Thanks to visit codestin.com
Credit goes to www.scribd.com

0% found this document useful (0 votes)
50 views167 pages

Energy - The Basics, 2nd Edition

Energy_ The Basics, 2nd Edition
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
0% found this document useful (0 votes)
50 views167 pages

Energy - The Basics, 2nd Edition

Energy_ The Basics, 2nd Edition
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
You are on page 1/ 167

ENERGY

T H E BA S I C S

Energy: The Basics offers a concise and engaging introduction to energy, answer­
ing critical questions and providing accessible definitions of essential concepts and
developments in the field.
People rarely stop to think about where the energy they use to power their everyday
lives comes from and when they do it is often to ask a worried question: is mankind’s
energy usage killing the planet? How do we deal with nuclear waste? What happens
when the oil runs out? Energy: The Basics answers these questions, but it also does
much more. In this engaging yet even-handed introduction, readers are introduced to:

 the concept of ‘energy’ and what it really means


 the ways energy is currently generated and the sources used
 new and emerging energy technologies such as solar power and biofuels
 the impacts of energy use on the environment including climate change

This new edition has been updated throughout and includes a new chapter on energy
storage, along with new material on transportation energy and batteries.
Featuring explanatory diagrams and an extensive further reading list, this
book is the ideal starting point for anyone interested in the impact and future
of the world’s energy supply.

Harold Schobert has nearly a half-century of experience in energy, primarily


research and development on coal science and technology. For the last ten years he
has been an independent consultant and provider of expert witness services. Before
partially retiring in 2013, he was Professor of Fuel Science at Penn State University
where, among many other activities, he served as director of the Earth and Mineral
Sciences Energy Institute and, for 25 years, taught an introductory course on energy
intended for students not pursuing science or engineering degrees. The second edi­
tion of Energy: The Basics is an outgrowth of that course. Previous to joining the
Penn State faculty, Schobert was manager of the Coal Science Division in the Uni­
versity of North Dakota Energy Research Center. Schobert is the author of 13
previous books, nearly 200 papers in peer-reviewed journals, and several hundred
additional papers published in conference preprints or proceedings.
ENERGY
THE BASICS
SECOND EDITION

H a rold S c hobert
Designed cover image: © Getty images
Second edition published 2025
by Routledge
4 Park Square, Milton Park, Abingdon, Oxon OX14 4RN
and by Routledge
605 Third Avenue, New York, NY 10158
Routledge is an imprint of the Taylor & Francis Group, an informa business
© 2025 Harold Schobert
The right of Harold Schobert to be identified as author of this work has been
asserted in accordance with sections 77 and 78 of the Copyright, Designs and
Patents Act 1988.
All rights reserved. No part of this book may be reprinted or reproduced or utilised in
any form or by any electronic, mechanical, or other means, now known or hereafter
invented, including photocopying and recording, or in any information storage or
retrieval system, without permission in writing from the publishers.
Trademark notice: Product or corporate names may be trademarks or registered
trademarks, and are used only for identification and explanation without intent to
infringe.
First edition published by Routledge 2013
British Library Cataloguing-in-Publication Data
A catalogue record for this book is available from the British Library

ISBN: 978-1-032-66554-2 (hbk)


ISBN: 978-1-032-66557-3 (pbk)
ISBN: 978-1-032-66560-3 (ebk)
DOI: 10.4324/9781032665603

Typeset in Times New Roman


by Taylor & Francis Books
CONTENTS

List of figures vii


Preface to the Second Edition ix

1 Energy and Us 1
2 Human Energy and the Energy Balance 10
3 Water, Wind, and Kinetic Energy 19
4 Heat, Steam, and Thermal Efficiency 28
5 Petroleum Fuels and Their Engines 39
6 Electricity 50
7 Electricity from Steam 60
8 Nuclear Energy 70
9 Energy from the Sun 81
10 Biomass Energy 90
11 Electricity from Falling Water 100
12 Electricity from Wind 110
13 Energy Storage 118
14 Energy and Climate 128
15 The 18-Terawatt Challenge 139
vi CONTENTS

Appendix 145
Suggested Further Reading 148
Index 150
FIGURES

1.1 This graph shows the relationship between per


capita energy use and gross domestic product for
various countries 3
3.1 A diagram of a waterfall and waterwheel. At point A
the water has a high amount of gravitational potential
energy. At B, a portion of the kinetic energy of the
water is converted to work in operating the
waterwheel. At C, the water still has some kinetic
energy, because it is still moving, and some potential
energy, because it could fall further somewhere
downstream 20
3.2 The energy diagram for a gravitational system. A
spontaneous change occurs in the direction of high
to low gravitational potential energy. With an
appropriate device, such as a waterwheel, some
energy can be captured and converted to useful
work 22
4.1 The energy diagram for a thermal system, which is
nearly identical to Figure 3.2, except for the kind of
potential energy involved. In this system, heat flows
from high to low thermal potential energy, giving us
the opportunity to capture some of the energy and
convert it into work 29
viii LIST OF ILLUSTRATIONS

5.1 A sketch of the global carbon cycle. Fossil fuels arise


from an interruption of the decay process that
organisms would normally experience in the
environment 40
5.2 A diagram of the four-stroke Otto cycle for
spark-ignition engines 42
6.1 A reverse energy diagram for an electrical system.
This diagram shows that a transition from low to high
potential energy is not spontaneous. To cause a
non-spontaneous change to occur, we must supply
work to the system, not extract work from it 56
8.1 The energy diagram for a nuclear process, very
similar to gravitational, thermal, and other systems 72
8.2 A simplified sketch of the curve of nuclear binding
energy. The higher a nucleus lies on this curve, the
greater its binding energy and the more stable it is 73
14.1 Data for the increase in atmospheric carbon dioxide
concentration, taken at Muana Loa, Hawaii 133
PREFACE TO THE SECOND
EDITION

Our use of energy pervades every activity that we undertake, and


every aspect of society. Agriculture, transportation, manufacturing,
and domestic activities all require, in various ways, electricity and
many other kinds of fuels. As we move deeper into the twenty-first
century, there seems to be a steadily increasing interest in non-pol­
luting energy sources that have little impact on the environment,
and that are unlikely to be depleted. Solar energy and wind energy
are strong contenders for meeting these requirements. They bring
with them the challenge of storing the energy they produce. Since
the latter decades of the twentieth century, there has been increasing
awareness that energy use is tightly linked with impacts on the
environment. The most massive of such impacts is the human con­
tribution to global climate change. This brings growing concern for
the development of energy sources that have low emissions of
carbon dioxide, and in an ideal case would be carbon-neutral. It is
important to understand that every energy source, regardless of
what it is, has some technological advantages and some dis­
advantages; has some negative impacts on the environment and
sometimes advantageous effects; and has some economic drawbacks
and some incentives. As a society, our challenge is to navigate our
way through this sequence of positive and negative issues, choosing
the optimum (or the least bad) solution for our locality. Despite the
claims of advocates of various specific energy sources, there is no
‘one size fits all’ answer.
This present book is a revised second edition of earlier book of
the same title, published in 2014. For about 25 years I taught a
course, Energy and Society, at Penn State University. The course
was intended for students who were not majoring in science or
engineering subjects, to fulfill part of their graduation requirement
to acquire some exposure to science. As the title of this book,
x PREFACE TO THE SECOND EDITION

Energy: The Basics, implies, I have tried to capture the essential


concepts from that course, and present them here in a book that can
be read by persons wanting to learn more about energy and its
impacts. I have assumed that most readers, whether students or
interested laypersons, will have minimal background in science and
mathematics, not beyond high school algebra and science. I hope
that readers without even that background can still enjoy, and learn
from, almost all the material in this book, picking up the needed
science or math ‘on the fly.’
It is always a great and special pleasure to acknowledge the con­
tributions of my wife, Nita, who assisted the preparation of this
book in many ways. Over the years I have benefitted from exchan­
ges of ideas and material with many colleagues at Penn State who
also taught this course, from useful discussions with friends and
colleagues at North-West University, from student feedback on
course evaluations, and certainly from my interactions with the
extraordinary variety of students who made their way through my
course. Certainly, any errors or shortcomings are entirely my
responsibility.
1

ENERGY AND US

OUR NEED FOR ENERGY


Everything we do depends on energy. At home, energy is important
for heating or cooling, lighting, and cooking. Very few of us are self-
sufficient in producing the food we need. Farming requires agricultural
machinery, industrially-produced fertilizers, or at the least the labor of
domesticated animals that require their own food sources. Food has to
be harvested, transported, processed, prepared, and packaged. Every
step in the chain from farm to consumer requires energy.
Our daily lives depend on manufactured articles. Not many of us
weave cloth, saw logs into boards, or make any of the other items
we use. Raw materials must be processed and fabricated into items
we want. Manufactured articles must be transported to stores. Daily
activity involves going out for work, shopping, or socializing.
Walking or cycling uses the energy of our own muscles. Carts or
carriages pulled by animals, cars or light trucks, buses, trains, and
airplanes all consume energy in some form.
Many people in the industrialized world are fortunate to be able
to surround themselves with electrical appliances and gadgets, often
more than one of each. When we’re buying one of these items,
nobody worries about whether there will be enough electricity to
operate them. Even asking whether there would be ‘enough’ elec­
tricity to operate the item we’re buying sounds silly. The question
never even comes up. We assume that we can use limitless electric
appliances. If there’s a problem, it’s whether there are enough elec­
tric sockets in the room.
The assumption about the eternal availability of unlimited quan­
tities of electricity is tested when the electricity supply is cut off
(commonly but incorrectly known as a power failure). When that
happens, many of us react with a feeling of annoyance or anger. No

