Mod Matterenergy
Mod Matterenergy
project
* SPICE ckt
v1 1 0 dc 12
v2 2 1 dc 15
r1 2 3 4700
r2 3 0 7100
.dc v1 12 12 1
.print dc v(2,3)
.print dc i(v2)
.end
V=IR
This is a copyrighted work, but licensed under the Creative Commons Attribution 4.0 International
Public License. A copy of this license is found in the last Appendix of this document. Alternatively,
you may visit http://creativecommons.org/licenses/by/4.0/ or send a letter to Creative
Commons: 171 Second Street, Suite 300, San Francisco, California, 94105, USA. The terms and
conditions of this license allow for free copying, distribution, and/or modification of all licensed
works by the general public.
ii
Contents
1 Introduction 3
2 Simplified Tutorial 5
3 Full Tutorial 11
3.1 Matter and atomic structure . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12
3.1.1 Atoms and particles . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13
3.1.2 Molecules and ions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15
3.1.3 Chemical formulae . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 16
3.2 Electronic structure . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 19
3.3 Work, energy, and power . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 27
3.3.1 Force and displacement vectors . . . . . . . . . . . . . . . . . . . . . . . . . . 28
3.3.2 Energy and power . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 30
3.3.3 Perpendicular force and displacement . . . . . . . . . . . . . . . . . . . . . . 31
3.3.4 More energy transfer . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 32
3.3.5 Rotational work . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 34
3.3.6 Energy and safety . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 38
3.3.7 Kinetic energy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 41
3.3.8 Dimensional analysis of energy . . . . . . . . . . . . . . . . . . . . . . . . . . 43
3.4 Elementary thermodynamics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 44
3.4.1 Heat versus Temperature . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 45
3.4.2 Temperature . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 46
3.4.3 Heat . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 48
3.4.4 Heat transfer . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 49
3.4.5 Specific heat and enthalpy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 60
3.4.6 Phase changes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 67
3.4.7 Phase diagrams and critical points . . . . . . . . . . . . . . . . . . . . . . . . 75
3.4.8 Thermodynamic degrees of freedom . . . . . . . . . . . . . . . . . . . . . . . 78
3.4.9 Applications of phase changes . . . . . . . . . . . . . . . . . . . . . . . . . . . 79
3.5 Unit conversions and physical constants . . . . . . . . . . . . . . . . . . . . . . . . . 90
3.5.1 Unity fractions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 91
3.5.2 Conversion formulae for temperature . . . . . . . . . . . . . . . . . . . . . . . 96
3.5.3 Conversion factors for distance . . . . . . . . . . . . . . . . . . . . . . . . . . 96
3.5.4 Conversion factors for volume . . . . . . . . . . . . . . . . . . . . . . . . . . . 96
iii
iv CONTENTS
5 Questions 123
5.1 Conceptual reasoning . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 127
5.1.1 Reading outline and reflections . . . . . . . . . . . . . . . . . . . . . . . . . . 128
5.1.2 Foundational concepts . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 129
5.1.3 Residential electric billing . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 130
5.1.4 Roller-coaster ride . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 131
5.1.5 Potential or kinetic? . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 131
5.1.6 BB guns and arrows . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 132
5.1.7 Applying foundational concepts to various systems . . . . . . . . . . . . . . . 133
5.2 Quantitative reasoning . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 134
5.2.1 Miscellaneous physical constants . . . . . . . . . . . . . . . . . . . . . . . . . 135
5.2.2 Introduction to spreadsheets . . . . . . . . . . . . . . . . . . . . . . . . . . . 136
5.2.3 Moving weights . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 139
5.2.4 Stopping distance . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 139
5.2.5 Potatoes and bullets . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 140
5.2.6 Antenna dimensions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 141
5.2.7 Analog-digital converter signals . . . . . . . . . . . . . . . . . . . . . . . . . . 142
5.2.8 Pressure unit conversions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 143
5.2.9 Flow rate conversion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 144
5.2.10 Frequency conversions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 144
5.3 Diagnostic reasoning . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 145
5.3.1 A motor that runs without energy? . . . . . . . . . . . . . . . . . . . . . . . . 146
5.3.2 Pulley and rope system . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 147
E References 169
Index 172
2 CONTENTS
Chapter 1
Introduction
This learning module introduces some basic concepts of matter and energy useful to students of
electricity and electronics. The Simplified Tutorial is very brief compared to the Full Tutorial, which
touches on basic chemistry and thermodynamics. Electricity is fundamentally regards the exchange
of energy using mobile electrical charges, and so a firm conceptual foundation in the concept of
energy, how it is conserved, as well as conservation laws related to electric charge and mass is very
important.
When first learning anything new, it is extremely helpful to challenge yourself to express what
you have learned into your own words, using your own examples, and to discover how these new
ideas connect to ideas previously learned. You can do this by writing, by speaking, and also by direct
application (by doing). Then, share your thoughts and actions with others to gain their perspective.
If you are a student in a formal educational environment, bringing this level of engagement to the
conversation will help you (as well as your fellow students) tremendously. You will find this not only
helps you remember, but it also gives you greater depth of understanding.
Important principles discussed in the Tutorial chapters include conservation laws (energy,
mass, electric charge), basic atomic structure, the relationship of energy to force and motion,
different forms of energy and how they relate to motion, and the distinction between energy versus
power. In particular, energy in its different forms is a unifying principle throughout the study of
electricity and electronics.
Here are some good questions to ask of yourself while studying this subject:
• How might an experiment be designed and conducted to explore the concept of energy
conservation? What hypothesis (i.e. prediction) might you pose for that experiment, and
what result(s) would either support or disprove that hypothesis?
• How might an experiment be designed and conducted to explore the concept of mass
conservation? What hypothesis (i.e. prediction) might you pose for that experiment, and
what result(s) would either support or disprove that hypothesis?
• How do we define energy, and how may we calculate its quantity?
• What does it mean for a physical property to be conserved ?
3
4 CHAPTER 1. INTRODUCTION
• What is work, and what does it mean in the active and passive form (i.e. what does it mean
to do work on something, versus have work done on something)?
• How does power differ from energy, and why are these concepts easy to confuse with one
another?
An excellent “active reading” strategy to apply in this text as well as all technical tutorials is to
actually perform the mathematical work shown by the author, for the purpose of reinforcing your
mathematical abilities as well as illuminating the author’s points. Look for opportunties in the text
to do this, where you can take the information given in the set-up of a mathematical problem, then
independently compute your own answer, checking against the text to see if your result matches the
author’s.
You will find many other “active reading” strategies in the “Reading outline and reflections”
subsection of the Conceptual reasoning section of the Questions chapter. Additionally, the
“Foundational concepts” subsection in that same chapter lists many of the important principles
contained in this learning module. A good self-check of your reading comprehension is to see if you
are able to define each of those listed concepts.
Common misconceptions include the following. For each of these, challenge yourself to prove
why the misconception is incorrect:
In the Derivations and Technical References chapter you will find resources for exploring
important concepts in greater depth. Examples include Newton’s Laws of Motion where we find Isaac
Newton’s expressions of fundamental principles related to the motion of matter, and Dimensional
analysis where we explore a powerful technique for validating the application of mathematics to
physical phenomena.
Chapter 2
Simplified Tutorial
Matter is anything possessing mass, “mass” being defined as the opposition to acceleration (changes
in velocity). Mass is related to weight, but not identical to it because “weight” is the downward force
exerted on a mass by a gravitational field. Matter in outer space, for example, would be weightless
since it experiences negligible gravity, but it would still have mass.
Isaac Newton, the famous scientist who lived from 1642-1727, formulated some laws of motion,
the first and second of which describe these properties of mass:
1. An object at rest tends to stay at rest; an object in motion tends to stay in motion1
2. The acceleration of an object is directly proportional to the net force acting upon it and
inversely proportional to the object’s mass
A very important property of mass is that it is conserved, which means it is eternal and cannot
be created or destroyed. Mass may change form, but it cannot be eliminated nor can it appear from
nothing. This is known as the Law of Mass Conservation, or simply the Conservation of Mass.
The standard metric unit of measurement for mass is the kilogram, one kilogram being equal
to one thousand grams (a much smaller measure of mass). In the Imperial (British) measurement
system mass is measured either in slugs or pounds.
Through many centuries of experimentation, people discovered that matter may be subdivided
into ever-smaller components. What may appear to be a completely solid and homogeneous object
is actually comprised of molecules, which are in turn formed of even smaller atoms, which in turn
are formed of even smaller sub-atomic particles. The separation of molecules into individual atoms,
as well as the combining of atoms to form molecules, is the domain of chemistry. The separation of
atoms into particles and combining particles to make atoms is the domain of nuclear physics.
5
6 CHAPTER 2. SIMPLIFIED TUTORIAL
A tremendously simplified model of a single atom of the element nitrogen is shown here, each
of its constituent particles represented by colored circles: protons, neutrons and electrons. Seven
protons and seven neutrons comprise the nucleus of a nitrogen atoms, and seven electrons “orbit”
around it:
N
P N
P
N P
e P N P e
P N
P
N N
e
e
e
The chemical identity of an atom is defined entirely by the number of protons (called the atomic
number ) within an atom’s nucleus. Those protons and neutrons comprising an atom’s nucleus are
very tightly bound together, so tight in fact as to be almost inseparable from each other. This is why
atoms are almost impossible to modify2 . The orbiting electrons are much easier to dislodge or to
add, and in fact it is precisely this process that results in atoms merging together to form molecules,
the following illustration showing a nitrogen molecule formed of two nitrogen atoms bonded together
by the sharing of three electrons:
An N2 molecule
e
e e
Chemical symbol
e
N N
P N P N
N
P P
N P N P
e P N P e e e P N P e
P N P N
P P
N N N N
2 e e
Number of atoms
bound together
in a molecule
e e
e
2 In fact, the word “atom” comes from the Greek word atomos, which means “indivisible”. Ancient philosophers
of nature speculated that atoms themselves were conserved (i.e. incapable of creation or destruction) but modern
scientific techniques have shown this to be untrue. Now it actually is possible to pluck particles out of or add particles
to atomic nuclei, thereby altering those atoms’ identities, but it is still very difficult process.
7
Electrons and protons are mutually attracted by a force known as an electrostatic force. This is
similar to gravity, but very different in one regard: whereas gravity can only attract matter together,
the electrostatic force can both attract and repel matter. The equivalent of mass for gravitational
fields is charge for electrostatic fields, and particles having the same “polarity” of charge repel each
other while particles having opposite-polarity charges attract. By convention, electrons are said to
have a negative charge and protons a positive charge. Electrons are attracted to protons, but repelled
by other electrons. Protons also tend to repel each other, but this repulsive electrostatic force is
much weaker than the strong nuclear force binding protons and neutrons together in the nucleus.
Neutrons have no charge.
In an electrically “balanced” atom, the number of electrons exactly equals the number of protons,
leaving the atom which no net charge. If an atom is stripped of one or more of its electrons, it will
have a net positive charge; if extra electrons join the atom, it will have a net negative charge. In
both cases we refer to the “imbalanced” atom as an ion, one example being a negatively-charged
nitrogen ion represented in the illustration below:
An N 3- ion
e
Chemical symbol e e
Ionic charge
(- means extra electrons,
+ means missing electrons)
3- e
P
N
N
N
P
N P
e P N P e
P N
P
N N
e
e
e
In some types of solid material, most notably metals, the atoms bind together in such a way that
their outer-most electrons are free to drift apart from their parent atoms and actually move like a
fluid within the solid. These mobile charges are the stuff of electricity. Ions existing within a liquid
or a gas – where whole atoms are free to drift independently of each other – may also serve as mobile
electric charges just like individual electrons moving about within a solid metal object.
Electric charge, like mass, is also a conserved quantity. That is to say, charges can neither be
created nor destroyed but are eternal. This is known as the Law of Electric Charge Conservation,
or simply the Conservation of Electric Charge.
To summarize, all matter has mass and experiences an attractive force (called the gravitational
force) toward other masses. Some types of matter also have charge and experience either an attractive
or a repulsive force with other charged matter (called the electrostatic force). Both mass and charge
are conserved quantities in our universe. Charges that are free to move about is what electricity is
made of.
8 CHAPTER 2. SIMPLIFIED TUTORIAL
A concept more abstract than either mass or charge is energy, and for the purposes of our
discussion will be defined as that which sets matter into motion. Imagine an electron and a proton
separated a short distance from each other in otherwise empty space, their electrostatic attraction
causing them to move toward each other:
F F
e P
In the absence of any external restraint, these two charged particles will move toward each other,
and thus we may conclude energy is at work here as per our definition of energy being that which
sets matter into motion. At first one might be tempted to identify the source of this energy as the
electrostatic force acting on these particles, but this is not entirely accurate. The next illustration
should make this clear:
F F
e P
Here, we see the electron and proton have moved together until they touch, at which point they
cannot travel any closer to each other. The attractive force of their opposing electrostatic charges
is still there, but since there is no longer room for motion they cannot move closer. If we define
energy as “that which sets matter into motion” we must conclude energy is no longer present in this
system, even though force still is. Therefore, energy cannot be the same thing as force.
Force is clearly involved in energy – for without a force attracting the two particles together they
would have no impetus to begin moving toward each other – but there must be more to the concept
of energy than force alone. What we need for there to be energy is both force and displacement
(travel, or at least the potential of travel).
9
In the study of physics, the combined action of force and displacement is called work, and it may
be calculated as the product (multiplication) of force and displacement and the cosine of the angle
between those two vectors3 .
Displacement
x Work (W) done by force is positive
Force W = Fx cos 0o
Object F W = (Fx) (1)
Displacement
x Work (W) done by force is zero
W = Fx cos 90o
Object Force
F W = (Fx) (0)
Displacement
x Work (W) done by force is negative
Force W = Fx cos 180o
Object F W = (Fx) (-1)
When work occurs – i.e. a force acting parallel with a displacement – energy is exchanged.
Returning to our electron-proton system, one could imagine some external force prying those two
particles apart from each other against the electrostatic force of their mutual attraction. In this
case that external force would be doing work on the particles, and in so doing that work those two
particles would end up having more energy (i.e. more potential to travel toward each other) than
before. If we could somehow harness both of those separated particles to make them motivate some
other matter as they fell toward each other again, the electron and proton would be doing work on
that other matter, and in so doing that work their energy would be transferred there.
Interestingly, energy – just like mass and just like charge – is a conserved quantity, incapable of
being created or destroyed. Any time work occurs in the physics sense of that word, energy merely
moves from one location or state to another, but the grand total of energy in the universe remains
exactly the same. This is called the Law of Energy Conservation, or simply the Conservation of
Energy.
3 A vector in physics is a quantity that has both a magnitude and a direction. This is true of both forces and
displacements.
10 CHAPTER 2. SIMPLIFIED TUTORIAL
Energy is usually defined as being one of two types, potential or kinetic. “Potential” energy is
capable of setting matter into motion, without any motion occurring at the moment. “Kinetic”
energy, by contrast, is what moving matter possesses – i.e. able to set other matter into motion by
its own motion (by collision, for example).
Both forms of energy are easy to calculate, potential energy being a function of force and potential
displacement parallel to that force, and kinetic being a function of mass and velocity:
1
Ep = F x Ek = mv 2
2
Where,
Ep = Potential energy in Joules or Newton-meters (metric), or foot-pounds (British)
F = Force in Newtons (metric) or pounds (British)
x = Potential displacement in meters (metric) or feet (British)
Ek = Kinetic energy in Joules or Newton-meters (metric), or foot-pounds (British)
m = Mass of object in kilograms (metric) or slugs (British)
v = Velocity of mass in meters per second (metric) or feet per second (British)
• A boulder with a weight of 500 Newtons (gravitational force) resting at the top of a 30 meter
hill possesses 15000 Newton-meters (or 15000 Joules) of potential energy
• A bullet having a mass of 2.9 grams (0.0029 kilograms) traveling at a velocity of 1100 meters
per second possesses 1754.5 Newton-meters (or 1754.5 Joules) of kinetic energy
Energy may freely exchange between these forms, and when such an exchange takes place the
sum total of energy remains constant due to Conservation. For example, if that bullet were fired
straight up4 it would leave the gun barrel with 1754.5 Joules of kinetic energy, which would convert
to potential energy as it climbed to maximum altitude and minimum velocity, then would convert
back into kinetic energy as the bullet lost altitude and fell toward Earth. Air friction during its
ascent would result in the bullet attaining less than 1754.5 Joules of potential energy at its apogee,
and air friction while falling would result in even less kinetic energy when it reached ground level
again. However, total energy would remain the same: 1754.5 Joules total in kinetic energy plus
potential energy plus thermal and acoustic energy5 in the air disturbed by the bullet’s flight.
Power is often confused with energy, but it means something different. Whereas “energy” is that
which sets matter into motion, “power” is defined as the rate at which energy gets exchanged from
one form to another. In the metric system the Newton-meter or Joule is used to measure energy,
but the Watt is the unit for power, with one Watt being equal to one Joule (or Newton-meter)
per second. Whereas a measurement of energy tells you how much work that energy could do, a
measurement of power tells you how rapidly work can get done.
Full Tutorial
Physics is really the study of matter and energy, and of the laws describing the interactions of both.
The movement of matter largely follows the laws described by Isaac Newton (so much so that this
category of physics is often referred to as Newtonian physics). The flow and transformation of energy
is a category called thermodynamics. In a sense, this Tutorial may be thought of as an extremely
condensed lesson in physics.
It may also be thought of as an extremely condensed lesson in chemistry, the study of how
different types of atoms interact with each other. Matter consists of atoms, but atoms rarely exist in
isolation. Usually atoms are found bonded to each other to form larger collections of matter called
molecules.
11
12 CHAPTER 3. FULL TUTORIAL
In everyday scenarios we find matter existing in the form of very large numbers of sub-microscopic
atoms and molecules. Common phases of matter include solid, liquid, and gas. Solids are materials
where the molecules are more or less locked together into a rigid form. In liquids, the molecules are
not locked together per se, but exist adjacent to one another while retaining individual mobility.
Gases consist mostly of empty space interspersed with molecules that are entirely mobile.
1 A good demonstration of this fact would be to attempt to shake a massive object such as an anvil if brought
about an orbiting space station or other spacecraft. Though weightless, that object would still resist being shaken,
and the more mass it has the more resistance it would offer to a shaking force.
3.1. MATTER AND ATOMIC STRUCTURE 13
2 In order for a wave of light to be influenced at all by an object, that object must be at least the size of the wave’s
length. To use an analogy with water waves, it would be comparing the interaction of a water wave on a beach against
a large rock (a disturbance in the wave pattern) versus the non-disturbance of that same wave as it encounters a small
buoy.
14 CHAPTER 3. FULL TUTORIAL
A tremendously simplified model of a common nitrogen atom is shown here, with 7 protons and
7 neutrons in the nucleus, and 7 electrons in “orbit” around the nucleus:
N
P N
P
N P
e P N P e
P N
P
N N
e
e
e
The atomic number of this atom (the number of protons in the nucleus) is seven, which is what
defines it as nitrogen. The atomic mass of this atom (the sum of protons and neutrons in the
nucleus) is fourteen. The chemical symbol for this atom is shown here:
14
Atomic number
7
N
(number of protons)
The atomic number is redundant to the letter “N” for nitrogen, since only the element nitrogen
can have an atomic number of seven. The atomic mass is only relevant when we need to distinguish
one isotope of nitrogen from another (variations of elements having the same number of protons but
different numbers of neutrons), and this is seldom because the chemical properties of isotopes are
identical – only their masses differ. For these reasons, you will usually find no left-hand subscripts
or superscripts placed near chemical symbols of elements in chemical expressions.
3.1. MATTER AND ATOMIC STRUCTURE 15
An N2 molecule
e
e e
Chemical symbol
e
N N
P N P N
N
P P
N P N P
e P N P e e e P N P e
P N P N
P P
N N N N
2 e e
Number of atoms
bound together
in a molecule
e e
e
An N3− ion is an atom of nitrogen having three more electrons than it normally would when
electrically balanced:
An N 3- ion
e
Chemical symbol e e
Ionic charge
(- means extra electrons,
+ means missing electrons)
3- e
P
N
N
N
P
N P
e P N P e
P N
P
N N
e
e
e
C2 H6 O
This is called a molecular formula, because it shows the proportions of atom types comprising
each molecule.
A more common way to write the formula for ethanol, though, is this:
C2 H5 OH
Here, an attempt is made to show the physical structure of the ethanol molecule, where one of
the hydrogen atoms is located further away from the others. This is called a structural formula.
If more detail of the bonds between atoms in a molecule is needed, a semi-graphic representation
called a displayed formula (also known as an expanded structural formula) may be used in lieu of a
structural formula:
H H
H C C O H
H H
Each letter in a displayed formula represents a single atom within that molecule, and each line
segment in a displayed formula represents a bond between two.
3.1. MATTER AND ATOMIC STRUCTURE 17
In organic chemistry – the study of molecules principally centered around carbon atoms – a
special type of notation is used to show the structural detail of the molecule with fewer lines and
letters than a displayed formula. This notation is called a line drawing, where each line segment
represents a single electron bond3 to a carbon atom, each vertex and line-end represents the location
of a carbon atom, and any hydrogen atoms directly bound to a carbon atom are simply omitted
for simplicity. Compare and contrast these displayed formulae and line drawings for a few different
organic compounds:
H C H
C C
Benzene (C6H6)
C C
H C H
O
H O
C
H C O O
γ-valerolactone
H C C H
H C
H
H H
An important principle in organic chemistry is that carbon atoms prefer to form exactly four
bonds with surrounding atoms4 . This fact is exploited in line-drawing notation where any bonds
not explicitly shown at a vertex are assumed to be single-bonds with hydrogen atoms, enough of
them to bring the total number of bonds with that carbon atom to four. Since a great many organic
compounds are principally comprised of carbon and hydrogen, the line-drawing symbols for these
molecules tend to be more lines than letters.
3 One line represents a single bond, which is one electron shared per bound atom. Two parallel lines represent a
double bond, where each carbon atom shares two of its valence electrons with the neighboring atom. Three parallel
lines represent a triple bond, where each atom shares three of its outer electrons with the neighboring atom.
4 Incidentally, nitrogen atoms preferentially form exactly three bonds, and oxygen atoms exactly two bonds. The
reason for this pattern is the particular patterns of electrons orbiting each of these atoms, and their respective energy
levels. For more information on this, see section 3.2 beginning on page 19.
18 CHAPTER 3. FULL TUTORIAL
Chemical engineers often perform mass and energy balance calculations for processes where
mixtures of similar compounds exist. Wastewater treatment is one example, where an array of
organic compounds must all be treated through oxidation (chemical reaction with oxygen). In such
cases, it is common for chemical engineers to write formulae expressing the average ratios of elements,
so that they may calculate the quantity of reactant(s) needed to complete the desired chemical
reaction with compounds in the mixture. Primary sludge clarified from municipal wastewater, for
example, may be represented by the compositional formula C22 H39 O10 N. This does not suggest the
existence of some monstrous molecule consisting of twenty-two carbon atoms, thirty-nine hydrogen
atoms, ten oxygen atoms, and a lone nitrogen atom somewhere in a sample of sludge, but rather
that the combined average carbon, hydrogen, oxygen, and nitrogen quantities in that sludge exist in
a variety of molecular forms in these approximate proportions. This aggregate formula expression
helps the engineer quantify the gross chemical characteristics of the sludge, and from that determine
how much oxygen will be necessary to completely oxidize it.
Sometimes, compositional formulae are written with non-integer subscripts. An example of this
would be the compositional formula C4.8 H8.4 O2.2 , which also happens to be an average composition
for municipal wastewater sludge (ignoring nitrogen content). The same formula could just as well
have been written C48 H84 O22 , or even C24 H42 O11 , because these subscript values all express the
exact same proportions.
3.2. ELECTRONIC STRUCTURE 19
Electrons situate themselves around the nucleus of any atom according to one basic rule: the
minimization of potential energy. That is, the electrons seek the lowest-energy positions available
around the nucleus. Given the electrostatic attraction between negative electrons and the positive
nucleus of an atom, there is potential energy stored in the “elevation” between an orbiting electron
and the nucleus, just as there is gravitational potential energy in any object orbiting a planet.
Electrons lose energy as they situate themselves closer to the nucleus, and it requires an external
input of energy to move an electron farther away from its parent nucleus.
In a sense, most of chemistry may be explained by this principle of minimized potential energy.
Electrons “want” to “fall” as close as they can to the positively-charged nucleus. However, there is
limited “seating” around the nucleus. As described by Pauli’s Exclusion Principle, electrons cannot
simply pile on top of each other in their quest for minimum energy, but rather must occupy certain
regions of space allowed by their unique quantum states.
An analogy6 for visualizing this is to picture an atom as if it were an amphitheater, with the
stage being the nucleus and the concentric array of seats being places where electrons may reside.
All spectators (electrons) desire to be as close to the stage (nucleus) in an amphitheater (atom)
as possible, but since everyone cannot occupy the best seat, people are forced to choose seats at
5 These orbitals just happen to be the 1s, 2p, 3d, and 4f orbitals, as viewed from left to right. In each case, the
nucleus lies at the geometric center of each shape. In a real atom, all orbitals share the same center, which means
any atom having more than two electrons (that’s all elements except for hydrogen and helium!) will have multiple
orbitals around one nucleus. This four-set of orbital visualizations shows what some orbitals would look like if viewed
in isolation.
6 Please understand that like all analogies, this one merely illustrates a complex concept in terms that are easier to
recognize. Analogies do not explain why things work, but merely liken an abstract phenomenon to something more
accessible to common experience.
20 CHAPTER 3. FULL TUTORIAL
different positions around the stage. As a result, the inner seats fill first, with most empty seats
being furthest away from the stage. The concept of energy fits neatly into this analogy as well: just
as electrons give up energy to “fall into” lower-energy regions around the nucleus, people must give
up money to purchase seats closest to the action on stage.
The energy levels available for orbiting electrons are divided into categories of shells and subshells.
A “shell” (or, principal quantum number, n) describes the main energy level of an electron. In our
amphitheater analogy, this is equivalent to a tier or seating level. A “subshell” (or, subsidiary
quantum number, l) further divides the energy levels within each electron shell, and assigns different
shapes to the electrons’ probability “clouds.” In the amphitheater analogy, a subshell would be a
row of seats within a particular tier. To make the analogy accurate, we would have to imagine each
row of seats in a tier having a different shape (not all arcs or straight lines), with varying degrees
of viewing comfort afforded by each shape. The first row in each tier faces uniformly toward the
stage, allowing easy viewing. Successive rows (subshells) in each tier (shell) contain more seats, but
are bent in such a way that the stage is not as easy to view, making these rows less desirable to
occupy. Electron subshells always have an even-numbered electron capacity, analogous to theater
rows containing even-numbered quantities of seats, because atomic electrons tend to gather in pairs
called orbitals.
Chemists identify electron shells both by number (the value of the quantum number n) and/or
by capital letters: the first shell by the letter K, the second by L, the third by M, and the fourth
by N. Higher-order shells exist for atoms requiring7 many electrons (high atomic number), and the
lettering pattern is alphabetic (fifth shell is O, sixth is P, etc.). Each successive shell has a greater
number of subshells available, like successive amphitheater tiers having more rows: the low-level
tiers closest to the stage having the fewest rows, and the high-level tiers furthest from the stage
having the most rows.
