
Spectrum 0.6 F'=4
0.4
0.2 F'=3
0.0
– 50 0 50 100
f (MHz) + 657,932,340 MHz
Fig. 9.5 Experimental spectroscopic results of the .6s 2 S1/2 F = 4 → 7p 2 P3/2 transition in
neutral cesium-133. This is an experimental result from my research group. A simplified Grotrian
diagram for the transition can be found in Fig. 5.12
Let’s find .Ir for all three real transitions in the cesium-133 example we have been
studying. To make things a bit easier, I included a copy of Fig. 9.4 on this page. As
a reminder, the ground state has quantum numbers .J = 1/2 and .F = 4, the excited
state has .J ' = 3/2 and .F ' = 3, 4, and 5, and the nuclear spin is .I = 7/2.
⎧1 7
⎫2
4
1. .F = 4 → F' = 3 : Ir = (2(4) + 1)(2(3) + 1) 2 2
3 = 7
16 = 0.4375
3 1 2
⎧1 7
⎫2
4
2. .F = 4 → F ' = 4 : Ir = (2(4) + 1)(2(4) + 1) 2 2
3 = 21
16 = 1.3125
4 1 2
⎧1 7
⎫2
4
3. .F = 4 → F ' = 5 : Ir = (2(4) + 1)(2(4) + 1) 2 2
3 = 11
4 = 2.75
5 1 2
From this math we see that the .F = 4 → F ' = 5 is the largest transition. Using
this largest transition as the reference, we can compare the size of the other two
transitions to it. The .F = 4 → F ' = 4 transition, which is the peak around 5 MHz
in Fig. 9.5, is . 21/16 '
11/4 = 44 = 0.47 times smaller than the .F = 4 → F = 5 transition.
21
you’ll see that these estimates are pretty close to what was measured experimentally.
While it is possible to calculate the amplitudes of the crossovers, it is much
more complicated. One reason is that the number of atoms with a particular
velocity, which creates the crossovers, depends on the temperature of your vapor
cell. However, we at least have a fairly straightforward way of finding the relative
amplitudes of the real transitions.
9.7 Example with Oxygen-17 191
.4.76 × 10
2 2 4
.1s 2s 2p
1
. D2 15,867.862 14
.1.01 × 10
2 2 4
.1s 2s 2p
1
. S0 33,792.583 15
.2.21 × 10
2 2 3
.1s 2s 2p 3s
5
. S2 73,768.200 15
.2.30 × 10
2 2 3
.1s 2s 2p 3s
3
. S1 76,794.978 15
For oxygen isotopes with no nuclear spin, such as oxygen-16, this table provides
all the information about their atomic states. There is a single ground state with
electronic configuration and term symbol .1s2 2s2 2p4 3 P2 . However, due to hyperfine
structure, the table does not tell the whole story for oxygen-17. Let’s examine the
ground state of oxygen-17. For this state, the total electronic angular momentum
quantum number is .J = 2 and the nuclear spin quantum number is .I = 5/2. The
possible values for the total atomic angular momentum quantum number are:
⎛ ⎞ | |
5 | 5 || 9 7 5 3 1
.F = 2+ , . . . , ||2 − = , , , , (9.11)
2 2| 2 2 2 2 2
We can also calculate the energy splitting using the hyperfine splitting formula. For
convenience, here is the hyperfine splitting formula:
3
K(K+1)−2I (I +1)J (J +1)
ΔE = 12 KA + 2 2I (2I −1)2J (2J −1) B
.
K = F (F + 1) − I (I + 1) − J (J + 1) (9.12)
A = 0 unless both I > 0 and J > 0
B = 0 unless both I > 1/2 and J > 1/2.
The first thing to notice is that for the .3 P2 ground state, both .I > 1/2 and .J > 1/2,
so we will need both the magnetic dipole term (the term with A) and the electric
192 9 Hyperfine Structure
quadrupole term (the term with B). For example, the hyperfine splitting for the
F = 9/2 hyperfine level is:
.
⎛ ⎞ ⎛ ⎞
K = F (F + 1) − I (I + 1) − J (J + 1) = 9
2
11
2 − 5
2
7
2 − 2(3) = 10
.
K = F (F + 1) − I (I + 1) − J (J + 1)
= I (I + 1) − I (I + 1) − 0(0 + 1)
. (9.14)
=0
→ ΔE = 0
Even though the nucleus has angular momentum, there is no energy shift when
J = 0. This hyperfine level has the same energy as the center of gravity. However,
.
we will still label this state with the F quantum number: .1s2 2s2 2p4 3 P0 F = 52 . You
will explore the hyperfine structure of oxygen-17 more in Problem 9.3.
Problems
9.1 Table 9.2 shows the energy shifts for the hyperfine levels with respect to the
center of gravity for the 7p 2 P3/2 state in cesium-133. All of the real transitions in
this problem are from the 6s 2 S1/2 F = 4 ground hyperfine level.
(a) In Sect. 9.4.1, we found ΔE for the F = 4 → F ' = 3 real transition. Confirm
ΔE for the other two real transitions.
(b) Find ΔE for the three V crossovers, similar to what we did in Sect. 9.4.2.
9.2 Below is table for four states in europium.8 The first row represents the ground
state, and the three subsequent rows are excited states. Europium has two stable
isotopes, europium-151 and europium-153.
Electron
configuration A151 (MHz) B151 (MHz) A153 (MHz) B153 (MHz)
4f7 6s2 8 S7/2 −20.0523±0.0002 −0.7012±0.0035 −8.8532±0.0002 −1.7852±0.0035
4f7 6s6p 8 P5/2 −157.01±0.03 74.5±0.4 −69.43±0.14 191±2.6
4f7 6s6p 8 P7/2 −218.66±0.04 −293.4±0.8 −97.15±0.13 −750±3
4f7 6s6p 8 P9/2 −228.84±0.02 226.9±0.5 −101.87±0.06 575.4±1.5
(a) Select either the 151 or 153 isotope and one of the three excited states from the
table above. Use the largest to smallest in integer steps rule to find the possible
F values for the ground state and your chosen excited state.
(b) Using the hyperfine splitting equation, calculate the energy splitting for the
ground state. Don’t worry about the uncertainties.
(c) Using the hyperfine splitting equation, calculate the energy splitting for the
excited state. Don’t worry about the uncertainties.
(d) Determine the transition frequency from a single ground hyperfine level (your
choice) to a single excited state hyperfine level (your choice). Report your
answer in MHz. Your final answer should include fcog in it.
Note: fcog is different for the two isotopes. This small difference is called an
isotope shift, which we will explore in Chap. 10.
(a) Find the possible F quantum numbers for all seven states. Hint: There are only
three calculations here!
(b) For the 1s2 2s2 2p4 3 P1 state, calculate the hyperfine energy shift ΔE for
each hyperfine level in terms of the hyperfine constants. Label your hyperfine
constants A3P 1 and B3P 1 .
(c) Suppose you were to excite an oxygen atom from the 1s2 2s2 2p4 3 P1 F = 7/2
state to the 1s2 2s2 2p3 3s 3 S1 F ' = 5/2 state. What is the transition frequency?
Your answer should look something like f = fcog + #A3P 1 + #B3P 1 + #A3S1 +
#B3S1 .
(d) Suppose you performed saturated absorption spectroscopy across all hyperfine
levels for the 1s2 2s2 2p4 3 P1 → 1s2 2s2 2p3 3s 3 S1 transition. Write the fit
function for this spectrum, assuming we have no crossovers.
(e) Challenge: Include crossovers!
(f) Optional: Find the relative intensities of all the real hyperfine transitions for the
spectrum in part (d). The relative intensities should be with respect to the largest
amplitude transition.
9.4 Consider the 3s 2 S1/2 → 3p 2 P◦3/2 transition in sodium-23. The little circle
on the excited state term symbol is the parity of the state, a topic we don’t cover
194 9 Hyperfine Structure
in this book. Parity is an advanced topic that is easy to say in words, but hard to
understand.9 Fig. 9.6 is a Grotrian diagram that shows the hyperfine structure of
both states. We will call the hyperfine constants for the ground state A2S and B2S
and the hyperfine constants for the excited state A2P and B2P .
References
1. Williams, W.D., Herd, M.T., Hawkins W.B.: Spectroscopic Study of the 7p1/2 and 7p3/2 States in
Cesium-133. Laser Phys. Lett. 15(9), 095702 (2018). https://doi.org/10.1088/1612-202X/aac97
2. Sandars P.G.H., Woodgate G.K.: Hyperfine structure in the ground state of the stable isotopes of
europium. Proc. R. Soc. Lond. A. 257, 269–276 (1960). http://doi.org/10.1098/rspa.1960.0149
3. Maruko, C., Cölmek, N., Herd, M.T., Ahrendsen, K., Cabrales, B., Cannon, G., Davis, E.,
Guo, X., Karani, T., Wallace, A., Wisnauckas, K., Williams, W.D.: Spectroscopic study of the
4f7 6s2 8 S◦7/2 − 4f7 (8 S◦ ) 6s6p(1 P◦ ) 8 P5/2,7/2 transitions in neutral europium-151 and europium-
9 You should, of course, do an internet search for “parity physics” if you’d like to learn more.
References 195
153: absolute frequency and hyperfine structure. J. Opt. Soc. Am. B. 41, 1217–1223 (2024).
https://doi.org/10.1364/JOSAB.521181
4. Herd, M.T., Maruko, C., Herzog, M.M., Brand, A., Cannon, G., Duah, B., Hollin, N., Karani,
T., Wallace, A., Whitmore, M., Williams, W.D.: Spectroscopic study of the 4f7 6s2 8 S◦7/2 −
4f7 (8 S◦ )6s6p(1 P ◦ )8 P9/2 transition in neutral europium-151 and europium-153: absolute fre-
quency and hyperfine structure. J. Opt. Soc. Am. B. 39, 2596–2602 (2022). https://doi.org/10.
1364/JOSAB.467968
Open Access This chapter is licensed under the terms of the Creative Commons Attribution 4.0
International License (http://creativecommons.org/licenses/by/4.0/), which permits use, sharing,
adaptation, distribution and reproduction in any medium or format, as long as you give appropriate
credit to the original author(s) and the source, provide a link to the Creative Commons license and
indicate if changes were made.
The images or other third party material in this chapter are included in the chapter’s Creative
Commons license, unless indicated otherwise in a credit line to the material. If material is not
included in the chapter’s Creative Commons license and your intended use is not permitted by
statutory regulation or exceeds the permitted use, you will need to obtain permission directly from
the copyright holder.
Isotope Shifts, Radioactive Decay,
and the Nuclear Forces 10
Abstract
In this chapter, we explore the nucleus, focusing on how the number of neutrons
in a nucleus influences transition frequencies and the stability of atoms. We
begin by examining the effect of neutrons on transition frequencies, then shift
into detailed discussions of the nuclear forces and principles governing atomic
stability and radioactive decay. Various modes of radioactive decay, including
neutron and proton emission, .α decay, spontaneous fission, .β − decay, .β + decay,
and electron capture, are investigated, with an emphasis on the conditions
under which each occurs. Additionally, we explore the nuclear shell model to
understand the energetics behind different types of decays and the stability of
isotopes.
Learning Goals
• that the weak nuclear force is the force that facilitates the annihilation and
creation of particles during radioactive decay.
• the nuclear shell model and the energy arguments behind radioactive decay.
Definitions:
• Mass number: The number of neutrons plus protons in an element,
represented by the variable A.
• Nucleon: A general word for either a neutron or a proton.
• Isotope shift: The change in the resonance frequency of a transition between
two states in an atom caused by changing the number of neutrons in the
nucleus.