DOI: 10.4324/9781032665603-1
2 ENERGY: THE BASICS

television, no cooking, no reading, no music, computer files may


have been lost…life’s dark in more ways than one. In some coun­
tries, we may have the same attitude toward gasoline. It seems to be
as easily and widely available as water. Gasoline shortages, even
sudden price spikes, produce the same anger as electricity outages.
Not everyone in the world enjoys the lifestyle that’s just been
described for the residents of prosperous, industrialized countries. In
too many places in the world, if electricity is available at all, it is
only for a limited time each day. Currently, the top ten countries in
which a substantial fraction (up to 30 percent) of the population
lacks ready access to electricity are all in sub-Saharan Africa.1
About 1.2 billion people, around 16 percent of the world’s popula­
tion, lack electricity.2 As with electricity, gasoline is not readily
available everywhere. But, for those people fortunate enough to live
in reasonably prosperous, industrialized nations, energy is readily
and always available at the flick of a switch, or the plugging of an
appliance into a wall socket, or a stop at a filling station. The key
idea is this: energy is ubiquitous in our lives and so common that we
seldom even think about it.
From the other perspective, how would we live if electricity, pet­
roleum products, and natural gas suddenly were not available any­
more? We would eat foods raised by ourselves or foraged in the
outdoors. We would get around by walking or by horseback. We
would stay warm with firewood or other combustible natural mate­
rials, if we had access to them, but only for as long as the supply
lasted. We would use clothes, tools, and utensils that we have now,
until they broke or wore out. Replacements would be made of wood
or natural fibers. We would rise at dawn and go to bed at dark. Our
existence would be about like that experienced during medieval
times. A very few, those competent at subsistence farming and at
the manufacturing or repairing of small tools and machinery, might
do fine.3

ENERGY AND NATIONS


Before the Industrial Revolution of the late-eighteenth century,
society relied on three energy sources: human and animal muscles,
firewood, and the energy of wind and water. Since then, there have
been three major historical transitions in our use of energy. First
came the steam engine, fueled mostly by coal. Second was the
development of electricity. Electricity was the first form of energy
for which its place of production could be separated by large
ENERGY AND US 3

distances from the place of its use. It is also the only energy source
easily converted into light, heat, or mechanical work wherever it is
used. Third was the internal combustion engine, which introduced
enormous mobility into society, at a cost of a significant depen­
dence on petroleum products.
In most countries, growth in energy consumption correlates to
income growth (e.g., Figure 1.1). Until about 1975 there was a
fairly good correlation between the economic well-being of a coun­
try, as measured by its total output of goods and services (gross
domestic product (GDP)), and its energy consumption. Countries
with the largest GDP tended to use the most energy, while undeve­
loped economies used less.
When thinking about this relationship, we should recognize that,
millennia ago, nobody first started taking advantage of the strength
of draft animals, or the energy of water and wind for mechanical
tasks, or the energy of wind to propel ships, for the purpose of get­
ting rich. They did it because using energy other than that of their
own bodies let them accomplish things more easily and provided
them a better quality of life.

Figure 1.1 This graph shows the relationship between per capita energy use
and gross domestic product for various countries
Source: ©2007 John E.J. Schmitz.
4 ENERGY: THE BASICS

As a country develops, GDP growth occurs through an increase


in population—demand for housing, transportation, consumer
goods, and services—and results in an increase in energy consump­
tion. The growth of GDP tends to parallel energy consumption, up
to a point. More precisely, the roughly linear relationship between
energy use and national GDP holds only to the point at which a
reasonably comfortable standard of living is achieved. Beyond that,
investment of money, energy, and resources in conspicuous con­
sumption4 results in energy use well above that needed for such
factors as health care, mass transit, and a good-quality education
system. But, abundant energy resources do not, by themselves,
guarantee quality of life. Consider Russia, the country with the
largest abundance and diversity of energy resources of any nation,
yet which seemingly is never able to achieve a quality of life con­
sistent with the possibilities of its energy consumption.
The last quarter of the twentieth century saw a change in the rela­
tionship between GDP and energy use, especially in developed coun­
tries. By 2000, energy use per dollar of GDP was dropping at about 2
percent per year, meaning that we are more efficient at generating
goods and services from the energy we use. The dollars of GDP pro­
duced for a given amount of energy consumed will reflect this.
The change came in part from the transition of developed coun­
tries to the so-called post-industrial society; that is, a transition from
an industry-based to a service-based economy, which requires much
less energy. The industries that remain in the post-industrial economy
tend to be ones that add considerable value to the materials or com­
ponents but require relatively little energy, such as making compu­
ters; this is, in comparison to an industrial society based on such
energy-intensive industries as steel and cement. As an example, ‘the
energy required to perform a hundred or so Google searches…would
heat the water needed to make a cup of tea.’5
Worldwide, enormous differences exist in levels of economic devel­
opment, standards of living, and access to energy. The richest one-fifth
of the world’s population consumes about four-fifths of the world’s
goods and services, and uses about half the world’s energy. The poor­
est one-fifth of the world’s people consumes 1 percent of its goods and
services, and gets by on about 5 percent of the world’s energy.

WHERE WE’RE GOING


Many people may feel that ‘science and technology’ have become
as foreign as another country, or at least another culture. The
ENERGY AND US 5

terminology about energy and the issues surrounding energy use


is all around us: getting up a head of steam, having a meltdown,
or experiencing the greenhouse effect are examples. It helps to
know what such terms mean, just as when visiting a country
where we do not speak the primary language, we can better
appreciate its culture and customs if we have at least a basic
vocabulary of its language. The world and the ways in which we
use energy are changing rapidly. Gaining an understanding of
energy is neither mysterious nor difficult, but it helps to know a
bit of the vocabulary of ‘energy.’
It’s also vital to understand that there are limits as to what can be
accomplished with energy. We will see that we can’t create energy
out of nothing. The best we can do is use whatever energy is avail­
able to us and convert it from one form to another. It makes sense
to make use of 100 percent of the available energy, with no waste or
losses. We’ll see that we will never really be able to attain this ideal.
The fundamental laws of nature that say that energy cannot be
created from nothing, and cannot be converted with 100 percent
efficiency to a different form, can be summarized in the statement
that, ‘You can’t win, and not only that, you can’t even break even.’
We must also understand that there are limits to what scientists
and engineers can accomplish, and what they can tell us. As an
example, reducing the environmental effects associated with a par­
ticular fuel could involve removing a pollutant from the fuel before
it is burned or capturing the pollutant before it can escape to the
environment after the fuel has been burned. Scientists can develop
methods for removing or capturing potential pollutants before they
impact the environment. Engineers can design plants for employing
these methods on a large scale. Economists can calculate the
increase in costs (in your electric bill, for example) resulting from
building and operating these plants. But none of these people can
tell or calculate, let alone dictate, your personal optimum trade-off
between increased electricity bills or fuel prices on one hand, versus
accepting greater environmental degradation on the other hand. It
is up to the individual citizen, or groups of citizens, to choose
between, e.g., less pollution but with increased electricity costs vs
forest destruction from loosely-regulated pollution. Some might
accept higher electric bills in exchange for protecting the environ­
ment. Others might adopt the ‘a tree’s a tree—how many do you
need to look at?’6 attitude. The choice does not come from scientific
or engineering calculations; it is a choice that each person must
make individually.
6 ENERGY: THE BASICS

THE LANGUAGE OF ENERGY


Understanding something of the culture and literature of science
requires knowing some of the language. This may seem peculiar,
because in any country scientists speak the same language as all the
other citizens—or at least they appear to. Confusion can be caused
because sometimes scientists use a common word—energy being a
fine example—with a meaning or connotation different from the
way the word is used in everyday speech. Many words used casually
in everyday conversation have more restricted or specialized
definitions when used in a scientific context.
All of us make and encounter comments like these: ‘I should wash
my car today, but I don’t have the energy.’; ‘I’ve converted my house
to run on solar energy.’; ‘I have to work all day tomorrow.’; ‘That test
was a lot of work!’; ‘A good golfer can hit a ball with a lot of power.’;
and ‘Oh no! The power just went off!’ Numerous similar examples
can be gleaned from everyday conversations and from the media. A
careful look will show that the same words are not being used in
quite the same way in each case. Surely the ‘power’ with which a
golfer hits a ball is not the ‘power’ that operates the television and
lamps in our home. (We hope!) The ‘energy’ to operate an entire
household is surely different from the ‘energy’ a person claims not to
have when he or she doesn’t want to wash the car.
Energy, work, and power are three key terms that we will use
throughout the remainder of this book. Their definitions are more
restricted than the ways in which we use them in everyday dis­
course. Let’s start with work. Work is causing an object to move
into, or out of, some position, especially when it moves against a
resistance. A few examples: carrying items upstairs against the
resistance of gravity; driving down the road against the resistance of
friction in the tires; fabricating materials against the resistance of
their natural strength or stiffness; and sending electricity from a
generating station to consumers against the electrical resistance of
the wires. Everything we try to accomplish in our daily lives—in
manufacturing, transportation, agriculture—represents doing work
in some way.
In a mechanical sense, a force of some sort must be applied to an
object, to move the object through a distance. This relationship
implies two things: first, if a force isn’t being exerted, nor a distance
traversed, then work isn’t being done; second, a large force applied
through a small distance can result in the same amount of work
being done as a small force applied through a large distance.
ENERGY AND US 7

To do work of any kind requires energy. The Nobel Laureate


Richard Feynman warns us that:

It is important to realize that in physics today, we have no knowledge


of what energy is. We do not have a picture that energy comes in little
blobs of a definite amount. It is not that way. However, there are for­
mulas for calculating some numerical quantity, and when we add it all
together it gives…always the same number. It is an abstract thing in
that it does not tell us the mechanism or the reasons for the various
formulas.7

Rather than trying to define energy in terms of what it is, we define


it in terms of what it does. Energy is the capacity for doing work.
Energy is measured in terms of the quantity of work that it is able
to perform.8 To do work, we need energy. Energy is the key to
accomplishing, or to having the capacity for doing, all of the activ­
ities that form the basis of modern society and of our lives.
We can think of concepts related to energy using analogies with
money. It’s obvious—sometimes painfully—that you can’t spend
unless you have money. That is, money is the capacity or ability to
spend. Money and spending have the same relationship as energy
and doing work. We do work to get something done, to accomplish
something, to make something happen. We need energy to do this.
We spend to acquire goods or services, to get what we need or want
for our daily lives. We need money to do this.
When buying something, money is transferred from ourselves to
the seller. The act of spending involves a transfer of money. In the
same way, the act of doing work involves a transfer of energy.
Though ordinarily there is one kind of money wherever we happen
to live, we will encounter many kinds of energy. As we go on, we
will examine what kinds of energy are being transferred, and how
that transfer takes place.
There’s another important aspect of what happens when we buy
something. The amount of your money drops, and the amount of
money that the seller has increases. But, the total amount of money
in circulation doesn’t change. That is, spending involves a transfer
of money but not a change in the total money supply. Similarly,
doing work involves a transfer of energy, but doesn’t change the
total amount of energy available in the world.
Power measures how quickly we can do work, or consume energy.
Power is the rate of doing work. This tells us that power always
includes some measure of time. When we say that a particular item
8 ENERGY: THE BASICS

(e.g., an automobile) is more powerful than another, what we mean


is either that it can do more work in a given amount of time or it
takes less time to do the same amount of work. A large amount of
work done in a short time, or a small amount of work done over a
long time, can represent the same amount of power.
Everything we want to do, or to have done for us, involves doing
work. We need energy to do that work. Power tells us how rapidly
we can accomplish a given amount of work or, from the other per­
spective, how much work we can do in a given amount of time.
Doing work lies at the heart of every activity that provides the food,
goods, and services we enjoy; energy is necessary to do that work;
and power tells us how fast we can do it.
These relationships represent the fundamentals of the tools and
machines used for helping us to do work. Imagine having to shift some
object that has a mass of 200 kg, such as a big rock or a large piece of
furniture—perhaps loading it from the ground into the back of a
truck. Very, very few of us can exert the force needed to lift and move
a 200-kg object straight upwards. Fortunately, our early ancestors
solved this problem intuitively. The secret of moving a 200-kg object a
short distance is to exert a force equivalent to moving, say, 20 kg over
ten times the distance. We can do this with a simple device that dates
from antiquity: a ramp, the so-called inclined plane. This and other
ubiquitous simple tools such as pulleys and levers derive from the
concept that an enormous amount of work can be done simply and
relatively easily by applying a small force through a large distance. We
roll the rock, or push the furniture, up a long, gentle slope.
For the overwhelming portion of human history, the only way
that humankind had of accomplishing work was to rely on the
muscles of the human body. Hominids, the family of primates of
which we are a member, are about 5,000,000 years old. Modern
humans (Homo sapiens, us) appeared in Africa about 200,000 years
ago. Domestication of horses first occurred about 5,000 years ago;
elephants and camels, some 3,000–4,000 years ago. For almost all
of human existence we relied on our own muscles to do work.
Waterwheels, used to do work such as grinding grain, came into
use about 3,000 years ago. Only since then—just 1.5 percent of
human existence—have we had any way of doing work other than
relying on human or animal muscles. Practical steam engines were
developed about 250 years ago. The first sustained generation of
electricity happened in 1831. The first oil well in the United States
was drilled in 1859. The first controlled nuclear reaction was
achieved in 1942.
ENERGY AND US 9

Imagine that the whole energy history of Homo sapiens was con­
densed into a single year. We have relied on our muscles to do work
for almost the entire span of history—all but nine ‘days’ of the
human ‘year.’ All but about four-and-a-half ‘days’ relied either on
human or animal muscles. For 97 percent of the entire time that
humans have existed, we have relied on only our own muscles for
doing work. Then, on December 31, the very last day, the steam
engine came into existence at about 11:00 a.m. The first nuclear
reactor was demonstrated at 9:00 p.m.

NOTES
1 https://worldpopulationreview.com/country-rankings/countries-without­
electricity.
2 https://www.visualcapitalist.com/mapped-billion-people-without-acces
s-to-electricity/.
3 A half-century ago, one of the founders of the science of ecology, Howard T.
Odum, said that ‘the citizen in the industrialized country thinks he can look
down upon the system of man, animals, and subsistence agriculture that pro­
vides some living from an acre or two in India when the monsoon rains are
favorable. Yet if fossil and nuclear fuels were cut off, we would have to recruit
farmers from India and other underdeveloped countries to show the now
affluent citizens how to survive on the land while the population was being
reduced a hundredfold to make it possible.’ (Odum, Howard T. Environment,
Power, and Society. Wiley-Interscience: New York, 1971, p. 120).
4 A term coined in 1899 by Thorstein Veblen, the Norwegian-American
economist and sociologist, generally meaning the purchasing, owning, and
showing off of items for the primary purpose of enhancing one’s prestige
and status.
5 From Sen, Paul. Einstein’s Fridge. Scribner: New York, 2021, p. 167.
6 Ronald W. Reagan, quoted in Borowitz, Andy. Profiles in Ignorance. Avid
Reader Press: New York, 2022, p. 19.
7 Feynman received the Nobel Prize in Physics in 1965. This quote is from
Feynman, Richard P., Leighton, Robert B., and Sands, Matthew. The
Feynman Lectures on Physics. Addison-Wesley Publishing Company: Read­
ing, MA, 1963, p. 4–2. He was undoubtedly one of the brightest stars in the
galaxy of twentieth-century physics.
8 This relationship was developed in the 1850s by the Scots physicist and
mathematician William Rankine. To have a break from his professional
work, Rankine amused himself by composing humorous songs. For his part,
Feynman loved to play the bongo drums. One can only wonder what a col­
laboration between the two would have been like.
2

HUMAN ENERGY AND THE


ENERGY BALANCE

HUMAN ENERGY
Studies in science and engineering are often facilitated by isolating—
either literally or conceptually—whatever it is that is being studied.
The object of the study could be almost anything: a microbe, a forest,
a machine, a rock formation, a chemical reaction. Let’s introduce the
concept of a system. A system is whatever portion of the universe is
under our specific observation or study. When we define a system,
sometimes we speak informally about ‘drawing a box’ around it. The
box can be very aptly thought of as a ‘window of attention.’1 The
rectangular outline of the imaginary box also frames an imaginary
window through which we focus on the object of study. This chapter
focuses on a system with which each of us is intimately familiar—
ourselves. We will make observations and draw inferences that we
can build upon further in subsequent chapters.
Everything that we do, every day of our lives, somehow involves
the doing of work as we defined it in Chapter 1. We also saw in
Chapter 1 that to have the capacity or ability for doing work, we
need energy. We might inquire, ‘What is “energy”—and have I got
any, or how much more could I call on in need, and where does it
come from, and what is it composed of ?’2
Prehistoric peoples likely recognized intuitively a relationship
between human energy and the intake of food. World history con­
tains dreadful tales of human suffering and death during famines or
deliberate starvation. Appalling experiences confirm that a person
who doesn’t, or can’t, eat reaches a point of not being able to fend
for, even care for, himself or herself. Death becomes inevitable. Life
requires the energy source of food.
How much energy is in food? Science is always on firmest ground
when accurate measurements can be made, when the measured

DOI: 10.4324/9781032665603-2
HUMAN ENERGY AND THE ENERGY BALANCE 11

results can be expressed numerically, and when numerical rela­


tionships can be established among the measurements. The great
nineteenth-century scientist William Thomson (Lord Kelvin) put it
more bluntly, ‘When you can measure what you are speaking
about and express it in numbers you know something about it, but
when you cannot your knowledge is of a meager and unsatisfac­
tory kind.’3
The device used to determine the energy released from foods, a
calorimeter, measures the heat given off when a weighed sample of
the food is burned under controlled, standardized conditions. (This
suggests that we will find relationships between food and substances
that we think of as fuels.) The energy content of various foods is
expressed in two systems of units. Most nations use the unit of
kilojoules (kJ). One kilojoule is energy equivalent to raising the
temperature of 239 grams of water by 1°C. The other unit, nutri­
tional calories (Calories), used mainly in the United States, is the
amount of energy that will raise the temperature of 1,000 grams of
water 1°C.4
The amount of energy derived from food depends not only on the
kind of food, but also on how much we eat. We derive more energy
from an apple than from a tiny grain of sugar. To be fair and useful,
a comparison of the amounts of energy available from different foods
needs to be made on the basis of equal quantities. The important
thing to remember is that whether using kilojoules or Calories, what
unit the ‘per’ refers to—liters, kilograms, cups, pints, whatever—
must always be the same for the comparison to be make sense.
Where does the energy in food come from? Many of us eat a varied
diet from both plant and animal sources. Plant products include
vegetables, grains, nuts, fruits, or berries. All foods derive ultimately
from plants. Animals consumed as food obtained their nourishment
by eating plants. So, the source of energy in food can be narrowed
down to asking about the origin of food energy in plants.
The vital process in the growth and development of plants is
photosynthesis. The prefix photo- signals that light is important in
the process. Plants absorb carbon dioxide from the atmosphere and,
by using energy from sunlight, combine it with water to form
molecules of sugars. The growing plant uses the sugar molecules to
form all of the many other chemical substances needed for its
growth and life processes. The solar energy that enabled the con­
version of two rather simple, but quite stable, molecules—carbon
dioxide and water—is now stored in bonds between the atoms in
the molecules in the components of plants. Put another way, plants
12 ENERGY: THE BASICS