A numbering and lettering system is also used by chemists to identify subshells within each shell
(the l quantum number value starting with zero, and lower-case letters beginning with “s”): the
first subshell (l = 0) in any shell represented by the letter s, the second (l = 1) by p, the third
(l = 2) by d, the fourth (l = 3) by f, and all others by successive lower-case letters of the alphabet8 .
Each subshell of each shell has an even-numbered capacity for electrons, since the electrons in each
subshell are organized in “orbital” regions, each orbital handling a maximum of two9 electrons. The
number of orbitals per shell is equal to twice the l value plus one. An “s” subshell has one orbital
holding up to two electrons. A “p” subshell has three orbitals holding up to six electrons total. A
“d” subshell has five orbitals holding up to ten electrons total. An “f” subshell has seven orbitals
holding up to 14 electrons total. A “g” subshell has nine orbitals holding up to 18 electrons total.
The number of subshells in any shell is the same as that shell’s n value. Thus, the first (K) shell
has only one subshell, “s”. The second (L) shell has two subshells, an “s” and a “p”. The third (M)
shell has three subshells available, an “s”, a “p”, and a “d”; and so on.
7 Truth be told, higher-order shells exist even in simple atoms like hydrogen, but are simply not occupied by that
atom’s electron(s) unless they are “excited” into a higher energy state by an external input of energy.
8 The letters s, p, d, and f refer to the words sharp, principal, diffuse, and fundamental, used to describe the
appearance of spectral lines in the early days of atomic spectroscopy research. Higher-order subshells are labeled
alphabetically after f : g, h, and i.
9 The two electrons of any orbital have opposite spin values.
3.2. ELECTRONIC STRUCTURE 21
This table shows the first few shells, their subshells, and electron capacities of each:
n=1;K l=0;s 2 × (2 × 0 + 1) = 2
n=2;L l=0;s 2 × (2 × 0 + 1) = 2
l=1;p 2 × (2 × 1 + 1) = 6
n=3;M l=0;s 2 × (2 × 0 + 1) = 2
l=1;p 2 × (2 × 1 + 1) = 6
l=2;d 2 × (2 × 2 + 1) = 10
n=4;N l=0;s 2 × (2 × 0 + 1) = 2
l=1;p 2 × (2 × 1 + 1) = 6
l=2;d 2 × (2 × 2 + 1) = 10
l=3;f 2 × (2 × 3 + 1) = 14
Reviewing our amphitheater analogy, atomic shells are like seating tiers (levels), subshells are
like rows of seats within each tier, and subshell electron capacity is like the number of seats in each
row. This simple illustration shows an atom with three shells (K, L, and M) with the respective
subshells (s, p, and d) represented by differently-shaded rings within each shell, having different
numbers of places for electrons within each one:
d
p
s
M
p
s L
s
K
Nucleus
This illustration is vastly over-simplified, failing to show the diverse shapes of each subshell,
serving only to show how each successive shell grows in subshells and electron capacities.
22 CHAPTER 3. FULL TUTORIAL
The complete electron configuration for an atom may be expressed using spectroscopic notation,
showing the shell numbers, subshell letters, and number of electrons residing within each subshell as
a superscript. For example, the element helium (with an atomic number of 2) would be expressed
as 1s2 , with just two electrons in the “s” subshell of the first shell. The following table shows the
electron structures of the first nineteen elements in the periodic table, from the element hydrogen
(atomic number10 = 1) to potassium (atomic number = 19):
In order to avoid having to write unwieldy spectroscopic descriptions of each element’s electron
structure, it is customary to write the notation only for subshells that are unfilled. For example,
instead of writing the electron structure of the element aluminum11 as 1s2 2s2 2p6 3s2 3p1 , we might
just as well write a condensed version showing only the last subshell (3p1 ), since all the previous
subshells are completely full.
Another way to abbreviate the spectroscopic notation for elements is to condense all the shells
below the newest (unfilled) shell as the corresponding noble element, in brackets. To use the example
of aluminum again, we could write its spectroscopic notation as [Ne]3s2 3p1 since its shell 1 and shell
2 configurations are completely described by the electron configuration of neon.
10 The atomic number is the quantity of protons found in an atom’s nucleus, and may only be a whole number.
Since any electrically balanced atom will have the same number of electrons as protons, we may look at the atomic
number of an element as being the number of electrons in each atom of that element.
11 Building on the amphitheater analogy for one atom of the element aluminum, we could say that there are two
electrons occupying the “s” seating row on the first level, plus two electrons occupying the “s” seating row on the
second level, plus six electrons occupying the “p” seating row on the second level, plus two electrons occupying the
“s” seating row on the third level, plus one electron occupying the “p” seating row on the third level.
3.2. ELECTRONIC STRUCTURE 23
Re-writing our electron shell table for the first nineteen elements using this condensed notation:
If we progress from element to element in increasing atomic number, we see that no new shell
begins to form until after we reach the noble element for that period12 at the far right-hand column.
With the beginning of each new period at the far-left end of the Table, we see the beginning of
the next higher-order electron shell. The shell(s) below are represented by whichever noble element
shares that same configuration13 , indicating a “noble core” of electrons residing in extremely stable
(low-energy) regions around the nucleus.
12 Recall the definition of a “period” in the Periodic Table being a horizontal row, with each vertical column being
called a “group”.
13 Building on the amphitheater analogy once again for one atom of the element aluminum, we could say that all
seats within levels 1 and 2 are occupied (just like an atom of neon), plus two electrons occupying the “s” seating row
on the third level, plus one electron occupying the “p” seating row on the third level.
24 CHAPTER 3. FULL TUTORIAL
The beginning of the next higher-order shell is what accounts for the periodic cycle of ionization
energies we see in elements of progressing atomic number. The first electron to take residence in a
new shell is very easy to remove, unlike the electrons residing in the “noble” configuration shell(s)
below:
Not only is the “noble core” notation convenient for tersely describing the electron structure of an
element, but it also reveals an important concept in chemistry: the idea of valence. Electrons residing
in lower-order shells are, by definition, at lower energy states than electrons residing in higher-order
shells and are therefore much more difficult to dislodge. Therefore, the electrons in unfilled shells,
being easier to dislodge, play a far greater role in chemical bonds than electrons residing in filled
shells below. These “outer” electrons are called valence electrons, and their number determines how
readily an atom will chemically interact with another atom. This is why elements found in the
same group (having similar outer-electron configurations) bear similar chemical characteristics: the
electrons lying below in the “noble core” configurations have little effect on how the atom will bond
with other atoms. A lithium atom, with its outer-most electron configuration being 2s1 , reacts in
much the same way as an atom of sodium having an outer-most configuration of 3s1 , and much the
same as a potassium atom having an outer-most configuration of 4s1 , all because those outer-shell
electrons are the most available for interaction with electrons of other atoms.
3.2. ELECTRONIC STRUCTURE 25
If we examine the electron structures of atoms with successively greater atomic numbers (more
protons in the nucleus, therefore more electrons in orbit to balance the electrical charge), we notice
that the shells and subshells fill up in an interesting pattern. One might think that all the lower-
order shells get completely filled before any electrons go into a higher-order shell – just as we might
expect people to fill every seat in all the lower tiers of an amphitheater before filling seats in any
of the higher tiers – but this is not always the case. Instead, the energy levels of subshells within
shells is split, such that certain subshells within a higher shell will have a lower energy value than
certain subshells within a lower shell. Referring back to our amphitheater analogy, where seating
tiers represented shells and seat rows of various shape represented subshells, it is as though people
choose to fill the more comfortable rows in higher-level tiers before sitting in the less-comfortable
rows in lowest available tiers, the desire for comfort trumping the desire for proximity to the stage.
A rule commonly taught in introductory chemistry courses called the Madelung rule (also referred
to as Aufbau order, after the German verb aufbauen meaning “to build up”) is that subshells fill
with increasing atomic number in such an order that the subshell with the lowest n + l value, in the
lowest shell, gets filled before any others.
The following graphic illustrates this ordering:
8s
7p 6d 5f
7s
6p 5d 4f
6s
5p 4d
Energy level
5s
4p 3d
4s
3p
3s
2p
2s
1s
It should be noted that exceptions exist for this rule. We see one of those exceptions with
the element chromium (24 Cr). Strictly following the Madelung rule in progressing from vanadium
(atomic number = 23, valence electron structure 3d3 4s2 ) to chromium (atomic number = 24), we
would expect the next electron to take residence in the “3d” subshell making chromium’s valence
structure be 3d4 4s2 , but instead we find two more electrons residing in chromium’s 3d subshell with
one less in the 4s subshell (3d5 4s1 ). The sequence resumes its expected progression with the next
element, manganese (atomic number = 25, valence electron structure 3d5 4s2 ). The general principle
26 CHAPTER 3. FULL TUTORIAL
of energy minimization still holds true . . . it’s just that the relative energies of succeeding subshells
do not follow a simple rule structure. In other words, the Aufbau order is an over-simplified view of
reality. To use the amphitheater analogy again, it’s as if someone gave up one of the nice chairs in
tier 4 to sit next to a friend who just occupied one of the less comfortable chairs in tier 3.
The practical importance of electron configurations in chemistry is the potential energy possessed
by electrons as they reside in different shells and subshells. This is extremely important in the
formation and breaking of chemical bonds, which occur due to the interaction of electrons between
two or more atoms. A chemical bond occurs between atoms when the outer-most (valence) electrons
of those atoms mutually arrange themselves in energy states that are collectively lower than they
would be individually. The ability for different atoms to join in chemical bonds completely depends
upon the default energy states of electrons in each atom, as well as the next available energy states
in the other atoms. Atoms will form stable bonds only if the union allows electrons to assume stable
(low-energy) levels. This is why different elements are very selective regarding which elements they
will chemically bond with to form compounds: not all combinations of atoms result in favorable
potential energy levels.
The amount of energy required to break a chemical bond (i.e. separate the atoms from each other)
is the same amount of energy required to restore the atoms’ electrons to their previous (default)
states before they joined. This is the same amount of energy released by the atoms as they come
together to form the bond. Thus, we see the foundation of the general principle in chemistry that
forming chemical bonds releases energy, while breaking chemical bonds requires an input of energy
from an external source. We also see in this fact an expression of the Conservation of Energy:
the amount of energy “invested” in breaking bonds is precisely the same as the amount of energy
“returned” when those same bonds re-form.
In summary, the whole of chemistry is a consequence of electrons not being able to assume
arbitrary positions around the nucleus of an atom. Instead, they seek the lowest possible energy
levels within a framework allowing them to retain unique quantum states. Atoms with mutually
agreeable electron structures readily bond together to form molecules, and they release energy in
the process of joining. Molecules may be broken up into their constituent atoms, if sufficient energy
is applied to overcome the bond. Atoms with incompatible electron structures do not form stable
bonds with each other.
3.3. WORK, ENERGY, AND POWER 27
W = F~ · ~x
Where,
W = Work, in newton-meters (metric) or foot-pounds (British)
F~ = Force vector (force and direction exerted) doing the work, in Newtons (metric) or pounds
(British)
~x = Displacement vector (distance and direction traveled) over which the work was done, in
meters (metric) or feet (British)
14 A common definition of energy is the “ability to do work” which is not always true. There are some forms of
energy which may not be harnessed to do work, such as the thermal motion of molecules in an environment where all
objects are at the same temperature. Energy that has the ability to do work is more specifically referred to as exergy.
While energy is always conserved (i.e. never lost, never gained), exergy is a quantity that can never be gained but
can be lost. The inevitable loss of exergy is closely related to the concept of entropy, where energy naturally diffuses
into less useful (more homogeneous) forms over time. This important concept explains why no machine can never be
perfectly (100.0%) efficient, among other things.
15 A vector is a mathematical quantity possessing both a magnitude and a direction. Force (F ), displacement (x),
and velocity (v) are all vector quantities. Some physical quantities such as temperature (T ), work (W ), and energy
(E) possess magnitude but no direction. We call these directionless quantities “scalar.” It would make no sense at
all to speak of a temperature being “79 degrees Celsius due North” whereas it would make sense to speak of a force
being “79 Newtons due North”. Physicists commonly use a little arrow symbol over the variable letter to represent
that variable as a vector, when both magnitude and direction matter. Thus F ~ represents a force vector with both
magnitude and direction specified, while plain F merely represents the magnitude of that force without a specified
direction. A “dot-product” is one way in which vectors may be multiplied, and the result of a dot-product is always
a scalar quantity.
28 CHAPTER 3. FULL TUTORIAL
W = F x cos θ
When the two vectors F~ and ~x point the same direction, the angle θ between them is zero and
therefore W = F x because cos 0o = 1. When the two vectors point in opposite directions, the angle
between them is 180o and therefore W = −F x because cos 180o = −1.
This means if a force acts in the same direction as a motion (i.e. force and displacement vectors
pointing in the same direction), the work being done by that force will be a positive quantity. If
a force acts in the opposite direction of a motion (i.e. force and displacement vectors pointing
in opposite directions), the work done by that force will be a negative quantity. If a force acts
perpendicularly to the direction of a motion (i.e. force and displacement vectors at right angles to
each other), that force will do zero work.
Displacement
x Work (W) done by force is positive
Force W = Fx cos 0o
Object F W = (Fx) (1)
Displacement
x Work (W) done by force is zero
W = Fx cos 90o
Object Force
F W = (Fx) (0)
Displacement
x Work (W) done by force is negative
Force W = Fx cos 180o
Object F W = (Fx) (-1)
3.3. WORK, ENERGY, AND POWER 29
Illustrations are helpful in explaining these concepts. Consider a crane hoisting a 2380 pound
weight 15 feet up into the air by means of an attached cable:
Crane does work on weight Wcrane = (2380 lb)(15 ft)(cos 0o) = +35700 ft-lbs
Work is done on the weight Wweight = (2380 lb)(15 ft)(cos 180o) = -35700 ft-lbs
The amount of work done from the crane’s perspective is Wcrane = +35700 ft-lbs, since the
crane’s force (Fcrane = 2380 lbs, up) points in the same upward direction as the cable’s motion (x =
15 ft, up) and therefore there is no angular difference between the crane force and motion vectors
(i.e. θ = 0). The amount of work done from weight’s perspective, however, is Wweight = −35700
ft-lbs because the weight’s force vector (Fweight = 2380 lbs, down) points in the opposite direction
as the cable’s motion vector (x = 15 feet, up), yielding an angular difference of θ = 180o . Another
way of expressing these two work values is to state the crane’s work in the active voice and the
weight’s work in the passive voice: the crane did 35700 ft-lbs of work, while 35700 ft-lbs of work
was done on the weight. This language is truly appropriate, as the crane is indeed the active agent
in this scenario, while the weight passively opposes the crane’s efforts. In other words, the crane is
the motive source of the work, while the weight is a load.
30 CHAPTER 3. FULL TUTORIAL
Power applies in this scenario to how quickly the weight rises. So far, all we know about the
weight’s lifting is that it took 35700 ft-lbs of energy to do that work. If we knew how long it took
the crane to do that work, we could calculate the crane’s power output. For example, a crane with
a power rating of 35700 ft-lbs per second could complete this 15-foot lift in only one second of time.
Likewise, a crane with a power rating of only 3570 ft-lbs per second would require 10 seconds of
time to execute the same lift.
3.3. WORK, ENERGY, AND POWER 31
x = 15 ft
Weight
2380 lbs
Fweight = 2380 lbs Crane
No work done on or by the crane Wcrane = (2380 lb)(15 ft)(cos 90o) = 0 ft-lbs
No work done on or by the weight Wweight = (2380 lb)(15 ft)(cos -90o) = 0 ft-lbs
It should be noted that the crane’s engine will do work as it overcomes rolling friction in the
wheels to move the crane along, but this is not work done on or by the hoisted weight. When we
calculate work – as with all other calculations in physics – we must be very careful to keep in mind
where the calculation(s) apply in the scenario. Here, the forces and displacement with regard to the
hoisted weight are perpendicular to each other, and therefore no work is being done there. The only
work done anywhere in this system as the crane rolls 15 feet horizontally involves the horizontal
force required to roll the crane, which is unspecified in this illustration.
Similarly, there is no transfer of energy to or from the hoisted weight while the crane rolls along.
Whatever energy comes through the crane’s engine only goes into overcoming rolling friction at the
wheels and ground, not to do any work with the weight itself. Generally this will be a very small
amount of energy compared to the energy required to hoist a heavy load.
32 CHAPTER 3. FULL TUTORIAL
x = 15 ft
Work is done on the crane Wcrane = (2380 lb)(15 ft)(cos 180o) = -35700 ft-lbs
Weight does work on crane Wweight = (2380 lb)(15 ft)(cos 0o) = +35700 ft-lbs
Notice how the weight is now the actively-working object in the system, doing work on the
crane. The crane is now the passive element, opposing the work being done by the weight. Both the
crane and the weight are still pulling the same directions as before (crane pulling up, weight pulling
down), but now the direction of displacement is going down which means the weight is “winning”
and therefore doing the work, while the crane is “losing” and opposing the work.
If we examine what is happening inside the crane as the weight descends, we see that energy is
being transferred from the descending weight to the crane. In most cranes, the descent of a load is
controlled by a brake mechanism, regulating the speed of descent by applying friction to the cable’s
motion. This brake friction generates a great deal of heat, which is a form of energy transfer: energy
stored in the elevated weight is now being converted into heat which exits the crane in the form of
hot air (air whose molecules are now vibrating at a faster speed than they were at their previous
temperature). If the crane is electric, we have the option of regenerative braking where we recapture
this energy instead of dissipating it in the form of heat. This works by switching the crane’s electric
motor into an electric generator on demand, so the weight’s descent turns the motor/generator shaft
to generate electricity to re-charge the crane’s battery or be injected back into the electric power
grid to do useful work elsewhere.
3.3. WORK, ENERGY, AND POWER 33
Referring back to the illustration of the crane hoisting the weight, it is clear that the weight
stored energy while it was being lifted up by the crane, and released this energy back to the crane
while it was being lowered down to the ground. The energy held by the elevated weight may therefore
be characterized as potential energy, since it had the potential to do work even if no work was being
done by that energy at that moment.
A special version of the general work formula W = F~ · ~x exists for calculating this gravitational
potential energy. Rather than express force and displacement as vectors with arbitrary directions,
we express the weight of the object as the product of its mass and the acceleration of gravity
(F = mg) and the vertical displacement of the object simply as its height above the ground (x = h).
The amount of potential energy stored in the lift is simply equal to the work done during the lift
(W = EP ):
Ep = mgh
Where,
Ep = Gravitational potential energy in newton-meters (metric) or foot-pounds (British)
m = Mass of object in kilograms (metric) or slugs (British)
g = Acceleration of gravity in meters per second squared (metric) or feet per second squared
(British)
h = Height of lift in meters (metric) or feet (British)
There is no need for vectors or cosine functions in the Ep = mgh formula, because gravity and
height are always guaranteed to act along the same axis. A positive height (i.e. above ground level)
is assumed to yield a positive potential energy value for the elevated mass.
Many different forms of potential energy exist, although the standard “textbook” example is of
a mass lifted to some height against the force of gravity. Compressed (or stretched) springs have
potential energy, as do pressurized fluids, chemical bonds (e.g. fuel molecules prior to combustion),
hot masses, electrically-charged capacitors, and magnetized inductors. Any form of energy with the
potential to be released into a different form at some later time is, by definition, potential energy.
34 CHAPTER 3. FULL TUTORIAL
Tow rope
. . . (to load) . . .
Ftruck = 1200 lb Fload = 1200 lb
Hitch
1.5 ft
Road
Clearly, the truck does work on the load by exerting a pulling force in the direction of motion.
If we wish to quantify this work, we may consider the work done by the truck as it tows the load a
distance of 40 feet:
Wtruck = F x cos θ
Wload = F x cos θ
In order to quantify this amount of work from the perspective of the truck’s rotating wheel, we
must cast the variables of pulling force and pulling distance into rotary terms. We will begin by first
examining the distance traveled by the wheel. Any circular wheel has a radius, and the wheel must
turn a certain number of revolutions in order to pull the load any distance. The obvious function
of a wheel is to convert between linear and rotary motion, and the common measure between these
two motions is the circumference of the wheel: each revolution of the wheel equates directly to one
circumference’s worth of linear travel. In our truck example, the wheel has a radius of 1.5 feet,
which means it must have a circumference of 9.425 feet (i.e. 2 × π × 1.5 feet, since C = πD = 2πr).
Therefore, in order to travel 40 linear feet this wheel must rotate 4.244 times (i.e. 40 feet ÷ 9.425
feet/revolution = 4.244 revolutions).
Next, we must relate pulling force to rotational force. The 1200 pounds of pulling force exerted
by the wheel16 originates from the twisting force exerted by the axle at the wheel’s center. The
technical term for this twisting force is torque (symbolized by the Greek letter “tau”, τ ), and it is a
function of both the linear pulling force and the wheel’s radius:
τ = rF
Solving for torque (τ ) in this application, we calculate 1800 pound-feet given the wheel’s linear
pulling force of 1200 pounds and a radius of 1.5 feet. Please note that the unit of “pound-feet” for
torque is not the same as the unit of “foot-pounds” for work. Work is the product of force and
displacement (distance of motion), while torque is the product of force and radius length. Work
requires motion, while torque does not17 .
The torque value of 1800 lb-ft and the turning of 4.244 revolutions should be sufficient to calculate
work done by the wheel, since torque equates directly to pulling force and the number of revolutions
equates directly to pulling distance. It should be obvious that the product of 1800 lb-ft and 4.244
revolutions does not, however, equal 48000 ft-lbs of work, and so our torque-revolutions-work formula
must contain an additional multiplication factor k:
W = kτ x
Substituting 48000 ft-lbs for work (W ), 1800 lb-ft for torque (τ ), and 4.244 for the number of
revolutions (x), we find that k must be equal to 6.283, or 2π. Knowing the value for k, we may
re-write our rotational work formula more precisely:
W = 2πτ x
This formula is correct no matter the wheel’s size. A larger-radius wheel will certainly travel
farther for each revolution, but that larger radius will proportionately reduce the pulling force for
any given amount of torque, resulting in an unchanged work value.
16 Note that this calculation will assume all the work of towing this load is being performed by a single wheel on
the truck. Most likely this will not be the case, as most towing vehicles have multiple driving wheels (at least two).
However, we will perform calculations for a single wheel in order to simplify the problem.
17 Consider the example of applying torque to a stubbornly seized bolt using a wrench: the force applied to the
wrench multiplied by the radius length from the bolt’s center to the perpendicular line of force yields torque, but
absolutely no work is done on the bolt until the bolt begins to move (turn).
36 CHAPTER 3. FULL TUTORIAL
Applying this work formula to another application, let us consider an electric winch where a
load is lifted against the force of gravity. A “winch” is a mechanism comprised of a tubular drum
and cable, the cable wrapping or unwrapping around the drum as the drum is turned by an electric
motor. For comparison we will use the same weight and lifting distance of the crane example, merely
replacing the crane with a winch located atop a tower. In this case we will make the winch drum 3
inches in diameter, and calculate both the torque required to rotate the drum as well as the number
of rotations necessary to lift the weight 15 feet:
cable
Tower
x = 15 ft
Weight
2380 lbs
From the crane example we already know the amount of work which must be done on the weight
to hoist it 15 feet vertically: 35700 ft-lbs. Our rotational work formula (W = 2πτ x) contains two
unknowns at this point in time, since we know the value of W but not τ or x. This means we cannot
yet solve for either τ or x using this formula. If we also knew the value of τ we could solve for x
and vice-versa, which means we must find some other way to solve for one of those unknowns.
A drum radius of 3 inches is equivalent to 0.25 feet, and since we know the relationship between
radius, force, and torque we may solve for torque in that manner:
τ = rF
τ = 595 lb-ft
Now that we have a value for τ we may substitute it into the rotational work formula to solve
for the number of necessary drum rotations:
W = 2πτ x
W
x=
2πτ
35700 ft-lb
x=
(2π)(595 lb-ft)
x = 9.549 revolutions
3.3. WORK, ENERGY, AND POWER 37
As always, energy is conserved in this electric winch system just as it is in any other system.
The 35700 foot-pounds of energy invested in the gravitational potential energy of the weight had
to originate from somewhere, which in the case of an electric winch is the electrical power source
feeding the winch motor. If that winch motor were allowed to turn in reverse and act as a generator,
it would convert the descending weight’s loss in potential energy into electrical power to be returned
to the source, whether that source by a rechargeable battery or an electrical power “grid” with other
electrical loads that may use that energy.
The same, of course, is true for the tow truck example. The energy expended in towing the
load must come from somewhere, and in the case of a combustion-type truck engine that source is
the fuel powering the engine. For electric vehicles, the energy source is a rechargeable battery, and
the reversible nature of electric motors means the vehicle may use its drive motor as a generator
to “brake” (slow down), returning the vehicle’s kinetic energy into electrical energy to recharge the
battery for later use. This is why all-electric and hybrid-electric vehicles are remarkably efficient
in stop-and-go traffic conditions: they utilize their drive motors as regenerative brakes to recover
the vehicle’s kinetic energy rather than dissipating that same energy in the form of heat using
friction-based brake mechanisms.
Many applications exist within the industrial world for rotational work and electric motors.
Conveyor belts, pumps, fans, compressors, and a host of other mechanisms may be powered via
electric motors, the speed and torque of those electric motors controlled using electronic circuits
called motor drives. Alternating-current (AC) electric motors are controlled by electronic devices
called Variable Frequency Drives, or VFDs, which achieve precise speed control by varying the
frequency of the AC power feeding the motor, and achieve precise torque control by varying the
voltage and current feeding the motor. VFDs are very important “final control” devices used in a
wide variety of industrial control systems.
To 480 VAC
3-phase
power source
AC induction motor
Stop L1 L2 L3
Fwd
Signals applied to VFD
inputs control the motor’s Rvs
starting, stopping, speed, Fwd jog
and torque. Com VFD
Modbus
RS-485
Analog T1 T2 T3
speed The VFD outputs electricity
command
at varying voltage, current,
and AC frequency values to
exert the desired control over
the motor.
Sophisticated VFDs control regeneration as well (i.e. using the motor as a generator to “brake”
a rotating machine), allowing energy to be extracted from the mechanism.
38 CHAPTER 3. FULL TUTORIAL
Multi-lock device
5
goes here 3
4
2
1
Procedures created and maintained at the worksite will identify the energy-flow devices in need
of securing prior to commencement of work on a piece of equipment. These procedures are literally
life-saving documents, as they ensure no energy-securing device is overlooked by personnel doing
work on the equipment or system.
3.3. WORK, ENERGY, AND POWER 39
Note the particular lock-out steps required in this procedure: switching the control mode to
the “off” position and tagging it, closing the fuel gas valve supplying fuel to the engine and
locking/tagging it, and finally closing the valve supplying high-pressure air for engine starting and
locking/tagging it. The closure of the starting air valve prevents the engine from being pneumatically
turned while personnel are performing work on it. The closure of the fuel gas valve eliminates
hazards resulting from the pressure of the fuel gas as well as its chemical energy (i.e. fire hazard)
and/or biological threats (poisoning or asphyxiation). Note also how this procedure lists steps of
notification to be taken prior to locking or tagging anything on the engine, as well as any other
procedures possibly necessary (e.g. inspecting the maintenance crane if that will be needed for work
on the engine).