The element name europium is just a placeholder for ‘the atom with 63 protons.’ To
display all the information at once, we often write an isotope as follows:
A
.Z XN (10.1)
Quick Quiz
Use Appendices A or B to find the proton number for the following atoms:
1 .A = Z + N → N = A − Z:
The number of neutrons in the nucleus slightly affects the energy of the states in
an atom. Neutrons affect the energy of the states in two ways: they change the
mass of the nucleus and slightly alter the charge distribution within the nucleus
(how the protons are distributed in the nucleus). For a given isotope, the energy
of each state in the atom will shift a different amount because each state has a
different electron configuration and term symbol (i.e. different angular momentum
and different distance from the nucleus). To study how changing the number of
neutrons affects the energy of a state, we typically select the same transition and
measure how the transition frequency changes as a function of neutron number.
Since the energy of each state in an atom shifts a different amount, the transition
frequency between those two states also changes. That shift is called the isotope
shift.
Table 10.1 shows experimental results from my research group for three tran-
sitions for two different isotopes of europium. The transitions are from the
7 2 8
.4f 6s S7/2 ground state to the .4f7 (8 S◦ )6s6p(1 P◦ ) 8 PJ states, where .J = 5/2,
.7/2, or .9/2. In the above electronic configurations, I did not write out the closed
subshells (e.g., .1s2 2s2 . . .). Notice that adding two neutrons to the nucleus decreased
the transition frequency for all three transitions. Because the two isotopes have a
different number of neutrons, all of the states in europium-151 have a different
energy than the states in europium-153. The difference in energy between the
ground states of the two isotopes is different from the difference in energy between,
say, the .J = 5/2 excited states. Therefore, the transition frequency from the ground
state to the .J = 5/2 excited state is slightly different for the two isotopes; this is
the isotope shift for this transition. Similarly, the difference in energy between the
.J = 7/2 excited states for the two isotopes is different from the difference in energy
between the .J = 5/2 excited states, so the isotope shift for this transition is different
from the isotope shift for the .J = 7/2 state. In the end, every transition in the atom
will have a different isotope shift.
Table 10.1 The isotope shifts for three transitions in the europium atom. All of the transitions
are from the atomic ground state. The labels for the excited state column are the total electronic
angular momentum quantum number J for the excited state: .4f7 (8 S◦ )6s6p(1 P◦ )8 PJ . .fr is the
frequency for the transition between the center of gravity of the ground state and the center of
gravity for an excited state. The numbers come from references [3] and [4]
Excited state (MHz) for 151
.fr 63 Eu88 (MHz) for 153
.fr 63 Eu90 (MHz) Isotope Shift (MHz)
.J = 9/2 652,389,757.16.±0.34 652,386,593.2.±0.5 3163.8.±0.6
.J = 7/2 647,708,930.6.±0.6 647,705,958.4.±2.6 2972.8.±0.5
.J = 5/2 642,894,493.3.±0.4 642,891,693.3.±0.9 2799.54.±0.20
200 10 Isotope Shifts, Radioactive Decay, and the Nuclear Forces
We accomplish this using Eq. 9.5. When you read studies of isotope
shifts from scientific papers, the numbers you are reading have already
accounted for hyperfine structure.
where A is the mass number for one isotope and .A' is the mass number of the other
isotope. Both the normal mass shift and the specific mass shift come about because
the two isotopes have different masses. The field shift comes about because the
nuclei for the two isotopes have a slightly different size. The different sizes change
the charge distribution (the distribution of protons) inside the nucleus and this leads
to a small shift in the energy levels.
The easiest way to think about normal mass shift is a classical analogy. The moon
orbiting the earth implies that the moon is orbiting about the center of the earth and
the earth is not orbiting around the moon. This is a bit of a misnomer. The moon and
the earth are actually orbiting around a point somewhere along a line between the
center of the earth and the center of the moon. Because the earth has so much more
mass than the moon, that point happens to be very close to the center of the earth.
If the moon and the earth had the same mass, they would both be orbiting about a
point directly between them. That point is called the center of mass. Analogously,
the electron is not orbiting around the center of the nucleus. Instead, the nucleus
and the electron are orbiting around the center of mass point. That point moves if
we change the mass of the nucleus. Accounting for this shift in the center of mass
point is the normal mass shift. The specific mass shift is present in atoms with more
than 1 electron. It comes about because the electrons are moving and interacting
with one another. Changing the mass of the nucleus has a small effect on those
interactions.
Out of the 3 contributions, only the normal mass shift has a simple formula:
' me A' − A
AA
δfNMS
. = fr , (10.3)
mp AA'
where .A' is the mass number of an isotope (for example the 151 in europium-151),
A is the mass number of a different isotope (for example the 153 in europium-153),
AA' '
.δf
NMS is the isotope shift of the isotope with mass number .A with respect to the
isotope with mass number A due to the normal mass shift, .me is the mass of an
electron, .mp is the mass of a proton, and .fr is the transition frequency from the
10.1 Isotope Shifts 201
center of gravity from the lower energy state to the center of gravity of the higher
energy state.2
Example
Rubidium, an element with 37 protons, has two naturally occurring isotopes:
rubidium-85 and rubidium-87.a There is a transition from the ground state to
an excited state with a center of gravity transition frequency of .377.107 THz.
What is the normal mass shift between the two isotopes?
For this example, let’s make .A' = 87 and .A = 85:
'
AA = me A' −A
δfNMS mp AA' fr
9.11×10−31 kg 87−85
. = 1.67×10−27 kg (85)(87)
(377.107 × 1012 Hz) (10.4)
= 55.6 MHz
Since we made .A' = 87 and .A = 85, a positive result means that rubidium-
87 would have a larger transition frequency, at least before we account for the
specific mass shift or the field shift.
a
On earth, about 72% of rubidium is rubidium-85 and about 28% is rubidium-87.
The experimental value of the isotope shift between rubidium-87 and rubidium-85
for this transition is .(77.583 ± 0.012) MHz, see reference [1]. We found the normal
mass contribution was 55.6 MHz. The remaining shift comes from the other two
contributions that are, unfortunately, difficult to calculate. However, we can try to
extract information about the nucleus by measuring a large number of isotopes for
a particular element.
To visually compare different isotopes, we can make a plot of the isotope shifts
for a particular transition as a function of neutron number. This plot is called a
King plot3 and is a nice way to visualize how changing the number of neutrons in
a nucleus affects the spectrum. Figure 10.1 shows an example of experimentally
measured isotope shifts for a transition in the krypton atom that is excited using a
laser near 760 nm. It was collected by Keim et al. in 1995, see reference [2]. All of
2 If you are paying close attention, you might ask the question, “Does .fr refer to the transition
frequency for the isotope with mass number A or .A' ?” The answer is neither. .fr is a calculated
transition frequency assuming a nucleus with infinite mass. However, in practice, you can use .fr
for either isotope or the calculated value for a nucleus with infinite mass. We can do this because
isotope shifts are usually on the order of MHz to a few GHz, which is typically more than 100,000
times smaller than .fr . The formula to calculate the transition frequency assuming a nucleus with
+M
infinite mass is . meM fr , where M is the mass of a nucleus for a particle isotope and .fr is the
transition frequency for that isotope. That fraction is very close to 1.
3 Named after the British physicist William H. King.
202 10 Isotope Shifts, Radioactive Decay, and the Nuclear Forces
0
– 100
f 86,A' (MHz)
– 200
– 300
– 400
– 500
75 80 85 90 95
A'
Fig. 10.1 The isotope shift of a transition in krypton near 760 nm using data from reference [2].
The isotope shifts are measured with respect to the isotope with mass number .A = 86
the isotope shifts are measured with respect to krypton-86. Notice there is a “kink”
in the graph at .A' = 86. As physicists, we want to know why! If we truly understand
the system, we should be able to calculate the isotope shifts using Eq. 10.2 and
predict such a graph. The reason for the “kink” at .A' = 86 is that nucleons fill shells
in the nucleus just like electrons fill shells. We discuss nuclear shells in Sect. 10.6.
There is a nuclear shell for neutrons that fills at 50 neutrons. Since krypton has 36
protons, this neutron shell fills at .A' = 86. The “kink” at .A' = 75 is not from a
shell filling. That “kink” is thought to come from the nucleus itself beginning to
deform due to lack of neutrons. There is so much one could ask and explore with
this data. What questions would you ask? The wonderful thing about physics is that
if we understand the system, we can interpret the data. If there is something we don’t
understand, this sort of data gives us a starting point to explore and learn more.
Inside the nucleus are neutrons and protons. Experimental evidence shows that if
the neutron-to-proton ratio isn’t “correct,” the nucleus becomes unstable and will
undergo radioactive decay. Radioactive decay is often written in the form of an
equation. The left hand side of the equation is the unstable isotope. We will call this
the parent system. The right hand side shows all of the decay particles. We will call
this the daughter system.4 The equation looks something like this:
4 Historically, parent and daughter are the names used to describe radioactive decay. The terminol-
ogy also shows up in biology during cell division (parent cell and daughter cells). Sometimes the
daughter system is called the decay system.
10.2 Radioactive Decay 203
parent → daughter
. (10.5)
Neutron Emission The nucleus will eject a neutron. An example of an isotope that
undergoes neutron emission is helium-5 (2 protons and 3 neutrons), which decays
to helium-4 (2 protons and 2 neutrons) and a neutron:
5
.2 He3 → 42 He2 + n
In the above equation, helium-5 is called the parent nucleus and helium-4 is called
the daughter nucleus. The general formula for neutron emission is:
A
.Z XN → A
Z XN −1 + n (10.6)
Proton Emission The nucleus will eject a proton. An example of an isotope that
undergoes proton emission is scandium-39 (21 protons and 18 neutrons), which
decays to calcium-38 (20 protons and 18 neutrons) and a proton:
39
21 Sc18
. → 38
20 Ca18 + p
In the above equation, scandium-39 is called the parent nucleus and calcium-38 is
called the daughter nucleus. The general formula for proton emission is:
A
.Z XN → Z−1
A
XN + p (10.7)
.β − Decay (Beta Minus Decay) A neutron will transform into a proton, an electron,
and a particle called an “anti-electron neutrino” (which has the label .ν̄e ). We haven’t
talked about neutrinos, but you can find more information about them with a quick
internet search. For our purposes, neutrinos are very light particles with no electric
charge. Both the electron and the anti-electron neutrino are ejected while the proton
remains in the nucleus. The electron that is ejected from the nucleus usually has a
ton of energy and is called a .β − particle. An example of an isotope that undergoes
.β
− decay is carbon-14:
.
14
6 C8 → 147 N7 + β − + ν̄e
In the above equation, carbon-14 (6 protons and 8 neutrons) is called the parent
nucleus and nitrogen-14 (7 protons and 7 neutrons) is called the daughter nucleus.
The general formula for .β − decay is:
A
.Z XN → Z+1
A
XN −1 + β − + ν̄e (10.8)
204 10 Isotope Shifts, Radioactive Decay, and the Nuclear Forces
β + Decay (Beta Plus Decay or Positron Decay) A proton will transform into a
.
neutron, a positron,5 and an electron neutrino (which has the label .νe ). Both the
positron and the electron neutrino are ejected while the neutron remains in the
nucleus. Like the electron in .β − decay, the positron has a ton of energy and is called
a .β + particle. An example of an isotope that undergoes .β + decay is potassium-40,
which is found in bananas and decays to argon-40:
+
40
.19 K21 → 40
18 Ar22 + β + νe
Electron Capture The nucleus will steal an electron from the atom. That electron
and a proton in the nucleus transform into a neutron and an electron neutrino. The
electron neutrino is ejected while the neutron stays in the nucleus. An example of an
isotope that undergoes electron capture is beryllium-7, which decays to lithium-7:
7
.4 Be3 + e → 73 Li4 + νe
α Decay (Alpha Decay) Heavy nuclei can emit a cluster of two protons and two
.
Spontaneous Fission While most heavy nuclei undergo .α decay, superheavy atoms
can emit nuclei of other elements. For example, californium-252 (98 protons
and 154 neutrons) can break apart into xenon-140 (54 protons and 86 neutrons),
5A positron is the antimatter partner to the electron. Antimatter is a topic in Chap. 11.