represent stored solar energy. Since we eat plants, or animals that


ate plants, to obtain the energy needed for our bodies, human
energy ultimately derives from solar energy.
We depend upon a sequence of energy-conversion steps: solar
energy is converted to plant energy, the plant energy is converted to
energy for animals, and the food that we consume is converted to
energy for us. As we go forward, we’ll see that almost every sort of
process for doing work breaks down into similar chains of energy
conversions. We humans are not very efficient solar energy con­
verters. Only about 1 to 2 percent of the solar energy reaching the
surface of the Earth is converted into the substance of living plants.
We consume only a tiny fraction of the plant or animal population
on Earth. Overall, human energy represents about two millionths of
the solar energy received on Earth.
In the nineteenth century, chemists studying the composition of
foods and the biological processes of life established three key facts:
(1) The major components of foods are carbon, hydrogen, and
oxygen. (2) Many foods contain small amounts of nitrogen and
sulfur, and various other chemical elements in still smaller quan­
tities. (3) We inhale air, a mixture of nitrogen and oxygen. When we
exhale, the amount of oxygen in the air is depleted, and our breath
contains increased amounts of carbon dioxide and water vapor.
Other chemists studying what happens when fuels such as wood
or coal are burned made comparable observations: (1) The major
components of fuels are carbon, hydrogen, and oxygen. (2) Some
fuels contain small amounts of nitrogen and sulfur, and various
other elements in still smaller quantities. (3) A fuel needs air to
burn. The air surrounding the burning fuel becomes depleted in
oxygen, and contains carbon dioxide and water vapor.
Rather similar lists! These similarities are not coincidental. We
recognize that energy is liberated when a fuel burns, because we can
feel the heat of the fire. The similarities between foods and fuels, in
terms of composition and of the processes that occur when each is
consumed, lead to a reasonable assumption: consuming food also
liberates energy.
About 250 years ago the great French chemist Antoine Lavoisier5
pointed out that, ‘Respiration is then a combustion, admittedly very
slow, but nevertheless completely analogous to that of charcoal.’
Just as burning wood in a fireplace produces heat that helps to keep
a house warm, energy from the consumption of food helps to
maintain our body temperature. In a warm region, we need to burn
less wood for comfort in the house than in colder regions. In the
HUMAN ENERGY AND THE ENERGY BALANCE 13

same way, we need to ‘burn’ less food to maintain a healthy body


temperature in warm climates. Needing to consume less food also
means needing less oxygen. This was observed by the German
physician Julius von Mayer. Serving on a vessel sailing in the Dutch
Indies (now Indonesia), and practicing in an era when bloodletting
was a therapeutic treatment, he observed that venous blood from
sailors was almost as bright red as arterial blood. This observation
indicated that the body was consuming less oxygen from the blood,
and hence needed less energy from food, than would have been
expected in colder climates. Regardless of climate, sailors were
obtaining warmth from two sources: the atmospheric heat around
them and heat generated internally from food. If a warmer climate
seemed to require less heat from food for maintaining the same
body temperature, then it seemed reasonable that the sum of the
two sources of energy (heat from the air and heat from food) must
be constant. Further development of this notion led to one of the
most fundamental laws of science: the conservation of energy.6
The same calorimeter used to measure the energy released by
foods can be used to test fuels as well. There is no difference in the
energy released from a substance whether it is eaten as a food or
burned as a fuel. Imagine being so extravagant as to burn sugar in a
home furnace or an automobile engine (supposing that we could
figure out how). Exactly the same amount of energy is released if
sugar were burned as a fuel as when the same amount of sugar is
consumed as food. The specific chemical pathways by which sugar
is converted to energy in the body or is converted to energy in a
furnace are quite different (consider the important observation that
we don’t set ourselves on fire by eating), but the energy output for a
given amount of sugar is exactly the same. If humans could eat
coal, the energy that would be released by eating an amount of coal
would be exactly the same as released by burning that same amount
in a furnace.
Burning fuels under the right conditions—or sometimes under
the wrong conditions—can produce spectacular fires. Some of us
have direct experiential evidence, from grills or kitchen stoves, that
foods can also produce remarkable flames if accidentally ignited.
We might ask why, when the energy is liberated from food in our
bodies, we don’t catch on fire ourselves. The overall chemical path­
ways of burning fuel and digesting food consume oxygen and lead
to the same products, carbon dioxide and water vapor,7 but the
details, at the molecular level, are quite different. Burning involves a
reaction with oxygen molecules from the surrounding air. Digestion
14 ENERGY: THE BASICS

proceeds through a sequence of biological reactions. Different che­


mical pathways, even leading to the same products, proceed at dif­
ferent rates. Digestion is much slower than burning. A bit of fat
accidentally ignited on your stove top burns quickly; that same
amount of fat taken into your body requires hours to digest. The
total energy released may be the same, the products may be the
same, but the rates are very different.
The difference in rates of burning and digestion has an important
implication for human activity. The slow release of energy from
digestion limits the amount of work we can do in a given time, i.e.,
limits our power. Since the work we can do in a given time is lim­
ited, so is our power. We humans are not very powerful.
Imagine the job of clearing a field of rocks. Perhaps in one
workday a person might shift a tonne of rock a distance of 100 m, if
the job had to be done manually. That person’s power output for
the day is a tonne per eight hours, i.e., 480 minutes.8 Few of us
would need to do this very long before getting the idea that there
must be a much easier way to move these rocks. Why not load them
all in the back of a truck, and drive the truck? If it takes the truck,
once loaded, a minute to drive off the field, the truck takes only 1/
480th of the time; from the other perspective, the truck is 480 times
as powerful as a human. Since a human is not as powerful as a
truck, a human does not use energy as rapidly as a truck. In a
human, the energy source, food, is converted rather slowly, and
indeed we don’t catch on fire. In a truck, the energy source, gasoline
or diesel fuel, is consumed quickly and most certainly does burn.
Fast consumption of energy leads to greater power.

THE ENERGY BALANCE


Chapter 1 introduced some analogies between energy and money.
Suppose that you just earned $100, and that you promptly buy an
item that costs $50. The remaining $50 didn’t disappear or ‘evapo­
rate’; this is money that you have saved, whether by keeping it on
your person, hiding it someplace, or putting it into a bank account.
That is, you have stored the remaining $50 someplace. We can
summarize this arithmetically by saying that the money paid in to
you, minus the money used for purchases, leaves an amount of
money that can be stored. This little bit of arithmetic is the basis of
balancing your finances.
Now suppose that you earn $100 but want to buy an item that
costs $150. As we all learn sooner or later, this transaction is not
HUMAN ENERGY AND THE ENERGY BALANCE 15

possible unless you are able to come up with the ‘missing’ $50. In
this case, the ‘money stored’ is actually a negative number, –$50.
The negative number means that the ‘missing’ $50 has to be pro­
vided by removing it (implied by the minus sign) from the money
that might have been saved or ‘stored’ earlier. In other words, the
total amount of money you have in savings or in ‘storage’ decreases.
Many of us develop an intuitive understanding of this approach
to keeping track of personal finances and the relationships among
the money we receive, the amount that we use, and what we have
saved almost as soon as we are first entrusted with money of our
own. Some of us learn the concept of negative numbers the hard
way. Human energy is exactly analogous: food energy taken in,
minus the food energy used, represents food energy stored some­
how. Storage of food energy occurs in several ways: in the short
term, by storing carbohydrates in the liver; in the long term, in a
reserve supply of energy of fatty tissue.
When the ‘food energy in’ is exactly equal to ‘food energy used,’
the difference is zero. We have neither increased nor decreased the
amount of energy stored; in practical terms, we would neither gain
nor lose weight. If the food energy taken in is greater than the
amount used, we increase the amount of energy stored in our ‘food
energy bank.’ We see this as a gain in weight. If the energy used for
doing work is greater than the amount taken in, the ‘food energy
stored’ becomes negative. We can do this much work with less
energy taken in only by withdrawing energy from the ‘energy bank,’
just like trying to purchase a $150 item with a $100 paycheck. The
analogy goes further. Some people are in such dire poverty that this
sort of financial transaction is not possible, because they do not
have the additional $50 saved anyplace, nor even a line of credit.
People in the last stages of starvation may no longer have extra
food energy stored in their bodies; for them, it is impossible to do
more work than they can ‘pay for’ with food energy taken in that
day. When even that little bit of work that can no longer be done
means only getting up from bed to obtain food, death is inevitable.
The food energy balance is one approach to regulating our body
weight.9 Keeping track of the amount of each food, and its associated
energy content, eaten during the course of a day gives a numerical
value for ‘food energy in.’ Information on the energy used in various
kinds of activities is widely available on the web. Anyone can use such
information to help regulate his or her body weight. We can keep track
of the energy that we use, and the energy that we take in as food, in
exactly the same way that we keep track of our personal finances.
16 ENERGY: THE BASICS