40 CHAPTER 3. FULL TUTORIAL
The following is a set of incomplete lists of various energy-securing devices and energy sources
which may be “locked out” for maintenance on a larger system:
Electrical energy
Mechanical energy
• Mechanical clutch (disengaged, and locked in that position) to prevent prime mover from
moving something
• Mechanical coupling (disassembled, and locked in that state) to prevent prime mover from
moving something
• Mechanical locking pin (inserted, and held in position by a padlock) to prevent motion
• Raised masses lowered to ground level, with lifting machines locked out
Chemical energy
• Vent valve (locked in “open” position) to prevent chemical fluid pressure buildup
With all these preventive measures, the hope is that no form of potential energy great enough
to pose danger may be accidently released.
3.3. WORK, ENERGY, AND POWER 41
Cable snaps!
Crane
The Conservation of Energy still (and always!) holds true: all the potential energy stored in the
elevated weight must be accounted for, even when it free-falls. What happens, of course, is that
the weight accelerates toward the ground at the rate determined by Earth’s gravity: 32.2 feet per
second per second (32.2 ft/s2 ). As the weight loses height, its potential energy decreases according
to the gravitational potential energy formula (Ep = mgh), but since we know energy cannot simply
disappear we must conclude it is taking some other form. This “other form” is based on the weight’s
velocity (v), and is calculated by the kinetic energy formula:
1
Ek = mv 2
2
Where,
Ek = Kinetic energy in Joules or newton-meters (metric), or foot-pounds (British)
m = Mass of object in kilograms (metric) or slugs (British)
v = Velocity of mass in meters per second (metric) or feet per second (British)
Thus, the Conservation of Energy explains why a falling object must fall faster as it loses height:
kinetic energy must increase by the same amount that potential energy decreases, if energy is to be
conserved. A very small amount of this falling weight’s potential energy will be dissipated in the
form of heat (as air molecules are disturbed) rather than get converted into kinetic energy. However,
the vast majority of the initial 35700 ft-lbs of potential energy gets converted into kinetic energy,
until the weight’s energy is all kinetic and no potential the moment it first touches the ground.
When the weight finally slams into the ground, all that (nearly 35700 ft-lbs) of kinetic energy once
again gets converted into other forms. Compression of the soil upon impact converts much of the
energy into heat (molecular vibrations). Chunks of soil ejected from the impact zone possess their
own kinetic energy, carrying that energy to other locations where they slam into other stationary
objects. Sound waves rippling through the air also convey energy away from the point of impact. All
in all, the 35700 ft-lbs of potential energy which turned into (nearly) 35700 ft-lbs of kinetic energy
at ground level becomes dispersed.
42 CHAPTER 3. FULL TUTORIAL
The Law of Energy Conservation is extremely useful in projectile mechanics problems, where
we typically assume a projectile loses no energy and gains no energy in its flight. The velocity of
a projectile, therefore, depends on its height above the ground, because the sum of potential and
kinetic energies must remain constant:
Ep + Ek = constant
As a projectile rises in altitude (i.e. its gravitational potential energy increases), its velocity
must slow down (i.e. its kinetic energy decreases) in order that its total energy remain the same in
accordance with the Law of Energy Conservation.
In free-fall problems, where the only source of energy for a projectile is its initial height, the
initial potential energy must be equal to the final kinetic energy:
Ep (initial) = Ek (final)
1
mv 2
mghi =
2 f
We can see from this equation that mass cancels out of both sides, leaving us with this simpler
form:
1 2
ghi =
v
2 f
It also leads to the paradoxical conclusion that the mass of a free-falling object is irrelevant to
its velocity. That is, both a heavy object and a light object in free fall hit the ground with the same
velocity, and fall for the same amount of time, if released from the same height under the influence
of the same gravity18 .
18 In practice, we usually see heavy objects fall faster than light objects due to the resistance of air. Energy losses
due to air friction nullify our assumption of constant total energy during free-fall. Energy lost due to air friction never
translates to velocity, and so the heavier object ends up hitting the ground faster (and sooner) because it had much
more energy than the light object did to start.
3.3. WORK, ENERGY, AND POWER 43
1
Ek = mv 2 Kinetic energy due to velocity
2
[kg][m2 ] hmi
2
= [kg] 2 [m] Potential energy due to elevation
[s ] s
[kg][m2 ] h m i2
= [kg] Kinetic energy due to velocity
[s2 ] s
[kg][m2 ] h m i2
2
= [kg] Mass-to-energy conversion
[s ] s
In all three cases, the unit for energy is the same: kilogram-meter squared per second squared.
This is the fundamental definition of a “joule” of energy (also equal to a “newton-meter” of energy),
and it is the same result given by all three formulae.
Power is defined as the rate at which work is being done, or the rate at which energy is transferred.
Mathematically expressed, power is the first time-derivative of work (W ):
dW
P =
dt
The metric unit of measurement for power is the Watt, defined as one joule of work performed
per second of time. The British unit of measurement for power is the Horsepower, defined as 550
foot-pounds of work performed per second of time.
Although the term “power” is often colloquially used as a synonym for force or strength, it is in
fact a very different concept. A “powerful” machine is not necessarily a machine capable of doing
a great amount of work, but rather (more precisely) a great amount of work in a short amount of
time. Even a “weak” machine is capable of doing a great amount of work given sufficient time to
complete the task. The “power” of any machine is the measure of how rapidly it may perform work.
44 CHAPTER 3. FULL TUTORIAL
3.4.2 Temperature
In an ideal, monatomic19 gas (one atom per molecule), the mathematical relationship between
average molecular velocity and temperature is as follows:
1 3
mv 2 = kT
2 2
Where,
m = Mass of each molecule
v = Velocity of a molecule in the sample
v = Average (“mean”) velocity of all molecules in the sample
v 2 = Mean-squared molecular velocities in the sample
k = Boltzmann’s constant (1.3806504 × 10−23 J / K)
T = Absolute temperature (Kelvin), 273.15 more than degrees Celsius
Non-ideal gases, liquids, and solids are more complex than this. Not only can the atoms of
complex molecules move to and fro, but they may also twist and oscillate with respect to each other.
No matter how complex the particular substance may be, however, the basic principle remains
unchanged: temperature is an expression of how rapidly molecules move within a substance.
There is a temperature at which all molecular motion ceases. At that temperature, the substance
cannot possibly become “colder,” because there is no more motion to halt. This temperature
is called absolute zero, equal to −273.15 degrees Celsius, or −459.67 degrees Fahrenheit. Two
temperature scales based on this absolute zero point, Kelvin and Rankine, express temperature
relative to absolute zero. That is, zero Kelvin and zero degrees Rankine is equal to absolute zero
temperature. Any temperature greater than absolute zero will be a positive value in either the
Kelvin or the Rankine scales. A sample of freezing water at sea level, for example, is 0 degrees
Celsius (32 degrees Fahrenheit) but could also be expressed as 273.15 Kelvin20 (0 plus 273.15) or
491.67 degrees Rankine (32 plus 459.67).
19 Helium at room temperature is a close approximation of an ideal, monatomic gas, and is often used as an example
A table of melting and boiling points (at sea-level atmospheric pressure) for various substances
appears in this table, labeled in these four different units of temperature measurement:
o o o
Melting or boiling substance C F K R
Melting point of water (H2 O) 0 32 273.15 491.67
Boiling point of water (H2 O) 100 212 373.15 671.67
Melting point of ammonia (NH3 ) −77.7 −107.9 195.45 351.77
Boiling point of ammonia (NH3 ) −33.6 −28.5 239.55 431.17
Melting point of gold (Au) 1063 1945 1336 2405
Melting point of magnesium (Mg) 651 1203.8 924.2 1663.5
Boiling point of acetone (C3 H6 O) 56.5 133.7 329.65 593.37
Boiling point of propane (C3 H8 ) −42.1 −43.8 231.05 415.87
Boiling point of ethanol (C2 H6 O) 78.4 173.1 351.55 632.77
Note how degrees Celsius and Kelvin for each point on the table differ by a constant (offset) of
273.15, while each corresponding degree Fahrenheit and degree Rankine value differs by 459.67 (note
that many of the figures in this table are slightly rounded, so the offset might not be exactly that
much). You might think of Kelvin as nothing more than the Celsius scale zero-shifted by 273.15
degrees, and likewise degrees Rankine as nothing more than the Fahrenheit scale zero-shifted by
459.67 degrees.
Note also how increments in temperature measured in degrees Fahrenheit are the same as
increments of temperature measured in degrees Rankine. The same is true for degrees Celsius
and Kelvin. The difference between the melting point of ammonia (−77.7 degrees C) and the
melting point of water (0 degrees C) is the same difference as that between the melting points of
ammonia and water expressed in Kelvin: 195.45 and 273.15, respectively. Either way, the difference
in temperature between these two substances’ melting points is 77.7 degrees (C or K). This is useful
to know when dealing with temperature changes over time, or temperature differences between points
in a system – if an equation asks for a temperature difference (∆T ) in Kelvin, it is the same value
as the temperature difference expressed in Celsius. Likewise, a ∆T expressed in degrees Rankine is
identical to a ∆T expressed in degrees Fahrenheit. This is analogous to differences between two fluid
pressures expressed in PSIG versus PSIA: the differential pressure value (PSID) will be the same.
Most people are familiar with the Fahrenheit and Celsius temperature scales used to express
temperature in common applications, but the absolute scales of Rankine and Kelvin have special
significance and purpose in scientific endeavors. The fact that Rankine and Kelvin are absolute scales
in the same manner that atmospheres and torr are units of absolute pressure measurement makes
them uniquely suited for expressing temperature (molecular motion) in relation to the absence of
thermal energy. Certain scientific laws such as the Ideal Gas Law and the Stefan-Boltzmann Law
relate other physical quantities to absolute temperature, and so require the use of these absolute
units of measurement.
48 CHAPTER 3. FULL TUTORIAL
3.4.3 Heat
Heat, being the transfer of energy in thermal (molecular motion) form, may be measured in the same
units as any other form of energy is measured: Joules (metric) and foot-pounds (British). However,
special units of measurement are often used for heat instead:
• calorie
• kilocalorie (or “dietary Calorie”)
• British Thermal Unit (BTU)
A calorie of heat is defined as the amount of thermal energy transfer required to change the
temperature of one gram of water by one degree Celsius (∆T = 1 o C = 1 K). One calorie is
equivalent to 4.186 Joules.
A British Thermal Unit, or BTU is defined as the amount of thermal energy transfer required
to change the temperature of one pound of water by one degree Fahrenheit (∆T = 1 o F = 1 o R).
One BTU is equivalent to 778.2 foot-pounds.
The unit of “dietary” calories is used to express the amount of thermal energy available in a
sample of food by combustion21 . Since the official unit of the “calorie” is so small compared to the
typical amounts of energy contained in a meal, nutritionists use the unit of the kilocalorie (1000
calories, or 4186 Joules) and call it “Calorie” (with a capital letter “C”).
Just as “Calories” are used to rate the energy content of food, the heat units of “calories” and
“BTU” are very useful in describing the potency of various industrial fuels. The following table
shows the heat of combustion for a few common fuels, in units of kilocalories per gram and BTU per
pound:
For example, if exactly one gram of methane gas were completely burnt, the resulting heat
liberated in the fire would be sufficient to warm 13.3 kilograms of water from 20 degrees Celsius to
21 degrees Celsius (a temperature rise, or ∆T , of one degree Celsius).
If a meal rated at 900 Calories (900 “dietary calories,” or 900 kilocalories) of energy were
completely metabolized, the resulting heat would be sufficient to warm a pool of water 900 kilograms
in mass (900 liters, or about 237 gallons) by one degree Celsius. This same amount of heat could
raise half the amount of water twice the temperature rise: 450 liters of water warmed two degrees
Celsius.
21 Animals process food by performing a very slow version of combustion, whereby the carbon and hydrogen atoms
in the food join with oxygen atoms inhaled to produce water and carbon dioxide gas (plus energy). Although it may
seem strange to rate the energy content of food by measuring how much heat it gives off when burnt, burning is just
a faster method of energy extraction than the relatively slow processes of biological metabolism.
3.4. ELEMENTARY THERMODYNAMICS 49
Practical examples of heat transfer often involve multiple modes rather than just one. For
example, the transfer of heat to a person’s body by sunlight obviously involves radiation from the
Sun, but it also involves conduction through layers of clothing and convection by air passing from
sun-warmed objects to the person.
Temperature sensors likewise rely on multiple heat-transfer modes to sample thermal energy from
matter. Infrared thermometers detect temperature by sensing the intensity of infrared light radiated
by hot objects. A contact-style sensor directly touching a hot object relies on conduction to sense
the temperature of that object. A similar sensor inserted into a pipe carrying a hot fluid relies on
convection to measure the average temperature of that fluid.
22 Heat may be forced to flow from cold to hot by the use of a machine called a heat pump, but this direction of
heat flow does not happen naturally, which is what the word “spontaneous” implies. In truth, the rule of heat flowing
from high-temperature to cold-temperature applies to heat pumps as well, just in a way that is not obvious from first
inspection. Mechanical heat pumps cause heat to be drawn from a cool object by placing an even cooler object (the
evaporator ) in direct contact with it. That heat is then transferred to a hot object by placing an even hotter object
(the condenser ) in direct contact with it. Heat is moved against the natural (spontaneous) direction of flow from the
evaporator to the condenser by means of mechanical compression and expansion of a refrigerant fluid.
23 In this context, we are using the word “radiation” in a very general sense, to mean thermal energy radiated away
from the hot source via photons. This is quite different from nuclear radiation, which is what some may assume this
term means upon first glance.
50 CHAPTER 3. FULL TUTORIAL
Radiation
If you have ever experienced the immediate sensation of heat from a large fire or explosion some
distance away, you know how radiation works to transfer thermal energy. Radiation is also the
method of heat transfer experienced in the Earth’s receiving of heat from the Sun (and also the
mechanism of heat loss from Earth to outer space). Radiation is the least efficient of the three heat
transfer mechanisms. It may be quantified by the Stefan-Boltzmann Law, which states the rate of
heat lost by an object ( dQ
dt ) is proportional to the fourth power of its absolute temperature, and
directly proportional to its radiating area:
dQ
= eσAT 4
dt
Where,
dQ
dt = Radiant heat loss rate (watts)
e = Emissivity factor (unitless)
σ = Stefan-Boltzmann constant (5.670400 × 10−8 W / m2 · K4 )
A = Surface area (square meters)
T = Absolute temperature (Kelvin)
Here is one of the scientific applications where temperature expressed in absolute units is truly
necessary. Radiant energy is a direct function of molecular motion, and so we would logically expect
objects to radiate energy at any temperature above absolute zero. The temperature value used in
this formula must be in units of Kelvin24 in order for the resulting dQ
dt value to be correct. If degrees
Celsius were used for T instead of Kelvin, the formula would predict zero thermal radiation at the
freezing point of water (0 o C) and negative radiation at any temperature below freezing, which is
not true. Remember that the “zero” points of the Celsius and Fahrenheit scales were arbitrarily set
by the inventors of those scales, but that the “zero” points of the Kelvin and Rankine scales reflect
a fundamental limit of nature.
The emissivity factor varies with surface finish and color, ranging from one (ideal) to zero (no
radiation possible). Dark-colored, rough surfaces offer the greatest emissivity factors, which is why
heater elements and radiators are usually painted black. Shiny (reflective), smooth surfaces offer
the least emissivity, which is why thermally insulating surfaces are often painted white or silver.
Like all heat-transfer modes, radiation is two-way. Objects may emit energy in the form of
radiation, and they may also receive energy in the form of radiation. Everyone knows white-colored
shirts are cooler than black-colored shirts worn on a hot, sunny day – this is an example of how
emissivity affects heat absorption by radiant transfer. A black-colored shirt (high emissivity value)
enhances the receiving of radiant energy by your body from the sun. What is not as obvious, though,
is that a white-colored shirt will keep you warmer than a black-colored shirt on a cold, dark day
because that same decreased emissivity inhibits body heat loss by radiation. Thus, high-emissivity
objects both heat and cool more readily by radiation than low-emissivity objects.
24 Or in degrees Rankine, provided a suitably units-corrected value for the Stefan-Boltzmann constant were used.
3.4. ELEMENTARY THERMODYNAMICS 51
Conduction
If you have ever accidently touched a hot iron or stove heating element, you possess a very vivid
recollection of heat transfer through conduction. In conduction, fast-moving molecules in the hot
substance transfer some of their kinetic energy to slower-moving molecules in the cold substance.
Since this transfer of energy requires collisions between molecules, it only applies when the hot and
cold substances directly contact each other.
Perhaps the most common application of heat conduction in industrial processes is through the
walls of a furnace or some other enclosure containing an extreme temperature. In such applications,
the desire is usually to minimize heat loss through the walls, so those walls will be “insulated” with
a substance having poor thermal conductivity.
Conductive heat transfer rate is proportional to the difference in temperature between the hot
and cold points, the area of contact, the distance of heat travel from hot to cold, and the thermal
conductivity of the substance:
dQ kA∆T
=
dt l
Where,
dQ
dt = Conductive heat transfer rate
k = Thermal conductivity
A = Surface area
∆T = Difference of temperature between “hot” and “cold” sides
l = Length of heat flow path from “hot” to “cold” side
Note the meaning of “∆T ” in this context: it refers to the difference in temperature between two
different locations in a system. Sometimes the exact same symbology (“∆T ”) refers to a change in
temperature over time in the study of thermodynamics. Unfortunately, the only way to distinguish
one meaning of ∆T from the other is by context.
52 CHAPTER 3. FULL TUTORIAL
An illustration showing heat conduction through a wall gives context to the variables in the
previous equation. As we see here, A refers to the surface area of the wall, ∆T refers to the
difference of temperature between either surface of the wall, and l refers to the thickness of the wall:
Tcold Thot
∆T = Thot - Tcold
A
Heat Heat
sink source
dQ dQ
dt dt
k
In the United States, a common measure of insulating ability used for the calculation of
conductive heat loss in shelters is the R-value. The greater the R-value of a thermally insulating
material, the less conductive it is to heat (lower k value). “R-value” mathematically relates to k
and l by the following equation:
l
R=
k
Rearranging this equation, we see that l = kR, and this allows us to substitute kR for l in the
conduction heat equation, then cancel the k terms:
dQ kA∆T
=
dt kR
dQ A∆T
=
dt R
R is always expressed in the compound unit of square feet · hours · degrees Fahrenheit per BTU.
This way, with a value for area expressed in square feet and a temperature difference expressed
in degrees Fahrenheit, the resulting heat transfer rate ( dQ
dt ) will naturally be in units of BTU per
hour, which is the standard unit in the United States for expressing heat output for combustion-type
heaters. Dimensional analysis shows how the units cancel to yield a heat transfer rate in BTUs per
hour:
The utility of R-value ratings may be shown by a short example. Consider a contractor trailer,
raised up off the ground on a mobile platform, with a total skin surface area of 2400 square feet
(walls, floor, and roof) and a uniform R-value of 4 for all surfaces. If the trailer’s internal temperature
must be maintained at 70 degrees Fahrenheit while the outside temperature averages 40 degrees
Fahrenheit, the required output of the trailer’s heater will be:
Convection
Most industrial heat-transfer processes occur through convection, where a moving fluid acts as an
intermediary substance to transfer heat from a hot substance (heat source) to a cold substance
(heat sink ). Convection may be thought of as two-stage heat conduction on a molecular scale:
fluid molecules come into direct contact with a hot object and absorb heat from that object through
conduction, then those molecules later release that heat energy through conduction by direct contact
with a cooler object. If the fluid is recycled in a piping loop, the two-stage conduction process
repeats indefinitely, individual molecules heating up as they absorb heat from the heat source and
then cooling down as they release heat to the heat sink.
Special process devices called heat exchangers perform this heat transfer function between two
different fluids, the two fluids circulating past each other on different sides of tube walls. A simple
example of a heat exchanger is the radiator connected to the engine of an automobile, being a water-
to-air exchanger, the engine’s hot water transferring heat to cooling air entering the grille of the car
as it moves.
54 CHAPTER 3. FULL TUTORIAL
Another example of a liquid-to-air heat exchanger is the condenser on a heat pump, refrigerator,
or air conditioner, a photograph appearing here:
Another common style of heat exchanger works to transfer heat between two liquids. A small
example of this design used to transfer heat from a boat engine is shown here:
The purpose for this heat exchanger is to exchange heat between the liquid coolant of the boat
engine and sea water, the latter being quite corrosive to most metals. An engine would soon be
damaged if sea water were used directly as the coolant fluid, and so heat exchangers such as this
provide a means to release excess heat to the sea without subjecting the engine block to undue
corrosion. The heat exchanger, of course, does suffer from the corrosive effects of sea water, but at
least it is less expensive and more convenient to replace than an entire engine when it reaches the
end of its service life.
3.4. ELEMENTARY THERMODYNAMICS 55
This marine engine heat exchanger is an example of a shell-and-tube design, where one fluid
passes inside small tubes and a second fluid passes outside those same tubes, the tube bundle being
contained in a shell. The interior of such an exchanger looks like this when cut away:
The tubes of this particular heat exchanger are made of copper, a metal with extremely high
thermal conductivity (k), to facilitate conductive heat transfer.
Liquid-to-liquid heat exchangers are quite common in industry, where a set of tubes carry one
process liquid while a second process liquid circulates on the outside of those same tubes. The metal
walls of the tubes act as heat transfer areas for conduction to occur. Metals such as copper with
very high k values (very low R values) encourage heat transfer, while long lengths of tube ensure
ample surface area for heat exchange.
56 CHAPTER 3. FULL TUTORIAL
Reactant "A"
feed
Reactor
TI 700 oF
430 oF 190 oF
TI TI
Reactant "B"
Heat feed
exchanger
Reaction product
out
TI 550 oF
25 In this diagram, each line tipped by an arrow is a tube or pipe conveying some fluid. Circle-symbols with “TI”
written inside represent temperature indicators (sensors). Each “bow-tie” symbol is a valve used to exert control over
the flow rate of a fluid. The reactor itself is simply a hollow, sealed vessel inside of which the chemical reaction ensues.
26 Jim Cahill of Emerson wrote in April 2010 (“Reducing Distillation Column Energy Usage” Emerson Process
Expert weblog) about a report estimating distillation column energy usage to account for approximately 6% of the
total energy used in the United States. This same report tallied the number of columns in US industry to be
approximately 40000 total, accounting for about 19% of all energy used in manufacturing processes!
3.4. ELEMENTARY THERMODYNAMICS 57
The following PFD shows a simple distillation process complete with heat exchangers for reboiling
(adding heat to the bottom of the distillation column), condensing (extracting heat from the
“overhead” product at the top of the column), and energy conservation (transferring heat from
the hot products to the incoming feed):
Cooling 85 oF
water
supply Condenser
130 oF
122 oF
o
Reflux 118 F 120 oF "Overhead"
product out
130 oF
Feed in
274 oF 242 oF
Product "A" out
337 oF Preheated feed
205 oF
Distillation
"tower" 341 oF 309 oF
Product "B" out
o
550 F Boil-up
Steam 646 oF
supply
295 oF
Reboiler 514 oF
Steam 530 oF 466 oF "Bottoms" out
return product
337 oF
Distillation “columns” (also called fractionating towers in the industry) are tall vessels containing
sets of “trays” where rising vapors from the boiling process contact falling liquid from the condensing
process. Temperatures increase toward the bottom of the column, while temperatures decrease
toward the top. In this case, steam through a “reboiler” drives the boiling process at the bottom
of the column (heat input), and cold water through a “condenser” drives the condensing process
at the top of the column (heat extraction). Products coming off the column at intermediate points
are hot enough to serve as pre-heating flows for the incoming feed. Note how the “economizing”
heat exchangers expose the cold feed flow to the cooler Product A before exposing it to the warmer
Product B, and then finally the warmest “Bottoms” product. This sequence of cooler-to-warmer
maximizes the efficiency of the heat exchange process, with the incoming feed flowing past products
of increasing temperature as it warms up to the necessary temperature for distillation entering the
58 CHAPTER 3. FULL TUTORIAL
column.
Some heat exchangers transfer heat from hot gases to cool(er) liquids An example of this type
of heat exchanger is the construction of a steam boiler, where hot combustion gases transfer heat to
water flowing inside metal tubes:
Exhaust stack
Steam
Steam drum
water
Riser
tubes
Downcomer
tubes
Mud drum
er
rn
Bu
Feedwater
Here, hot gases from the combustion burners travel past the metal “riser” tubes, transferring heat
to the water within those tubes. This also serves to illustrate an important convection phenomenon:
a thermal siphon (often written as thermosiphon). As water heats in the “riser” tubes, it becomes
less dense, producing less hydrostatic pressure at the bottom of those tubes than the colder water in
the “downcomer” tubes. This difference of pressure causes the colder water in the downcomer tubes
to flow down to the mud drum, and hot water in the riser tubes to flow up to the steam drum. This
natural convection current will continue as long as heat is applied to the riser tubes by the burners,
and an unobstructed path exists for water to flow in a loop.
3.4. ELEMENTARY THERMODYNAMICS 59
Natural convection also occurs in heated air, such as in the vicinity of a lit candle:
Natural convection
near a candle flame
Air motion
Candle
This thermally forced circulation of air helps convect heat from the candle to all other points
within the room it is located, by carrying heated air molecules to colder objects.
60 CHAPTER 3. FULL TUTORIAL
Q ∝ m∆T
Where,
Q = Heat gain or loss
m = Mass of sample
∆T = Temperature change (rise or fall) over time
The next logical question to ask is, “How does the relationship between heat and temperature
change work for substances other than water?” Does it take the same amount of heat to change the
temperature of one gram of iron by one degree Celsius as it does to change the temperature of one
gram of water by one degree Celsius? The answer to this question is a resounding no! Different
substances require vastly different amounts of heat gain/loss to alter their temperature by the same
degree, even when the masses of those substances happen to be identical.
We have a term for this ability to absorb or release heat, called heat capacity or specific
heat, symbolized by the variable c. Thus, our heat/mass/temperature change relationship may
be described as a true formula instead of a mere proportionality:
Q = mc∆T
Where,
Q = Heat gain or loss (metric calories or British BTU)
m = Mass of sample (metric grams or British pounds)
c = Specific heat of substance
∆T = Temperature change (metric degrees Celsius or British degrees Fahrenheit)
Pure water, being the standard by which all other substances are measured, has a specific heat
value of 1. The smaller the value for c, the less heat gain or loss is required to alter the substance’s
temperature by a set amount. That substance (with a low value of c) has a low “heat capacity”
because each degree of temperature rise or fall represents a relatively small amount of energy gained
or lost. Substances with low c values are easy to heat and cool, while substances having high c
values require much heat in order to alter their temperatures, assuming equal masses.
3.4. ELEMENTARY THERMODYNAMICS 61
A table of specific heat values (at room temperature, 25 degrees Celsius27 ) for common substances
appears here:
If a liquid or a gas is chosen for use as a coolant (a substance to efficiently convect heat away
from an object), greater values of c are better. Water is one of the best liquid coolants with its
relatively high c value of one: it has more capacity to absorb heat than other liquids, for the same
rise in temperature. The ideal coolant would have an infinite c value, being able to absorb an infinite
amount of heat without itself rising in temperature at all.