10.2 Radioactive Decay 205
ruthenium-108 (44 protons and 64 neutrons), and 4 neutrons (notice that .140 +
108 + 4 = 252). This is called spontaneous fission:
252
.98 Cf154 → 140
54 Xe86 + 44 Ru64 + 4n
108
In the above equation, both xenon-140 and ruthenium-108 are considered daughters.
If the balance between neutrons and protons deviates too far from a stable
equilibrium, the nucleus will undergo radioactive decay. Some isotopes can undergo
two or more forms of decay. For example, francium-220 undergoes .α decay 99.65%
of the time and .β − decay for the remaining 0.35%. Radium-226 almost always
undergoes .α decay, but .3.2 × 10−9 % of the time it will undergo spontaneous fission
by emitting a carbon-14 nucleus. Exploring the physics behind the balance between
neutrons and protons is the topic of Sect. 10.6.
Will’s Rant
It is important to understand that during .β − decay, .β + decay, and electron
capture, particles cease to exist and new particles come into existence. Some
books imply, for example, that a neutron is composed of a proton, an electron,
and an anti-electron neutrino, as if you could look inside the neutron and
find those three particles squished together. This is incorrect and drives me
a little crazy, hence the rant ☺. The neutron is actually composed of three
smaller particles called quarks, a topic in Chap. 11. What is important here is
that during .β − decay, the neutron ceases to exist. The neutron is present one
moment and gone the next. The proton, electron, and anti-electron neutrino
did not exist before the decay; these particles are created during the decay. In
contrast, no new particles are created or destroyed in neutron emission, proton
emission, or .α decay.
There is a general rule about whether or not an isotope will undergo radioactive
decay. In fact, this rule is more general than radioactive decay. As a general rule of
nature, things try to get to the place of lowest energy.
As an analogy, think about a ball that is sitting halfway up a bowl. When released,
the ball will try to get to the lowest point in the bowl and, with friction present, will
eventually settle at this lowest point. The reason the ball settled at the lowest point
is because this is the place of lowest energy.
Consider the radioactive decay of carbon-14:
206 10 Isotope Shifts, Radioactive Decay, and the Nuclear Forces
14
.6 C8 → 147 N7 + β − + ν̄e
The daughter nucleus, nitrogen-14, has lower energy than the parent, carbon-14.
The decay happens because carbon-14 is a higher energy state than nitrogen-14,
see Fig. 10.2. The higher energy state has a lifetime of about 5700 years. After that
characteristic time, it will decay to nitrogen-14. We will discuss the mechanisms
behind radioactive decay in Sects. 10.4, 10.5 and 10.6.
All unstable isotopes have a characteristic time for decay, known as the half-life,
represented by the parameter .t1/2 . Radioactive decay is a random process, but we
can state probabilities for whether an isotope has decayed. Half-life is defined as the
time it takes for there to be a 50% probability that the isotope has decayed. If the
radioactive sample is large enough, the half-life can also be thought of as the time it
takes for half of the parent nuclei to decay to the daughter nucleus. The number of
remaining parents as a function of time is given by the formula:
⎛1⎞ t
t1/2
N (t) = N0
. , (10.12)
2
where .N0 is the number of radioactive atoms we start with, t is the time elapsed,
and .t1/2 is the half-life. For example, rubidium-84 has a half-life of .t1/2 =
32.82 days. If we had a sample of 1,000,000,000 rubidium-84 atoms, half would
decay after 32.82 days leaving about 500,000,000 rubidium-84 atoms. Of those
remaining atoms 500,000,000 rubidium-84 atoms, half of those will decay in the
next 32.82 days. This continues until there are no radioactive atoms remaining, see
Table 10.2.
Extra Fun
Carbon-14 is a radioactive isotope of carbon with a half-life of about
5700 years. It is created in the atmosphere when high energy neutrons from
space, also known as cosmic rays, slam into the stable isotope nitrogen-14.
The high energy neutron basically crashes into a proton in the nitrogen-14
nucleus, pushing the proton out while remaining behind. This radioactive
carbon-14 attaches to an oxygen molecule to form radioactive carbon diox-
ide. Since there is a source (cosmic rays hitting nitrogen-14) and a sink
(radioactive decay), the atmosphere reaches an equilibrium between the
carbon-12 (the stable isotope) and carbon-14 forms of carbon dioxide. This
equilibrium results in a constant isotopic abundance. If you collected a bunch
of carbon dioxide from the air, about 1 molecule in every .1012 is radioactive.
Living organisms in contact with the atmosphere breathe in, absorb, or
ingest this radioactive carbon dioxide, thereby reaching equilibrium with the
atmosphere. If the organism ceases interaction with the atmosphere,6 the
isotopic abundance decreases (the radioactive carbon decays without being
replaced). If sometime later you dig up the formally living being and you
know the starting isotopic abundance, you can quickly calculate how long
that being has been removed from the atmosphere. This process is known
as radiocarbon dating. There are other types of radioactive dating including
radioargon dating and radiokrypton dating.
6
RIP ☹.
The table of isotopes, also known as the table of nuclides, is similar to the periodic
table, but includes all known elements and isotopes. Figure 10.3 is a screenshot of
the table of isotopes from a website maintained by the International Atomic Energy
Agency (IAEA).7 Figure 10.4 is a zoom in for the lighter elements. The vertical axis
represents the number of protons in the isotope, while the horizontal axis represents
the number of neutrons in the isotope. The entire row with 2 protons consists of
helium isotopes; the entire row with 63 protons are europium isotopes. The black
7I consider the Live Chart of Nuclides website one of three essential tools for an atomic and
nuclear physicist: https://www-nds.iaea.org/relnsd/vcharthtml/VChartHTML.html. The other two
are the IAEA app for your phone, known as ‘Isotope Browser’, and the NIST spectral database,
https://www.nist.gov/pml/atomic-spectra-database.
208 10 Isotope Shifts, Radioactive Decay, and the Nuclear Forces
boxes indicate stable isotopes, and the colors representing the various types of
radioactive decay are shown as insets in both Figs. 10.3 and 10.4. The interactive
table is a lot of fun to play with.
Stable
or electron capture
Proton emission
Neutron emission
Spontaneous fission
Unknown
Fig. 10.3 A screenshot of the entire table of isotopes from the Live Chart of Nuclides maintained
by the IAEA: https://www-nds.iaea.org/relnsd/vcharthtml/VChartHTML.html Note: I added the
axes and the legend. Some isotopes have more than one type of decay. The legend shows the
dominate decay
Stable
or electron capture
10He
Proton emission
6H
Neutron emission
Spontaneous fission
Unknown
Fig. 10.4 A zoom in screenshot of the lighter elements from the table of isotopes. Note: I added
the axes and the legend
10.4 The Strong Nuclear Force 209
Important Reminder
When two particles are interacting, there is a force between them. For
example, an electron and a proton are attracted to each other because they have
opposite charges. That attractive force is called the Coulomb force, which is
also known as the electrostatic force.
annihilation and creation of particles, the weak nuclear force is involved. Finally,
we will tie everything together in Sect. 10.6 using energy principles to explore why
some isotopes are stable and some are not.
The helium-4 nucleus has 2 protons and 2 neutrons. Since both protons have
positive charge, they repel each other due to the Coulomb force. The protons in
helium-4 are separated by about .1 × 10−15 m. This incredibly small distance is
called a femtometer, which is given the unit fm: .1 × 10−15 m = 1 fm. The repulsive
Coulomb force between the two protons has a magnitude of about 230 newtons
(about 50 pounds of force). Considering each proton has a mass of only .1.67 ×
10−27 kg, this is huge!! If we took a free proton and acted on it with 230 Newtons
of force, the proton would accelerate at .1 × 1029 m/s2 !!!! For comparison, freefall
on earth is only about .9.8 m/s2 . The relative comparison is so enormous, I have
absolutely no idea how to convey how large that repulsive force is.
However, the protons inside a helium-4 nucleus do not fly apart.8 Therefore,
there must be some other attractive force that is larger than the repulsive Coulomb
force to stop the protons from being pushed out of the nucleus. An intuitive guess
is that maybe gravity is keeping the protons in the nucleus. After all, gravity is the
attractive force that we interact with everyday. Remarkably, gravity is an extremely
weak force. You will explore just how weak gravity is in Problem 10.6. In addition,
neutrons have no charge, so they don’t feel any Coulomb force. What is keeping
neutrons in the nucleus?
1
Force (104 N)
Distance (fm)
1 2 3 Strong nuclear force
Coulomb force x 10
–1
–2
Repulsive Attractive
Region Region
–3
Fig. 10.5 An illustrative plot showing the forces between nucleons inside the nucleus. The strong
nuclear force is attractive (negative force) above about 0.85 fm but falls off very quickly. The
Coulomb force between two protons, which is increased by a factor of 10 for visibility, is always
repulsive (positive force)
The force keeping the nucleons “glued” to one another is the strong nuclear
force.9 The strong nuclear force acts on both neutrons and protons, and it can be
either an attractive or repulsive force. At a distance of 1 femtometer, the strong
nuclear force is attractive and approximately 100 times stronger than the repulsive
Coulomb force. Figure 10.5 shows a graph of both the Coulomb force and the strong
nuclear force.10 At short distances, both the strong nuclear force and the Coulomb
force are repulsive. Somewhere around 0.85 fm, the nucleons start to be attracted
by the strong nuclear force. The really important thing to take from this graph is
that at around 1 fm, the strong nuclear force is more attractive than the Coulomb
force is repulsive. The total force is the sum of the two, which for small distances,
is attractive. Outside of about 5 fm, the strong nuclear force is basically 0 and the
Coulomb force will dominate, pushing the protons away from each other.
Figure 10.5 is also useful for understanding why we cannot have a stable atom
with more than about 200 nucleons. A proton near the edge of a large nucleus is far
away from the nucleons that are on the other side of the nucleus. Because the strong
nuclear force goes to zero so quickly, that proton does not feel any attractive strong
nuclear force from the distant nucleons. It does, however, still feel the repulsive
9 In Chap. 11, we are going to talk about quarks, which are the subparticles that make up protons
and neutrons. The strong nuclear force is also what keeps the quarks glued to one another inside
the nucleon. However, the strong nuclear force extends outside of a nucleon to interact with other
nucleons. The force that extends out of the nucleon is often called the “strong residual force”.
Technically, it is still the same strong nuclear force, but there are subtle and important differences
for how it behaves inside a nucleon and how it behaves between nucleons. Figure 10.5 is an
illustrative plot for the strong residual force between nucleons.
10 This strong nuclear force plot is calculated from the Reid potential, see Reference [3], which
was developed and named after American physicist Roderick V Reid Jr. This is a popular model
for the strong nuclear force first used in 1968.
10.4 The Strong Nuclear Force 211
Coulomb force from all the other protons. Consequently, this proton near the edge
will be repelled away. This is why the periodic table does not have any stable
elements above lead-208. The next atom on the periodic table is bismuth. Bismuth
has no stable isotopes, although bismuth-209 is super long-lived with a half-life of
.2 × 10
19 years.
An Analogy
The Coulomb force is a long-range force. Two protons will weakly repel each
other even if they are 10 meters apart. The strong nuclear force is similar to
a contact force such as two pieces of Velcro. When the two pieces of Velcro
are touching, they stick together really well. The moment they are no longer
touching, the force between the two is zero.
So, why do we need neutrons for a stable nucleus? Neutrons add additional attractive
strong nuclear force to the nucleus without adding any additional repulsive Coulomb
forces. A nucleus with only 2 protons, which would be helium-2, is extremely
unstable. The Coulomb force, while small compared to the strong nuclear force, is
still large enough to push the two protons apart. So, the decay of helium 2 would be
two protons flying away in different directions. Adding a neutron to the nucleus adds
additional strong nuclear force (attractive) without adding any additional repulsive
Coulomb forces. Both helium-3 (2 protons and 1 neutron) and helium-4 (2 protons
and 2 neutrons) are stable. Helium-5 (2 protons and 3 neutrons) is unstable. But why
is helium-5 unstable? An extra neutron should just add additional attractive strong
nuclear force with no additional repulsive Coulomb force. In the end, it comes down
to the energy analogy: there is a daughter system with lower energy than helium-5.