The importance of the balance of food energy is that the concept


is actually relevant to any kind of energy. When we are concerned
about balancing human energy, the system is our body. Just as we
balance our bank accounts, or see whether we’ve gained or lost
weight, we can speak of the energy balance for a system. For any
system whatsoever, Energy in-Energy out = Energy stored. If more
energy is put into a system than is taken (or comes) out, some
energy has to be stored somehow inside the system. If we take more
energy out of a system than we put into it, then some energy that
had previously been stored in the system must have been used. If
there is no energy previously stored in the system, the transaction is
impossible. This leads to a general, extremely important, statement:
if no energy is already stored in a system, a process that produces
more energy than was originally put in is impossible.
Outdoors on a cold day, we do a bit of exercising because it helps
us to feel warmer. After a car has been driven for a time, the hood
will feel perceptibly warm. An electric motor running steadily will
feel slightly warm. What does your body’s getting hot have to do
with helping you do work? What does the hood of a car getting
warm have to do with the work of moving the car down the road?
What does the casing of an electric motor getting warm have to do
with helping it turn a fan? The answer to each of these questions is
exactly the same: nothing. These examples are actually profound
observations about the world and how it works: whenever energy is
converted into work, a portion is wasted (usually as heat). Another
way of saying the same thing is that: energy cannot be converted
completely to work. Some energy is inevitably wasted.
Where does the portion of energy that is apparently wasted go?
Some can be absorbed by nearby objects, but eventually it becomes
radiated away from Earth into outer space. There’s no way for us to
get it back or to make any further use of it. The waste heat becomes
a part of an enormous reservoir of heat energy in the universe, but
the crucial point is that such energy is no longer available to us.
This is not because we lack the appropriate kind of technology to
bring it back. Rather, the reason is much more fundamental. The
kind of energy associated with heat can only be made to move (in
other words, be used) if there is a difference in temperature between
two objects, as we will see in Chapter 4. The waste heat radiated
away from the Earth becomes part of the vast reservoir of heat
energy in the universe, which happens to be at a uniform tempera­
ture. There is no way it can ever be used again.
HUMAN ENERGY AND THE ENERGY BALANCE 17

Our planet is home to more than 8.2 billion people,10 several


billion vehicles, and millions of factories, airplanes, railway loco­
motives, ships, and electricity-generating plants. All of these things,
including us, generate heat every day. If a mechanism did not exist
to get rid of excess heat, Earth would have become uninhabitable
long ago. But now we might finally be on the brink of doing just
that, because we are changing the composition of the atmosphere in
such a way as to retain more of this heat on Earth, rather than
allowing it to radiate away into space. This problem is popularly
known as the greenhouse effect, though more properly referred to
as global climate change or global warming. We will examine it in
more detail in Chapter 14, and will find that understanding the
concept of the energy balance is one of the keys to understanding
the energy balance of our planet—the only one we’ve got.

NOTES
1 Odum, Howard T. and Odum, Elisabeth C. A Prosperous Way Down.
University Press of Colorado: Boulder, 2001, p. 61.
2 This is a paraphrase of a question concerning morale, originally proposed by
an Englishwoman, Nella Last, during the darkest days of World War II. She
is quoted in Downing, Taylor. 1942. Pegasus Books: New York, 2022, p. 54.
3 Quoted in Jorgensen, Timothy J. Strange Glow. Princeton University Press:
Princeton, 2016, p. 187.
4 There is an unfortunate opportunity for confusion with these terms. In sci­
entific work, other than nutrition, 1 calorie is the amount of heat energy
required to raise the temperature of 1 gram of water by 1°C. A nutritional
calorie, or Calorie, is therefore 1,000 calories.
5 The quotation is from Sen, Paul. Einstein’s Fridge. Scribner: New York,
2021, p. 44. Lavoisier, one of the greatest of the eighteenth-century scien­
tists, is best known for elucidating the role of oxygen in combustion. He
was executed during the French Revolution, the judge in his case approving
the sentence by supposedly saying that the Republic needed neither scho­
lars nor chemists, an attitude that tragically still persists in some parts of
the world. As it turned out, a mere three months later the Republic had no
use for the judge either, and he met the same fate.
6 This fundamental rule, also known as the First Law of Thermodynamics,
is usually credited to a contemporary German physicist, Hermann von
Helmholtz.
7 The First Law of Thermodynamics tells us that when we start with the
same material, and wind up with the exact same products, the energy
released in that conversion must be the same, regardless of the pathway by
which the conversion occurred. More formally, ‘the change in energy
accompanying a change of a system from an initial state to a final state…is
independent of the path by way of which this change is effected.’ (From
Pauling, Linus. General Chemistry. W.H. Freeman and Company: San
Francisco, 1970, p. 345.)
18 ENERGY: THE BASICS

8 This is equivalent to a power output of 34 joules per second (1/22 horsepower).


This is very close to the average sustained human power output of 37 J/s or 1/20
horsepower (from Heinberg, Richard. The Party’s Over. New Society
Publishers: Gabriola Island, BC, 2005, p. 30).
9 A useful and informative source is Roberts, Susan B. and Flaherman, Val­
erie. ‘Dietary energy.’ Advances in Nutrition, Vol. 13, pp. 2681–2685, 2022.
10 In the 2014 edition of this book, the number was slightly more than six
billion. That’s about a 33 percent increase in a little over a decade, which
should be terrifying to anyone who is concerned about the future of the
planet or of humankind. Check the website https://www.worldometers.info/
world-population/#google_vignette and watch in horror.
3

WATER, WIND, AND KINETIC


ENERGY

OBSERVING WATER
Think about just watching a waterfall.1 Let it be our system, our
window of attention. A first observation might be that, of course,
the water coming down the fall is moving. The moving water can
exert a force. Something caught in the water will be swept along. A
person standing at the bottom of the fall, in the rushing water,
could be knocked down. It’s easy to experience the force of moving
water by cupping your hand and holding it in the stream flowing
from a wide-open kitchen or bathroom tap—feel the force of water
trying to push your hand away. (It might be better to perform this
experiment in someone else’s kitchen.) Moving water has a force
that can move a resistance (you) through a distance, which is the
same as saying that moving water is capable of doing work.
Water that is not moving can’t do this. We don’t worry about
being knocked down when standing in a swimming pool or a pond.
The difference is that the water in these examples isn’t moving. Only
moving water has the capacity for doing work. Water standing still,
as in a pool, does not. This means that moving water must possess
some particular kind of energy that allows it to do work.
The energy associated with motion, as in water moving in a
waterfall, is called kinetic energy. The energy associated with moving
bodies is called kinetic energy. Kinetic energy can be used to do
work. The amount of kinetic energy depends on the mass of the
water that is flowing, and on its velocity.
Imagine a simple wheel with paddles on it, mounted on a shaft.
When it is partially immersed in a flowing stream, as in a waterfall,
water will push against the paddles, just as it pushes against us if we
are standing at the bottom. The ‘push’ of the water against the
paddles will cause the wheel to turn. Water is doing work on this

DOI: 10.4324/9781032665603-3
20 ENERGY: THE BASICS

special kind of wheel, which we can call a waterwheel. As the


waterwheel and the shaft on which it is mounted turn, the kinetic
energy of water does work on the waterwheel; by connecting the
turning shaft to some sort of device, that device can do work for us.
The kinetic energy of falling water was one of the first energy
sources put to work (literally) by humans that did not involve our
muscles or those of animals. (The other was the kinetic energy of
wind, discussed later in this chapter.) This system—waterwheel,
shaft, and operating device—fits the simplest definition of a
machine: ‘any device that transmits or modifies energy.’2
Let’s now think about this system: a waterfall that happens to have
a waterwheel mounted somewhere in the fall. Water that is in the
stream before the fall will not be able to do work, because it is not
able to come in contact with the waterwheel. For our purposes, that
upstream water has no useful kinetic energy.3 But potentially it could
do work, as soon as it reaches the lip of the fall and starts down­
ward, acquiring kinetic energy. We can speak of the upstream water
as having potential energy, energy that we can’t use just yet. It starts
to fall downward as it flows off the edge of the waterfall because it is
being pulled toward the center of the Earth by the force of gravity.
Figure 3.1 illustrates the key concepts. Water at A, upstream of a
waterfall, is not able to do work on the waterwheel (at least not

Figure 3.1 A diagram of a waterfall and waterwheel. At point A the water has
a high amount of gravitational potential energy. At B, a portion of
the kinetic energy of the water is converted to work in operating the
waterwheel. At C, the water still has some kinetic energy, because it
is still moving, and some potential energy, because it could fall
further somewhere downstream
WATER, WIND, AND KINETIC ENERGY 21

yet), because it does not have kinetic energy in the direction down­
ward along the waterfall. But, water at A could potentially do work.
All that’s needed is for that bit of water to acquire kinetic energy, by
falling off the cliff over the waterfall. We say that the water at A has
gravitational potential energy because its potential for doing work is
due to its position in Earth’s gravitational field.
B is the point at which the water is hitting the blades of the
waterwheel. The kinetic energy of the water imparts kinetic energy
to the wheel; with an appropriate mechanism, the motion of the
wheel can be used to do work. It might be tempting to conclude
that the water no longer has potential energy at point C. However,
it could—in principle, at least—encounter another waterfall some­
where further downstream. Water at C still has some remaining
gravitational potential energy because it has the potential of falling
still further. In fact, it will always have some gravitational potential
energy unless it happens to be at the center of the Earth. The idea
of extracting more work from moving water downstream of our
waterwheel system is often used to advantage. A spectacular exam­
ple is the Cunene River in Angola, which had a series of 29 artifi­
cial waterfalls (dams) for generating electricity by the technology we
will see in Chapter 11.4
We know from many personal, sometimes painful, observations
that things always fall down—never up. Anything able to move will
fall spontaneously from a position of high gravitational potential
energy to a new position of lower gravitational potential energy. We
can extract kinetic energy—in this case, by using a waterwheel—
when that change from high to low potential energy occurs. The
gravitational potential energy at C is lower than at A because C is
closer to the center of the Earth.
The essentials of the system shown in Figure 3.1 can be repre­
sented by a simpler diagram, Figure 3.2. The central concept that
goes along with Figure 3.2 is this: when a system undergoes a spon­
taneous change from high potential to low potential, we can extract
energy (or work) from the system.
We don’t know where or when waterwheels were first used, or
who developed them. Likely, there were multiple, independent
discoveries at various places around the world at various times.
The first clear description was given by the Roman engineer and
architect Marcus Vitruvius Pollio, in his book De architectura,
published about 27 BCE.5 Vitruvius claimed that the machines
he described were in common use, and attributed their invention
to Greek engineers.
22 ENERGY: THE BASICS