As you can see from the table, the light gases (hydrogen and helium) have extraordinarily high
c values. Consequently, they function as excellent media for convective heat transfer. This is why
large electric power generators often use hydrogen gas as a coolant: hydrogen has an amazing
ability to absorb heat from the wire windings of a generator without rising much in temperature.
In other words, hydrogen absorbs a lot of heat while still remaining “cool” (i.e. remains at a low
temperature). Helium, although not quite as good a coolant as hydrogen, has the distinct advantage
of being chemically inert (non-reactive), in stark contrast to hydrogen’s extreme flammability. Some
nuclear reactors use helium gas as a coolant rather than a liquid such as water or molten sodium
metal.
Lead has an extraordinarily low c value, being a rather “easy” substance to heat up and cool
down. Anyone who has ever cast their own lead bullets for a firearm knows how quickly a new lead
bullet cools off after being released from the mold, especially if that same person has experience
casting other metals such as aluminum.
27 An important detail to note is that specific heat does not remain constant over wide temperature changes. This
complicates calculations of heat required to change the temperature of a sample: instead of simply multiplying the
temperature change by mass and specific heat (Q = mc∆T or Q = mc[T2 − T1 ]), we must integrate specific heat
R T2
over the range of temperature (Q = m c dT ), summing up infinitesimal products of specific heat and temperature
T1
change (c dT ) over the range starting from temperature T1 through temperature T2 then multiplying by the mass to
calculate total heat required. So, the specific heat values given for substances at 25 o C only hold true for relatively
small temperature changes deviating from 25 o C. To accurately calculate heat transfer over a large temperature
change, one must incorporate values of c for that substance at different temperatures along the expected range.
62 CHAPTER 3. FULL TUTORIAL
Numerical examples are helpful to better understand specific heat. Consider a case where a
copper pot filled with water receives heat from a small gas burner operating at an output of 5000
BTU per hour (350 calories per second):
Water
c = 1.00
m = 3700 grams
Starting temperature = 20 oC
Copper pot
c = 0.092
m = 1100 grams
dQ
= 5000 BTU/h = 350 cal/s
dt
Time of heating = 40 seconds
A reasonable question to ask would be, “How much will the temperature of this water-filled pot
rise after 40 seconds of heating?” With the burner’s heat output of 350 calories per second and a
heating time of 40 seconds, we may assume28 the amount of heat absorbed by the water-filled pot
will be the simple product of heat rate times time:
dQ 350 cal
Q= t= 40 s = 14000 calories
dt s
This amount of heat not only goes into raising the temperature of the water, but it also raises
the temperature of the copper pot. Each substance (water, copper) has its own specific heat and
mass values (c and m), but they will share the same temperature rise (∆T ), so we must sum their
heats as follows:
28 In reality, the amount of heat actually absorbed by the pot will be less than this, because there will be heat losses
from the warm pot to the surrounding (cooler) air. However, for the sake of simplicity, we will assume all the burner’s
heat output goes into the pot and the water it holds.
3.4. ELEMENTARY THERMODYNAMICS 63
∆T = 3.68 o C
So, if the water and pot began at a temperature of 20 degrees Celsius, they will be at a
temperature of 23.68 degrees Celsius after 40 seconds of heating over this small burner.
Another example involves the mixing of two substances at different temperatures. Suppose a
heated mass of iron drops into a cool container29 of water. Obviously, the iron will lose heat energy
to the water, causing the iron to decrease in temperature while the water rises in temperature.
Suppose the iron’s mass is 100 grams, and its original temperature is 65 degrees Celsius. Suppose
the water’s mass is 500 grams, and its original temperature is 20 degrees Celsius:
Water
c = 1.00 Iron
m = 500 grams c = 0.108
Tstart = 20 oC m = 100 grams
Tstart = 65 oC
Styrofoam cup
(negligible mass and specific heat)
29 We will assume for the sake of this example that the container holding the water is of negligible mass, such as a
Styrofoam cup. This way, we do not have to include the container’s mass or its specific heat into the calculation.
64 CHAPTER 3. FULL TUTORIAL
What will the equilibrium temperature be after the iron falls into the water and both their
temperatures equalize? We may solve this by setting two heat equations equal to each other30 : the
heat lost by the iron and the heat gained by the water, with the final equilibrium temperature being
T:
Qiron = Qwater
miron ciron (65) − miron ciron T = mwater cwater T − mwater cwater (20)
miron ciron (65) + mwater cwater (20) = miron ciron T + mwater cwater T
miron ciron (65) + mwater cwater (20) = T (miron ciron + mwater cwater )
T = 20.95 o C
Thus, the iron’s temperature falls from 65 degrees Celsius to 20.95 degrees Celsius, while the
water’s temperature rises from 20 degrees Celsius to 20.95 degrees Celsius. The water’s tremendous
specific heat value compared to the iron (nearly 10 times as much!), as well as its superior mass (5
times as much) results in a much larger temperature change for the iron than for the water.
30 An alternative way to set up the problem would be to calculate ∆T for each term as T
f inal − Tstart , making
the iron’s heat loss a negative quantity and the water’s heat gain a positive quantity, in which case we would have to
set up the equation as a zero-sum balance, with Qiron + Qwater = 0. I find this approach less intuitive than simply
saying the iron’s heat loss will be equal to the water’s heat gain, and setting up the equation as two positive values
equal to each other.
3.4. ELEMENTARY THERMODYNAMICS 65
An analogy to help grasp the concept of specific heat is to imagine heat as a fluid31 that may be
“poured” into vessels of different size, those vessels being objects or substances to be heated. The
amount of liquid held by any vessel represents the total amount of thermal energy, while the height
of the liquid inside any vessel represents its temperature:
Q T
T Q T Q Q T
The factor determining the relationship between liquid volume (heat) and liquid height
(temperature) is of course the cross-sectional area of the vessel. The wider the vessel, the more
heat will be required to “fill” it up to any given temperature. In this analogy, the area of the
vessel is analogous to the term mc: the product of mass and specific heat. Objects with larger
mass require more heat to raise their temperature to any specific point, specific heats being equal.
Likewise, objects with large specific heat values require more heat to raise their temperature to any
specific point, masses being equal.
In the first numerical calculation example where we determined the temperature of a pot of water
after 40 seconds of heating, the analogous model would be to determine the height of liquid in a
vessel after pouring liquid into it for 40 seconds at a fixed rate. A model for the second numerical
example would be to calculate the equilibrium height (of liquid) after connecting two vessels together
at their bottoms with a tube. Although the liquid heights of those vessels may be different at first,
the levels will equalize after time by way of liquid passing through the tube from the higher-level
vessel to the lower-level vessel.
31 This is not far from the hypotheses of eighteenth-century science, where heat was thought to be an invisible fluid
called caloric.
66 CHAPTER 3. FULL TUTORIAL
Many industrial processes use fluids to convectively transfer thermal energy from one object (or
fluid) to another. In such applications, it is important to know how much thermal energy will be
carried by a specific quantity of that fluid over a specified temperature drop. One common way to
express this quantity is called enthalpy. Enthalpy is the amount of heat lost by a unit mass (one
gram metric, or one pound British) of a substance as it cools from a given temperature all the way
down to the freezing point of water (0 degrees Celsius, or 32 degrees Fahrenheit). In other words,
enthalpy is a measure32 of a substance’s thermal energy using the freezing temperature of water as
a reference. A sample of water at a temperature of 125 degrees Fahrenheit, for example, has an
enthalpy of 93 BTU per pound (or 93 calories per gram), because 93 BTU of thermal energy would
be lost if one pound of that water happened to cool from its given temperature (125 o F) down to
32 o F:
Q = mc∆T
BTU
Q = (1 lb) 1 o (125 o F − 32 o F)
lb F
Q = 93 BTU
Even if the substance in question does not cool down to the freezing temperature of water,
enthalpy is a useful figure for comparing the thermal energy “content” of hot fluids (per unit mass).
For example, if one were given the enthalpy values for a substance before and after heat transfer, it
would be easy to calculate the amount of heat transfer that transpired simply by subtracting those
enthalpy values33 . If water at 125 o F has an enthalpy value of 93 BTU/lb and water at 170 o F
has an enthalpy of value 138 BTU/lb, we may calculate the amount of heat needed to increase the
temperature of a sample of water from 125 o F to 170 o F simply by subtracting 93 BTU/lb from 138
BTU/lb to arrive at 45 BTU/lb.
In this rather trivial example, it would have been just as easy for us to calculate the heat
necessary to increase water’s temperature from 125 o F to 170 o F by using the specific heat formula
(Q = mc∆T )34 , and so it might appear as though the concept of enthalpy sheds no new light on the
subject of heat transfer. However, the ability to calculate heat transfer based on a simple subtraction
of enthalpy values proves quite useful in more complex scenarios where substances change phase, as
we will see next.
32 A useful analogy for enthalpy is the maximum available balance of a bank account. Suppose you have a bank
account with a minimum balance requirement of $32 to maintain that account. Your maximum available balance at
any time would be the total amount of money in that account minus $32, or to phrase this differently your maximum
available balance is the most money you may spend from this account while still keeping that account open. Enthalpy
is much the same: the amount of thermal energy a sample may “spend” (i.e. lose) before its temperature reaches 32
degrees Fahrenheit.
33 Appealing to the maximum available balance analogy, if we compared the maximum available balance in your
bank account before and after a transaction, we could determine how much money was deposited or withdrawn from
your account simply by subtracting those two values.
34 Following the formula Q = mc∆T , we may calculate the heat as (1)(1)(170 − 125) = 45 BTU. This is obviously
the same result we obtained by subtracting enthalpy values for water at 170 o F and 125 o F.
3.4. ELEMENTARY THERMODYNAMICS 67
Q = mc∆T
cal
Q = (70 g) 1 o (100 o C − 24 o C)
g C
Q = 5320 cal
However, actually boiling the 70 grams of water into 70 grams of steam (both at 100 degrees
Celsius) requires a comparatively enormous input of heat: 37734 calories – over seven times as
much heat to turn the water to steam as what is required to warm the water to its boiling point!
Furthermore, this additional input of 37734 calories does not increase the temperature of the water
at all: the resulting steam is still at a temperature of (only) 100 degrees Celsius. If further heat is
added to the 70 gram steam sample, its temperature will rise, albeit at a rate proportional to the
value of steam’s specific heat (0.476 calories per gram degree Celsius, or BTU per pound degree
Fahrenheit).
What we see here is a fundamentally different phenomenon than we saw with specific heat. Here,
we are looking at the thermal energy required to transition a substance from one phase to another,
not to change its temperature. We call this quantity latent heat rather than specific heat, because no
temperature change occurs35 . Conversely, if we allow the steam to condense back into liquid water,
it must release the same 37734 calories of heat energy we invested in it turning the water into steam
before it may cool at all below the boiling point (100 degrees Celsius).
Latent heat has the effect of greatly increasing a substance’s enthalpy. Recall that “enthalpy”
is the amount of heat lost by one pound (mass) of a substance if it happened to cool from its given
temperature all the way down to the freezing temperature of water (0 o C, or 32 o F). Hot water has
an enthalpy of 1 BTU/lb for every degree of temperature above freezing. Steam, however, possesses
far greater enthalpy because of the latent heat released in the phase change from vapor to liquid
before it releases heat as water cooling down to 32 o F.
35 The word “latent” refers to something with potential that is not yet realized. Here, heat exchange takes place
without there being any realized change in temperature. By contrast, heat resulting in a temperature change
(Q = mc∆T ) is called sensible heat.
68 CHAPTER 3. FULL TUTORIAL
As with specific heat, there is a formula relating mass, latent heat, and heat exchange:
Q = mL
Where,
Q = Heat of transition required to completely change the phase of a sample (metric calories or
British BTU)
m = Mass of sample (metric grams or British pounds)
L = Latent heat of substance
Each substance has its own set of latent heat values, one36 for each phase-to-phase transition.
Water, for example, exhibits a latent heat of vaporization (boiling/condensing) of 539.1 calories per
gram, or 970.3 BTU per pound, at atmospheric pressure (boiling point = 100 o C = 212 o F). Water
also exhibits a latent heat of fusion (melting/freezing) of 79.7 calories per gram, or 143.5 BTU per
pound. Both figures are enormous compared to water’s specific heat value of 1 calorie per gram-
degree Celsius (or 1 BTU per pound-degree Fahrenheit37 ): it takes only one calorie of heat to warm
one gram of water one degree Celsius, but it takes 539.1 calories of heat to boil that same gram of
water into one gram of steam, and 79.7 calories of heat to melt one gram of ice into one gram of
water. The lesson here is simple: phase changes involve huge amounts of heat.
A table showing various latent heats of vaporization (all at room temperature, 70 degrees
Fahrenheit) for common industrial fluids appears here, contrasted against their specific heat values
(as liquids). In each case you will note how much larger L is than c:
One of the most important, and also non-intuitive, consequences of latent heat is the relative
stability of temperature during the phase-change process. Referencing the table of latent heats of
vaporization, we see how much more heat is needed to boil a liquid into a vapor than is needed to
warm that same liquid by one degree of temperature. During the process of boiling, all heat input
to the liquid goes into the task of phase change (latent heat) and none of it goes into increased
temperature. In fact, until all the liquid has been vaporized, the liquid’s temperature cannot rise
above its boiling point! The requirement of heat input to vaporize a liquid forces temperature to
stabilize (not rise further) until all the liquid has evaporated from the sample.
36 Latent heat of vaporization also varies with pressure, as different amounts of heat are required to vaporize a liquid
depending on the pressure that liquid is subject to. Generally, increased pressure (increased boiling temperature)
results in less latent heat of vaporization.
37 The reason specific heat values are identical between metric and British units, while latent heat values are not, is
because latent heat does not involve temperature change, and therefore there is one less unit conversion taking place
between metric and British when translating latent heats. Specific heat in both metric and British units is defined in
such a way that the three different units for heat, mass, and temperature all cancel each other out. With latent heat,
we are only dealing with mass and heat, and so we have a proportional conversion of 59 or 59 left over, just the same
as if we were converting between degrees Celsius and Fahrenheit alone.
3.4. ELEMENTARY THERMODYNAMICS 69
If we take a sample of ice and add heat to it over time until it melts, warms, boils, and then
becomes steam, we will notice a temperature profile that looks something like this:
All ice Ice/water mix All water Water/steam mix All steam
Steam
heating
Temperature
Water Boiling
100oC heating
Ice
Melting
0o C heating
The flat areas of the graph during the melting and boiling periods represents times where the
sample’s temperature does not change at all, but where all heat input goes into the work of changing
the sample’s phase. Only where we see the curve rising does the temperature change. So long as
there is a mixture of different phases, the temperature remains “locked” at one value. Only when
there is a single phase of material is the temperature “allowed” to rise or fall.
The sloped areas of the graph reveal the specific heat of the substance in each particular phase.
Note how the liquid (water) portion of the graph has a relatively shallow slope, due to the specific
heat value (c) of water being equal to 1. Both the ice and the steam portions of the graph have
steeper slopes because both of those phases possess smaller values of specific heat (c = 0.5 and
c = 0.476, respectively). The smaller the value of c, the more a sample’s temperature will rise for
any given input of thermal energy. For any given rate of heat transfer, smaller c values result in
more rapid temperature changes.
70 CHAPTER 3. FULL TUTORIAL
We may employ our liquid-filled vessel analogy to the task of explaining latent heat. Any point
of phase change is analogous to a point along the vessel’s height equipped with a large expansion
chamber, so that the vessel “acts” as if its area were much larger at one point, requiring much more
fluid volume (heat) to change height (temperature) past that one point:
Expansion
chamber
T
Q
Liquid poured into this vessel will fill it at a rate proportional to the volume added and inversely
proportional to the vessel’s cross-sectional area at the current liquid height. As soon as the liquid
level reaches the expansion chamber, a great deal more liquid must be added to cause level to
increase, since this chamber must completely fill before the liquid level may rise above it. Once that
happens, the liquid level rises at a different rate with addition introduced volume, because now the
phase is different (with a different specific heat value).
Remember that the filling of a vessel with liquid is merely an analogy for heat and temperature,
intended to provide an easily visualized process mimicking another process not so easily visualized.
The important concept to realize with latent heat and phase change is that it constitutes a
discontinuity in the temperature/heat function for any given substance.
A vivid demonstration of this phenomenon is to take a paper38 cup filled with water and place
it in the middle of a roaring fire39 . “Common sense” might tell you the paper will burn through
with the fire’s heat, so that the water runs out of the cup through the burn-hole. This does not
happen, however. Instead, the water in the cup will rise in temperature until it boils, and there
it will maintain that temperature no matter how hot the fire burns. The boiling point of water
happens to be substantially below the burning point of paper, and so the boiling water keeps the
paper cup too cool to burn. As a result, the paper cup remains intact so long as water remains in
the cup. The rim of the cup above the water line will burn up because the steam does not have
the same temperature-stabilizing effect as the water, leaving a rimless cup that grows shorter as the
water boils away.
38 Styrofoam and plastic cups work as well, but paper exhibits the furthest separation between the boiling point of
water and the burning point of the cup material, and it is usually thin enough to ensure good heat transfer from the
outside (impinging flame) to the inside (water).
39 This is a lot of fun to do while camping!
3.4. ELEMENTARY THERMODYNAMICS 71
The point at which a pure substances changes phase not only relates to temperature, but to
pressure as well. We may speak casually about the boiling point of water being 100 degrees Celsius
(212 degrees Fahrenheit), but that is only if we assume the water and steam are at atmospheric
pressure (at sea level). If we reduce the ambient air pressure40 , water will boil at a lesser temperature.
Anyone familiar with cooking at high altitudes knows you must generally cook for longer periods of
time at altitude, because the decreased boiling temperature of water is not as effective for cooking.
Conversely, anyone familiar with pressure cooking (where the cooking takes place inside a vessel
pressurized by steam) knows how comparatively little cooking time is required because the pressure
raises water’s boiling temperature. In either of these scenarios, where pressure influences41 boiling
temperature, the latent heat of water acts to hold the boiling water’s temperature stable until all
the water has boiled away. The only difference is the temperature at which the water begins to boil
(or when the steam begins to condense).
Many industrial processes use boiling liquids to convectively transfer heat from one object (or
fluid) to another. In such applications, it is important to know how much heat will be carried by
a specific quantity of the vapor as it condenses into liquid over a specified temperature drop. The
quantity of enthalpy (heat content) used for rating the heat-carrying capacity of liquids applies to
condensing vapors as well. Enthalpy is the amount of heat lost by a unit mass (one gram metric, or
one pound British) of the fluid as it cools from a given temperature all the way down to the freezing
point of water (0 degrees Celsius, or 32 degrees Fahrenheit)42 . When the fluid’s initial state is vapor,
and it condenses into liquid as it cools down to the reference temperature (32 o F), the heat content
(enthalpy) is not just a function of specific heat, but also of latent heat.
Water at its atmospheric boiling point has an enthalpy of approximately 180 BTU per pound.
Steam at atmospheric pressure and 212 o F, however, has an enthalpy of about 1150 BTU per pound :
more than six times as much heat as water at the same temperature. 970 of that 1150 BTU/lb is
due to the phase change from steam to water, while the rest is due to water’s specific heat as it cools
from 212 o F to 32 o F.
Many technical reference books contain a set of data known as a steam table showing various
properties of steam at different temperatures and pressures. Enthalpy is one of the most important
parameters given in a steam table, showing how much available energy resides in steam under
different pressure and temperature conditions. For this reason, enthalpy is sometimes referred to
as total heat (hg ). Steam tables also show saturation temperature (the condensing temperature for
steam at that pressure) and steam density.
40 This may be done in a vacuum jar, or by traveling to a region of high altitude where the ambient air pressure is
Molecules in a liquid reside close enough to each other that they cohere, whereas molecules in a vapor or gas are
relatively far apart and act as independent objects. The process of boiling requires that cohesion between liquid
molecules to be broken, so the molecules may drift apart. Increased pressure encourages cohesion in liquid form
by helping to hold the molecules together, while decreased pressure encourages the separation of molecules into a
vapor/gas.
42 As mentioned previously, a useful analogy for enthalpy is the maximum available balance for a bank account with
a $32 minimum balance requirement: that is, how much money may be spent from that account without closing it
out.
72 CHAPTER 3. FULL TUTORIAL
If the vapor in question is at a temperature greater than its boiling point at that pressure, the
vapor is said to be superheated. The enthalpy of superheated vapor comes from three different
heat-loss mechanisms:
• Cooling the vapor down to its condensing temperature (specific heat of vapor)
Using steam as the example once more, a sample of superheated steam at 500 o F and atmospheric
pressure (boiling point = 212 o F) has an enthalpy of approximately 1287 BTU per pound. We may
calculate the heat lost by one pound of this superheated steam as it cools from 500 o F to 32 o F in
each of the three steps previously described. Here, we will assume a specific heat for steam of 0.476,
a specific heat for water of 1, and a latent heat of vaporization for water of 970:
Enthalpy values are very useful43 in steam engineering to predict the amount of thermal energy
delivered to a load given the steam’s initial temperature, its final (cooled) temperature, and the mass
flow rate. Although the definition of enthalpy – where we calculate heat value by supposing the vapor
cools all the way down to the freezing point of water – might seem a bit strange and impractical (how
common is it for steam to lose so much heat to a process that it reaches freezing temperature?), it is
not difficult to shift the enthalpy value to reflect a more practical final temperature. Since we know
the specific heat of liquid water is very nearly one, all we have to do is offset the enthalpy value by
the amount that the final temperature differs from freezing in order to calculate how much heat the
steam will lose (per pound) to its load44 .
Furthermore, the rate at which heat is delivered to a substance by steam (or conversely, the rate
at which heat is required to boil water into steam) may be easily calculated if we take this heat
value in units of BTU per pound and multiply it by the mass flow rate in pounds per minute: as the
unit of “pound” cancels in the multiplication, we arrive at a result for heat transfer rate in units of
BTU per minute.
43 At first it may seem as though the enthalpy of steam is so easy to calculate it almost renders steam tables useless.
If the specific heats of water and steam were constant, and the latent heat of vaporization for water likewise constant,
this would be the case. However, both these values (c and L) are not constant, but rather change with pressure and
with temperature. Thus, steam tables end up being quite valuable to engineers, allowing them to quickly reference
heat content of steam across a broad range of pressures and temperatures without having to account for changing c
R T2
and L values (performing integral calculus in the form of Q = m c dT for specific heat) in their heat calculations.
T1
44 This is not unlike calculating the voltage dropped across an electrical load by measuring the voltage at each of
the load’s two terminals with respect to ground, then subtracting those two measured voltage values. In this analogy,
electrical “ground” is the equivalent of water at freezing temperature: a common reference point for energy level.
3.4. ELEMENTARY THERMODYNAMICS 73
For example, suppose we were to employ the same 500 o F superheated steam used in the previous
example to heat a flow of oil through a heat exchanger, with the steam condensing to water and
then cooling down to 170 degrees Fahrenheit as it delivers heat to the flowing oil. Here, the steam’s
enthalpy value of 1287 BTU per pound may simply be offset by 138 (170 degrees minus 32 degrees)
to calculate how much heat (per pound) this steam will deliver to the oil: 1287 − 138 = 1149 BTU
per pound:
1149 BTU/lb
Cold oil Hot oil (Heat liberated by steam as
1287 BTU/lb it cools from 500 oF to 170 oF)
o
(Enthalpy of 500 F steam)
170 oF water
138 BTU/lb
(138 BTU/lb) (Enthalpy of 170 oF water)
o
o
170 F water 32 F water
Here we see how 500 o F steam has an enthalpy (total heat) of 1287 BTU/lb, but since the steam
does not in fact cool all the way down to 32 o F in the act of heating oil in the heat exchanger, we
must subtract the enthalpy of the 170 o F water (138 BTU/lb) to determine45 the amount of heat
actually delivered to the oil by the steam (1149 BTU/lb). Calculating heat transfer rate is a simple
matter of multiplying this heat per pound of steam by the steam’s mass flow rate: for example, if
the mass flow rate of this steam happened to be 2 pounds per minute, the heat transfer rate would
be 2298 BTU per minute.
If we happen to be dealing with a situation where steam gives up some heat energy to a process
fluid but not enough to cool to the point of condensation, all we need to do to calculate the amount
of heat liberated by the superheated steam as it cools is subtract the enthalpy values between its
hot and cool(er) states.
For example, suppose we have a heat-exchange process where superheated steam enters at 105
PSIG and 600 o F, exiting at 75 PSIG and 360 o F. The enthalpy of the steam under those two sets of
conditions as given by a superheated steam table are 1328 BTU/lb and 1208 BTU/lb, respectively.
Thus, the heat lost by the steam as it goes through this heat exchanger is the difference in enthalpy
values: 1328 BTU/lb − 1208 BTU/lb = 120 BTU/lb. Once again, calculating heat transfer rate is
a simple matter of multiplication: if the mass flow rate of this steam happened to be 80 pounds per
hour, the heat transfer rate would be 120 BTU/lb × 80 lb/hr = 9600 BTU/hr.
45 Applying the maximum available balance analogy to this scenario, it would be as if your bank account began
with a maximum available balance of $1287 and then finished with a maximum available balance of $138 after an
expenditure: the amount of money you spent is the different between the initial and final maximum available balances
($1287 − $138 = $1149).
74 CHAPTER 3. FULL TUTORIAL
By encompassing both specific heat and latent heat into one quantity, enthalpy figures given
in steam tables greatly simplify heat transfer calculations, as compared to evaluating specific heat
and latent heat formulae (Q = mc∆T and Q = mL, respectively) for water. Calculations based on
steam tables are also more accurate than those derived from the specific and/or latent heat formulae,
because steam tables take into account the changing values of c and L over wide temperature and
pressure ranges. This is the power of empirical data: steam tables were developed by taking actual
calorimetric measurements of steam under those temperature and pressure conditions, and as such
are a record of water’s true behavior rather than a prediction of water’s theoretical behavior.
3.4. ELEMENTARY THERMODYNAMICS 75
Pcritical
218 atm Critical point
Freezing
Melting
Liquid
Pressure
Solid
(P)
Condensing
Boiling
Ptriple Vapor
0.006 atm
Sublimation
Temperature
Ttriple (T) Tcritical
0.01 deg C 374 deg C
This phase diagram (for water) illustrates some of the features common to all phase diagrams:
curved lines define the boundaries between solid, liquid, and vapor phases; the point of intersection
of these three curves is where the substance may exist in all three phases simultaneously (called the
triple point 46 ) and points where a curve simply ends within the span of the graph indicate critical
points, where the certain phases cease to exist.
The curved line from the triple point up and to the right defines the boundary between liquid
water and water vapor. Each point on that line represents a set of unique pressure and temperature
conditions for boiling (changing phase from liquid to vapor) or for condensation (changing phase
from vapor to liquid). As you can see, increased pressure results in an increased boiling point (i.e.
at higher pressures, water must be heated to greater temperatures before boiling may take place).
46 When H O is at its triple point, vapor (steam), liquid (water), and solid (ice) of water will co-exist in the same
2
space. One way to visualize the triple point is to consider it the pressure at which the boiling and freezing temperatures
of a substance become the same.
76 CHAPTER 3. FULL TUTORIAL
In fact, the whole concept of a singular boiling point for water becomes quaint in the light of a
phase diagram: boiling is seen to occur over a wide range of temperatures47 , the exact temperature
varying with pressure.