We will explore this more in Sect. 10.6.
Fun Fact
If the strong nuclear force were only 2% larger, helium-2 could undergo .β −
decay to hydrogen-2 (also known as deuterium or heavy hydrogen) instead of
breaking apart into 2 protons.11
11
See reference [4].
The strong nuclear force holds nucleons in the nucleus. If the balance between pro-
tons and neutrons is not correct, the nucleus will be unstable and undergo radioactive
decay (more on this in Sect. 10.6). We have hinted at possible mechanisms behind
proton emission and .α decay (the Coulomb force overwhelms the strong nuclear
force), but nothing we have talked about so far explains the forces behind the
annihilation or creation of particles (i.e., .β − , .β + , and electron capture; see the
beginning of Sect. 10.2 for a recap of the types of radioactive decays). To explore
those decays, we need to understand the weak nuclear force.
Important
If the radioactive decay process involves the annihilation and creation of
particles, the weak nuclear force is involved.
Let’s start by reviewing a few important concepts. The Coulomb force acts only
on charged particles. It is a long-range force and can be either an attractive force
(opposite charges) or a repulsive force (same charges). The strong nuclear force
acts on protons and neutrons but not on electrons. It is a short-range force with a
range on the order of a few femtometers and is, for our purposes, an attractive force.
The weak nuclear force, sometimes referred to simply as “the weak force,”
acts on electrons, protons, and neutrons. Like the Coulomb force, it can be either
attractive or repulsive. However, unlike the Coulomb force, it is an extremely short-
range force. While the strong nuclear force has a range on the order of a few
femtometers, the weak nuclear force has a range on the order of .1 × 10−18 m, or
about 0.1% of the diameter of a proton. This distance is called an attometer (am):
.1 × 10
−18 m = 1 am.
As you can probably guess from the name, the weak nuclear force is weak
compared to the strong nuclear force and the Coulomb force. Consider two protons
about 1 fm apart. The weak nuclear force is about 1 million times weaker than the
strong nuclear force. For stable atoms, the strong nuclear force dominates the weak
10.5 The Weak Nuclear Force 213
nuclear force, but for atoms with a few too many neutrons or protons, the weak
nuclear force can play an important role.
To explore the importance of the weak nuclear force, let’s discuss the neutron. A
neutron outside of a nucleus, called a free neutron, will undergo radioactive decay
with a half-life of about 10.2 minutes according to the equation:
.n → p + β − + ν̄e , (10.13)
where n is the neutron that decays into a proton p, a .β − particle (a high energy
electron), and an anti-electron neutrino .ν̄e .
The Analogy Think about a ball that is sitting half way up a bowl. When
released, the ball will try to get to the lowest point in the bowl and, with
friction present, will eventually settle at this lowest point. The reason the ball
settled at the lowest point is because this is the place of lowest energy. Both
gravity and friction were the mechanisms (forces) that got the ball to settle at
the bottom of the bowl. Gravity pulled the ball down, and friction dissipated
the extra energy into heat. The ball went from a higher energy state to a lower
energy state.
The proton, .β − particle, and anti-electron neutrino (i.e., the daughter system) have
lower energy than the free neutron (the parent system). Something had to facilitate
this system going from higher energy to lower energy. The force behind this
interaction is the weak nuclear force. So, according to the general rule of nature,
a free neutron “wants” to get to a place of lower energy, and the weak nuclear force
is the force that helps it get there.
A free neutron will undergo radioactive decay facilitated by the weak nuclear
force. For a neutron inside a nucleus, there is also the strong nuclear force from
other nucleons interacting with the neutron. The question to ask is, “Is there a place
of lower energy for the system to go?” If the answer is yes, the weak nuclear force
will eventually get the system to a lower energy. If the answer is no, then the weak
force has no place to push the system. So, the weak nuclear force will cause the free
neutron to cease existing and create a proton, .β − particle, and anti-electron neutrino.
In contrast, there is no place of lower energy for, for example, a helium-4 nucleus.
So, the helium-4 nucleus is stable.
If a neutron in a helium-4 nucleus were to undergo .β − decay, the equation would
be:
214 10 Isotope Shifts, Radioactive Decay, and the Nuclear Forces
4 He → 43 Li1 + β − + ν̄e
2
.2
This does not happen!
The system on the right of this equation (the lithium-4 atom, the .β − particle, and
the anti-electron neutrino) has a higher energy than the helium-4 atom. Therefore,
this process does not occur.
Current Summary
As a general rule of thumb:
• If there are way too many neutrons in the nucleus, a neutron will escape
(more on this in Sect. 10.6) to get the system to a point of lower energy.
• If there are a few too many neutrons in the nucleus, the weak nuclear force
will facilitate .β − decay to get the system to a point of lower energy.
• If there are way too many protons in the nucleus, the Coulomb force will
push a proton out of the nucleus to get the system to a point of lower energy.
A free proton is stable9 because a free proton is already the system of lowest energy.
However, a proton can decay inside the nucleus. How?!? The answer always goes
back to the energy argument: a system wants to get to the place of lowest energy. If
a nucleus has a few too many protons, the weak nuclear force will find a way to that
place of lower energy. In this case, a proton can transform into a neutron, a positron
(more on positrons in Chap. 11), and an electron neutrino. This is called .β + decay.
As an example, consider the following decay for oxygen-15:
15
.8 O7 →157 N8 + β + + νe (10.14)
This decay, which has a half-life of about 2 minutes, is used in Positron Emission
Tomography (PET) scans to measure blood flow and oxygen metabolism.10 A
proton in the oxygen-15 nucleus transforms into a neutron, a .β + particle (positron),
and an electron neutrino. The right-hand side of that equation is lower energy than
the left-hand side, so the weak nuclear force will make this decay happen.
The daughter particle, nitrogen-15, is the lowest energy state for this system, so
it does not undergo any further type of decay. Therefore, neither of the following
occurs:
9 Asfar as we know.
10 Due to its short half-life, it is made in an accelerator on site (usually in the basement of the
hospital) and brought immediately to the patient.
10.5 The Weak Nuclear Force 215
→ 15 +
6 C9 + β + νe
15 N
7 8
.
15 N
7 8→ 8 O7 + β − + ν̄e
15
Both of these processes have daughter systems with higher energy than the parent
system, so they will never occur.
There is one more process in which the weak nuclear force plays an important
role. For some atoms, the electrons “orbiting” the nucleus can come too close. If
that happens and there is a daughter system with a lower energy, that electron and
a proton in the nucleus will transform into a neutron and an electron neutrino. This
is called electron capture. The weak nuclear force, once again, is the force that is
facilitating this process. I want to emphasize that the proton and electron do not
combine to form a neutron. The proton and electron literally stop existing, and a
neutron and an electron neutrino start existing (see Will’s rant in Sect. 10.2). An
atom that undergoes electron capture decay is the beryllium-7 atom, which has a
half-life of about 53 days. The equation to describe beryllium-7 decay is:
7
.4 Be3 + e → 73 Li4 + ve , (10.15)
where .73 Li4 is the daughter lithium-7 atom. Once again, this process can only occur
because the daughter system has lower energy than the parent system.
Current Summary
As a general rule of thumb:
• If there are way too many neutrons in the nucleus, a neutron will escape
(more on this in Sect. 10.6) to get the system to a point of lower energy.
• If there are a few too many neutrons in the nucleus, the weak nuclear force
will facilitate .β − decay to get the system to a point of lower energy.
• If there are a few too many protons in the nucleus, the weak nuclear force
will facilitate either .β + decay or electron capture to get the system to a
point of lower energy.
• If there are way too many protons in the nucleus, the Coulomb force will
push a proton out of the nucleus to get the system to a point of lower energy.
216 10 Isotope Shifts, Radioactive Decay, and the Nuclear Forces
When we talked about the strong nuclear force, we had an unanswered question:
“Why is a nucleus with too many neutrons unstable?” After all, neutrons only add
an attractive force to the system. There was an explanation about the existence of
a lower energy state, but that explanation, at least to me, felt unsupported. In this
section, we will explore what happens when more and more nucleons are added
to a nucleus. We will also see why, from an energy viewpoint, different types of
decays occur and why stable lighter atoms have about equal numbers of protons
and neutrons while stable heavier atoms have more neutrons than protons, see the
black boxes in Fig. 10.3. To do so, we will rely on the general rule of nature and the
Pauli Exclusion Principle, see Chap. 8, Sect. 8.5 for a review of the Pauli Exclusion
Principle.
Important Reminder
In Chap. 8, we learned that electrons fill shells. The .1s subshell can hold two
electrons. Using the ket notation .|n 𝓁 m𝓁 s ms 〉, the two electrons have quantum
numbers .|1 0 0 12 21 〉 and .|1 0 0 12 - 12 〉. Notice they have different quantum numbers
as required by the Pauli Exclusion Principle. All six electrons in a filled .2p subshell
also have different quantum numbers. The .1s electrons are in a lower energy state
than the .2p electrons, but the Pauli Exclusion Principle prevents any more electrons
from being in the .1s subshell. The dominant interaction between the electrons
and the nucleus is the electromagnetic force, so this force determines how many
Table 10.3 Electron Shell Subshell Subshell max electrons Shell max electrons
configuration for shells and
1 1s 2 2
subshells
2 2s 2 8
2p 6
3 3s 2 18
3p 6
3d 10
4 4s 2 32
4p 6
4d 10
4f 14
10.6 The Nuclear Shell Model and Energetics 217
subshells there are for a particular shell. Table 10.3 shows the maximum number
of electrons that can fit into each subshell and shell. I want to emphasize that every
electron has a unique set of quantum numbers. Neon has a total of ten electrons to
completely fill the .n = 1 and .n = 2 shells. Each of those ten electrons has a unique
set of quantum numbers. As a reminder, electrons do not have to completely fill a
shell before starting to fill the next shell (see Fig. 8.3). For example, potassium has a
single electron in the 4s subshell with no electrons in the 3d subshell. This ordering
is determined by the electromagnetic force.
Protons and neutrons are also fermions with a spin quantum number of .1/2.
Therefore, they also fill nuclear shells and subshells. Since protons and neutrons are
different particles, they fill individual nuclear shells; protons fill the proton shells
and neutrons fill the neutron shells. Isotopes with filled nuclear shells and subshells
tend to be more stable. This stability occurs at so-called “magic numbers.” The
magic numbers for protons indicate how many protons are in the nucleus when this
extra stability occurs. The magic numbers for protons are 2, 8, 20, 28, 50, 82, and
114. The magic numbers for neutrons are 2, 8, 20, 28, 50, 82, and 126. Oxygen-16
(8 protons and 8 neutrons) is called doubly magic because it has both a filled proton
shell and a filled neutron shell. Oxygen-16 is an extra stable nucleus.
The configuration of the nuclear shells is very different from the electron shells
because the nuclear shells are determined by both the strong nuclear force and
the electromagnetic force. However, every proton (or neutron) in the nucleus has
a unique set of quantum numbers and fills the proton (or neutron) shells. The last
magic number for protons (114) is smaller than the last magic number for neutrons
(126). This is because protons have an additional Coulomb force acting on them, and
that additional force causes a different structure for the shells. I want to emphasize
that these are not electron shells! Protons and neutrons do not fill up their nuclear
shells according to the Madelung rule (see Fig. 8.3).
Let’s use this as a starting point to understand nuclear stability. Figure 10.6
shows a planetary model for protons and neutrons filling their respective shells. As
a reminder, the planetary model is not correct. Protons and neutrons are quantum
mechanical particles that have wave-like properties. Nonetheless, this nuclear shell
model can help us understand the stability of the nucleus.