Figure 3.2 The energy diagram for a gravitational system. A spontaneous


change occurs in the direction of high to low gravitational potential
energy. With an appropriate device, such as a waterwheel, some
energy can be captured and converted to useful work

The Domesday Book, William the Conqueror’s inventory of his


new conquest, lists nearly 6,000 watermills in England in 1086.6 For
medieval Europe, and likely China and Persia, water-operated
machinery was used for making flour, cloth, leather, beer, paper,
and sawn lumber. Waterwheels were raising ore from mines, oper­
ating hammers for crushing ore, operating the bellows on furnaces,
hammering hot iron into shape, and operating grindstones for
sharpening tools. There may have been a half-million throughout
Europe by 1800. Some were in use for 700 years; a few were mod­
ified and still working into the twentieth century. Will people in the
year 2725 be using any device built in this century?
Nonetheless, even massive waterwheels provided remarkably little
power. Only 10 percent of the kinetic energy of the water was con­
verted to work. Ideally, we would like all of the kinetic energy of the
water to be converted into work, and none wasted. We never seem
to be able to get there. Not only that, for a given mass flow of water,
there is an upper limit to how much work can be produced by even
the finest designs of waterwheels. That upper limit is fixed by the
height through which the water falls—the change in its gravitational
potential energy—as it passes over the wheel. If we have a wheel
installed in a waterfall, we can’t change the height of the waterfall,
but what we can change is the position in which we install the
wheel. We could choose to install a wheel near the top of the fall, or
near the bottom. In the near-the-top case, not much potential
energy of the water will be converted to kinetic energy; in the near-
the-bottom case, much more potential energy will be converted to
kinetic energy. The latter case produces a greater amount of useful
work, but we still can’t convert all of the kinetic energy to work.
WATER, WIND, AND KINETIC ENERGY 23

This relatively simple system demonstrates the conversion of


potential to kinetic energy, and of kinetic energy to work. What
appear to be the limitations in those conversions put us on the track
of some fundamental principles of energy and its applications,
which we will elaborate further in the coming chapters.
Water is relatively easy to control; many of the basic principles were
understood in classical or medieval times. Though most waterwheels
provide low power, the best designs convert 80 to 90 percent of kinetic
energy in water to useful work. This is excellent in comparison to
many other energy systems. An advantage of waterwheels is that run­
ning water is usually free; the efficiency of converting energy to work is
a greater concern when the fuel has to be purchased.
Sometimes practical use of a waterwheel requires changing the
speed or the direction of rotation, or both. For example, there
might be a need to exploit the energy of a fast-moving stream to
operate a heavy, slow-moving hammer rock crusher. Or, we might
want to connect a wheel having a horizontal shaft to a vertical shaft
turning millstones for grinding grain to flour. Speed or direction of
motion can be changed using gears, exactly as happens millions of
times each day in the transmissions and differentials of modern
vehicles. The development of gears and their use to change the
speed of rotation, roughly 2,000 years ago, was the first major
achievement in the design of continuously operated machinery.

WATER CYCLES
Where does the gravitational potential energy of water come from
in the first place? Imagine a river that flows over a waterfall and,
from there, into the ocean. Given enough time, sooner or later all
the water in the river will have to wind up in the ocean—unless the
water in the upper portion of the river can somehow be replenished.
As water in the ocean is heated by the sun, some evaporates into the
atmosphere. Moisture-laden warm air rises into the atmosphere;
some will be carried over land. Somewhere, the air will cool, caus­
ing the water vapor to condense and to fall as rain or snow. Such
water will accumulate eventually in streams, which run into larger
streams, which run into a river, which, in this example, runs into the
ocean. The cycle will begin all over again, and will run for eons,
unless perturbed by some major climatic change.
The model describing the transport and fate of water
throughout the world is known as the hydrologic cycle. The
energy that drives the operation of the hydrologic cycle comes
24 ENERGY: THE BASICS

from the sun—solar energy. Energy engineers who exploit water


for doing work are using solar energy in an indirect way, relying on the
sun to drive the hydrologic cycle to ensure a steady supply of water
with the necessary potential energy. Water ‘power’ is an indirect form
of solar energy.

WORK FROM WIND


Wind also possesses kinetic energy. We all occasionally experience
the force of the wind in various ways—scattering papers, blowing a
hat off, or just the difficulty of walking into a strong wind. Such
experiences have much in common with ways in which we person­
ally experience the force of moving water. Just as moving water can
be harnessed to do work, we can also harness the wind.
Wind is a mass of air in motion relative to the surface of the
Earth. Like any other body in motion, wind possesses kinetic
energy. Wind loses some kinetic energy due to friction with the
Earth’s surface and due to friction within the wind itself. If these
losses are not restored, eventually so much kinetic energy would be
lost that the wind would stop blowing. Of course, this does not
happen. Wind blows continually, and, in some parts of the world
almost all the time. From the energy balance, we know that we
cannot extract unlimited energy from a system, because at some
point all of the stored energy would become depleted and any fur­
ther removal of energy would be impossible. Energy lost from the
wind by friction effects, or what we extract from it to do work, must
be replenished.
The energy source that causes wind, and that replenishes the
energy lost from wind, is the sun. Wind is caused by the unequal
heating of various parts of Earth’s atmosphere. Energy from the sun
is absorbed by the land and water surfaces. Soil, green plants, and
oceans absorb the sun’s radiation to different extents. Not all loca­
tions on Earth’s surface are heated to the same extent by the sun.
The heated land and water heat the air above them. Because of the
different absorption levels of solar radiation by different substances,
and the different amounts of solar radiation supplied to various
places on Earth, air is not heated to the same extent everywhere in
the world. Air above the hotter locations expands and, as it does, rises.
As warm air rises into the atmosphere, air from cooler locations flows
toward the area where the warm air is rising, establishing wind cur­
rents. Winds are a manifestation of solar energy.
WATER, WIND, AND KINETIC ENERGY 25

Wind represents the spontaneous flow from higher to lower


potential. In this case, air flows from regions of high atmospheric
pressure to lower pressure. A high-to-low pressure flow is true of
any gas. The greater the pressure difference, the higher the speed of
the air flow.
Regions of pressure differences in the atmosphere are actually due
to temperature differences. The sun warms the atmosphere, but not
to the same extent everywhere. Even with no other complications,
sunlight does not strike the surface of the Earth to the same extent at
different latitudes. But there are other complications: oceans, moun­
tains, and large tracts of land that do not warm to the same extent.
The sailing ship represents the first time that energy available in
nature was successfully harnessed to do work.7 Cloth sails were
known at least by 3500 BCE. The first practical development of
windmills came in the Islamic countries of the Middle East during
the seventh century. The stimulus was the need to grind grain in
areas lacking steadily flowing streams of water. An abundant supply
of wind energy occurs in many parts of the Middle East, sometimes
at speeds up to 160 km/hr.
Wind differs in a crucial aspect from water: water flows in a
single direction, determined by the bed of the stream or river, but
wind direction can shift. In many places, wind direction can change
even sometimes during a single day.
A windmill has several advantages relative to a waterwheel. No
investment is needed for water-control equipment. The windmill can
operate in arid areas, or in areas where there are no fast-moving
streams, or even in freezing winter weather. The fact that no one can
‘shut off’ the wind means that windmills could be used in besieged
castles or towns for grinding grain to continue feeding the population.8
Leonardo da Vinci developed what is the stereotypical windmill.
He realized that it was only the top of the mill, with the sails or
vanes, that needed to turn with the wind. In this design, called a
turret mill or tower mill, the main portion is a non-moving, solid
structure. With this design, mills could be made much larger and
more efficient than earlier designs that required turning the whole
mill when wind shifted.
The sixteenth-century Flemish mathematician Simon Stevin per­
formed what was likely the first detailed engineering analysis of any
mechanical device for these windmills. He designed mills with larger
sails, which revolved more slowly, and that used more efficient
designs of gears. He worked out the minimum wind pressure
required on a given area of each vane to lift water to the height
26 ENERGY: THE BASICS

required for dumping it. Stevin also determined how much water
would be raised with each complete revolution of the vanes.9

WIND AND WATER


Both wind and water occur readily in nature and are replenished
through natural cycles. They don’t have to be extracted from the
ground or produced in a factory. They will never run out—if
Earth’s water dries up or the wind stops blowing, humans are
doomed anyway. For these reasons, wind and water are categorized
as renewable energy resources—a concept that we will see several
more times in later chapters.
Windmills and waterwheels share the limitation that the mechan­
ical work they provide has to be used on, or near, the spot where it is
produced. It took nearly to the end of the nineteenth century to
figure out how to transmit energy over long distances, as we will see
later. As a rule, water is a better source of energy than wind. Wind­
mills can be badly affected by fluctuations in the wind. This is not
good when a steady, reliable output of mechanical work was needed
all day. Even though the steam engine (Chapter 4) eventually became
a dominant source of industrial work, improvements were still being
made to windmills and watermills well into the nineteenth century.
Windmills and watermills share another limitation. In a windmill,
the wind moving past the blades of the mill still has some kinetic
energy. If it didn’t, the air would have to be perfectly still (i.e., have
zero kinetic energy) just downwind of the mill. Since wind moving
past the mill still possesses some kinetic energy, the kinetic energy
of the wind could not have been converted completely to work. The
same argument applies to waterwheels. Water leaving the wheel still
has some kinetic energy (because it is moving). The kinetic energy
of the water has not been converted completely to work. These
observations lead to a conclusion of enormous importance, which
we will explore in more detail later: it appears that it is impossible to
convert energy completely (i.e., 100 percent) into work.