Something interesting happens when the temperature is raised above a certain value called the
critical temperature. At this value (approximately 374 degrees Celsius for water), no amount of
pressure will maintain it in a liquid state. Water, once heated beyond 374 degrees Celsius, is no
longer a liquid and may only exist in a stable condition as a vapor. The critical pressure of any
substance is the pressure value at the liquid/vapor boundary at the point of critical temperature.
A vivid example of critical temperature is this photograph of an ultra-high pressure storage vessel
for oxygen gas, at a rocket engine testing facility:
The critical temperature for oxygen is 154.58 Kelvin, which is equal to −118.57 degrees Celsius
or −181.43 degrees Fahrenheit. Since this pressure vessel is completely uninsulated, we know the
temperature of the oxygen inside of it will be the same (or nearly the same) as ambient temperature,
which is obviously much warmer than −118.57 o C. Since the oxygen’s temperature is well above the
critical temperature for the element oxygen, we may safely conclude that the oxygen inside this
vessel must exist as a gas. Even at the extremely high pressure this vessel is capable of holding
(15000 PSIG), the oxygen cannot liquefy.
The slightly curved line from the triple point up and to the left defines the boundary between
solid ice and liquid water. As you can see, the near-vertical pitch of this curve suggests the freezing
temperature of water is quite stable over a wide pressure range.
47 Anywhere between the triple-point temperature and the critical temperature, to be exact.
3.4. ELEMENTARY THERMODYNAMICS 77
Carbon dioxide exhibits a different set of curves than water on its phase diagram, with its own
unique critical temperature and pressure values:
Pcritical
72.9 atm
Freezing Critical point
Melting
Liquid
Pressure
Solid
(P)
Condensing
Boiling
Ptriple Vapor
5.11 atm
Sublimation
Temperature
Ttriple (T) Tcritical
-56.6 deg C 31 deg C
Note how the triple-point pressure of carbon dioxide is well above ambient conditions on Earth.
This means carbon dioxide is not stable in its liquid state unless put under substantial pressure (at
least 60.4 PSIG). This is why solid carbon dioxide is referred to as dry ice: it does not liquefy with
the application of heat, rather it sublimates directly into its vapor phase.
Another interesting difference between the carbon dioxide and water phase diagrams is the slope
of the solid/liquid boundary line. With water, this boundary drifts to the left (lower temperature)
as pressure increases. With carbon dioxide, this boundary drifts to the right (higher temperature) as
pressure increases. Whether the fusion temperature increases or decreases with increasing pressure
marks whether that substance contracts or expands as it transitions from liquid to solid. Carbon
dioxide, like most pure substances, contracts to a smaller volume when it goes from liquid to solid,
and its fusion curve drifts to the right as pressure increases. Water is unusual in this regard,
expanding to a larger volume when freezing, and its fusion curve drifts to the left as pressure
increases.
78 CHAPTER 3. FULL TUTORIAL
Pressure gauge
Thermometer
To water Vessel
pump
discharge
Valve
To gas Burner
fuel supply
So long as the water is all liquid (one phase), we may adjust its pressure and temperature
independently. In this state, the system has two thermodynamic degrees of freedom.
However, if the water becomes hot enough to boil, creating a system of two phases in direct
contact with each other (equilibrium), we will find that pressure and temperature become linked:
one cannot alter one without altering the other. For a steam boiler, operation at a given steam
3.4. ELEMENTARY THERMODYNAMICS 79
pressure thus defines the temperature of the water, and vice-versa. In a single-component, two-
phase system, there is only one degree of thermodynamic freedom.
Our freedom to alter pressure and temperature becomes even more restricted if we ever reach the
triple point 48 of the substance. For water, this occurs (only) at a pressure of −14.61 PSIG (0.006
atmospheres) and a temperature of 0.01 degrees Celsius: the coordinates where all three phase
transition curves intersect on the phase diagram. In this state, where solid (ice), liquid (water), and
vapor (steam) coexist, there are zero degrees of thermodynamic freedom. Both the temperature and
pressure are locked at these values until one or more of the phases disappears.
The relationship between degrees of freedom and phases is expressed neatly by Gibbs’ Phase Rule
– the sum of phases and degrees of freedom equals the number of substances (“components”) plus
two:
48 The triple point for any substance is the pressure at which the boiling and freezing temperatures become one and
the same.
49 The non-freedom of both pressure and temperature for a pure substance at its triple point means we may
exploit different substances’ triple points as calibration standards for both pressure and temperature. Using suitable
laboratory equipment and samples of sufficient purity, anyone in the world may force a substance to its triple point
and calibrate pressure and/or temperature instruments against that sample.
80 CHAPTER 3. FULL TUTORIAL
same pressure
Propane
Propane vapor
vapor
Propane
vapor Liquid
Liquid propane
propane
Liquid
propane
This is counter-intuitive, as most people tend to think the fullest tank should register the highest
pressure (having the least space for the vapor to occupy). However, since the interior of each tank
is a liquid/vapor system in equilibrium, the pressure is defined by the point on the liquid/vapor
transition curve on the phase diagram for pure propane matching the tanks’ temperature. Thus,
the pressure gauge on each tank actually functions as a thermometer 50 , since pressure is a direct
function of temperature for a saturated liquid/vapor system and therefore cannot change without
a corresponding change in temperature. This is a thermodynamic system with just one degree of
freedom.
Storage tanks containing liquid/vapor mixtures in equilibrium present unique safety hazards. If
ever a rupture were to occur in such a vessel, the resulting decrease in pressure causes the liquid to
spontaneously boil, halting any further decrease in pressure. Thus, a punctured propane tank does
not lose pressure in the same manner than a punctured compressed air tank loses pressure. This
gives the escaping vapor more “power” to worsen the rupture, as its pressure does not fall off over
time the way it would in a simple compressed-gas application. As a result, relatively small punctures
can and often do grow into catastrophic ruptures, where all liquid previously held inside the tank
escapes and flashes into vapor, generating a vapor cloud of surprisingly large volume51 .
Compounding the problem of catastrophic tank rupture is the fact that propane happens to be
highly flammable. The thermodynamic properties of a boiling liquid combined with the chemical
50 To be more precise, a propane tank acts like a Class II filled-bulb thermometer, with liquid and vapor coexisting
in equilibrium.
51 Steam boilers exhibit this same explosive tendency. The expansion ratio of water to steam is on the order of a
thousand to one (1000:1), making steam boiler ruptures very violent even at relatively low operating pressures.
3.4. ELEMENTARY THERMODYNAMICS 81
property of flammability in air makes propane tank explosions particularly violent. Fire fighters
often refer to this as a BLEVE : a Boiling Liquid Expanding Vapor Explosion.
82 CHAPTER 3. FULL TUTORIAL
Vapor
Vapor Bulb Bulb
Volatile
liquid
Volatile
liquid
Heat applied to the bulb literally “boils” the liquid inside until its pressure reaches the equilibrium
point with temperature. As the bulb’s temperature increases, so does the pressure throughout the
sealed system, indicating at the operator display where a bellows (or some other pressure-sensing
element) moves a pointer across a calibrated scale.
The only difference between the two filled-bulb thermometers shown in the illustration is which
end of the instrument is warmer. The Class IIA system on the left (where liquid fills the pressure-
indicating element) is warmer at the bulb than at the indicating end. The Class IIB system on
the right (where vapor fills the indicating bellows) has a cooler bulb than the indicating bellows.
The long length and small internal diameter of the connecting tube prevents any substantial heat
transfer from one end of the system to the other, allowing the sensing bulb to easily be at a different
temperature than the indicating bellows. Both types of Class II thermometers work the same52 ,
the indicated pressure being a strict function of the bulb’s temperature where the liquid and vapor
coexist in equilibrium.
52 Class IIA systems do suffer from elevation error where the indicator may read a higher or lower temperature than
it should due to hydrostatic pressure exerted by the column of liquid inside the tube connecting the indicator to the
sensing bulb. Class IIB systems do not suffer from this problem, as the gas inside the tube exerts no pressure over an
elevation.
3.4. ELEMENTARY THERMODYNAMICS 83
Control rods
Steam "bubble"
Pressurizer 2100 PSIG
644 oF
Boiling water
Steam to
Non-boiling water turbines
Reactor
pressure Steam
vessel Generator
Water from
575 oF 2100 PSIG
condensers
Core
(contains
fuel pellets) Cooling water
in (cold)
In order to maintain a liquid-only cooling environment for the reactor core, the water is held at
a pressure too high for boiling to occur inside the reactor vessel. Typical operating conditions for a
pressurized water reactor are 575 o F and 2100 PSIG. A steam table shows the boiling point of water
at 2100 PSIG to be over 640 o F, which means the water inside the reactor cannot boil if the reactor
only operates at 575 o F. Referencing the phase diagram for water, the operating point of the reactor
core is maintained above the liquid/vapor phase transition line by an externally supplied pressure.
This excess pressure comes from a device in the primary coolant loop called a pressurizer. Inside
the pressurizer is an array of immersion-style electric heater elements. The pressurizer is essentially
an electric boiler, purposely boiling the water inside at a temperature greater54 than that reached
53 Circulation pumps and a multitude of accessory devices are omitted from this diagram for the sake of simplicity.
54 This is another example of an important thermodynamic concept: the distinction between heat and temperature.
While the temperature of the pressurizer heating elements exceeds that of the reactor core, the total heat output
of course does not. Typical comparative values for pressurizer power versus reactor core power are 1800 kW versus
84 CHAPTER 3. FULL TUTORIAL
by the reactor core itself. For the example figure of 2100 PSIG, the pressurizer elements would have
to operate at a temperature of approximately 644 o F to maintain a boiling condition inside the
pressurizer.
By maintaining the water temperature inside the pressurizer greater than at the reactor core,
the water flowing through the reactor core literally cannot boil. The water/vapor equilibrium inside
the pressurizer is a system with one degree of freedom (pressure and temperature linked), while
the water-only environment inside the reactor core has two degrees of freedom (temperature may
vary to any amount below the pressurizer’s temperature without water pressure changing at all).
Thus, the pressurizer functions like the temperature-sensing bulb of a gigantic Class IIA filled-bulb
thermometer, with a liquid/vapor equilibrium inside the pressurizer vessel and liquid only inside the
reactor vessel and all other portions of the primary coolant loop. Reactor pressure is then controlled
by the temperature inside the pressurizer, which is in turn controlled by the amount of power applied
to the heating element array55 .
Steam boilers
Boilers in general (the nuclear reactor system previously described being just one example of a large
“power” boiler) are outstanding examples of phase change applied to practical use. The purpose of
a boiler is to convert water into steam, sometimes for heating purposes, sometimes as a means of
producing mechanical power (through a steam engine), sometimes for chemical processes requiring
pressurized steam as a reactant, sometimes for utility purposes (maintenance-related cleaning,
process vessel purging, sanitary disinfection, fire suppression, etc.) or all of the above. Steam
is a tremendously useful substance in many industries, so you will find boilers in use at almost every
industrial facility.
3800 MW, respectively: a ratio exceeding three orders of magnitude. The pressurizer heating elements don’t have to
dissipate much power (compared to the reactor core) because the pressurizer is not being cooled by a forced convection
of water like the reactor core is.
55 In this application, the heaters are the final control element for the reactor pressure control system.
3.4. ELEMENTARY THERMODYNAMICS 85
Exhaust stack
Saturated
steam out
Steam drum
Feedwater
in
Economizer
Water
tubes
Flame
"Mud" drum
Blowdown valve
Water enters the boiler through a heat exchanger in the stack called an economizer. This allows
cold water to be pre-heated by the warm exhaust gases before they exit the stack. After pre-heating
in the economizer, the water enters the boiler itself, where water circulates by natural convection
(“thermosiphon”) through a set of tubes exposed to high-temperature fire. Steam collects in the
“steam drum,” where it is drawn off through a pipe at the top. Since this steam is in direct contact
with the boiling water, it will be at the same temperature as the water, and the steam/water
environment inside the steam drum represents a two-phase system with only one degree of freedom.
With just a single degree of freedom, steam temperature and pressure are direct functions of each
other – coordinates at a single point along the liquid/vapor phase transition line of water’s phase
diagram. One cannot change one variable without changing the other.
Consulting a steam table56 , you will find that the temperature required to boil water at a pressure
of 120 PSIG is approximately 350 degrees Fahrenheit. Thus, the temperature of the steam drum will
be fixed at 350 o F while generating steam pressure at 120 PSIG. The only way to increase pressure
in that boiler is to increase its temperature, and vice-versa.
When steam is at the same temperature as the boiling water it came from, it is referred to as
saturated steam. Steam in this form is very useful for heating and cleaning, but not as much for
operating mechanical engines or for process chemistry. If saturated steam loses any temperature at
all (by losing its latent heat), it immediately condenses back into water. Liquid water can cause
major mechanical problems inside steam engines (although “wet” steam works wonderfully well as
a cleaning agent!), and so steam must be made completely “dry” for some process applications.
56 Since the relationship between saturated steam pressure and temperature does not follow a simple mathematical
formula, it is more practical to consult published tables of pressure/temperature data for steam. A great many
engineering manuals contain steam tables, and in fact entire books exist devoted to nothing but steam tables.
86 CHAPTER 3. FULL TUTORIAL
The way this is done is by a process known as superheating. If steam exiting the steam drum
of a boiler is fed through another heat exchanger inside the firebox so it may receive more heat, its
temperature will rise beyond the saturation point. This steam is now said to be superheated :
Exhaust stack
Saturated Superheated
steam out steam out
Steam drum
Feedwater
in
Super-
Economizer heater
Water
tubes
Flame
"Mud" drum
Blowdown valve
Superheated steam is absolutely dry, containing no liquid water at all. It is therefore safe to
use as a fluid medium for engines (piston and turbine alike) and as a process reactant where liquid
water is not tolerable. The difference in temperature between superheated steam and saturated
steam at any given pressure is the amount of superheat. For example, if saturated steam at 350
degrees Fahrenheit and 120 PSI drawn from the top of the steam drum in a boiler is heated to a
higher temperature of 380 degrees Fahrenheit (at the same pressure of 120 PSI), it is said to have
30 degrees (Fahrenheit) of superheat.
3.4. ELEMENTARY THERMODYNAMICS 87
57 An experiment illustrative of this point is to maintain an ice-water mixture in an open container, then to insert
a sealed balloon containing liquid water into this mixture. The water inside the balloon will eventually equalize in
temperature with the surrounding ice-water mix, but it will not itself freeze. Once the balloon’s water reaches 0
degrees Celsius, it stops losing heat to the surrounding ice-water mix, and therefore cannot make the phase change
to solid form.
88 CHAPTER 3. FULL TUTORIAL
Warm, moist
air out
Motor
Fan
Hot water
in
Fill Fill
Cool water
out Cooled water
Smaller evaporative cooling towers use fans to force air upward through the tower, employing
inert “fill” material to provide large amounts of surface area for the liquid water and the air to
contact each other. Some large evaporative cooling towers are self-drafting, the heat of the water
providing enough convective force to the air that no fans are needed.
3.4. ELEMENTARY THERMODYNAMICS 89
The following photograph shows a pair of induced-draft evaporative cooling towers used at a
beer brewery:
This next photograph shows a forced-draft evaporative cooling tower used at a coal-fired electric
power plant. Note the large air fans located around the circumference of the cooling tower, blowing
cool air into the tower from outside. This fan placement eliminates problems arising from having
the fan blades and motor located within the moist air stream:
90 CHAPTER 3. FULL TUTORIAL
58 An interesting point to make here is the United States did get something right when they designed their monetary
system of dollars and cents. This is essentially a metric system of measurement, with 100 cents per dollar. The
founders of the USA wisely decided to avoid the utterly confusing denominations of the British, with their pounds,
pence, farthings, shillings, etc. The denominations of penny, dime, dollar, and eagle ($10 gold coin) comprised a
simple power-of-ten system for money. Credit goes to France for first adopting a metric system of general weights
and measures as their national standard.
3.5. UNIT CONVERSIONS AND PHYSICAL CONSTANTS 91
P = IV
Substituting units of measurement for each variable in this formula (i.e. Watts for power,
Amperes for current, and Volts for voltage), using bracket symbols to denote these as unit
abbreviations rather than variables, we get this result:
35 qt = ??? gal
Now, most people know there are four quarts in one gallon, and so it is tempting to simply
divide the number 35 by four to arrive at the proper number of gallons. However, the purpose of
this example is to show you how the technique of unity fractions works, not to get an answer to a
problem.
92 CHAPTER 3. FULL TUTORIAL
To demonstrate the unity fraction technique, we will first write the original quantity as a fraction,
in this case a fraction with 1 as the denominator:
35 qt
1
Next, we will multiply this fraction by another fraction having a physical value of unity (1) so
that we do not alter59 the quantity. This means a fraction comprised of equal measures in the
numerator and denominator, but having different units of measurement. This “unity” fraction must
be arranged in such a way that the undesired unit cancels out and leaves only the desired unit(s)
in the product. In this particular example, we wish to cancel out quarts and end up with gallons,
so we must arrange a fraction consisting of quarts and gallons having equal quantities in numerator
and denominator, such that quarts will cancel and gallons will remain:
35 qt 1 gal
1 4 qt
Now we see how the unit of “quarts” cancels from the numerator of the first fraction and the
denominator of the second (“unity”) fraction, with “gallons” being the only unit remaining:
35 qt 1 gal
= 8.75 gal
1 4 qt
One reason this conversion technique is so powerful is it allows one to perform the largest range
of unit conversions while memorizing the smallest possible set of conversion factors. Another reason
is that when written out like this, showing canceled units, anyone examining this work will be able to
tell how and why this conversion is correct. In other words, the “unity fraction” technique literally
teaches how to convert between units using a bare minimum of memorized equivalent measures.
59 A basic mathematical identity is that multiplication of any quantity by 1 does not change the value of that original
quantity. If we multiply some quantity by a fraction having a physical value of 1, no matter how strange-looking
that fraction may appear, the value of the original quantity will be left intact. The goal here is to judiciously choose
a fraction with a physical value of 1 but with its units of measurement so arranged that we cancel out the original
quantity’s unit(s) and replace them with the units we desire.
3.5. UNIT CONVERSIONS AND PHYSICAL CONSTANTS 93
Here is a set of six equal volumes, each one expressed in a different unit of measurement:
1 gallon (gal) = 231.0 cubic inches (in3 ) = 4 quarts (qt) = 8 pints (pt) = 128 fluid ounces (fl. oz.)
= 3.7854 liters (l)
Since all six of these quantities are physically equal, it is possible to build a “unity fraction” out
of any two, to use in converting any of the represented volume units into any of the other represented
volume units. Shown here are a few different volume unit conversion problems, using unity fractions
built only from these factors (all canceled units shown using strike-out lines):
40 gallons converted into fluid ounces (using 128 fl. oz. = 1 gal in the unity fraction):
40 gal 128 fl. oz.
= 5120 fl. oz
1 1 gal
5.5 pints converted into cubic inches (using 231 in3 = 8 pt in the unity fraction):
231 in3
5.5 pt
= 158.8 in3
1 8 pt
By contrast, if we were to try to memorize a 6 × 6 table giving conversion factors between any
two of six volume units, we would have to commit 30 different conversion factors to memory! Clearly,
the ability to set up “unity fractions” is a much more memory-efficient and practical approach.
94 CHAPTER 3. FULL TUTORIAL
This economy of conversion factors is very useful, and may also be extended to cases where linear
units are raised to powers to represent two- or three-dimensional quantities. To illustrate, suppose
we wished to convert 5.5 pints into cubic feet instead of cubic inches: with no conversion equivalence
between pints and cubic feet included in our string of six equalities, what do we do?
We should know the equality between inches and feet: there are exactly 12 inches in 1 foot.
This simple fact may be applied by incorporating another unity fraction in the original problem to
convert cubic inches into cubic feet. We will begin by including another unity fraction comprised of
12 inches and 1 foot,just to see how this might work:
231 in3
5.5 pt 1 ft
= 13.23 in2 · ft
1 8 pt 12 in
1 ft
Unfortunately, this yields a non-sensical unit of square inch-feet. Even though 12 in is a valid
unity fraction, it does not completely cancel out the unit of cubic inches in the numerator of the
first unity fraction. Instead, the unit of “inches” in the denominator of the unity fraction merely
cancels out one of the “inches” in the “cubic inches” of the previous fraction’s numerator, leaving
square inches (in2 ). What we need for full cancellation of cubic inches is a unity fraction relating
1 ft
cubic feet to cubic inches. We can get this, though, simply by cubing the 12 in unity fraction:
5.5 pints converted into cubic feet (our second attempt! ):
3
231 in3
5.5 pt 1 ft
1 8 pt 12 in
Distributing the third power to the interior terms of the last unity fraction:
3 3
231 in3
5.5 pt 1 ft
1 8 pt 123 in3
Calculating the values of 13 and 123 inside the last unity fraction, then canceling units and
solving:
Once again, this unit conversion technique shows its power by minimizing the number of
conversion factors we must memorize. We need not memorize how many cubic inches are in a
cubic foot, or how many square inches are in a square foot, if we know how many linear inches are in
a linear foot and we simply let the fractions “tell” us whether a power is needed for unit cancellation.
Unity fractions are also useful when we need to convert more than one unit in a given quantity.
For example, suppose a flowmeter at a wastewater treatment facility gave us a flow measurement
of 205 cubic feet per minute but we needed to convert this expression of water flow into units of
cubic yards per day. Observe the following unit-fraction conversion to see how unity fractions serve
the purpose of converting cubic feet into cubic yards, and minutes into days (by way of minutes to
hours, and hours to days):
205 ft3
3 3
1 yd 60 min 24 hr
= 10933.3 yd3 /day
min 33 ft3 1 hr 1 day
Note how the only units left un-canceled on the left-hand side of the “equals” symbol are cubic
yards (yd3 ) and days, which therefore become the units of measurement for the final result.
A major caveat to this method of converting units is that the units must be directly proportional
to one another, since this multiplicative conversion method is really nothing more than an exercise
in mathematical proportions. Here are some examples (but not an exhaustive list!) of conversions
that cannot be performed using the “unity fraction” method:
• Absolute / Gauge pressures, because one scale is offset from the other by 14.7 PSI (atmospheric
pressure).
• Celsius / Fahrenheit, because one scale is offset from the other by 32 degrees.
• Wire diameter / gauge number, because gauge numbers grow smaller as wire diameter grows
larger (inverse proportion rather than direct) and because there is no proportion relating the
two.
The following subsections give sets of physically equal quantities, which may be used to create
unity fractions for unit conversion problems. Note that only those quantities shown in the same line
(separated by = symbols) are truly equal to each other, not quantities appearing in different lines!
96 CHAPTER 3. FULL TUTORIAL
1 gallon (gal) = 231.0 cubic inches (in3 ) = 4 quarts (qt) = 8 pints (pt) = 16 cups = 128 fluid
ounces (fl. oz.) = 3.7854 liters (l)
1 mile per hour (mi/h) = 88 feet per minute (ft/m) = 1.46667 feet per second (ft/s) = 1.60934
kilometer per hour (km/h) = 0.44704 meter per second (m/s) = 0.868976 knot (knot – international)
1 acre = 43560 square feet (ft2 ) = 4840 square yards (yd2 ) = 4046.86 square meters (m2 )
3.5.9 Conversion factors for pressure (either all gauge or all absolute)
Note: all conversion factors shown in bold type are exact, not approximations.
1 pounds per square inch (PSI) = 2.03602 inches of mercury at 0 o C (in. Hg) = 27.6799 inches of
water at 4 o C (in. W.C.) = 6.894757 kilopascals (kPa) = 0.06894757 bar
1 bar = 100 kilopascals (kPa) = 14.504 pounds per square inch (PSI)
1 standard atmosphere (Atm) = 14.7 pounds per square inch absolute (PSIA) = 101.325
kilopascals absolute (kPaA) = 1.01325 bar absolute = 760 millimeters of mercury absolute
(mmHgA) = 760 torr (torr)
1 horsepower = 550 foot-pounds per second (ft-lbf/s) = 745.7 watts (W) = 2544.43 British
thermal units per hour (Btu/h) = 0.0760181 boiler horsepower (hp – boiler)
98 CHAPTER 3. FULL TUTORIAL
Chapter 4
This chapter is where you will find mathematical derivations too detailed to include in the tutorial,
and/or tables and other technical reference material.
99
100 CHAPTER 4. DERIVATIONS AND TECHNICAL REFERENCES
P = IV
Where,
P = Power (Watts)
I = Current (Amperes)
V = Voltage (Volts)
Each of the units of measurement in the above formula (Watt, Ampere, Volt) are actually
comprised of more fundamental physical units. One Watt of power is one Joule of energy exchanged
per second of time. One Ampere of current is one Coulomb of electric charge moving per second.
One Volt of potential is one Joule of energy per Coulomb of electric charge. When we write the
equation showing these units in their proper orientations, we see that the result (power in Watts,
or Joules per second) actually does agree with the units for Amperes and Volts because the unit
of electric charge (Coulombs) cancels out. In dimensional analysis we customarily distinguish unit
symbols from variables by using non-italicized letters and surrounding each one with square brackets:
P = IV
[kg][m2 ]
[J] =
[s2 ]
Within the metric system of measurements, an international standard exists for which units
are considered fundamental and which are considered “derived” from the fundamental units. The
modern standard is called SI, which stands for Système International. This standard recognizes
seven fundamental, or base units, from which all others are derived1 :
An older standard existed for base units, in which the centimeter, gram, and second comprised
the first three base units. This standard is referred to as the cgs system, in contrast to the SI
system2 . You will still encounter some derived cgs units used, including the poise and the stokes
(both used to express fluid viscosity). Then of course we have the British engineering system which
uses such wonderful3 units as feet, pounds, and (thankfully) seconds. Despite the fact that the
majority of the world uses the metric (SI) system for weights and measures, the British system is
sometimes referred to as the Customary system.
1 The only exception to this rule being units of measurement for angles, over which there has not yet been full
agreement whether the unit of the radian (and its solid counterpart, the steradian) is a base unit or a derived unit.
2 The older name for the SI system was “MKS,” representing meters, kilograms, and seconds.
3 I’m noting my sarcasm here, just in case you are immune to my odd sense of humor.
102 CHAPTER 4. DERIVATIONS AND TECHNICAL REFERENCES
1. An object at rest tends to stay at rest; an object in motion tends to stay in motion
2. The acceleration of an object is directly proportional to the net force acting upon it and
inversely proportional to the object’s mass
Newton’s first law may be thought of as the law of inertia, because it describes the property
of inertia that all objects having mass exhibit: resistance to change in velocity. This law is quite
counter-intuitive for many people, who tend to believe that objects require continual force to keep
moving. While this is true for objects experiencing friction, it is not for ideal (frictionless) motion.
This is why satellites and other objects in space continue to travel with no mode of propulsion: they
simply “coast” indefinitely on their own inertia because there is no friction in space to dissipate their
kinetic energy and slow them down.
Newton’s second law is the verbal equivalent of the force/mass/acceleration formula: F = ma.
This law elaborates on the first, in that it mathematically relates force and motion in a very precise
way. For a frictionless object, the change in velocity (i.e. its acceleration) is proportional to force.