The number of nucleons that can occupy a subshell is indicated on the left. The
lowest energy proton shell can hold two protons (one spin up and one spin down)
and the lowest energy neutron shell can hold two neutrons. This is the first magic
number, 2. The next shell can hold a total of six nucleons distributed among two
subshells. The sum of all protons (or neutrons for the neutron shells) that completely
fill the first two shells is the second magic number, .2 + 6 = 8. The third grouping of
subshells holds a total of twelve nucleons, a magic number of .2 + 6 + 12 = 20, and
so on. Notice that as the energy increases, the neutron shells are lower in energy than
the proton shells. This is because the Coulomb force acts on the protons, but not on
the neutrons, and also because neutrons have a slightly larger mass than protons. It
should be noted that the energy spacings are not correct; they are greatly simplified
for us to explore concepts. There are also higher energy shells that are not shown in
these diagrams.
218 10 Isotope Shifts, Radioactive Decay, and the Nuclear Forces
Fig. 10.6 The nuclear shells for protons and neutrons. Starting from the left, helium-4 is a stable
isotope. The system is in its lowest energy state. The middle diagram is an example of .β− decay.
The right diagram is an example of proton emission
Starting from the left diagram in Fig. 10.6, we see that helium-4 is a stable
isotope. In fact, this is a super-stable, doubly magic nucleus because it perfectly
fills the lowest nuclear shell for both the protons and the neutrons. There isn’t any
place of lower energy for any of the nucleons. The helium-4 nucleus is also the .α
particle from .α decay. The middle diagram is an example of .β − decay. The system
has a place of lower energy if the highest-energy neutron transforms into a proton
(plus a .β − particle and anti-electron neutrino). One could ask the question, “Why is
this .β − decay and not neutron emission?” The answer always come back to energy.
For this scenario, it is energetically more favorable for the weak nuclear force to
facilitate .β − decay than the neutron to just leave the nucleus. The right diagram
is an example of proton emission. The highest-energy proton has a place to go to
bring the system to a point of lower energy. It can leave the nucleus or undergo
.β
+ decay. In this scenario, the Coulomb force makes proton emission energetically
more favorable than .β + decay, so that is what happens. Finally, the left diagram
in Fig. 10.7 shows an example of .β + decay. In this scenario, it is energetically
more favorable for the weak nuclear force to facilitate the decay of scandium-41
to calcium-41. And this is the essence of stable versus unstable nuclei. If there is
a state of lower energy, the system will try to get there. Sometimes the mechanism
is proton emission, neutron emission, .α decay, or spontaneous fission. Other times,
the weak nuclear force facilitates the annihilation and creation of particles. In the
end, if there is a state of lower energy, the nucleus will decay to this lower state.
There is another interesting concept I’d like to explore with you. For smaller
mass, stable nuclei, there are roughly equal numbers of neutrons and protons. This
can be seen by the black boxes (stable nuclei) in Fig. 10.3, which make a linear line
with a slope of one for smaller mass nuclei. This is because, for smaller masses,
the nuclear energy levels for the protons and neutrons are about the same. This
10.6 The Nuclear Shell Model and Energetics 219
Fig. 10.7 Starting from the left, scandium-41 is an example of .β + decay. The middle diagram and
right diagram are examples of stable nuclei. Oxygen-16 is stable and has 8 protons and 8 neutrons.
Germanium-72 is also stable, but it needs 40 neutrons to balance the energy of the 32 protons. The
three dots on the subshell that can hold 8 nucleons are indicating the shell is filled; I couldn’t fit 8
nucleons on the line
is shown in the left diagram in Fig. 10.6 for helium-4 and the middle diagram in
Fig. 10.7 for oxygen-16. However, as more protons are packed into the nucleus,
the Coulomb force becomes larger and larger, increasing the nuclear energy levels
for the protons. As such, more neutrons are needed to balance the energy of the
protons and neutrons. This is an equivalent way of saying more neutrons are needed
to provide a larger strong nuclear force to overcome this increase in the Coulomb
force repulsion. An example of a heavier nucleus is shown in the right diagram
of Fig. 10.7. Germanium-72 needs eight more neutrons than protons to create a
stable balance between the total proton energy and the total neutron energy. In fact,
germanium-70, -72, -73, and -74 are all stable.11 Notice how the diagrams for all
the stable nuclei have approximately equal energy for the highest-energy neutrons
and protons while the unstable nuclei do not.
If the nucleus gets too big, there is always a state of lower energy. The largest
mass nucleus that is stable is lead-208. Lead-208 has a proton number of 82 and
a neutron number of 126, which are both magic numbers. There are also isotopes
that have two or more decay paths to states of lower energy. This is more likely for
larger mass nuclei. For example, bismuth-212 can undergo both .α decay (36% of
the time) and .β − decay (64% of the time). For the heaviest of atoms, the dominant
decay is not the annihilation and creation of particles, but particles simply leaving
the nucleus to reach a lower energy state, see Fig. 10.3 or the online table of isotopes.
The nucleus is an amazing and wonderful system to study. It is such a rich
system to explore and is needed to understand the world of the super small. Atomic
11 Germanium-71 decays via electron capture, which is facilitated by the weak nuclear force.
220 10 Isotope Shifts, Radioactive Decay, and the Nuclear Forces
physicists tend to use lasers to excite electrons to higher energy atomic states. We
care about the nucleus, but usually in the context of how the structure of the nucleus
affects electronic energy levels, for example, isotope shifts discussed in Sect. 10.1.
Nuclear physicists conduct similar experiments on nuclei with the goal of
understanding the internal forces at play. Just like electrons, protons and neutrons
can be excited to higher energy states. Compared to the Coulomb force and the
mass of the electron, the strong nuclear force and the mass of the nucleons are much
larger! This results in nuclear excited states with much higher energies than atomic
energy levels. For the most part, we can’t use lasers to excite nucleons to higher
energy states. Thorium-229 is the one exception. The lowest energy nuclear excited
state of thorium-229 requires a laser with a wavelength .λ = 150 nm, or a frequency
of .2.00 × 1015 Hz. That laser isn’t easy to make or use, but it is possible! The next
isotope with the lowest lying nuclear excited state is protactinium-234, which would
require a laser with .λ = 16.8 nm ☹. However, nuclear physicists are clever and use
other methods to study nuclear excited states. For example, nuclear physicists can
use accelerators and high-energy collisions to excite a nucleus and then measure
the energy of the high-energy photons that are emitted. They also use radioactive
decay. For this technique, some parent systems decay to excited nuclear levels that
then decay via high-energy photons. If you find the nucleus as fascinating as I do,
nuclear physics might be the field for you!
Problems
10.1 Consider the transition in the beryllium atom 1s2 s2 1 S0 → 1s2 s2p 1 P1 . The
resonance frequency for this transition is fr = 1, 276, 080, 100 MHz.
(a) Calculate the normal mass shift as a function of mass number for A' = 7 to
A' = 12 with respect to the stable isotope beryllium-9.
(b) Make a (modified) King plot of the normal mass shift versus A' .
10.2 Write the decay equation for the following radioactive decays:
• β − decay:
– Potassium-40 (40 19 K21 ): Potassium-40 is the largest source of natural radioac-
tivity in animals, including humans. About 89% of all potassium-40 decay is
β − decay.
– Rubidium-87 (87 37 Rb50 ): Rubidium-87 is used in rubidium-strontium dating, a
radiometric dating technique used to determine the age of rocks and minerals.
• β + decay:
– Sodium-22 (22 11 Na11 ): Sodium-22 is used as a calibration source for positron
emission tomography (PET) scans.
– Carbon-11 (11 6 C5 ): Carbon-11 is used in PET scans to detect sites of prostate
cancer.
10.6 The Nuclear Shell Model and Energetics 221
• Electron capture:
– Potassium-40 (40 19 K21 ): Potassium-40 is used in potassium-argon dating.
About 11% of all potassium-40 decay is β − decay.
– Beryllium-7 (74 Be3 ): Beryllium-7 is used in cosmogenic isotope studies to
understand solar activity and atmospheric processes. It is also the lightest
element to undergo electron capture.
• Alpha decay:
– Uranium-238 (23892 U146 ): Uranium-238 produces about 40% of the radioactive
heat produced in the earth.
– Thorium-232 (23290 Th142 ): Thorium-232 is used in thorium reactors and in
dating geological formations through thorium-lead dating.
(a) Using the table of isotopes or online resources, find the half-life of carbon-14.
(b) A mummy was recently found in what was the ancient city of Memphis.
The egyptologist who led the expedition sends a sample of the mummy to
a radiocarbon dating specialist for carbon dating. The results came back as
isotopic abundance, and not a date. The isotopic abundance of the sample is
−13 . Assuming the atmospheric isotopic abundance stays
carbon=12 = 5.8 × 10
carbon-14
constant in time,12 estimate how long ago the individual who is now a mummy
died?
10.5 A decay chain is a series of radioactive decays that shows the sequential
process of decays. It is also known as a “radioactive cascade.” For example, oxygen-
20 undergoes β − decay to fluorine-20, which then undergoes β − decay to neon-20,
which is stable. So, the decay chain is:
β − 20 β − 20
8 O12 −→ 9 F11 −→ 10 Ne10
20
.
Construct a decay chain for thorium-232, considering only the dominant decay
mode at each step. When reaching bismuth-212, follow the decay path to thallium-
208. The decay chain should terminate at lead-208.
12 It
is not constant, but we have ways to calibrate how the ratio has changed in times as well as
geographical variations.
222 10 Isotope Shifts, Radioactive Decay, and the Nuclear Forces
Gm1 m2
FG =
. , (10.16)
r2
2
where G = 6.674 × 10−11 N m2 is a constant of nature, known as the gravitational
kg
constant, and r is the distance between the two objects. The gravitational force
between two objects is always attractive.
Coulomb’s law is a model of the force between two objects with charge.
Mathematically, the force between two objects with charge q1 and q2 is:
kq1 q2
FC =
. , (10.17)
r2
2
where k = 8.988×109 N m2 is a constant of nature, known as the Coulomb constant,
C
r is the distance between the two objects, and the unit C stands for Coulomb. A
proton has a charge of +1.602 × 10−19 C and an electron has the same magnitude
charge but opposite sign, −1.602 × 10−19 C.
(a) Consider two protons in a nucleus. How much larger is the Coulomb force
compared to the gravitational force?
Hint: Take the ratio of the two forces first. The distance between the two protons
will cancel out.
(b) For an object with mass m1 at the surface of the earth, Newton’s law of universal
gravitation is:
Gm1 mE
FG =
.
2
, (10.18)
RE
where mE = 5.972 × 1024 kg is the mass of the earth, RE = 6.378 × 106 m is the
radius of the earth. This formula can be simplified to:
FG = m1 g,
. (10.19)
GmE
where g = 2 . What is the numerical value of g? The units should simplify to
RE
m/s2 .
kg m
Hint: A newton N is the same thing as s2
.
(c) Calculate the force of gravity between you and the earth. The conversion
between pounds and kilograms is 1 kg = 2.205 lbs.
References 223
(d) Find the ratio of the Coulomb force of two protons separated by 1 fm to your
answer in part (c). For context, you are much, much more massive than a proton!
(e) The magnitude of the strong nuclear force between two nucleons separated by
1 fm is about 24,000 N. Find the ratio of this strong nuclear force compared to
your answer in part (c).
References
1. Barwood, G.P., Gill, P., Rowley, W.R.C.: Frequency measurements on optically narrowed Rb-
stabilised laser diodes at 780 nm and 795 nm. Appl. Phys. B 53, 142–147 (1991). https://doi.
org/10.1007/BF00330229
2. Keim, M., Arnold, E., Borchers, W., Georg, U., Klein, A., Neugart, R., Vermeeren, L., Silverans,
R.E., Lievens, P.: Laser-spectroscopy measurements of 72–96Kr spins, moments and charge
radii. Nucl. Phys. A 586(2), 219–239 (1995). https://doi.org/10.1016/0375-9474(94)00786-M
3. Reid, R.V.: Local phenomenological nucleon-nucleon potentials. Ann. Phys. 50(3), 411–448
(1968). https://doi.org/10.1016/0003-4916(68)90126-7
4. Bradford, R.A.W.: The effect of hypothetical diproton stability on the universe. J. Astrophys.
Astron. 30, 119–131 (2009). https://doi.org/10.1007/s12036-009-0005-x
Open Access This chapter is licensed under the terms of the Creative Commons Attribution 4.0
International License (http://creativecommons.org/licenses/by/4.0/), which permits use, sharing,
adaptation, distribution and reproduction in any medium or format, as long as you give appropriate
credit to the original author(s) and the source, provide a link to the Creative Commons license and
indicate if changes were made.