NOTES
1 Those who have never had the opportunity to watch one in person can find
vast numbers of videos on the web, by searching for a topic such as ‘water­
fall videos.’
2 Morris, Christopher (ed.). Academic Press Dictionary of Science and
Technology. Academic Press: London, p. 1289.
WATER, WIND, AND KINETIC ENERGY 27

3 Of course it has kinetic energy of some sort, because it is moving—just not


in a direction that is of any use to us.
4 Kapuściński, Ryszard. Another Day of Life. Penguin Books: London, 1987,
p. 132.
5 This book is still available in a print edition as volume 251 in the Loeb
Classical Library, now published by Harvard University Press. It can also be
downloaded from the web. Lucky is the author whose book is in print for
two millennia.
6 Wyn Jones, R. Gareth. Energy: The Great Driver. University of Wales Press:
Cardiff, 2019, p. 78.
7 However, water was likely the first energy source exploited for work in what
we would recognize as some sort of industrial setting.
8 In the siege warfare of the medieval and early modern eras, a standard tactic
was to try to starve the besieged population into submission. One way of
accomplishing this was to stop or at least divert the flow of rivers through
the besieged city, effectively shutting down any watermills that might be
used to produce flour.
9 While he was at it, Stevin also invented the ‘land yacht,’ essentially a wind-
powered automobile. In 1600 he, along with almost 30 passengers, achieved
a successful test run on a smooth, level stretch of beach sands.
4

HEAT, STEAM, AND THERMAL


EFFICIENCY

HEAT AND TEMPERATURE


A cup of hot beverage, allowed to stand for some time, becomes
noticeably cooler. It has become closer to the temperature of the
room. But, a cold beverage becomes warmer, though it too has
become closer to the temperature of the room. Such observations
are part of our normal experience. No one has ever observed a glass
of water sitting on a table spontaneously beginning to boil or
spontaneously to freeze. As with a waterfall: there is a spontaneous
flow in one direction, and a flow in the opposite direction never
happens by itself.
In these thermal systems heat flows spontaneously from higher to
lower thermal potential energy; i.e., hot to cold. We know that
something is hot or cold, and how hot or cold, by measuring its
temperature. Heat spontaneously flows from an object at high tem­
perature to one at low temperature. Heat never spontaneously flows
from low to high temperature just as water never spontaneously
flows from a low to high gravitational potential energy. A simple
energy diagram (Figure 4.1), analogous to the one for a gravita­
tional system, can help to explain the flow of heat. Temperature is a
measure of thermal potential energy.
As in a gravitational system, there ought to be a way of converting
the flow of thermal energy into useful work as well. We will see how to
do that in this chapter. First, we need to consider the nature of heat.
All forms of matter are made of atoms or molecules in constant
motion. Thus, the atoms or molecules have kinetic energy. Adding
heat to a material increases the kinetic energy of its constituent
particles. Their mass doesn’t change, so the increase in kinetic
energy means that the particles must be moving faster. ‘The hotter,
the faster; the slower, the colder.’1

DOI: 10.4324/9781032665603-4
HEAT, STEAM, AND THERMAL EFFICIENCY 29

Figure 4.1 The energy diagram for a thermal system, which is nearly identical
to Figure 3.2, except for the kind of potential energy involved. In
this system, heat flows from high to low thermal potential energy,
giving us the opportunity to capture some of the energy and convert
it into work

Temperature and heat are not the same things, though the
words are sometimes loosely used as if they were. Temperature
measures the average kinetic energy of the particles in an object.
Heat is the energy transferred between matter as a result of dif­
ferences in temperature. The ability of matter to transfer heat
energy depends on how much matter there is—its mass—and on
its temperature. Two objects, regardless of the mass of each, are
at the same temperature when the average kinetic energy of the
particles in each is the same.2 Heat moves in the direction from
the hotter to the colder object. Temperature allows us to quan­
tify the relative hotness or coldness of the objects, and is worth
measuring because it indicates whether heat will be transferred
between objects.
In the thermal system, the device that we use to extract useful
work from a spontaneous change is a heat engine. Obtaining 100
percent conversion of kinetic energy to work in a waterwheel would
require the kinetic energy of the water to be zero when it leaves the
mill. Water would have to come to a dead stop, which cannot
happen in any realistic device. A kinetic energy of zero would cor­
respond to complete conversion of kinetic energy into work. In a
thermal system, we could obtain 100 percent conversion of thermal
energy into work by reaching a point of zero thermal energy. This
‘zero point’ of thermal energy would be the lowest temperature that
could be attained, and correspond to zero kinetic energy of the
particles.
30 ENERGY: THE BASICS

WORK FROM STEAM


Several virtues of steam gave it a central role in the development of
energy technology and help maintain its importance today. It is
produced from water, readily available in most parts of the world. It
is easy to generate, needing temperatures only a bit above 100°C
and simple equipment. Steam itself is not a source of pollution
(although the methods for producing it certainly can be).
Water is a liquid at ordinary temperatures because attractive
forces between the molecules prevent most of the molecules from
escaping into the vapor phase. Increasing the kinetic energy of
water molecules gets them moving fast enough to overcome the
intermolecular attractive forces and make their getaway.3
In most steam-generating systems, the heat necessary for adding
kinetic energy to water molecules derives from burning a fuel. Thermal
energy liberated in the burning of a fuel comes from the chemical
transformation of fuel molecules, which contain a characteristic
amount of chemical energy in the bonds within their molecules. The
energy in the atom-to-atom bonds of fuel molecules is chemical
potential energy. The products of this transformation are most com­
monly carbon dioxide and water vapor. These substances are more
stable—have a lower amount of energy in their bonds—than are the
fuel molecules. The difference between the relatively high chemical
potential energy of the fuel molecules and the lower energy of the pro­
duct molecules appears as heat. We could use this heat to boil water.
In the late 1600s, the French inventor Denis Papin reasoned that
water boiling inside a cylinder would cause a piston to be pushed
upward on the column of steam. Allowing the steam to condense
would create a partial vacuum inside the cylinder. The pressure of
the atmosphere would push the piston down. If the piston were
connected to some device, then useful work could be done, e.g.,
raising a heavy load. The device is in essence a machine, generally
defined as any device that helps in the doing of work.4
An English hardware dealer, Thomas Newcomen, developed an
engine around 1712 that used a heavy counterweight to pull the
piston upward, and condensed steam by spraying cold water into
the cylinder. It’s called an engine rather than a machine because an
engine operates in a cyclic manner, returning to its original state at
the end of each cycle. Condensing only part of the steam keeps the
cylinder warm, and lets the engine work more quickly. Since power
is work done per unit time, decreasing the amount of time to do a
given amount of work makes the engine more powerful.
HEAT, STEAM, AND THERMAL EFFICIENCY 31

Newcomen’s engine was the first new, widely-accepted source of


work in almost a millennium—since the extensive adoption of water­
wheels. It was used throughout Europe, despite numerous flaws. Still,
40 years after its introduction more than a hundred were in use, mostly
pumping water from mines. A Newcomen engine used at a mine in
Derbyshire (England) was in operation from 1791 to 1918. The last
survivor (in Yorkshire) was not scrapped until 1934, after operating
for more than a century without needing extensive maintenance.
While repairing a model Newcomen engine at the University of
Glasgow, the university’s instrument maker, James Watt, recognized
its major problem: a single cylinder had to alternate between being
hot and being cold, causing inefficient operation that wasted con­
siderable fuel. Watt saw the benefit of having two chambers—one,
the cylinder with piston, that is always hot, and one that is always
cold, to condense the steam.
Watt’s device used the expansion of the steam to do the work and
so, strictly speaking, represents the first true steam engine, an appara­
tus or device for converting heat (in steam) into kinetic energy. He
developed a system of gears that converted the reciprocating motion
of the piston into the rotary motion of a shaft. Watt’s engine included
a governor that automatically adjusted the amount of steam passing
into the cylinders according to how rapidly the machinery was going.
This is the first example of feedback control of a machine. The Greek
word for ‘governor’ has come into English as ‘cybernetics.’5
The steam engine freed humans from relying on energy sources
supplied by nature (wind, water, or animal muscles). It was the first
device that provided useful work from an energy source that could
be created on demand by humans. It also was the first step in
removing the constraints of geography. It was no longer necessary
to build factories by water that had high kinetic energy, or locations
offering reasonably steady wind. A steam engine could be erected
any place that work was needed. However, the engine itself needed
to be close to the machine that it was intended to operate. Every
factory had to have its own engine close by, and its own supply of
fuel. About a century after Watt’s great invention, the second lib­
erating step began to emerge from the study of electricity. This we
will explore in future chapters.