This is why a frictionless object may continue to move without any force applied: once moving,
force would only be necessary for continued acceleration. If zero force is applied, the acceleration
will likewise be zero, and the object will maintain its velocity indefinitely (again, assuming no friction
at work).
Newton’s third law describes how forces always exist in pairs between two objects. The
rotating blades of a helicopter, for example, exert a downward force on the air (accelerating the
air), but the air in turn exerts an upward force on the helicopter (suspending it in flight). A spider
hanging on the end of a thread exerts a downward force (weight) on the thread, while the thread
exerts an upward force of equal magnitude on the spider (tension). Force pairs are always equal in
magnitude but opposite in direction.
4.4. ROTATIONAL MOTION 103
Right-hand rule
for vector cross-products
right angle C A
right angle B
Radius vector
(r) (r)
radius
C=A×B
This relationship may be expressed mathematically as a vector cross-product, where the vector
directions are shown by the right-hand rule (the first vector ~r is the direction of the index finger,
the second vector F~ is the direction of the middle finger, and the product vector ~τ is the direction
of the thumb, with all three vectors perpendicular to each other):
~τ = ~r × F~
Labeling force, radius, and torque as vectors is the formally correct way of noting the variables
in a mechanical system such as this, and is the way college students studying physics typically learn
the calculation of torque. In less academic settings, the force vector (F~ ) is typically labeled as a
force along the line of action, and the radius vector (~r) is called the moment arm, with the line of
action and moment arm always being perpendicular to each other.
The proper unit of measurement for torque is the product of the force unit and distance unit.
In the metric system, this is customarily the Newton-meter (N-m). In the British system, this is
customarily the foot-pound (ft-lb) or alternatively the pound-foot (lb-ft). Note that while these are
the exact same units as those used to express work, they are not the same types of quantities. Torque
is a vector cross-product, while work is a dot-product (W = F~ · ~x). The cross-product of two vectors
104 CHAPTER 4. DERIVATIONS AND TECHNICAL REFERENCES
is always another vector4 , while the dot-product of two vectors is always a scalar (direction-less)
quantity. Thus, torque always has a direction, whereas work or energy does not.
Direction of torque
F = 58 lb
Axis of rotation
r=
Mo 0.61
me fee
nt t
arm 90o
τ=r×F
τ = (0.61 ft) × (58 lb)
τ = 35.38 lb-ft
ac e of
n
tio
Lin
With the radius and force vectors at right angles to each other, torque is simply the product of
both. In many non-academic settings, torque is calculated this way as a scalar quantity, with the
direction of rotation determined by observation rather than by strict adherence to the right-hand
rule of vector cross products. In this example, we see the magnitude of torque as the simple product
of 58 pounds force and 0.61 feet of moment arm (35.38 lb-ft of torque), with the torque direction
obviously counter-clockwise as viewed from the head of the bolt.
4 Technically, it is a pseudovector, because it does not exhibit all the same properties of a true vector, but this is a
If we apply the same force to the wrench handle at a different angle (not perpendicular to the
handle), the resulting torque will be less. The radius vector (moment arm), however, will still
remain perpendicular to the force vector (line of action) – it just decreases in length. To determine
the placement of the radius vector, all one must do is draw a line straight from the axis of rotation
perpendicular to the line of action, then use trigonometry to calculate its magnitude:
Direction of torque
F = 58 lb
Axis of rotation r=0
15o .59 f
eet
Wr Mom 90o
en
ch
e n t arm
ha
nd
le
=0
.61
τ=r×F fee
t
actioof
n
τ = 34.22 lb-ft
Line
A very practical example of torque is in the action of meshing gears, transferring mechanical
power from one gear to another. Each gear effectively acts as a wheel, the point of contact between
gear teeth acting to transfer force perpendicular to the radius of each gear (wheel). Thus, torque
applied to one gear becomes a linear force at the meshing teeth, which translates into another torque
at the second gear:
Gear #1 Gear #2
The ratio of torques between two meshing gears is equal to the ratio of gear teeth:
τ1 n1
=
τ2 n2
Where,
τ1 = Torque of first gear
τ2 = Torque of second gear
n1 = Number of teeth on first gear
n2 = Number of teeth on second gear
106 CHAPTER 4. DERIVATIONS AND TECHNICAL REFERENCES
For example, if a small gear having 28 teeth meshes with a larger gear having 75 teeth, the torque
multiplication factor from the small gear to the large gear will be 75:28, or 2.679 to 1. A torque
of 40 lb-ft applied to the small gear will result in a torque of 107.1 lb-ft or torque generated at the
large gear. This ratio of gear teeth is called the gear ratio.
As gears multiply torque (τ ), they divide rotational speed (ω). Thus, the 75:28 tooth gear set
creates a multiplication of torque from the small gear to the large gear, and an identical reduction
ratio of speed from the small gear to the large gear. Given this ratio, the small gear will have to be
turned 2.679 revolutions in order to make the large gear turn just one revolution.
We may express gear speeds as another ratio of gear teeth, reciprocated in relation to torque:
ω1 n2
=
ω2 n1
Where,
ω1 = Rotational speed of first gear
ω2 = Rotational speed of second gear
n1 = Number of teeth on first gear
n2 = Number of teeth on second gear
In a set of meshed gears, the smaller gear will have the least torque and the greatest speed; the
larger gear will have the greatest torque and the least speed.
This is precisely how gear sets are used in industry: to transform torque and speed in mechanical
power systems. The complementary effects of a gear set on torque and speed is analogous to the
complementary effects that a transformer has on AC voltage and current: a step-up transformer
(having more turns of wire in the secondary coil than in the primary coil) will multiply voltage but
reduce (divide) current, both by the same turns ratio.
4.4. ROTATIONAL MOTION 107
Every quantity of force and motion which may be expressed in linear form has a rotational
equivalent. As we have seen, torque (τ ) is the rotational equivalent of force (F ). The following table
contrasts equivalent quantities for linear and rotational motion (all units are metric, shown in italic
font):
Linear quantity, symbol, and unit Rotational quantity, symbol, and unit
Force (F ) N Torque (τ ) N-m
Linear displacement (x) m Angular displacement (θ) radian
Linear velocity (v) m/s Angular velocity (ω) rad/s
Linear acceleration (a) m/s2 Angular acceleration (α) rad/s2
Mass (m) kg Moment of Inertia (I) kg-m2
Familiar equations for linear motion have rotational equivalents as well. For example, Newton’s
Second Law of motion states, “The acceleration of an object is directly proportional to the net
force acting upon it and inversely proportional to the object’s mass.” We may modify this law
for rotational motion by saying, “The angular acceleration of an object is directly proportional to
the net torque acting upon it and inversely proportional to the object’s moment of inertia.” The
mathematical expressions of both forms of Newton’s Second Law are as follows:
F = ma τ = Iα
The calculus-based relationships between displacement (x), velocity (v), and acceleration (a) find
parallels in the world of angular motion as well. Consider the following formula pairs, linear motion
on the left and angular motion on the right:
dx dθ
v= (Velocity as the time-derivative of displacement) ω=
dt dt
dv dω
a= (Acceleration as the time-derivative of velocity) α=
dt dt
d2 x d2 θ
a= (Acceleration as the second time-derivative of displacement) α=
dt2 dt2
108 CHAPTER 4. DERIVATIONS AND TECHNICAL REFERENCES
An object’s “moment of inertia” represents its angular inertia (opposition to changes in rotational
velocity), and is proportional to the object’s mass and to the square of its radius. Two objects having
the same mass will have different moments of inertia if there is a difference in the distribution of
their mass relative to radius. Thus, a hollow tube will have a greater moment of inertia than a solid
rod of equal mass, assuming an axis of rotation in the center of the tube/rod length:
axis of rotation
axis of rotation
m = 300 kg m = 300 kg
This is why flywheels 5 are designed to be as wide as possible, to maximize their moment of inertia
with a minimum of total mass.
The formula describing the amount of work done by a torque acting over an angular displacement
is remarkably similar to the formula describing the amount of work done by a force acting over a
linear displacement:
W = Fx W = τθ
The formula describing the amount of kinetic energy possessed by a spinning object is also similar
to the formula describing the amount of energy possessed by a linearly-traveling object:
1 1 2
Ek = mv 2 Ek = Iω
2 2
5 A “flywheel” is a disk on a shaft, designed to maintain rotary motion in the absence of a motivating torque for the
function of machines such as piston engines. The rotational kinetic energy stored by an engine’s flywheel is necessary
to give the pistons energy to compress the gas prior to the power stroke, during the times the other pistons are not
producing power.
4.5. PERIODIC TABLE OF THE ELEMENTS 109
H 1 He 2
Hydrogen
1.00794
Periodic Table of the Elements Helium
4.00260
Metalloids Nonmetals
1s1 1s2
Li 3 Be 4 Symbol K 19 Atomic number B 5 C 6 N 7 O 8 F 9 Ne 10
Lithium Beryllium Potassium Boron Carbon Nitrogen Oxygen Fluorine Neon
6.941 9.012182 Name 39.0983 10.81 12.011 14.0067 15.9994 18.9984 20.179
Atomic mass
2s1 2s2 4s1 (averaged according to 2p1 2p2 2p3 2p4 2p5 2p6
Electron occurence on earth)
Na 11 Mg 12 configuration Al 13 Si 14 P 15 S 16 Cl 17 Ar 18
Sodium Magnesium Aluminum Silicon Phosphorus Sulfur Chlorine Argon
22.989768 24.3050 26.9815 28.0855 30.9738 32.06 35.453 39.948
Metals
3s1 3s2 3p1 3p2 3p3 3p4 3p5 3p6
K 19 Ca 20 Sc 21 Ti 22 V 23 Cr 24 Mn 25 Fe 26 Co 27 Ni 28 Cu 29 Zn 30 Ga 31 Ge 32 As 33 Se 34 Br 35 Kr 36
Potassium Calcium Scandium Titanium Vanadium Chromium Manganese Iron Cobalt Nickel Copper Zinc Gallium Germanium Arsenic Selenium Bromine Krypton
39.0983 40.078 44.955910 47.88 50.9415 51.9961 54.93805 55.847 58.93320 58.69 63.546 65.39 69.723 72.61 74.92159 78.96 79.904 83.80
4s1 4s2 3d14s2 3d24s2 3d34s2 3d54s1 3d54s2 3d64s2 3d74s2 3d84s2 3d104s1 3d104s2 4p1 4p2 4p3 4p4 4p5 4p6
Rb 37 Sr 38 Y 39 Zr 40 Nb 41 Mo 42 Tc 43 Ru 44 Rh 45 Pd 46 Ag 47 Cd 48 In 49 Sn 50 Sb 51 Te 52 I 53 Xe 54
Rubidium Strontium Yttrium Zirconium Niobium Molybdenum Technetium Ruthenium Rhodium Palladium Silver Cadmium Indium Tin Antimony Tellurium Iodine Xenon
85.4678 87.62 88.90585 91.224 92.90638 95.94 (98) 101.07 102.90550 106.42 107.8682 112.411 114.82 118.710 121.75 127.60 126.905 131.30
5s1 5s2 4d15s2 4d25s2 4d45s1 4d55s1 4d55s2 4d75s1 4d85s1 4d105s0 4d105s1 4d105s2 5p1 5p2 5p3 5p4 5p5 5p6
Cs 55 Ba 56 57 - 71 Hf 72 Ta 73 W 74 Re 75 Os 76 Ir 77 Pt 78 Au 79 Hg 80 Tl 81 Pb 82 Bi 83 Po 84 At 85 Rn 86
Cesium Barium Lanthanide Hafnium Tantalum Tungsten Rhenium Osmium Iridium Platinum Gold Mercury Thallium Lead Bismuth Polonium Astatine Radon
132.90543 137.327 series 178.49 180.9479 183.85 186.207 190.2 192.22 195.08 196.96654 200.59 204.3833 207.2 208.98037 (209) (210) (222)
6s1 6s2 5d26s2 5d36s2 5d46s2 5d56s2 5d66s2 5d76s2 5d96s1 5d106s1 5d106s2 6p1 6p2 6p3 6p4 6p5 6p6
Fr 87 Ra 88 89 - 103 Unq 104 Unp 105 Unh 106 Uns 107 108 109
Francium Radium Actinide Unnilquadium Unnilpentium Unnilhexium Unnilseptium
(223) (226) series (261) (262) (263) (262)
7s1 7s2 6d27s2 6d37s2 6d47s2
La 57 Ce 58 Pr 59 Nd 60 Pm 61 Sm 62 Eu 63 Gd 64 Tb 65 Dy 66 Ho 67 Er 68 Tm 69 Yb 70 Lu 71
Lanthanide Lanthanum Cerium Praseodymium Neodymium Promethium Samarium Europium Gadolinium Terbium Dysprosium Holmium Erbium Thulium Ytterbium Lutetium
series 138.9055 140.115 140.90765 144.24 (145) 150.36 151.965 157.25 158.92534 162.50 164.93032 167.26 168.93421 173.04 174.967
5d16s2 4f15d16s2 4f36s2 4f46s2 4f56s2 4f66s2 4f76s2 4f75d16s2 4f96s2 4f106s2 4f116s2 4f126s2 4f136s2 4f146s2 4f145d16s2
Multiple attributes appear for each element in the table. Two of these attributes – atomic number
and atomic mass – are directly related to the number of particles in the nucleus of each atom. We
will examine the table’s entry for the element potassium (K) to explore these concepts.
Potassium has an atomic number (number of protons in the nucleus of each potassium atom)
of 19. This number defines the element. If we were somehow to add or subtract protons from
the nucleus of a potassium atom6 , it would cease being potassium and transmutate into a different
element. Note how every element in the table has its own unique atomic number, and how each of
these numbers is whole (no fractions or decimals).
The atomic mass or atomic weight shown for potassium is 39.0983. This quantity is the sum of
protons and neutrons found in the nucleus of each potassium atom. Like the atomic number (19),
6 The amount of energy required to rearrange particles in the nucleus for even just a single atom is tremendous, lying
well outside the energy ranges of chemical reactions. Such energy levels are the exclusive domain of nuclear reactions
and high-energy radiation (subatomic particles traveling at high velocity). The extremely large energy “investment”
required to alter an atom’s nucleus is why atomic identities are so stable. This is precisely why alchemists of antiquity
utterly failed to turn lead into gold: no materials, processes, or techniques they had at their disposal were capable
of the targeted energy necessary to dislodge three protons from a nucleus of lead (82 Pb) to that it would turn into a
nucleus of gold (79 Au). That, and the fact the alchemists had no clue about atomic structure to begin with, made
their endeavor fruitless.
110 CHAPTER 4. DERIVATIONS AND TECHNICAL REFERENCES
we would logically expect the atomic mass to be a whole number as well, since protons and neutrons
only come in whole quantities. The primary reason we see a non-whole number for potassium’s
atomic mass is that this table reflects the average atomic mass of potassium atoms as found in
nature. Some potassium atoms have atomic masses greater than 39, and some have atomic masses
less than 39. We know that the number of protons in every potassium atom is fixed (which is what
gives potassium its elemental identity), which means the only quantity that may cause the atomic
mass to vary is the number of neutrons in the nucleus. The most common form of potassium (39 K)
atom possesses 19 protons and 20 neutrons in its nucleus, giving it an atomic mass of 39 (19 + 20).
The next most common form of potassium found on Earth is (41 K), possessing 19 protons and 22
neutrons.
We refer to atoms of the same element with differing atomic masses as isotopes. From a chemical
perspective, isotopes are identical. That is to say, they engage in the exact same chemical reactions
in the exact same manner. To use potassium as an example, an atom of 39 K will join with a chlorine
atom (Cl) to form the compound potassium chloride (KCl) just as readily as an atom of 41 K will
join with a chlorine atom to form the same compound. The three isotopes of hydrogen (1 H, 2 H,
and 3 H: hydrogen, deuterium, and tritium, respectively) are all chemically identical: all are highly
flammable, combining with oxygen to create water (H2 O). However, deuterium (2 H) has twice the
density of normal hydrogen (1 H), while tritium (3 H) has three times the density of normal hydrogen
and is highly radioactive! Isotopes only differ in their mass and in their nuclear properties (such as
radioactivity: the tendency for a nucleus to spontaneously decay, usually resulting in a loss or gain
of protons that subsequently alters the identity of the decayed atom).
The Periodic Table is called “periodic” because its configuration reveals a repeating pattern of
chemical behaviors approximately following atomic number. Horizontal rows in the table are called
periods, while vertical columns are called groups. Elements in the same group (vertical column) share
similar chemical reactivities – that is, they tend to engage in the same types of chemical reactions
– despite having different masses and physical properties such as melting point, boiling point, etc.
This periodicity is a function of how electrons are arranged around the nucleus of each atom, a
subject we will explore in more detail later in this chapter. As mentioned previously, chemistry is
the study of how atoms bond together to form molecules, and this bonding takes place through the
interaction of the electrons surrounding the atoms’ nuclei. It makes perfect sense, then, that the
configuration of those electrons determine the chemical (bonding) properties of atoms.
4.5. PERIODIC TABLE OF THE ELEMENTS 111
Some periodic tables show the first ionization energy value for each element – the amount of
energy required to force the first electron of an electrically balanced atom to separate from that atom
– in addition to other attributes such as atomic number and atomic mass. If we note the ionization
energies of the elements, reading each element in turn from left-to-right, starting with period 1
(hydrogen and helium) and progressing to subsequent periods, we see an interesting pattern:
First ionization energy represents the relative stability of the last electron balancing the electrical
charge of an atom. We see from this table that 24.5874 electron-Volts of energy is needed to remove
one electron from an electrically-balanced atom of helium (changing He into He1+ ), while only
13.5984 electron-Volts of energy is required to do the same to an atom of hydrogen. This tells us
the electron configuration of helium is at a lower energy (and therefore more stable) than that of
hydrogen.
112 CHAPTER 4. DERIVATIONS AND TECHNICAL REFERENCES
The ionization energies increase with increasing atomic number (with an occasional down-step)
until the last column of the period is reached, and then there is a comparatively enormous down-step
in energy at the first column of a new period. This pattern is clearly evident when the first ionization
energies are plotted against atomic number:
25
20
Ionization 15
energy (eV)
10
5
0
1 5 10 15 20 25 30 35
Atomic number
This periodicity suggests that as atoms grow in atomic number, the additional electrons do not
simply pile on in random fashion or in a plain and simple progression from inner orbits to outer
orbits. Rather, they “fill in” a structured energy pattern, with major changes in structure at the
start of each new period. More details of this structured pattern will be explored later in this
chapter.
The low ionization energy values for all the “Group 1” elements (far left-hand column) suggest
they are relatively easy to positively ionize, and indeed we find this to be the case through
experimentation. Hydrogen, lithium, sodium, potassium, and the rest all readily become positively-
charged ions upon interaction with other atoms, since their low ionization energy values means they
may easily lose an electron.
The high ionization energy values for all the “Group 18” elements (far right-hand column) suggest
they possess a very stable electron structure, which is also verified by experiment. These are the
noble elements, possessing very little reactive potential7 .
Looking at the “Group 17” column, just to the left of the noble elements, we notice that they are
all just one electron shy of the stable electron structure enjoyed by the noble atoms when in their
electrically-balanced states. This suggests it might be easy to add one more electron to atoms of these
elements, which (once again!) is a principle validated by experiment. Fluorine, chlorine, bromine,
iodine, and even astatine8 all readily ionize negatively, readily accepting an extra electron from
surrounding atoms. As one might expect from this tendency, these elements readily bond through
electrostatic attraction with the “Group 1” elements (hydrogen, lithium, sodium, potassium, etc.),
each “Group 17” atom accepting an extra electron from each “Group 1” atom which readily provides
it. Ordinary table salt (sodium chloride, or NaCl) is an example of a compound formed by this sort
of bond.
7 It used to be believed that these elements were completely inert: incapable of forming molecular bonds with other
atoms. However, this is not precisely true, as some compounds are now known to integrate noble elements.
8 All isotopes of astatine (At) are radioactive with very short half-lives, making this element difficult to isolate and
study.
4.5. PERIODIC TABLE OF THE ELEMENTS 113
Thus, Group 1 and Group 17 elements are both highly reactive in a chemical sense, but in
different ways. Group 1 elements easily form bonds with Group 17 elements, but Group 1 elements
do not generally bond (solely) with other Group 1 elements, and likewise Group 17 elements do not
generally bond (solely) with other Group 17 elements. It is the structure of the electrons surrounding
each atom’s nucleus that determines how those atoms will bond with other atoms.
114 CHAPTER 4. DERIVATIONS AND TECHNICAL REFERENCES
Molar quantities make it convenient to relate macroscopic samples of elements and compounds
with each other. We know, for instance, that one mole of naturally occurring iron (Fe) atoms will
have a mass of 55.8 grams, and that one mole of naturally occurring oxygen (O) atoms will have a
mass of 16.0 grams, because the average atomic mass of naturally occurring iron is 55.8 amu, and
the average atomic mass of naturally occurring oxygen is 16.0 amu. One mole of naturally occurring
oxygen molecules (O2 ) will have a mass of 32.0 grams, since each molecule is a pair of oxygen atoms
at 16 amu each, and “moles” counts the number of discrete entities which in the case of molecular
oxygen is the number of O2 molecules rather than the number of O atoms. Applying the same
reasoning, one mole of ozone (O3 ) molecules will have a mass of 48.0 grams.
The same mathematical proportions apply to compounds as they do to elements, since
compounds are nothing more than different elements bound together in whole-number ratios, and
the Conservation of Mass tells us a molecule cannot have a mass greater or less than the sum total of
the constituent elements’ masses. To illustrate this principle, we may calculate the mass of one mole
of iron oxide (Fe2 O3 ), the principal component of rust: 55.8×2 + 16.0×3 = 159.6 grams. Likewise,
we may calculate the mass of five moles of pure glucose (C6 H12 O6 ): 5×(12.01×6 + 1.01×12 +
16.0×6) = 900.9 grams. The sum of the atomic masses of a molecule’s constituent atoms is called
the molecular weight or formula weight for that molecule. In the case of iron oxide, the molecular
9 Truth be told, a “mole” is 602,200,000,000,000,000,000,000 counts of literally any discrete entities. Moles do not
represent mass, or volume, or length, or area, but rather a quantity of individual units. There is nothing wrong with
measuring the amount of eggs in the world using the unit of the mole, or the number of grains of sand in moles, or
the number of bits in a collection of digital data. Think of “mole” as nothing more than a really big dozen, or more
precisely, a really big half -dozen!
10 Another way to define one mole is that it is the number of individual nucleons (i.e. protons and/or neutrons)
necessary to comprise one gram of mass. Since protons and neutrons comprise the vast majority of an atom’s mass,
we may essentially ignore the mass of an atom’s electrons when tabulating its mass and pay attention only to the
nucleus. This is why one mole of Hydrogen atoms, each atom having just one lone proton in its nucleus, will have a
combined mass of one gram. By extension, one mole of Carbon-12 atoms, each atom with 6 protons and 6 neutrons,
will have a combined mass of twelve grams.
4.6. MOLECULAR QUANTITIES 115
weight is 159.6 (typically rounded up to 160 grams per mole). In the case of glucose, the molecular
weight is 180.18 (typically rounded down to 180 grams per mole).
When referring to liquid solutions, the concentration of a solute is often expressed as a molarity,
defined as the number of moles of solute per liter of solution. Molarity is usually symbolized by an
italicized capital letter M. It is important to bear in mind that the volume used to calculate molarity
is that of the total solution (solute plus solvent) and not the solvent alone.
Suppose we had a solution of salt-water, comprised of 33.1 grams of table salt thoroughly mixed
with pure water to make a total volume of 1.39 liters. In order to calculate the molarity of this
solution, we first need to determine the equivalence between moles of salt and grams of salt. Since
table salt is sodium chloride (NaCl), and we know the atomic masses of both sodium (23.0 amu)
and chlorine (35.5 amu), we may easily calculate the mass of one mole of table salt:
Another common expression for the concentration of a solute in either a liquid or a gas solution is
related to the concept of percent, expressing the presence of the solute as a ratio of how many “parts”
of solute exist per “parts” of solution. Earth’s atmosphere, for example, contains approximately
20.9% oxygen gas by volume. This means that for every 100 molecules found in a sample of air,
approximately 21 of those are oxygen molecules. When the concentration of a solute is very small,
however, percent becomes an awkward unit of measurement. In such cases it is common to see
low concentrations of solute expressed as parts per million (ppm) or even parts per billion (ppb).
The volumetric concentration of methane gas in Earth’s atmosphere is a good example where parts-
per-million is a more appropriate expression than percent: for every million molecules found in a
sample of air, approximately 2 of them are methane molecules (i.e. methane has an atmospheric
concentration of 2 ppm). As a percentage, this equates to only 0.0002%.
We may use parts-per-unit concentration values as unity fractions just like molecular weights and
just like molarity values, to relate total solution quantity to solute quantity. For example, if we need
to calculate the total mass of hydrogen gas in a compressed air cylinder storing 47000 standard cubic
feet of air, we could multiply the total volume of that air sample (47000 SCF) by the volumetric
concentration of hydrogen naturally found in Earth’s atmosphere which is 0.5 ppm:
47000 SCF air 0.5 parts hydrogen
= 0.0235 SCF hydrogen
1 1000000 parts air
Note how the units of “parts” and “air” cancel out to leave “SCF hydrogen”.
116 CHAPTER 4. DERIVATIONS AND TECHNICAL REFERENCES
An important caveat when using percent, ppm, or ppb is that we must clearly define the “parts”
proportion as either being volume or mass. The concentration of hydrogen gas in the atmosphere
(0.5 ppm) was specified as a volumetric concentration, and so it is appropriate to use this 0.5 ppm
figure to calculate a proportion of the 47000 SCF total volume. If, however, we were given a ppm
mass concentration for hydrogen, we could only use that figure in conjunction with a total mass
quantity for the air sample.
The following photograph illustrates this concept, showing the label on a calibration gas
bottle containing a certified mixture of gases used to check the accuracy of air-safety monitoring
instruments. Note the concentrations of each gas type within the mixture – some expressed in
percent, others in ppm – and how the label states “MOLE %” in the upper-right corner using large
bold print to let you know the concentration values refer to molar quantities (e.g. 18% oxygen
means 18% of the molecules contained in this bottle are oxygen molecules), which for gases closely
corresponds to volumetric quantity rather than mass:
4.7. SATURATED STEAM TABLE 117
By definition a saturated steam table does not describe steam at temperatures greater than the
boiling point. For such purposes, a superheated steam table is necessary.
118 CHAPTER 4. DERIVATIONS AND TECHNICAL REFERENCES
Data for this saturated steam table was taken from Thermal Properties of Saturated and
Superheated Steam by Lionel Marks and Harvey Davis, published in 1920 by Longmans, Green,
and Company.
liquid densities taken from table on page F-3 and solid densities taken from table on page F-1. Some liquid densities
taken from tables on pages E-27 through E-31. All temperatures at or near 20 o C.
4.8. WEIGHT DENSITIES OF COMMON MATERIALS 121
Questions
This learning module, along with all others in the ModEL collection, is designed to be used in an
inverted instructional environment where students independently read1 the tutorials and attempt
to answer questions on their own prior to the instructor’s interaction with them. In place of
lecture2 , the instructor engages with students in Socratic-style dialogue, probing and challenging
their understanding of the subject matter through inquiry.