The images or other third party material in this chapter are included in the chapter’s Creative
Commons license, unless indicated otherwise in a credit line to the material. If material is not
included in the chapter’s Creative Commons license and your intended use is not permitted by
statutory regulation or exceeds the permitted use, you will need to obtain permission directly from
the copyright holder.
The Standard Model of Particle Physics
11
Abstract
In this chapter, we explore the fascinating world of particle physics and the Stan-
dard Model of Particle Physics. We discuss the constraints of the Schrödinger
equation and the necessity of quantum field theory, introducing key concepts
like antimatter, vacuum fluctuations, and Feynman diagrams. The chapter details
the fundamental particles and forces in the Standard Model and highlights
unresolved questions, such as dark matter, dark energy, and the integration of
gravity. Additionally, we examine the role of virtual particles and the impact
of vacuum fluctuations on our understanding of particle interactions. We aim to
provide a comprehensive overview of the current state of particle physics and the
exciting challenges that lie ahead.
Learning Goals
. that quantum mechanics does a great job of describing the world of the super
small, but it is not perfect.
. that the Standard Model of Particle Physics is a conceptual model of the world
of the super small; the mathematical framework for it is called Quantum Field
Theory.
. that the Standard Model of Particle Physics is not complete. There are problems
with it, but we aren’t entirely sure how to make a more complete model. There
are lots of ideas, but so far, no one has been able to experimentally confirm a
better model.
We spent a lot of this book conceptually thinking about atoms. Many of the
conceptual ideas come from quantum mechanics, which is mathematically described
by the Schrödinger equation. Whenever you work with a model, it is important to
know the limitations of your model. For example, you can’t use classical physics,
like Newton’s second law, to describe the world of the super small. For things that
are moving very fast,1 we need Einstein’s theory of special relativity. The goal of
fundamental science is to develop a model that accurately describes nature. Since
we now have a good conceptual background on the world of the super small, let’s
discuss the limitations of the quantum mechanical model.
1 By very fast, we mean something like 10% of the speed of light or faster.
2 This is also called conservation of probability.
3 At least you should. It is super important to always know the limitations of your model.
11.2 The Uncertainty Principle Part 4 227
important thing here is that the mass of the three individual quarks is only about
1% of the mass of the proton. Think about that for a minute. Imagine I hand you
3 steel balls. You weigh each one individually and determine two of the balls
have a mass of 2.2 grams and one of the balls has a mass of 4.7 grams. Now you
put all 3 steel balls on the scale. A logical guess would say the total mass should
be .2.2 grams + 2.2 grams + 4.7 grams = 9.1 grams. Instead, you find that the
mass is 938 grams. Replace grams with a much smaller unit for mass and you
have what actually happens . . .
. The speed of causality: Imagine you had two charged particles that are separated
by a few meters. Next you nudge one of the charged particles. According to
the Schrödinger equation (and also classical electromagnetic theory), the other
particle reacts instantaneously. You could repeat this experiment with one of the
charged particles on earth and the other on Mars and get the same result. The two
charged particles feel a force from the other charged particle and if you move one
particle, the other will instantaneously feel a force that pushes or pulls on it. This
violates special relativity. Special relativity says the fastest anything can move,
including information about where a particle is located, is 299,792,458 m/s. This
speed is known as the speed of light or the speed of causality. If the two charged
particles were 299,792,458 meters apart and you nudged one of them, the other
particle shouldn’t know about that for an entire second.
. Spectroscopy: Simply put, the Schrödinger equation gets pretty close to calcu-
lating the correct energy for atomic states, but it is still wrong.
h̄
(Δx)(Δp) ≥
. (11.1)
2
where .Δx is the uncertainty (or range of possible measurements) in the position
of the particle and .Δp is the uncertainty (or range of possible measurements) in
the momentum of the particle. The greater than or equal sign is important as well.
For some systems, the product of the two uncertainties is close to .h̄/2 while other
systems are not. For example, the hydrogen atom with the electron in the ground √
state has a range of possible position values that can be calculated to be .Δx = 23 a0 ,
where .a0 = 5.29 × 10−11 m is a constant called the Bohr radius, named after Danish
228 11 The Standard Model of Particle Physics
h̄
(ΔE)(Δt) ≥
. . (11.2)
2
This is a very useful uncertainty principle! It says that a state that only exists for a
short time cannot have a definite energy. This is why an excited state in an atom has
a natural linewidth (Chap. 3). The lifetime of an excited state tells us how long, on
average, an electron stays in that state, see Fig. 3.6 and Eq. 3.7. Some electrons move
to a lower energy state quickly, while some hang around for a while. This uncertainty
principle helps us understand why the lifetime is connected to the natural linewidth
of a state, see Eq. 3.8 and Fig. 3.8. In fact, the summary at the end of Chap. 3.3 is
stating in words this uncertainty principle!
Let’s work with Eq. 3.8. For ease, here is the equation again:
1
.τ= . (11.3)
𝚪
τ is the lifetime of an excited state, but this is a characteristic time, see the graph
.
in Fig. 11.1. Some states decay quickly (.t < τ ) while others decay more slowly
(.t > τ ). There is a range or uncertainty in the time it takes for that state to decay.
So, let’s call the uncertainty in time .τ : .Δt = τ . Now let’s think about the energy of
the excited state, see the energy level diagram in Fig. 11.1. The energy of the state is
centered at the resonance frequency, but there is a width to that resonance. We call
that width the natural linewidth. In energy units, the width is .h̄𝚪. This is the range
or uncertainty of the excited state: .ΔE = h̄𝚪. Let’s multiply the two uncertainties
Fig. 11.1 An electron in an excited state decays with a characteristic lifetime .τ . That lifetime is
related to the linewidth of the excited state by Eq. 11.3. If .τ is small, .𝚪 is large. Conversely, if .τ is
large, .𝚪 is small
11.3 Antimatter 229
This is really neat! We now know why a spectral feature has a width and an excited
state has a lifetime: time and energy are incompatible observables. While that was
fun, you might be thinking, “Why is this in Chap. 11, a chapter devoted to the
issues and problems in quantum mechanics?” We will need it in Sect. 11.4.2 to
understand something called vacuum fluctuations. But first, we need to discuss one
more concept: antimatter.
11.3 Antimatter
E = mc2
.
That is a lot of energy!a This equation also provides us with a brand new unit
for mass. The mass of the proton is 938.27 MeV/c2 .
a
Converting mass to energy is the basic idea behind nuclear reactors.
In nature, there are a number of conservation laws that seem to always be true. For
example, conservation of energy tells us that for an isolated system (i.e., nothing
comes in or out), the total energy of the system is a constant value. Other conserved
quantities include charge and angular momentum. As we will discuss in the next
section, atom number is not a conserved quantity. For example, we can, and have,
created particles from high-energy photons. This process is called pair production.
To accomplish this, we send a high-energy photon towards a nucleus. The nucleus
“nudges” the photon a little bit and, if the photon has enough energy, it transforms
230 11 The Standard Model of Particle Physics
into an electron and a positron.4 Before the transformation, the photon existed while
the electron and positron did not. After the transformation, the photon no longer
exists.
A positron, which we introduced at various times throughout the book, is known
as the antimatter partner to the electron. It has the same mass and spin as the
electron, but it has opposite (positive) charge. Conservation laws are really helpful
at telling us the properties of the positron. Since the photon has spin 1 and the
electron has a spin of 1/2, conservation of spin tells us the positron also has a spin
of 1/2. A photon has no charge. Likewise, the charges of the electron and positron
are the same magnitude, but with opposite signs. This is called conservation of
charge. Since the photon has no electric charge, if the photon is to be converted into
particles, those particles must together have no charge. There are other differences
between the electron and positron as well, but we aren’t going to go through all of
them.5 If a positron collides with an electron, the two particles “annihilate” each
other back into photons. Before the annihilation, the electron and positron existed
while the photons did not. After the annihilation, the electron and positron no longer
exist.
Other examples of pair production include creating a muon and an antimuon (see
Sect. 11.5), a proton and an antiproton, and a neutron and an antineutron. For this
book, the exact details of antimatter don’t really matter. What matters is that these
antimatter particles can exist, and we have created a lot of antimatter in accelerators
or through studying .β + decay. If an antimatter particle collides with its matter
partner, they will annihilate each other creating photons.6
Fun Fact
The nuclei of an atom can also have excited states similar to the electrons in
the atom. These nuclear excited states are much higher energy compared to
electron excited states. Almost all of these excited states emit a high energy
photon called an x-ray or a gamma ray to transition back to the nuclear
ground state. However, there are nuclei that emit matter and antimatter pairs
to transition back to the nuclear ground state! As an example, oxygen-16 has
an excited state about .6050 × 103 eV above the nuclear ground state.b When
(continued)
4 The nucleus is the external interaction needed to start the process. As a general rule, you always
easy to say but hard to understand. Electrons and positrons have opposite parity. Fun fact: we used
to think parity was a conserved quantity. However, in 1956, Chinese American physicist Tsung-
Dao Lee and Chinese physicist Chen-Ning Yang proposed a theory that the weak force does not
conserve parity. This idea was experimentally confirmed by Chinese American physicist Chien-
Shiung Wu and her collaborators.
6 For completeness, other particles besides photons can also be created during annihilation.
11.4 Going from Quantum Mechanics to Quantum Field Theory 231
Reminder
Quantum mechanics conserves particle number. However, unstable atoms
decay, and we create and destroy particles all the time at particle accelerators.
To fix this incompatibility with nature (i.e., the theory says no, but experiment says
yes), physicists had to remove the constraint of conservation of particle number
from their models. Particle number is clearly not conserved in nature, so we should
try to make a model that doesn’t conserve particle number. However, this causes a
really interesting problem. The Schrödinger equation models an individual particle.
What do we do if the particle doesn’t always exist? The solution was to form a
mathematical model that doesn’t try to model a particle, but something called a field.
But what is a field? The easiest way to understand a field is with a few examples.
Figure 11.2 is a temperature field of the minimum temperatures across the Nordic
countries Norway, Sweden, and Finland from January 27, 1999. The region had an
unusually cold day.7 For reference, .−50◦ C = −58◦ F and .0◦ C = +32◦ F. This
type of field is known as a scalar field. Every point on the map has a temperature. If
I told you a coordinate, you could tell me the temperature at that coordinate. In other
words, this scalar field is really some function .T (x, y). If you knew the function,
you could quickly find the temperature for any coordinate values of x and y.
Figure 11.3 is another type of field known as a vector field. This is a plot that
shows the speed of wind in a simulated tornado. This is, again, a function, but the
answer that you get from this function is a vector. It tells you both the speed of the
T
Temp erature (°C)
– 50 – 40 – 30 – 20 – 10 0
wind and the direction. Other vector fields you might have heard about are electric
fields and magnetic fields.
Instead of following around a particle, quantum field theory keeps track of a field.
Particles are “excitations” of the fields. As an analogy, let’s think about the standing
waves on a quantum mechanical string, see Fig. 6.2. The background field in this
Fig. 11.3 A (simulated) vector field showing the wind speed of a tornado that has a diameter of
about 300 m and a top speed of about 20 m/s. This type of plot gives us both direction (the flow is
counterclockwise) and the speed of the air
11.4 Going from Quantum Mechanics to Quantum Field Theory 233
analogy would be the state with a single loop. The particle would be an excitation of
this field, or the state with two loops. A quantum particle can go into the field (in our
analogy, the system went from two loops to one) or come out of it (one loop to two).