Answers are not provided for questions within this chapter, and this is by design. Solved problems
may be found in the Tutorial and Derivation chapters, instead. The goal here is independence, and
this requires students to be challenged in ways where others cannot think for them. Remember
that you always have the tools of experimentation and computer simulation (e.g. SPICE) to explore
concepts!
The following lists contain ideas for Socratic-style questions and challenges. Upon inspection,
one will notice a strong theme of metacognition within these statements: they are designed to foster
a regular habit of examining one’s own thoughts as a means toward clearer thinking. As such these
sample questions are useful both for instructor-led discussions as well as for self-study.
1 Technical reading is an essential academic skill for any technical practitioner to possess for the simple reason
that the most comprehensive, accurate, and useful information to be found for developing technical competence is in
textual form. Technical careers in general are characterized by the need for continuous learning to remain current
with standards and technology, and therefore any technical practitioner who cannot read well is handicapped in
their professional development. An excellent resource for educators on improving students’ reading prowess through
intentional effort and strategy is the book textitReading For Understanding – How Reading Apprenticeship Improves
Disciplinary Learning in Secondary and College Classrooms by Ruth Schoenbach, Cynthia Greenleaf, and Lynn
Murphy.
2 Lecture is popular as a teaching method because it is easy to implement: any reasonably articulate subject matter
expert can talk to students, even with little preparation. However, it is also quite problematic. A good lecture always
makes complicated concepts seem easier than they are, which is bad for students because it instills a false sense of
confidence in their own understanding; reading and re-articulation requires more cognitive effort and serves to verify
comprehension. A culture of teaching-by-lecture fosters a debilitating dependence upon direct personal instruction,
whereas the challenges of modern life demand independent and critical thought made possible only by gathering
information and perspectives from afar. Information presented in a lecture is ephemeral, easily lost to failures of
memory and dictation; text is forever, and may be referenced at any time.
123
124 CHAPTER 5. QUESTIONS
• Summarize as much of the text as you can in one paragraph of your own words. A helpful
strategy is to explain ideas as you would for an intelligent child: as simple as you can without
compromising too much accuracy.
• Simplify a particular section of the text, for example a paragraph or even a single sentence, so
as to capture the same fundamental idea in fewer words.
• Where did the text make the most sense to you? What was it about the text’s presentation
that made it clear?
• Identify where it might be easy for someone to misunderstand the text, and explain why you
think it could be confusing.
• Identify any new concept(s) presented in the text, and explain in your own words.
• Identify any familiar concept(s) such as physical laws or principles applied or referenced in the
text.
• Did the text reveal any misconceptions you might have harbored? If so, describe the
misconception(s) and the reason(s) why you now know them to be incorrect.
• Identify where any fundamental laws or principles apply to the solution of this problem,
especially before applying any mathematical techniques.
• Devise a thought experiment to explore the characteristics of the problem scenario, applying
known laws and principles to mentally model its behavior.
• Describe in detail your own strategy for solving this problem. How did you identify and
organized the given information? Did you sketch any diagrams to help frame the problem?
• Is there more than one way to solve this problem? Which method seems best to you?
• Show the work you did in solving this problem, even if the solution is incomplete or incorrect.
• What would you say was the most challenging part of this problem, and why was it so?
• Was any important information missing from the problem which you had to research or recall?
• Was there any extraneous information presented within this problem? If so, what was it and
why did it not matter?
• Examine someone else’s solution to identify where they applied fundamental laws or principles.
• Simplify the problem from its given form and show how to solve this simpler version of it.
Examples include eliminating certain variables or conditions, altering values to simpler (usually
whole) numbers, applying a limiting case (i.e. altering a variable to some extreme or ultimate
value).
• For quantitative problems, identify the real-world meaning of all intermediate calculations:
their units of measurement, where they fit into the scenario at hand. Annotate any diagrams
or illustrations with these calculated values.
• Were there any assumptions you made while solving this problem? Would your solution change
if one of those assumptions were altered?
• Identify where it would be easy for someone to go astray in attempting to solve this problem.
• Formulate your own problem based on what you learned solving this one.
• Show how thorough documentation assisted in the completion of this experiment or project.
It is far more important that you convey your reasoning than it is to simply convey a correct
answer. For this reason, you should refrain from researching other information sources to answer
questions. What matters here is that you are doing the thinking. If the answer is incorrect, your
instructor will work with you to correct it through proper reasoning. A correct answer without an
adequate explanation of how you derived that answer is unacceptable, as it does not aid the learning
or assessment process.
You will note a conspicuous lack of answers given for these conceptual questions. Unlike standard
textbooks where answers to every other question are given somewhere toward the back of the book,
here in these learning modules students must rely on other means to check their work. The best way
by far is to debate the answers with fellow students and also with the instructor during the Socratic
dialogue sessions intended to be used with these learning modules. Reasoning through challenging
questions with other people is an excellent tool for developing strong reasoning skills.
Another means of checking your conceptual answers, where applicable, is to use circuit simulation
software to explore the effects of changes made to circuits. For example, if one of these conceptual
questions challenges you to predict the effects of altering some component parameter in a circuit,
you may check the validity of your work by simulating that same parameter change within software
and seeing if the results agree.
3 Analytical thinking involves the “disassembly” of an idea into its constituent parts, analogous to dissection.
Synthetic thinking involves the “assembly” of a new idea comprised of multiple concepts, analogous to construction.
Both activities are high-level cognitive skills, extremely important for effective problem-solving, necessitating frequent
challenge and regular practice to fully develop.
128 CHAPTER 5. QUESTIONS
Francis Bacon’s advice is a blueprint for effective education: reading provides the learner with
knowledge, writing focuses the learner’s thoughts, and critical dialogue equips the learner to
confidently communicate and apply their learning. Independent acquisition and application of
knowledge is a powerful skill, well worth the effort to cultivate. To this end, students should read
these educational resources closely, journal their own reflections on the reading, and discuss in detail
their findings with classmates and instructor(s). You should be able to do all of the following after
reading any instructional text:
√
Briefly SUMMARIZE THE TEXT in the form of a journal entry documenting your learning
as you progress through the course of study. Share this summary in dialogue with your classmates
and instructor. Journaling is an excellent self-test of thorough reading because you cannot clearly
express what you have not read or did not comprehend.
√
Demonstrate ACTIVE READING STRATEGIES, including verbalizing your impressions as
you read, simplifying long passages to convey the same ideas using fewer words, annotating text
and illustrations with your own interpretations, working through mathematical examples shown in
the text, cross-referencing passages with relevant illustrations and/or other passages, identifying
problem-solving strategies applied by the author, etc. Technical reading is a special case of problem-
solving, and so these strategies work precisely because they help solve any problem: paying attention
to your own thoughts (metacognition), eliminating unnecessary complexities, identifying what makes
sense, paying close attention to details, drawing connections between separated facts, and noting
the successful strategies of others.
√
Identify IMPORTANT THEMES, especially GENERAL LAWS and PRINCIPLES, expounded
in the text and express them in the simplest of terms as though you were teaching an intelligent
child. This emphasizes connections between related topics and develops your ability to communicate
complex ideas to anyone.
√
Form YOUR OWN QUESTIONS based on the reading, and then pose them to your instructor
and classmates for their consideration. Anticipate both correct and incorrect answers, the incorrect
answer(s) assuming one or more plausible misconceptions. This helps you view the subject from
different perspectives to grasp it more fully.
√
Devise EXPERIMENTS to test claims presented in the reading, or to disprove misconceptions.
Predict possible outcomes of these experiments, and evaluate their meanings: what result(s) would
confirm, and what would constitute disproof? Running mental simulations and evaluating results is
essential to scientific and diagnostic reasoning.
√
Specifically identify any points you found CONFUSING. The reason for doing this is to help
diagnose misconceptions and overcome barriers to learning.
5.1. CONCEPTUAL REASONING 129
Mass
Weight
Conservation of Mass
Charge
Energy
Conservation of Energy
Electrons
Protons
Neutrons
Ion
Electricity
130 CHAPTER 5. QUESTIONS
Work
Potential energy
Kinetic energy
Power
First, identify the various forms that electrical energy becomes converted into while it is used.
Next, if the Law of Energy Conservation is true, why do customers have to keep paying for their
electricity? If energy can never be destroyed, why not just pay for electricity once and then keep
using that same purchased electricity forever?
Electricity bills are often expressed in units of kiloWatt-hours (kW-h). How many Joules are
represented by one kW-h?
Challenges
• Why do you suppose kiloWatt-hours are used in electric billing statements instead of Joules?
5.1. CONCEPTUAL REASONING 131
• At which point(s) in the roller-coaster track is the car’s potential energy greatest?
• At which point(s) in the roller-coaster track is the car’s kinetic energy greatest?
• Can the car ever reach a height greater than the point it was initially hoisted before let loose
to coast?
• What happens to the kinetic energy of the car when it comes to a stop at the end of the ride?
Challenges
• Compressed spring =
• Stretched spring =
• Falling stone =
• Speeding train =
• Hot coal =
• Charged battery =
Challenges
• For each case, identify how to convert that form of energy into the other form.
132 CHAPTER 5. QUESTIONS
Challenges
• Suppose we wished to double the speed of a given BB (or arrow) with fixed mass. How much
more energy would we need to put into the weapon to achieve this goal?
• Identify a way we double the speed of the projectile without “charging” the weapon with
additional energy.
5.1. CONCEPTUAL REASONING 133
• If a whole atom becomes ionized and then later rejoins with the stripped electrons, it will once
again have zero net electric charge
• More energy is required to hoist a weight rapidly than to hoist a weight slowly (assuming the
same lifting height)
• An automobile crashing into a stationary barrier at 60 km/hr will suffer twice the damage of
an automobile crashing at 30 km/hr, all other factors being equal
• A crate accidently dropped from a height of 6 meters will suffer twice the damage of a crate
dropped from a height of 3 meters, all other factors being equal
Challenges
Mental arithmetic and estimations are strongly encouraged for all calculations, because without
these abilities you will be unable to readily detect errors caused by calculator misuse (e.g. keystroke
errors).
You will note a conspicuous lack of answers given for these quantitative questions. Unlike
standard textbooks where answers to every other question are given somewhere toward the back
of the book, here in these learning modules students must rely on other means to check their work.
My advice is to use circuit simulation software such as SPICE to check the correctness of quantitative
answers. Refer to those learning modules within this collection focusing on SPICE to see worked
examples which you may use directly as practice problems for your own study, and/or as templates
you may modify to run your own analyses and generate your own practice problems.
Completely worked example problems found in the Tutorial may also serve as “test cases4 ” for
gaining proficiency in the use of circuit simulation software, and then once that proficiency is gained
you will never need to rely5 on an answer key!
4 In other words, set up the circuit simulation software to analyze the same circuit examples found in the Tutorial.
If the simulated results match the answers shown in the Tutorial, it confirms the simulation has properly run. If
the simulated results disagree with the Tutorial’s answers, something has been set up incorrectly in the simulation
software. Using every Tutorial as practice in this way will quickly develop proficiency in the use of circuit simulation
software.
5 This approach is perfectly in keeping with the instructional philosophy of these learning modules: teaching students
to be self-sufficient thinkers. Answer keys can be useful, but it is even more useful to your long-term success to have
a set of tools on hand for checking your own work, because once you have left school and are on your own, there will
no longer be “answer keys” available for the problems you will have to solve.
5.2. QUANTITATIVE REASONING 135
Magnetic permeability of free space (µ0 ) = 1.25663706212(19) × 10−6 Henrys per meter (H/m)
Electric permittivity of free space (ǫ0 ) = 8.8541878128(13) × 10−12 Farads per meter (F/m)
Gravitational constant (G) = 6.67430(15) × 10−11 cubic meters per kilogram-seconds squared
(m3 /kg-s2 )
Molar gas constant (R) = 8.314462618... Joules per mole-Kelvin (J/mol-K) = 0.08205746(14)
liters-atmospheres per mole-Kelvin
Speed of light in a vacuum (c) = 299,792,458 meters per second (m/s) = 186282.4 miles per
second (mi/s)
Note: All constants taken from NIST data “Fundamental Physical Constants – Complete Listing”,
from http://physics.nist.gov/constants, National Institute of Standards and Technology
(NIST), 2018 CODATA Adjustment.
136 CHAPTER 5. QUESTIONS
A B C D
1 Distance traveled 46.9 Kilometers
4
5
Text labels contained in cells A1 through A3 and cells C1 through C3 exist solely for readability
and are not involved in any calculations. Cell B1 contains a sample distance value while cell B2
contains a sample time value. The formula for computing speed is contained in cell B3. Note how
this formula begins with an “equals” symbol (=), references the values for distance and speed by
lettered column and numbered row coordinates (B1 and B2), and uses a forward slash symbol for
division (/). The coordinates B1 and B2 function as variables 6 would in an algebraic formula.
When this spreadsheet is executed, the numerical value 39.74576 will appear in cell B3 rather
than the formula = B1 / B2, because 39.74576 is the computed speed value given 46.9 kilometers
traveled over a period of 1.18 hours. If a different numerical value for distance is entered into cell
B1 or a different value for time is entered into cell B2, cell B3’s value will automatically update. All
you need to do is set up the given values and any formulae into the spreadsheet, and the computer
will do all the calculations for you.
Cell B3 may be referenced by other formulae in the spreadsheet if desired, since it is a variable
just like the given values contained in B1 and B2. This means it is possible to set up an entire chain
of calculations, one dependent on the result of another, in order to arrive at a final value. The
arrangement of the given data and formulae need not follow any pattern on the grid, which means
you may place them anywhere.
6 Spreadsheets may also provide means to attach text labels to cells for use as variable names (Microsoft Excel
simply calls these labels “names”), but for simple spreadsheets such as those shown here it’s usually easier just to use
the standard coordinate naming for each cell.
5.2. QUANTITATIVE REASONING 137
Common7 arithmetic operations available for your use in a spreadsheet include the following:
• Addition (+)
• Subtraction (-)
• Multiplication (*)
• Division (/)
• Powers (^)
Parentheses may be used to ensure8 proper order of operations within a complex formula.
Consider this example of a spreadsheet implementing the quadratic formula, used to solve for roots
of a polynomial expression in the form of ax2 + bx + c:
√
−b ± b2 − 4ac
x=
2a
A B
1 x_1 = (-B4 + sqrt((B4^2) - (4*B3*B5))) / (2*B3)
3 a = 9
4 b = 5
5 c = -2
This example is configured to compute roots9 of the polynomial 9x2 + 5x − 2 because the values
of 9, 5, and −2 have been inserted into cells B3, B4, and B5, respectively. Once this spreadsheet has
been built, though, it may be used to calculate the roots of any second-degree polynomial expression
simply by entering the new a, b, and c coefficients into cells B3 through B5. The numerical values
appearing in cells B1 and B2 will be automatically updated by the computer immediately following
any changes made to the coefficients.
7 Modern spreadsheet software offers a bewildering array of mathematical functions you may use in your
computations. I recommend you consult the documentation for your particular spreadsheet for information on
operations other than those listed here.
8 Spreadsheet programs, like text-based programming languages, are designed to follow standard order of operations
by default. However, my personal preference is to use parentheses even where strictly unnecessary just to make it
clear to any other person viewing the formula what the intended order of operations is.
9 Reviewing some algebra here, a root is a value for x that yields an overall value of zero for the polynomial. For
this polynomial (9x2 + 5x − 2) the two roots happen to be x = 0.269381 and x = −0.82494, with these values displayed
in cells B1 and B2, respectively upon execution of the spreadsheet.
138 CHAPTER 5. QUESTIONS
Alternatively, one could break up the long quadratic formula into smaller pieces like this:
p
y = b2 − 4ac z = 2a
−b ± y
x=
z
A B C
1 x_1 = (-B4 + C1) / C2 = sqrt((B4^2) - (4*B3*B5))
3 a = 9
4 b = 5
5 c = -2
Note how the square-root term (y) is calculated in cell C1, and the denominator term (z) in cell
C2. This makes the two final formulae (in cells B1 and B2) simpler to interpret. The positioning of
all these cells on the grid is completely arbitrary10 – all that matters is that they properly reference
each other in the formulae.
Spreadsheets are particularly useful for situations where the same set of calculations representing
a circuit or other system must be repeated for different initial conditions. The power of a spreadsheet
is that it automates what would otherwise be a tedious set of calculations. One specific application
of this is to simulate the effects of various components within a circuit failing with abnormal values
(e.g. a shorted resistor simulated by making its value nearly zero; an open resistor simulated by
making its value extremely large). Another application is analyzing the behavior of a circuit design
given new components that are out of specification, and/or aging components experiencing drift
over time.
10 My personal preference is to locate all the “given” data in the upper-left cells of the spreadsheet grid (each data
point flanked by a sensible name in the cell to the left and units of measurement in the cell to the right as illustrated
in the first distance/time spreadsheet example), sometimes coloring them in order to clearly distinguish which cells
contain entered data versus which cells contain computed results from formulae. I like to place all formulae in cells
below the given data, and try to arrange them in logical order so that anyone examining my spreadsheet will be able
to figure out how I constructed a solution. This is a general principle I believe all computer programmers should
follow: document and arrange your code to make it easy for other people to learn from it.
5.2. QUANTITATIVE REASONING 139
A 15 kg mass is dragged along a level surface with a steady pull of 25 Newtons for 10 meters.
How much work is done?
15 kg 25 N pull
Challenges
• If the 50 Newton weight were dropped after being hoisted up 4 meters above ground level, how
much energy would it have just before striking the ground?
• Work represents a transfer of energy from one system to another, or from one location to
another. In each of these cases, where is energy being transferred to and from?
Challenges
Next, calculate the kinetic energy of a bullet with a mass of 150 grains (9.7198 grams) traveling
at a velocity of 2820 feet per second, expressing your answer in both British and metric units.
Challenges
• Demonstrate how to estimate numerical answers for this problem without using a calculator.
• Describe the various energy-transfer operations taking place when a crane lifts a heavy weight
and then sets it down in a different location.
• If a crane lifts a weight 10 feet above the ground, then sets that weight down on top of an
object 6 feet above the ground, is energy still conserved? Explain in detail.
• Which has a greater effect on a moving object’s kinetic energy: an increase in mass or an
increase in velocity?
5.2. QUANTITATIVE REASONING 141
Knowing this, convert the length of a 0.089223 meter antenna into centimeters, feet, and inches.
Challenges
• Convert 40 miles into kilometers using nothing but the following ratios: 2.54 cm to 1 inch,
5280 feet to 1 mile, 12 inches to 1 foot, and of course metric prefixes (centi = 100, kilo = 1000,
etc.).
142 CHAPTER 5. QUESTIONS
Explain how the equivalence of 5 Volts DC = 255 counts could be used as a unity fraction for
converting any input voltage value into its corresponding count value.
Next, convert between the following Vin and output count values for this same ADC:
Challenges
1.000 pound per square inch (PSI) = 2.036 inches of mercury (in. Hg) = 27.68 inches of water (in.
W.C.) = 6.895 kilo-pascals (kPa) = 0.06895 bar
25 PSI
= ??? kPa
1
40 ”WC
= ??? PSI
1
5.6 bar
= ??? PSI
1
1200 ”Hg
= ??? ”WC
1
12 ”WC
= ??? bar
1
110 kPa
= ??? ”WC
1
982 ”Hg
= ??? kPa
1
50 PSI
= ??? bar
1
250 kPa
= ??? ”Hg
1
31 bar
= ??? ”Hg
1
Challenges
• Demonstrate how to estimate numerical answers for these conversion problems without using
a calculator.
144 CHAPTER 5. QUESTIONS
Challenges
• From your work, derive a single factor converting CFS into GPM (or vice versa), so that
anyone else needing to do this conversion (who isn’t familiar with unity fractions) may easily
convert from one to the other.
• 500 Hz to rad/sec =
Challenges
• What purpose does natural frequency (ω) expressed in radians per second serve?
5.3. DIAGNOSTIC REASONING 145
As always, your goal is to fully explain your analysis of each problem. Simply obtaining a
correct answer is not good enough – you must also demonstrate sound reasoning in order to
successfully complete the assignment. Your instructor’s responsibility is to probe and challenge
your understanding of the relevant principles and analytical processes in order to ensure you have a
strong foundation upon which to build further understanding.
You will note a conspicuous lack of answers given for these diagnostic questions. Unlike standard
textbooks where answers to every other question are given somewhere toward the back of the book,
here in these learning modules students must rely on other means to check their work. The best way
by far is to debate the answers with fellow students and also with the instructor during the Socratic
dialogue sessions intended to be used with these learning modules. Reasoning through challenging
questions with other people is an excellent tool for developing strong reasoning skills.
Another means of checking your diagnostic answers, where applicable, is to use circuit simulation
software to explore the effects of faults placed in circuits. For example, if one of these diagnostic
questions requires that you predict the effect of an open or a short in a circuit, you may check the
validity of your work by simulating that same fault (substituting a very high resistance in place of
that component for an open, and substituting a very low resistance for a short) within software and
seeing if the results agree.
146 CHAPTER 5. QUESTIONS
Explain what makes this claim suspicious at best, and how we could truly put it to the test just
in case the inventor has indeed made a revolutionary discovery.
Challenges
• Identify some similar claims you have heard, and then identify ways you could comprehensively
test those claims.
• Interestingly, you never encounter people claiming to have made machines that output more
mass than they take in. Why not?
5.3. DIAGNOSTIC REASONING 147
Rope Scale
(registers 8 pounds
of pulling force)
Pulley
16 pound
mass
Suppose someone tells you this machine multiplies your energy. Identify the conceptual error at
work here, and correct it.
Challenges
• Address this misconception in the converse: how would this machine actually behave if it did
multiply your energy?
148 CHAPTER 5. QUESTIONS
Appendix A
Problem-Solving Strategies
The ability to solve complex problems is arguably one of the most valuable skills one can possess,
and this skill is particularly important in any science-based discipline.
• Study principles, not procedures. Don’t be satisfied with merely knowing how to compute
solutions – learn why those solutions work.
• Identify what it is you need to solve, identify all relevant data, identify all units of measurement,
identify any general principles or formulae linking the given information to the solution, and
then identify any “missing pieces” to a solution. Annotate all diagrams with this data.
• Sketch a diagram to help visualize the problem. When building a real system, always devise
a plan for that system and analyze its function before constructing it.
• Follow the units of measurement and meaning of every calculation. If you are ever performing
mathematical calculations as part of a problem-solving procedure, and you find yourself unable
to apply each and every intermediate result to some aspect of the problem, it means you
don’t understand what you are doing. Properly done, every mathematical result should have
practical meaning for the problem, and not just be an abstract number. You should be able to
identify the proper units of measurement for each and every calculated result, and show where
that result fits into the problem.
• Perform “thought experiments” to explore the effects of different conditions for theoretical
problems. When troubleshooting real systems, perform diagnostic tests rather than visually
inspecting for faults, the best diagnostic test being the one giving you the most information
about the nature and/or location of the fault with the fewest steps.
• Simplify the problem until the solution becomes obvious, and then use that obvious case as a
model to follow in solving the more complex version of the problem.
• Check for exceptions to see if your solution is incorrect or incomplete. A good solution will
work for all known conditions and criteria. A good example of this is the process of testing
scientific hypotheses: the task of a scientist is not to find support for a new idea, but rather
to challenge that new idea to see if it holds up under a battery of tests. The philosophical
149
150 APPENDIX A. PROBLEM-SOLVING STRATEGIES
principle of reductio ad absurdum (i.e. disproving a general idea by finding a specific case
where it fails) is useful here.
• Add quantities to problems that are qualitative in nature, because sometimes a little math
helps illuminate the scenario.
• Sketch graphs illustrating how variables relate to each other. These may be quantitative (i.e.
with realistic number values) or qualitative (i.e. simply showing increases and decreases).
• Treat quantitative problems as qualitative in order to discern the relative magnitudes and/or
directions of change of the relevant variables. For example, try determining what happens if a
certain variable were to increase or decrease before attempting to precisely calculate quantities:
how will each of the dependent variables respond, by increasing, decreasing, or remaining the
same as before?
• Consider limiting cases. This works especially well for qualitative problems where you need to
determine which direction a variable will change. Take the given condition and magnify that
condition to an extreme degree as a way of simplifying the direction of the system’s response.
• Check your work. This means regularly testing your conclusions to see if they make sense.
This does not mean repeating the same steps originally used to obtain the conclusion(s), but
rather to use some other means to check validity. Simply repeating procedures often leads to
repeating the same errors if any were made, which is why alternative paths are better.
Appendix B
Instructional philosophy
“The unexamined circuit is not worth energizing” – Socrates (if he had taught electricity)
These learning modules, although useful for self-study, were designed to be used in a formal
learning environment where a subject-matter expert challenges students to digest the content and
exercise their critical thinking abilities in the answering of questions and in the construction and
testing of working circuits.
The following principles inform the instructional and assessment philosophies embodied in these
learning modules:
• The first goal of education is to enhance clear and independent thought, in order that
every student reach their fullest potential in a highly complex and inter-dependent world.
Robust reasoning is always more important than particulars of any subject matter, because
its application is universal.
• Literacy is fundamental to independent learning and thought because text continues to be the
most efficient way to communicate complex ideas over space and time. Those who cannot read
with ease are limited in their ability to acquire knowledge and perspective.
• Faulty assumptions and poor reasoning are best corrected through challenge, not presentation.
The rhetorical technique of reductio ad absurdum (disproving an assertion by exposing an
absurdity) works well to discipline student’s minds, not only to correct the problem at hand
but also to learn how to detect and correct future errors.
• Important principles should be repeatedly explored and widely applied throughout a course
of study, not only to reinforce their importance and help ensure their mastery, but also to
showcase the interconnectedness and utility of knowledge.
151
152 APPENDIX B. INSTRUCTIONAL PHILOSOPHY
An inspection of these learning modules reveals certain unique characteristics. One of these is
a bias toward thorough explanations in the tutorial chapters. Without a live instructor to explain
concepts and applications to students, the text itself must fulfill this role. This philosophy results in
lengthier explanations than what you might typically find in a textbook, each step of the reasoning
process fully explained, including footnotes addressing common questions and concerns students
raise while learning these concepts. Each tutorial seeks to not only explain each major concept
in sufficient detail, but also to explain the logic of each concept and how each may be developed
1 In a traditional teaching environment, students first encounter new information via lecture from an expert, and
then independently apply that information via homework. In an “inverted” course of study, students first encounter
new information via homework, and then independently apply that information under the scrutiny of an expert. The
expert’s role in lecture is to simply explain, but the expert’s role in an inverted session is to challenge, critique, and
if necessary explain where gaps in understanding still exist.
2 Socrates is a figure in ancient Greek philosophy famous for his unflinching style of questioning. Although he
authored no texts, he appears as a character in Plato’s many writings. The essence of Socratic philosophy is to
leave no question unexamined and no point of view unchallenged. While purists may argue a topic such as electric
circuits is too narrow for a true Socratic-style dialogue, I would argue that the essential thought processes involved
with scientific reasoning on any topic are not far removed from the Socratic ideal, and that students of electricity and
electronics would do very well to challenge assumptions, pose thought experiments, identify fallacies, and otherwise
employ the arsenal of critical thinking skills modeled by Socrates.