Consider a free neutron moving through space. We know from Chap. 10 that a free
neutron is unstable and will decay into a proton, an electron, and an anti-electron
neutrino with a half-life of about 11 minutes. From a mathematical viewpoint, that
neutron is described as moving through a quantum field. When the neutron decays,
the neutron goes into the field and a proton, electron, and anti-electron neutrino
come out of the field.8
Important Summary
Removing the constraint on conservation of particle number forced physicists
to develop a model that doesn’t rely on a particle always existing. Instead,
quantum field theory models fields where particles can go into or out of these
fields.
8 To
be completely correct, only a quark inside the neutron goes into the field and another quark
comes out, but we haven’t talked about quarks yet.
234 11 The Standard Model of Particle Physics
massive the particle that pops into existence (a big .ΔE), the quicker that particle
must return to the field (a small .Δt). Removing conservation of particle number
combined with the energy-time uncertainty principle allows this odd behavior to
happen. It is important to reiterate that this behavior is a natural consequence and
not something forced into the math. The temporarily existing particles are known as
virtual particles.
These fluctuations can be used to explain why an electron in an excited state
decays back to the ground state. These virtual particles, when they exist, push on the
atom from the outside. This is the external interaction that causes the electron to fall
back to the ground state. Remarkably, theorists can use these vacuum fluctuations
to calculate the lifetime of an excited state, and their math correctly predicts the
lifetime of excited states.
These virtual particles also explain why the mass of a proton or neutron is
so much heavier than the three quarks that make them up. Inside the proton (or
neutron), the three quarks are bound tightly together by the strong nuclear force,
mediated by virtual particles known as gluons. We will discuss quarks and gluons
more in Sect. 11.5. Because of this strong binding, quantum field theory predicts
that virtual particles are continuously generated and annihilated within the field.
The time-averaged mass of the proton thus includes not only the mass of the three
quarks but also the contributions from these virtual particles. Ready to have your
mind blown: 99% of the mass of a proton (or neutron) comes from particles popping
into and out of the field. That means that 99% of your mass has now gone back into
the field and been replaced with new particles from the field.9
I recognize that this all probably seems very, very odd. But, every experiment
seems to confirm that these surprising predictions are true.
Einstein’s theory of special relativity put a speed limit on the universe. Nothing,
including information, can travel faster than the speed of light. However, the
Schrödinger equation has instantaneous information transfer. How do we solve this
problem of instantaneous information transfer?
The answer is force carriers. These force carriers are particles that temporarily
come from the field to transmit information from one particle to another. These ideas
are nicely illustrated in Feynman diagrams, which are named after the American
physicist Richard Feynman.
9 For completeness, there are other ways to think about this extra mass. Some physicists prefer to
use something called binding energy or simply “fluctuations of the field” to account for the extra
mass. All of these ideas are mathematically correct. I have found that virtual particles are a very
accessible way for first-time learners to think about the world of the super small. They will also be
helpful when we get to force carriers, which are virtual particles, in the next section.
11.4 Going from Quantum Mechanics to Quantum Field Theory 235
c
Some physicists like to rotate their Feynman diagrams so that time goes left to right instead
of bottom to top.
Fig. 11.4 An example of a Feynman diagram. This diagram illustrates a possible interaction
between two electrons
236 11 The Standard Model of Particle Physics
Fig. 11.5 A Feynman diagram depicting how a virtual photon can momentarily split into an
electron-positron pair while being exchanged between two electrons
apart, the virtual photons need to exist for a longer time to mediate the interaction,
which means they have less energy. Consequently, the force between the two
charged particles decreases with distance. This is why, according to quantum field
theory, the repulsive or attractive force between two charged particles diminishes as
the distance between them increases.
Other things can happen as well. For example, Fig. 11.5 describes how a virtual
photon can split into an electron-positron pair when traveling between the two
charged particles. This doesn’t happen very often compared to the simple exchange
of a virtual photon (the electron-positron pair have much more energy, so the
energy-time uncertainty principle makes it harder for this to happen), but when the
electron-position pair pop into existence from the field, they have a real impact on
the system.
Consider an electron in an atom. That electron is constantly interacting with the
protons in the nucleus via virtual photons. Every once in a while, that virtual photon
breaks apart into an electron-positron pair similar to Fig. 11.5 (just replace one of the
electrons with a proton). The effect of the electron-positron pair is that the transition
frequency between any two states is shifted by just a little bit, but this small shift is
still big enough for us to experimentally measure!
There are other things besides virtual photons and electron-positron pairs that
can happen, and theorists calculate the effect that all of these different scenarios
will have on an atom. Adding up all of the different possible scenarios produces a
single overall shift to the transition frequency. Experimentalists then go measure the
transition frequency.
Quantum field theory uses all these ideas to make predictions about the atom.
Remarkably, these predictions have almost always been confirmed by experiment.
In fact, the theory has been confirmed so many times that if there is a disagreement
between theory and experiment, we all assume someone made an error.
There are two major subfields to quantum field theory: quantum electrodynamics
and quantum chromodynamics. Quantum electrodynamics (QED) is a mathematical
framework that describes interactions between charged particles and light (including
virtual photons). Quantum chromodynamics (QCD) is a mathematical framework
11.5 The Standard Model of Particle Physics 237
that describes the interactions inside the nucleus including the interactions of the
quarks inside the proton and neutron.
You may have heard some form of the expression, “Particle accelerators make
particles.” This is a true statement, but we now have the background to understand
how they make new particles. According to quantum field theory, we have virtual
particles popping into existence from the field and hanging out for a short amount
of time before returning to the field.
Important Reminder
Einstein showed us that energy and mass are the same thing
.E = mc2
Let’s take two protons and give them a ton of kinetic energy. Next, we are going to
smash the two protons together. If the kinetic energy of the protons is larger than
the energy equivalent of the particle’s mass coming from the field (.E = mc2 ), the
production of that particle does not violate conservation of energy. In this case, the
energy-time uncertainty principle is not the limiting factor, allowing the particle to
exist as a real particle. This is exactly what particle accelerators do. Almost all of
the particles created in accelerators are unstable and undergo radioactive decay, but
they exist long enough for us to determine important properties of the particles, such
as mass, charge, and spin.
Quantum field theory is currently the best mathematical model that we have to
describe the world of the super small. The conceptual model is called the Standard
Model of Particle Physics, or just the Standard Model for short, see Fig. 11.6.
A Little History Before the 1960s, the world of particle physics was really con-
fusing. Physicists at particle accelerators were creating hundreds of new particles.
It was a little overwhelming; could there really be hundreds upon hundreds of
elementary particles? There were so many new particles that in 1956, an American
physicist named Robert Oppenheimer coined the term “subnuclear zoo.”
In 1964, American physicists Murray Gell-Mann and George Zweig indepen-
dently realized that the whole crazy picture could be simplified to just a few
elementary particles: quarks. Despite particle accelerators seemingly finding new
238 11 The Standard Model of Particle Physics
particles almost every week, what they figured out was that all of the particles
were made from twelve elementary particles and their antimatter counterparts. Six
of these are now called the up quark, the down quark, the charm quark, the strange
quark, the top quark, and the bottom quark,10 and all of their antimatter counterparts.
At the moment, our best conceptual model of the world of the super small says
that everything in the universe is made up of the elementary particles shown in
Fig. 11.6. There are four major groupings in the Standard Model.
. Leptons are elementary particles that do not experience the strong force and
include electrons, muons, taus, neutrinos, and all their antimatter counterparts.
They are colored green in Fig. 11.6.
. Quarks are elementary particles that experience all fundamental forces, includ-
ing the strong force. Each quark has an antimatter counterpart. Any particle made
from quarks, including protons and neutrons, is called a hadron. They are colored
blue in Fig. 11.6.
. Force Carriers are particles that mediate the fundamental forces. They include
the photon (electromagnetic force), W and Z bosons (weak force), gluons (strong
force), and the hypothetical graviton (gravity). They are colored red in Fig. 11.6.
. Higgs Boson is an elementary particle associated with the Higgs field, which
gives mass to other particles. It is colored black in Fig. 11.6.
10 Some of these names are odd, but it was the 1960s and 1970s .. . . .
11.5 The Standard Model of Particle Physics 239
Higgs field through the Higgs boson. This interaction is, according to the Standard
Model, the thing that gives the particles mass. This was the last particle to be
discovered by particle physicists. The Higgs boson itself was theoretically predicted
back in 1964 and finally created in a particle accelerator in 2012. Peter Higgs and
Belgian physicist François Englert won the Nobel Prize for the Higgs boson in 2013.
This is exactly the question physicists are always asking! We want to understand
nature. To do so, we develop models and test the models to make sure they
accurately describe nature. There are different ways to test the Standard Model, and
Part 1 of this book discusses how atomic and nuclear physicists use spectroscopy
as a tool to test these models. The Standard Model of Particle Physics, which is
mathematically described by quantum field theory, is our current best model for the
world of the super small. The theory has been tested over and over again, and it has
succeeded almost every single time. However, we know the theory is not complete.
There are plenty of things in the universe that are not included in the Standard
Model. That means that we might be able to create a more complete theory. Here is
a list of some of the phenomena that are not in the Standard Model.
Gravity Believe it or not, quantum field theory does not include gravity. Einstein’s
theory of general relativity is our best model of gravity. General relativity has
predicted fascinating phenomena such as gravitational lensing (the theoretically
predicted and experimentally measured phenomenon of light bending around
massive astronomical objects like large stars and galaxies) and gravitational waves
(the theoretically predicted and experimentally measured phenomenon that creates
“ripples” in space when two black holes merge; quite literally space is compressed
just a little bit and this compression wave ripples out to be detected on earth).
Like quantum field theory, general relativity is an incredibly successful theory.
Interestingly, the two theories are incompatible with one another. There are a
number of reasons why, but one of the big reasons is that general relativity requires
“spacetime” to be smooth and continuous while quantum field theory discretizes
the fields. In other words, quantum says no to smooth and continuous while general
relativity says no to discrete and bumpy. Quantum field theory also needs a force
carrier to transmit gravitational information between two objects with mass, which
we call a graviton. However, we have yet to create and measure a graviton in the lab.
Where Is All the Antimatter? Quantum field theory predicts there should be about
equal parts matter and antimatter. For example, Fig. 11.8 shows a photon that has
enough energy to create particles (we have done this at accelerators!). The particles
that are created are always a matter/antimatter pair. However, when we look out in
the universe, we only see matter. So, where is all the antimatter? We don’t know.
This problem is called baryon asymmetry. This isn’t a bad thing. If the universe
was equal parts matter and antimatter, they would have annihilated each other. So,
11.6 So, What’s Next? 241
in a way, this is one experimental measurement that the Standard Model failed to
predict. The Standard Model predicts equal amounts of matter and antimatter, while
observations show a universe dominated by matter.
Why Do Neutrinos Have Mass? This one is short. The Standard Model says that
neutrinos don’t have mass. However, they do. This is why I use the phrase “The
theory has been tested over and over again, and it has succeeded almost every single
time.” The Standard Model says neutrinos don’t have mass, but we have found that
they do.
What Is Dark Matter and Dark Energy? This one is also short. From astronomi-
cal observations, the universe is thought to be about 5% atoms, 26% dark matter, and
69% dark energy. However, we don’t even know what dark matter or dark energy
is . . . so it is a little hard to include them in a model. Still, it is kinda weird that our
most successful model ever tested (quantum field theory) doesn’t know how to deal
with 95% of the universe.
Even though quantum field theory has been incredibly successful, there is
still a lot to learn! For physicists, this is the best thing ever. There have been
some incredibly interesting and ingenious ideas to make the Standard Model more
complete. Some of these ideas include string theory, supersymmetry, and loop
quantum gravity. However, no one has come up with any way to definitively test
if any of these ideas are correct. But this is what makes everything so exciting! We
still have more to learn, more to understand, and more to explore. To learn more and
push our understanding always onward, we use the most important equation in all
of physics:
Problems
11.1 Explain why the Schrödinger equation is not suitable for describing particles
moving at speeds close to the speed of light.
11.2 Describe the process of pair production and explain why it violates the
conservation of particle number in quantum mechanics.
11.4 In Sect. 11.3, we stated the mass of a proton can be written as .938.27 MeV/c2 .
11.5 Draw a Feynman diagram for .β + decay. The force carrier will be a .W+ boson.
11.6 Draw a Feynman diagram for electron capture. The force carrier will be a .W+
boson.
11.7 Explain how vacuum fluctuations can cause an electron in an excited state to
decay back to its ground state.
11.8 An accelerator accelerates two protons to very high energy. The two protons
travel in opposite directions with the same speed and collide. The goal is to create a
top quark and an anti-top quark.
(a) What is the minimum energy each proton must have to make a top quark-
antiquark pair?
(b) The relationship between the energy of a particle and its velocity comes from
special relativity. The formula is
mc2
E=/
.
2
1 − vc2
Using your answer from part (a), what is the speed of a proton? Express your
answer in units of c. For example, an incorrect answer is .v = 0.92c, or 92% the
speed of light.
Reference 243
Fun Fact The Tevatron was a particle accelerator at Fermi National Accelerator
Laboratory (Fermilab), located in Batavia, Illinois, near Chicago. It accelerated
protons and antiprotons to energies of 980 GeV, producing proton-antiproton col-
lisions with energies of up to 1.96 TeV. Sadly, the Tevatron shut down in 2011
because the Large Hadron Collider (LHC) at CERN, which became operational
in 2008, surpassed the Tevatron in terms of energy levels. The LHC can collide
protons at energies of 7 TeV per beam (14 TeV in total). CERN is located in Geneva,
Switzerland, and it is currently the world’s largest particle accelerator. CERN is an
acronym for the French name “Conseil Européen pour la Recherche Nucléaire,”
which translates to “The European Organization for Nuclear Research.”
Reference
1. Tveito, O.E., Førland, E.J., Heino, R., Hanssen-Bauer, I., Alexandersson, H., Dahlström, B.,
Drebs, A., Kern-Hansen, C., Jónsson, T., Vaarby-Laursen E., Westman, Y.: Nordic Temperature
Maps DNMI Klima 9/00 KLIMA. Norwegian Meteorological Institute, Oslo (2000). Copernicus
Climate Change Service, Climate Data Store, (2021): Nordic gridded temperature and precipi-
tation data from 1971 to present derived from in-situ observations. Copernicus Climate Change
Service (C3S) Climate Data Store (CDS). https://doi.org/10.24381/cds.e8f4a10c, Accessed Jan
25, 2024
Open Access This chapter is licensed under the terms of the Creative Commons Attribution 4.0
International License (http://creativecommons.org/licenses/by/4.0/), which permits use, sharing,
adaptation, distribution and reproduction in any medium or format, as long as you give appropriate
credit to the original author(s) and the source, provide a link to the Creative Commons license and
indicate if changes were made.
The images or other third party material in this chapter are included in the chapter’s Creative
Commons license, unless indicated otherwise in a credit line to the material. If material is not
included in the chapter’s Creative Commons license and your intended use is not permitted by
statutory regulation or exceeds the permitted use, you will need to obtain permission directly from
the copyright holder.
The Periodic Table
A
Below is a list of all of the known elements sorted by the number of protons in the
nucleus (Z). The last column lists the mass number (A) for all of the stable isotopes.
For example, the stable isotope magnesium-25 has 12 protons and 13 neutrons for a
total of 25 nucleons.
The following transition rules (also known as selection rules) must be satisfied for
an electron to transition between two atomic states:
• .ΔL = ±1: The electronic orbital angular momentum quantum number must
change by 1.
• .ΔS = 0: The spin quantum number must remain unchanged.
• .ΔJ = −1, 0, + 1; J = 0 /→ J = 0: The total electronic angular momentum
quantum number must change by 1, 0, or -1. However, transitions between .J =
0 → J = 0 are forbidden.
• Parity must change: A transition must occur between states of opposite parity.
Parity is indicated by a circle in the term symbol. For example, a state with odd
parity has a small circle, like .1 P◦1 . A state with even parity does not have a circle,
like .1 S0 .
• .ΔmJ = −1, 0, +1: The z-component (projection) of the total electronic angular
momentum (.mJ ) can change by 1, 0, or -1. However, if .ΔJ = 0, the transition
between .mJ = 0 → mJ = 0 is forbidden.
• Photon: The force carrier of the electromagnetic force. Photons are massless
and mediate interactions between charged particles.
• W and Z Bosons: The force carriers of the weak nuclear force. These bosons
are massive and are responsible for processes such as beta decay in radioactive
materials.
• Gluons: The force carriers of the strong nuclear force. Gluons are massless
and mediate the force that holds quarks together within protons and neutrons.
• Graviton (hypothetical): The proposed force carrier of gravity in quantum
gravity theories. Gravitons are hypothesized to be massless and mediate the
gravitational force, although they have not been experimentally observed.
Higgs Boson: An elementary particle associated with the Higgs field, responsible
for giving mass to other particles.
Hyperfine level: When a nucleus has angular momentum, an atomic state splits
into multiple closely spaced states known as hyperfine levels.
Hyperfine structure splitting: The splitting of energy levels in an atom due to
nuclear spin.
Incompatible observables: Observables that cannot be precisely measured simul-
taneously. Measuring one observable alters the system, making subsequent
measurements of the other unpredictable.
Intensity: The intensity of a laser beam is the power of the laser divided by the
cross sectional area of the laser beam, denoted as .I = P /A.
Ion: An atom with a net charge due to an unequal number of protons and electrons.
An ion with more protons than electrons is called a positive ion, while an ion with
more electrons than protons is called a negative ion.
Ionization Threshold: The minimum energy required to remove an electron from
an atom or molecule, thereby ionizing it. Each atom has a unique ionization
threshold. Most atoms have multiple ionization thresholds depending on the
angular momentum and energy state of the remaining electrons.
Isotope shift: The change in the resonance frequency of a transition between two
states in an atom due to a change in the number of neutrons in the nucleus. The
'
variable for isotope shift is .δf AA , where A is the mass number for one isotope
and .A' is the mass number of the other isotope.
Kinetic energy: The energy of motion. The variable K represents kinetic energy.
Leptons: Elementary particles that do not experience the strong force, including
electrons, muons, taus, neutrinos, and their antimatter counterparts.
Mass number: The total number of protons and neutrons in an atomic nucleus,
represented by the variable A.
Molecule: A group of two or more atoms bonded together by chemical bonds.
Momentum: A property of a moving object. For a classical object, momentum is
given by .p = mv, where p is the object’s momentum, m is the object’s mass,
and v is the object’s velocity. An object’s momentum changes if acted upon by
an external force. The unit of momentum is .kgm/s.
Natural linewidth: The minimum possible full width at half maximum (FWHM)
of a spectral feature. The natural linewidth is a property of a transition, with each
transition in an atom having a unique natural linewidth. The variable for natural
linewidth is the lowercase Greek letter gamma .γ when using linear frequency
units and the uppercase Greek letter gamma .𝚪 when using angular frequency
units.
Neutral atom: An atom with no net charge, having an equal number of electrons
and protons.
Neutron: A particle found inside the nucleus of an atom. A neutron has no charge,
a mass of .1.675 × 10−27 kg, and a radius of .0.8 × 10−15 m = 0.8 fm. A neutron
is composed of three quarks: 1 up quark and 2 down quarks. A free neutron is
unstable and undergoes .β − decay with a half life of about 11 minutes.
Glossary 257
Saturation parameter: The ratio of the laser intensity to the saturation intensity.
The variable s represents the saturation parameter.
Scattering rate: The number of photons per second an atom absorbs and re-emits.
The variable for scattering rate is .rγ (δ, s) when using linear frequency units and
.r𝚪 (Δ, s) when using angular frequency units.
Doppler shift formula, .v‖ is negative if the atom is traveling towards the laser and
positive if it is traveling away from the laser.
Virtual particles: Temporary fluctuations in quantum fields that arise from the
energy-time uncertainty principle. They can mediate forces, like virtual photons
in electromagnetic interactions, or appear as vacuum fluctuations, but they are
not directly observable.
Wave interference: A phenomenon occurring when two or more waves overlap,
resulting in a wave with greater (constructive interference) or lower (destructive
interference) amplitude. Interference effects can be observed with all types of
waves, including light waves, water waves, and electron waves.
Waist of a laser: The half-width of the intensity profile where the intensity is
13.5% of the maximum intensity, as shown in Fig. 3.7. The variable w represents
waist.
Wavelength: The distance between two consecutive “like” points on a wave, such
as two adjacent wave peaks. The variable that represents wavelength is the
lowercase Greek letter lambda, .λ.
Wave-particle duality: The concept that particles, like electrons and photons,
exhibit both wave-like and particle-like behavior depending on the experiment,
meaning they can act as waves in some situations and as particles in others.
Index
A Detuning, 51
Absorption plot, 44, 47, 72 Diffraction, 26–27
Absorption spectroscopy, 34 Dispersive element, 26
A general rule of nature, 164, 205, 213 Doppler broadening, 76
Angular momentum, 137–139 Doppler effect, 65
adding quantum mechanical angular Doppler profile, 76, 87
momentum, 145–150, 177–181 Doppler shift, 70, 87
orbitals, 172 astronomy, 79
Angular units, 47 Doppler width, 76, 77
Antimatter, 229–231, 238 Double-slit experiment, 8–13
Atomic physics, 4
E
B Electromagnetic force/interaction, 166, 216
Basis sets, 121, 142 Electron, 5, 138, 238
Beyond the standard model, 240 Electron configuration, 164
baryon asymmetry, 240 Electron shells and subshells, 161–165
dark matter and dark energy, 241 Equipartition Theorem, 77–78
gravity, 240 Excited state fraction, 55
neutrino mass, 241
Blackbody radiation, 30–34
the ultraviolet catastrophe, 31 F
Bosons, 172–173 Fermions, 172–173
Bra-ket notation, 58, 117, 152, 172, 180, Feynman diagrams, 234–237, 239
216 Fine structure, 167–169
Fluorescence, 44
Force, 136
C Newton’s 2nd Law, 136
Center of gravity, 58, 179, 185 Force carriers, 234, 238
Coulomb force/interaction, 160, 209 Frequency, 10
Crossover-free spectroscopy, 103–106
Crossovers, 90–97, 189
V crossovers, 91, 189 G
X crossovers, 94 Gluons, 234, 238
.Λ crossovers, 94 Graviton, 238
H
D Higgs boson, 238
Degenerate/degeneracy, 169–170
K
Kinetic energy, 77 Q
Quantum field theory, 232–237
quantum chromodynamics, 236
L quantum electrodynamics, 236
Leptons, 238 Quantum harmonic oscillator, 127–128
Lifetime, 49, 228 Quantum mechanics, 7, 114
Copenhagen interpretation, 17, 114
limitations, 226
M Quantum numbers, 137, 162, 178
Madeling energy ordering rule, 161 nuclear spin, 150, 177
Mass number, 198 nuclear spin: z-component, 150
Maxwell-Boltzmann velocity distribution, orbital angular momentum, 137, 139,
73–75 148–150
Molecules, 5 orbital angular momentum: z-component,
Momentum, 114, 136 139, 148–150
Muon, 238 principal quantum number, 116, 162
spin, 148–150
spin: z-component, 148–150
N total electronic angular momentum,
Natural linewidth, 44, 47, 228 148–150
Neutrino, 203, 238 total electronic angular momentum:
Neutron, 5, 138, 207, 234 z-component, 148–150
Nuclear notation, 198 Quantum states, 116–119
Nuclear shells, 217–219 energy states, 15, 116, 128, 158
magic numbers, 217 position states, 117
Nucleon, 198 Quarks, 205, 226, 239
O R
Observables, 114 Radioactive decay, 203–207, 226
compatible/incompatible observables, 120, .α decay, 204
141, 151–153, 180 −
.β decay, 203, 213, 239
the measurement game, 132 +
.β decay, 204, 214
Index 263