3 This rhetorical technique is known by the Latin phrase reductio ad absurdum. The concept is to expose errors by
counter-example, since only one solid counter-example is necessary to disprove a universal claim. As an example of
this, consider the common misconception among beginning students of electricity that voltage cannot exist without
current. One way to apply reductio ad absurdum to this statement is to ask how much current passes through a
fully-charged battery connected to nothing (i.e. a clear example of voltage existing without current).
153
from “first principles”. Again, this reflects the goal of developing clear and independent thought in
students’ minds, by showing how clear and logical thought was used to forge each concept. Students
benefit from witnessing a model of clear thinking in action, and these tutorials strive to be just that.
Another characteristic of these learning modules is a lack of step-by-step instructions in the
Project and Experiment chapters. Unlike many modern workbooks and laboratory guides where
step-by-step instructions are prescribed for each experiment, these modules take the approach that
students must learn to closely read the tutorials and apply their own reasoning to identify the
appropriate experimental steps. Sometimes these steps are plainly declared in the text, just not as
a set of enumerated points. At other times certain steps are implied, an example being assumed
competence in test equipment use where the student should not need to be told again how to use
their multimeter because that was thoroughly explained in previous lessons. In some circumstances
no steps are given at all, leaving the entire procedure up to the student.
This lack of prescription is not a flaw, but rather a feature. Close reading and clear thinking are
foundational principles of this learning series, and in keeping with this philosophy all activities are
designed to require those behaviors. Some students may find the lack of prescription frustrating,
because it demands more from them than what their previous educational experiences required. This
frustration should be interpreted as an unfamiliarity with autonomous thinking, a problem which
must be corrected if the student is ever to become a self-directed learner and effective problem-solver.
Ultimately, the need for students to read closely and think clearly is more important both in the
near-term and far-term than any specific facet of the subject matter at hand. If a student takes
longer than expected to complete a module because they are forced to outline, digest, and reason
on their own, so be it. The future gains enjoyed by developing this mental discipline will be well
worth the additional effort and delay.
Another feature of these learning modules is that they do not treat topics in isolation. Rather,
important concepts are introduced early in the series, and appear repeatedly as stepping-stones
toward other concepts in subsequent modules. This helps to avoid the “compartmentalization”
of knowledge, demonstrating the inter-connectedness of concepts and simultaneously reinforcing
them. Each module is fairly complete in itself, reserving the beginning of its tutorial to a review of
foundational concepts.
This methodology of assigning text-based modules to students for digestion and then using
Socratic dialogue to assess progress and hone students’ thinking was developed over a period of
several years by the author with his Electronics and Instrumentation students at the two-year college
level. While decidedly unconventional and sometimes even unsettling for students accustomed to
a more passive lecture environment, this instructional philosophy has proven its ability to convey
conceptual mastery, foster careful analysis, and enhance employability so much better than lecture
that the author refuses to ever teach by lecture again.
Problems which often go undiagnosed in a lecture environment are laid bare in this “inverted”
format where students must articulate and logically defend their reasoning. This, too, may be
unsettling for students accustomed to lecture sessions where the instructor cannot tell for sure who
comprehends and who does not, and this vulnerability necessitates sensitivity on the part of the
“inverted” session instructor in order that students never feel discouraged by having their errors
exposed. Everyone makes mistakes from time to time, and learning is a lifelong process! Part of
the instructor’s job is to build a culture of learning among the students where errors are not seen as
shameful, but rather as opportunities for progress.
154 APPENDIX B. INSTRUCTIONAL PHILOSOPHY
To this end, instructors managing courses based on these modules should adhere to the following
principles:
• Student questions are always welcome and demand thorough, honest answers. The only type
of question an instructor should refuse to answer is one the student should be able to easily
answer on their own. Remember, the fundamental goal of education is for each student to learn
to think clearly and independently. This requires hard work on the part of the student, which
no instructor should ever circumvent. Anything done to bypass the student’s responsibility to
do that hard work ultimately limits that student’s potential and thereby does real harm.
• It is not only permissible, but encouraged, to answer a student’s question by asking questions
in return, these follow-up questions designed to guide the student to reach a correct answer
through their own reasoning.
• All student answers demand to be challenged by the instructor and/or by other students.
This includes both correct and incorrect answers – the goal is to practice the articulation and
defense of one’s own reasoning.
• No reading assignment is deemed complete unless and until the student demonstrates their
ability to accurately summarize the major points in their own terms. Recitation of the original
text is unacceptable. This is why every module contains an “Outline and reflections” question
as well as a “Foundational concepts” question in the Conceptual reasoning section, to prompt
reflective reading.
• No assigned question is deemed answered unless and until the student demonstrates their
ability to consistently and correctly apply the concepts to variations of that question. This is
why module questions typically contain multiple “Challenges” suggesting different applications
of the concept(s) as well as variations on the same theme(s). Instructors are encouraged to
devise as many of their own “Challenges” as they are able, in order to have a multitude of
ways ready to probe students’ understanding.
• No assigned experiment or project is deemed complete unless and until the student
demonstrates the task in action. If this cannot be done “live” before the instructor, video-
recordings showing the demonstration are acceptable. All relevant safety precautions must be
followed, all test equipment must be used correctly, and the student must be able to properly
explain all results. The student must also successfully answer all Challenges presented by the
instructor for that experiment or project.
155
Students learning from these modules would do well to abide by the following principles:
• No text should be considered fully and adequately read unless and until you can express every
idea in your own words, using your own examples.
• You should always articulate your thoughts as you read the text, noting points of agreement,
confusion, and epiphanies. Feel free to print the text on paper and then write your notes in
the margins. Alternatively, keep a journal for your own reflections as you read. This is truly
a helpful tool when digesting complicated concepts.
• Never take the easy path of highlighting or underlining important text. Instead, summarize
and/or comment on the text using your own words. This actively engages your mind, allowing
you to more clearly perceive points of confusion or misunderstanding on your own.
• A very helpful strategy when learning new concepts is to place yourself in the role of a teacher,
if only as a mental exercise. Either explain what you have recently learned to someone else,
or at least imagine yourself explaining what you have learned to someone else. The simple act
of having to articulate new knowledge and skill forces you to take on a different perspective,
and will help reveal weaknesses in your understanding.
• Perform each and every mathematical calculation and thought experiment shown in the text
on your own, referring back to the text to see that your results agree. This may seem trivial
and unnecessary, but it is critically important to ensuring you actually understand what is
presented, especially when the concepts at hand are complicated and easy to misunderstand.
Apply this same strategy to become proficient in the use of circuit simulation software, checking
to see if your simulated results agree with the results shown in the text.
• Above all, recognize that learning is hard work, and that a certain level of frustration is
unavoidable. There are times when you will struggle to grasp some of these concepts, and that
struggle is a natural thing. Take heart that it will yield with persistent and varied4 effort, and
never give up!
Students interested in using these modules for self-study will also find them beneficial, although
the onus of responsibility for thoroughly reading and answering questions will of course lie with
that individual alone. If a qualified instructor is not available to challenge students, a workable
alternative is for students to form study groups where they challenge5 one another.
Tony R. Kuphaldt
4 As the old saying goes, “Insanity is trying the same thing over and over again, expecting different results.” If
you find yourself stumped by something in the text, you should attempt a different approach. Alter the thought
experiment, change the mathematical parameters, do whatever you can to see the problem in a slightly different light,
and then the solution will often present itself more readily.
5 Avoid the temptation to simply share answers with study partners, as this is really counter-productive to learning.
Always bear in mind that the answer to any question is far less important in the long run than the method(s) used to
obtain that answer. The goal of education is to empower one’s life through the improvement of clear and independent
thought, literacy, expression, and various practical skills.
156 APPENDIX B. INSTRUCTIONAL PHILOSOPHY
Appendix C
Tools used
I am indebted to the developers of many open-source software applications in the creation of these
learning modules. The following is a list of these applications with some commentary on each.
You will notice a theme common to many of these applications: a bias toward code. Although
I am by no means an expert programmer in any computer language, I understand and appreciate
the flexibility offered by code-based applications where the user (you) enters commands into a plain
ASCII text file, which the software then reads and processes to create the final output. Code-based
computer applications are by their very nature extensible, while WYSIWYG (What You See Is What
You Get) applications are generally limited to whatever user interface the developer makes for you.
There is so much to be said about Linus Torvalds’ Linux and Richard Stallman’s GNU
project. First, to credit just these two individuals is to fail to do justice to the mob of
passionate volunteers who contributed to make this amazing software a reality. I first
learned of Linux back in 1996, and have been using this operating system on my personal
computers almost exclusively since then. It is free, it is completely configurable, and it
permits the continued use of highly efficient Unix applications and scripting languages
(e.g. shell scripts, Makefiles, sed, awk) developed over many decades. Linux not only
provided me with a powerful computing platform, but its open design served to inspire
my life’s work of creating open-source educational resources.
Writing code for any code-based computer application requires a text editor, which may
be thought of as a word processor strictly limited to outputting plain-ASCII text files.
Many good text editors exist, and one’s choice of text editor seems to be a deeply personal
matter within the programming world. I prefer Vim because it operates very similarly to
vi which is ubiquitous on Unix/Linux operating systems, and because it may be entirely
operated via keyboard (i.e. no mouse required) which makes it fast to use.
157
158 APPENDIX C. TOOLS USED
Developed in the late 1970’s and early 1980’s by computer scientist extraordinaire Donald
Knuth to typeset his multi-volume magnum opus The Art of Computer Programming,
this software allows the production of formatted text for screen-viewing or paper printing,
all by writing plain-text code to describe how the formatted text is supposed to appear.
TEX is not just a markup language for documents, but it is also a Turing-complete
programming language in and of itself, allowing useful algorithms to be created to control
the production of documents. Simply put, TEX is a programmer’s approach to word
processing. Since TEX is controlled by code written in a plain-text file, this means
anyone may read that plain-text file to see exactly how the document was created. This
openness afforded by the code-based nature of TEX makes it relatively easy to learn how
other people have created their own TEX documents. By contrast, examining a beautiful
document created in a conventional WYSIWYG word processor such as Microsoft Word
suggests nothing to the reader about how that document was created, or what the user
might do to create something similar. As Mr. Knuth himself once quipped, conventional
word processing applications should be called WYSIAYG (What You See Is All You
Get).
Like all true programming languages, TEX is inherently extensible. So, years after the
release of TEX to the public, Leslie Lamport decided to create a massive extension
allowing easier compilation of book-length documents. The result was LATEX, which
is the markup language used to create all ModEL module documents. You could say
that TEX is to LATEX as C is to C++. This means it is permissible to use any and all TEX
commands within LATEX source code, and it all still works. Some of the features offered
by LATEX that would be challenging to implement in TEX include automatic index and
table-of-content creation.
This wonderful program is what I use to create all the schematic diagrams and
illustrations (but not photographic images or mathematical plots) throughout the ModEL
project. It natively outputs PostScript format which is a true vector graphic format (this
is why the images do not pixellate when you zoom in for a closer view), and it is so simple
to use that I have never had to read the manual! Object libraries are easy to create for
Xcircuit, being plain-text files using PostScript programming conventions. Over the
years I have collected a large set of object libraries useful for drawing electrical and
electronic schematics, pictorial diagrams, and other technical illustrations.
159
Essentially an open-source clone of Adobe’s PhotoShop, I use Gimp to resize, crop, and
convert file formats for all of the photographic images appearing in the ModEL modules.
Although Gimp does offer its own scripting language (called Script-Fu), I have never
had occasion to use it. Thus, my utilization of Gimp to merely crop, resize, and convert
graphic images is akin to using a sword to slice bread.
This amazing project is a C++ library you may link to any C/C++ code for the purpose
of generating PostScript graphic images of mathematical functions. As a completely
free and open-source project, it does all the plotting I would otherwise use a Computer
Algebra System (CAS) such as Mathematica or Maple to do. It should be said that
ePiX is not a Computer Algebra System like Mathematica or Maple, but merely a
mathematical visualization tool. In other words, it won’t determine integrals for you
(you’ll have to implement that in your own C/C++ code!), but it can graph the results, and
it does so beautifully. What I really admire about ePiX is that it is a C++ programming
library, which means it builds on the existing power and toolset available with that
programming language. Mr. Hwang could have probably developed his own stand-alone
application for mathematical plotting, but by creating a C++ library to do the same thing
he accomplished something much greater.
160 APPENDIX C. TOOLS USED
Both Python and C++ find extensive use in these modules as instructional aids and
exercises, but I’m listing Python here as a tool for myself because I use it almost daily
as a calculator. If you open a Python interpreter console and type from math import
* you can type mathematical expressions and have it return results just as you would
on a hand calculator. Complex-number (i.e. phasor ) arithmetic is similarly supported
if you include the complex-math library (from cmath import *). Examples of this are
shown in the Programming References chapter (if included) in each module. Of course,
being a fully-featured programming language, Python also supports conditionals, loops,
and other structures useful for calculation of quantities. Also, running in a console
environment where all entries and returned values show as text in a chronologically-
ordered list makes it easy to copy-and-paste those calculations to document exactly how
they were performed.
Appendix D
By exercising the Licensed Rights (defined below), You accept and agree to be bound by the terms
and conditions of this Creative Commons Attribution 4.0 International Public License (“Public
License”). To the extent this Public License may be interpreted as a contract, You are granted the
Licensed Rights in consideration of Your acceptance of these terms and conditions, and the Licensor
grants You such rights in consideration of benefits the Licensor receives from making the Licensed
Material available under these terms and conditions.
Section 1 – Definitions.
a. Adapted Material means material subject to Copyright and Similar Rights that is derived
from or based upon the Licensed Material and in which the Licensed Material is translated, altered,
arranged, transformed, or otherwise modified in a manner requiring permission under the Copyright
and Similar Rights held by the Licensor. For purposes of this Public License, where the Licensed
Material is a musical work, performance, or sound recording, Adapted Material is always produced
where the Licensed Material is synched in timed relation with a moving image.
b. Adapter’s License means the license You apply to Your Copyright and Similar Rights in
Your contributions to Adapted Material in accordance with the terms and conditions of this Public
License.
c. Copyright and Similar Rights means copyright and/or similar rights closely related to
copyright including, without limitation, performance, broadcast, sound recording, and Sui Generis
Database Rights, without regard to how the rights are labeled or categorized. For purposes of this
Public License, the rights specified in Section 2(b)(1)-(2) are not Copyright and Similar Rights.
d. Effective Technological Measures means those measures that, in the absence of proper
authority, may not be circumvented under laws fulfilling obligations under Article 11 of the WIPO
Copyright Treaty adopted on December 20, 1996, and/or similar international agreements.
e. Exceptions and Limitations means fair use, fair dealing, and/or any other exception or
161
162 APPENDIX D. CREATIVE COMMONS LICENSE
limitation to Copyright and Similar Rights that applies to Your use of the Licensed Material.
f. Licensed Material means the artistic or literary work, database, or other material to which
the Licensor applied this Public License.
g. Licensed Rights means the rights granted to You subject to the terms and conditions of
this Public License, which are limited to all Copyright and Similar Rights that apply to Your use of
the Licensed Material and that the Licensor has authority to license.
h. Licensor means the individual(s) or entity(ies) granting rights under this Public License.
i. Share means to provide material to the public by any means or process that requires
permission under the Licensed Rights, such as reproduction, public display, public performance,
distribution, dissemination, communication, or importation, and to make material available to the
public including in ways that members of the public may access the material from a place and at a
time individually chosen by them.
j. Sui Generis Database Rights means rights other than copyright resulting from Directive
96/9/EC of the European Parliament and of the Council of 11 March 1996 on the legal protection
of databases, as amended and/or succeeded, as well as other essentially equivalent rights anywhere
in the world.
k. You means the individual or entity exercising the Licensed Rights under this Public License.
Your has a corresponding meaning.
Section 2 – Scope.
a. License grant.
1. Subject to the terms and conditions of this Public License, the Licensor hereby grants You a
worldwide, royalty-free, non-sublicensable, non-exclusive, irrevocable license to exercise the Licensed
Rights in the Licensed Material to:
2. Exceptions and Limitations. For the avoidance of doubt, where Exceptions and Limitations
apply to Your use, this Public License does not apply, and You do not need to comply with its terms
and conditions.
4. Media and formats; technical modifications allowed. The Licensor authorizes You to exercise
the Licensed Rights in all media and formats whether now known or hereafter created, and to make
technical modifications necessary to do so. The Licensor waives and/or agrees not to assert any right
or authority to forbid You from making technical modifications necessary to exercise the Licensed
Rights, including technical modifications necessary to circumvent Effective Technological Measures.
163
For purposes of this Public License, simply making modifications authorized by this Section 2(a)(4)
never produces Adapted Material.
5. Downstream recipients.
A. Offer from the Licensor – Licensed Material. Every recipient of the Licensed Material
automatically receives an offer from the Licensor to exercise the Licensed Rights under the terms
and conditions of this Public License.
B. No downstream restrictions. You may not offer or impose any additional or different terms
or conditions on, or apply any Effective Technological Measures to, the Licensed Material if doing
so restricts exercise of the Licensed Rights by any recipient of the Licensed Material.
b. Other rights.
1. Moral rights, such as the right of integrity, are not licensed under this Public License, nor
are publicity, privacy, and/or other similar personality rights; however, to the extent possible, the
Licensor waives and/or agrees not to assert any such rights held by the Licensor to the limited extent
necessary to allow You to exercise the Licensed Rights, but not otherwise.
2. Patent and trademark rights are not licensed under this Public License.
3. To the extent possible, the Licensor waives any right to collect royalties from You for the
exercise of the Licensed Rights, whether directly or through a collecting society under any voluntary
or waivable statutory or compulsory licensing scheme. In all other cases the Licensor expressly
reserves any right to collect such royalties.
Your exercise of the Licensed Rights is expressly made subject to the following conditions.
a. Attribution.
1. If You Share the Licensed Material (including in modified form), You must:
A. retain the following if it is supplied by the Licensor with the Licensed Material:
i. identification of the creator(s) of the Licensed Material and any others designated to receive
attribution, in any reasonable manner requested by the Licensor (including by pseudonym if
designated);
B. indicate if You modified the Licensed Material and retain an indication of any previous
modifications; and
C. indicate the Licensed Material is licensed under this Public License, and include the text of,
or the URI or hyperlink to, this Public License.
2. You may satisfy the conditions in Section 3(a)(1) in any reasonable manner based on the
medium, means, and context in which You Share the Licensed Material. For example, it may be
reasonable to satisfy the conditions by providing a URI or hyperlink to a resource that includes the
required information.
3. If requested by the Licensor, You must remove any of the information required by Section
3(a)(1)(A) to the extent reasonably practicable.
4. If You Share Adapted Material You produce, the Adapter’s License You apply must not
prevent recipients of the Adapted Material from complying with this Public License.
Where the Licensed Rights include Sui Generis Database Rights that apply to Your use of the
Licensed Material:
a. for the avoidance of doubt, Section 2(a)(1) grants You the right to extract, reuse, reproduce,
and Share all or a substantial portion of the contents of the database;
b. if You include all or a substantial portion of the database contents in a database in which
You have Sui Generis Database Rights, then the database in which You have Sui Generis Database
Rights (but not its individual contents) is Adapted Material; and
c. You must comply with the conditions in Section 3(a) if You Share all or a substantial portion
of the contents of the database.
For the avoidance of doubt, this Section 4 supplements and does not replace Your obligations
under this Public License where the Licensed Rights include other Copyright and Similar Rights.
a. Unless otherwise separately undertaken by the Licensor, to the extent possible, the Licensor
offers the Licensed Material as-is and as-available, and makes no representations or warranties of
any kind concerning the Licensed Material, whether express, implied, statutory, or other. This
includes, without limitation, warranties of title, merchantability, fitness for a particular purpose,
non-infringement, absence of latent or other defects, accuracy, or the presence or absence of errors,
165
whether or not known or discoverable. Where disclaimers of warranties are not allowed in full or in
part, this disclaimer may not apply to You.
b. To the extent possible, in no event will the Licensor be liable to You on any legal theory
(including, without limitation, negligence) or otherwise for any direct, special, indirect, incidental,
consequential, punitive, exemplary, or other losses, costs, expenses, or damages arising out of this
Public License or use of the Licensed Material, even if the Licensor has been advised of the possibility
of such losses, costs, expenses, or damages. Where a limitation of liability is not allowed in full or
in part, this limitation may not apply to You.
c. The disclaimer of warranties and limitation of liability provided above shall be interpreted in
a manner that, to the extent possible, most closely approximates an absolute disclaimer and waiver
of all liability.
a. This Public License applies for the term of the Copyright and Similar Rights licensed here.
However, if You fail to comply with this Public License, then Your rights under this Public License
terminate automatically.
b. Where Your right to use the Licensed Material has terminated under Section 6(a), it reinstates:
1. automatically as of the date the violation is cured, provided it is cured within 30 days of Your
discovery of the violation; or
For the avoidance of doubt, this Section 6(b) does not affect any right the Licensor may have to
seek remedies for Your violations of this Public License.
c. For the avoidance of doubt, the Licensor may also offer the Licensed Material under separate
terms or conditions or stop distributing the Licensed Material at any time; however, doing so will
not terminate this Public License.
a. The Licensor shall not be bound by any additional or different terms or conditions
communicated by You unless expressly agreed.
b. Any arrangements, understandings, or agreements regarding the Licensed Material not stated
herein are separate from and independent of the terms and conditions of this Public License.
Section 8 – Interpretation.
a. For the avoidance of doubt, this Public License does not, and shall not be interpreted to,
reduce, limit, restrict, or impose conditions on any use of the Licensed Material that could lawfully
166 APPENDIX D. CREATIVE COMMONS LICENSE
b. To the extent possible, if any provision of this Public License is deemed unenforceable, it shall
be automatically reformed to the minimum extent necessary to make it enforceable. If the provision
cannot be reformed, it shall be severed from this Public License without affecting the enforceability
of the remaining terms and conditions.
c. No term or condition of this Public License will be waived and no failure to comply consented
to unless expressly agreed to by the Licensor.
Creative Commons is not a party to its public licenses. Notwithstanding, Creative Commons
may elect to apply one of its public licenses to material it publishes and in those instances will
be considered the “Licensor.” Except for the limited purpose of indicating that material is shared
under a Creative Commons public license or as otherwise permitted by the Creative Commons
policies published at creativecommons.org/policies, Creative Commons does not authorize the
use of the trademark “Creative Commons” or any other trademark or logo of Creative Commons
without its prior written consent including, without limitation, in connection with any unauthorized
modifications to any of its public licenses or any other arrangements, understandings, or agreements
concerning use of licensed material. For the avoidance of doubt, this paragraph does not form part
of the public licenses.
References
Chase, Malcolm W. Jr., NIST-JANAF Thermochemical Tables, Fourth Edition, Part I, Al-Co,
Journal of Physical and Chemical Reference Data, Monograph No. 9, American Institute of Physics,
American Chemical Society, 1998.
Considine, Douglas C., Energy Technology Handbook, McGraw-Hill Book Company, New York, NY,
1977.
Control Valve Handbook, Third Edition, Fisher Controls International, Inc., Marshalltown, IA, 1999.
Dolmalski, Eugene S., Selected Values of Heats of Combustion and Heats of Formation of Organic
Compounds Containing the Elements C, H, N, O, P, and S, Chemical Thermodynamics Data Center,
National Bureau of Standards, Washington, D.C., 1972.
Giancoli, Douglas C., Physics for Scientists & Engineers, Third Edition, Prentice Hall, Upper Saddle
River, NJ, 2000.
Haug, Roger Tim, The Practical Handbook of Compost Engineering, CRC Press, LLC, Boca Raton,
FL, 1993.
Hicks, Tyler G., Standard Handbook of Engineering Calculations, McGraw-Hill, Inc., New York, NY,
1972.
Marks, Lionel Simeon and Davis, Harvey Nathaniel, Thermal Properties of Saturated and
Superheated Steam, tenth impression, Longmans, Green, and Company, New York, NY, 1920.
Mills, Ian; Cvitas̆, Tomislav; Homann, Klaus; Kallay, Nikola; Kuchitsu, Kozo, Quantities, Units and
Symbols in Physical Chemistry (the “Green Book”), Second Edition, International Union of Pure
and Applied Chemistry (IUPAC), Blackwell Science Ltd., Oxford, England, 1993.
169
170 APPENDIX E. REFERENCES
“NIOSH Pocket Guide to Chemical Hazards”, DHHS (NIOSH) publication # 2005-149, Department
of Health and Human Services (DHHS), Centers for Disease Control and Prevention (CDC), National
Institute for Occupational Safety and Health (NIOSH), Cincinnati, OH, September 2005.
Pauling, Linus, General Chemistry, Dover Publications, Inc., Mineola, NY, 1988.
Rosman, K.J.R. and Taylor, P.D.P, Isotopic Compositions of the Elements 1997, International Union
of Pure and Applied Chemistry (IUPAC), 1997.
Scerri, Eric R., “How Good Is the Quantum Mechanical Explanation of the Periodic System?”,
Journal of Chemical Education, Volume 75, Number 11, pages 1384-1385, 1998.
Thompson, Ambler and Taylor, Barry N., Guide for the Use of the International System of Units
(SI), special publication 811 (second printing), National Institute of Standards and Technology,
Gaithersburg, MD, 2008.
Weast, Robert C.; Astel, Melvin J.; and Beyer, William H., CRC Handbook of Chemistry and
Physics, 64th Edition, CRC Press, Inc., Boca Raton, FL, 1984.
Whitten, Kenneth W.; Gailey, Kenneth D.; and Davis, Raymond E., General Chemistry, Third
Edition, Saunders College Publishing, Philadelphia, PA, 1988.
Appendix F
Version history
This is a list showing all significant additions, corrections, and other edits made to this learning
module. Each entry is referenced by calendar date in reverse chronological order (newest version
first), which appears on the front cover of every learning module for easy reference. Any contributors
to this open-source document are listed here as well.
23 December 2023 – added a new Quantitative Reasoning question for frequency unit conversions.
29 November 2022 – placed questions at the top of the itemized list in the Introduction chapter
prompting students to devise experiments related to the tutorial content.
6-9 March 2022 – added a Full Tutorial section on unit conversions, as well as some Quantitative
Reasoning problems.
18 July 2021 – added an “Applying foundational concepts...” question to the Conceptual Reasoning
section of the Questions chapter.
22 March 2021 – added more digits to Boltzmann’s constant, and clarified that Kelvin is 273.15
more than Celsius.
18 March 2021 – corrected multiple instances of “volts” that should have been capitalized “Volts”.
24 September 2020 – added more content to the Introduction chapter, including references to
the “Reading outline and reflections” and “Foundational concepts” subsections. This is intended
for helping students new to inverted instruction adapt to the expectations of courses based on these
modules, where daily reading of these texts is fundamental to learning. Many students enter college
unfamiliar with how to outline texts, and so they need guidance on how to do so.
29 August 2020 – minor edits to the Tutorial regarding Electric Charge Conservation.
171
172 APPENDIX F. VERSION HISTORY
23 August 2020 – expanded Introduction chapter, and made a minor edit to the Full Tutorial for
typesetting purposes.
173
174 INDEX
Unit conversions, 91
Units of measurement, 149
Unity fraction, 91
Watt, 10
Weight, 5
Winch, 36
Work, 9, 27
Work in reverse to solve a problem, 150
WYSIWYG, 157, 158