THE NOTION OF EFFICIENCY


After the initial investment in a windmill or waterwheel, there was
no cost for the water or wind itself. Not so with the steam engine!
32 ENERGY: THE BASICS

Generating steam required burning a fuel, usually coal, which had


to be purchased. This represented a continuous expense for the
owner of a steam engine. An engine that required less coal to
accomplish a set amount of work (pumping a certain quantity of
water, for example) had a lower cost to operate. In modern termi­
nology we refer to the amount of work done per amount of fuel
consumed as the efficiency of the engine.
Sadi Carnot, a captain in the French army corps of engineers,
recognized that an engine functions because heat flows from a
hot body to a colder one. During the spontaneous, one-way-only
flow of heat from a hotter to a colder body, heat can be made to
do work. A temperature difference in any system, with a flow of
heat from higher to lower thermal energy, provides the possibility
for generating work. The greater the temperature difference, the
greater the work for the same quantity of heat, just as water
falling 200 m will do twice the work of the same amount of
water falling 100 m.
Heat will not create work if there’s no temperature difference to
establish a flow. As Carnot said, ‘The production of heat alone is
not sufficient to give birth to the impelling power: it is necessary
that there should also be cold; without it, the heat would be use­
less.’6 It’s not the value of the high or low temperature that dictates
the amount of energy converted to work, but rather the temperature
difference. The available energy (not the total energy) can be
expressed in terms of the temperature difference within the heat
engine, as Th – Tl, where h and l stand for high and low.
Water at the bottom of a waterfall still has kinetic energy. Con­
ceptually, more work could be extracted from the water at the base
of a waterfall until all the available gravitational potential energy
was used up—at the center of the Earth, the zero point of gravita­
tional energy. In a thermal system the point at which no more work
could be extracted would be the zero point of thermal potential
energy below which it is impossible to go—absolute zero. The total
energy of the thermal system would be the difference between the
temperature of the hot side and absolute zero: Th – 0, or, simply Th.

THE QUEST FOR ABSOLUTE ZERO


Between the seventeenth and early-nineteenth centuries, relationships
were developed among the temperature, volume, and pressure of
gases. (The relevant equations are given in the Appendix.) A fixed
quantity of gas, held at a fixed pressure, changes its volume as the
HEAT, STEAM, AND THERMAL EFFICIENCY 33

temperature changes. When a constant amount of gas is kept at con­


stant pressure, its volume is directly proportional to its temperature.
For every 1°C drop in temperature, a gas contracts by 1/273 of its
volume. Reducing its temperature to –273°C would cause the gas to
contract by the fraction 273/273. The gas would lose all its volume
and vanish. The matter of which the gas is made would have been
destroyed—a physical impossibility confirmed by innumerable
experiments the world over. Therefore, –273°C represents the lowest
possible temperature.7
Quantitatively comparing the amounts of thermal energy in a
system requires a definite, immutable, fixed point on which a tem­
perature scale can be based. Since –273°C is the lowest possible
temperature, it represents the point of zero thermal energy, absolute
zero. This allows establishing a temperature scale with its zero point
at absolute zero. This temperature scale is the Kelvin scale. Its units
are kelvins and have the symbol K. Absolute zero is the lowest pos­
sible temperature (the point of zero thermal energy). At absolute
zero, the component molecules of a substance also have zero kinetic
energy. All molecular motion has stopped.
Heat engine efficiency is a measure of the fraction of total energy
that can be converted to work. (Equations for calculating efficiency
are also given in the Appendix.) The efficiency of any device will be
a fraction having a value between 0 and 1, or expressed as percen­
tages by multiplying the fraction by 100. The ratio of available
energy to total energy represents the maximum efficiency of a heat
engine. This efficiency, sometimes called the Carnot efficiency, is
reached only if the engine is mechanically perfect; e.g., if there are
no losses of energy through friction, or losses of heat to the outside
world. We can never convert all of the thermal energy of any fuel
completely into work in any engine.
The arithmetic of calculating efficiency doesn’t say anything
about what substance is passing through the engine, i.e., the work­
ing fluid. Though here we focus on steam engines, when considering
efficiency, the steam is irrelevant. It’s the temperature drop that’s
important. The work of the engine is due to the heat flow.
Complete conversion represents an efficiency of 1.000 (or 100
percent), which requires that the low temperature be absolute zero
(0 K). But, it is impossible to attain a temperature of absolute zero
(0 K). This brings us to the unfortunate conclusion that, it is
impossible to convert thermal energy to work completely (i.e. with
100 percent efficiency). Evidence suggesting this conclusion has
already surfaced in other systems. We get hot when we exercise or
34 ENERGY: THE BASICS

work vigorously; wind or water leaving a mill still has kinetic


energy left. Now we have a thermal system where we can show that
the complete conversion of thermal energy to work is impossible.
Extending the ‘unfortunate conclusion’ for a thermal system to other
systems requires a connection between heat and other kinds of energy.
The inventor and scientist Benjamin Thompson, also known as Count
Rumford, observed that drilling the ‘bore’ into cannon barrels made
them very hot. If the barrel and the drill were covered with water, the
water would get hot enough to boil. Even if the water boiled away
completely, a new batch of water would soon boil, and if that boiled
away too, a third batch of water would boil. Water would boil for as
long as the drill could be turned. Where was the heat coming from?8
The only thing going on in Rumford’s system (cannon barrel, drill bit,
and water) was the turning motion of the drill bit inside the barrel. As a
result, this led to the hypothesis that heat is a form of motion.
Heat is a special kind of kinetic energy—the kinetic energy of
atoms or molecules. Temperature scales measure the average ther­
mal kinetic energy per atom in an object. The more kinetic energy
each atom has, the more vigorous its thermal motion. The atomic-
scale work that one vibrating atom does on another is what passes
heat through an object, or from one object to another. The object
with more average thermal kinetic energy per atom will pass heat to
an object with less.
James Prescott Joule, the English son of a brewer,9 conducted the
key experiment relating heat to work. A paddle immersed in water
was turned by the action of a falling weight, i.e., a spontaneous
change in a gravitational system. During its fall, the weight does
work by causing a vigorous stirring of the water. As this happens, the
temperature of the water rises. This experiment established the
mechanical equivalent of heat, a fixed mathematical relationship
between a quantity of mechanical work done and a quantity of heat.
Since energy from a mechanical system is equivalent to thermal
energy (as Joule proved), the conclusion reached about the effi­
ciency of a thermal system must extend to mechanical systems as
well. Later workers showed that all forms of energy can be inter-
converted, not just heat and mechanical energy. The conclusion
regarding efficiencies and the production of work in heat engines is
even worse than it seemed at first. It is impossible to convert any
kind of energy to work completely (with 100 percent efficiency). No
matter what we do to convert energy into work in any system, some
of the energy is going to be wasted! Very often (but not always) the
wasted energy is lost as heat.
HEAT, STEAM, AND THERMAL EFFICIENCY 35

Every investigator has found that a fixed amount of one kind of


energy can be converted into a fixed amount of another kind of
energy. Regardless of the kinds of energy involved, no energy is lost
or created when these transformations take place. Very careful
measurements showed no ‘extra’ energy appeared nor disappeared
in every case that has ever been studied. The growing body of data
led to an important statement summarizing the results. The total
amount of energy in the universe is constant. It can be converted
from one form to another, but is neither created nor destroyed. This
statement is the basis of the law of conservation of energy, or the
First Law of Thermodynamics.
Heat never spontaneously flows from a cold object to a hot object,
just as water never spontaneously flows uphill. To move water uphill,
we have to supply the work needed to operate a pump. In the same
way, to get heat to flow ‘uphill’ we have to supply work. To reduce
the temperature of food, the work of pumping heat uphill is done by
a refrigerator. (And the surrounding room gets warm; just feel behind
your refrigerator.) These observations are consequences of the Second
Law of Thermodynamics. This can be stated as, ‘heat cannot flow
from a low temperature to a higher temperature unless aided by
work.’10 Another consequence is that practical machines operating
with 100 percent efficiency are impossible.

HOW HEAT FLOWS


Heat can move by three mechanisms: conduction, convection, and
radiation. Conduction transfers heat directly through the body of
an object. As one end of an object gets hot, atoms in that portion
of the object increase in kinetic energy. As the atoms gain kinetic
energy by being heated, the atomic vibrations become more rapid,
and the atomic movements extend further from their fixed position.
The atoms at the hot end will vibrate the most because they have
the most kinetic energy. They will jostle neighboring atoms in the
slightly cooler portion. Those atoms in turn will acquire more
kinetic energy from these impacts. They vibrate more energetically
themselves, and jostle atoms in an adjoining cooler portion of the
object. This goes on and on until the kinetic energy bumps its way
through the whole length of the object.
Convection heat transfer can occur in gases or liquids. We can see
convection, for example, when we’re cooking and watch a pot of
water come to a ‘rolling boil.’ We can feel convection when we’re
warmed by a hot summer breeze or cooled by a cold draft.
36 ENERGY: THE BASICS

An example of radiation heat transfer is sensing the heat from an


object that is red hot, even though we are not in direct contact with
it for conduction, nor feeling currents of heated air from convec­
tion. The most vital example of heat transfer by radiation is the
process that brings heat from the sun to the Earth. While both
conduction and convection involve the transfer of heat between
particles of matter, radiation does not. The energy is transferred by
electromagnetic wave motion.
Regardless of how heat is transferred, it will flow from a hot
object to a cool object until the temperatures are equal. The tem­
perature scale classifies or ‘orders’ objects according to the direction
in which thermal energy will flow between them. This lets us create
a straightforward definition of the concept of heat: the way in
which thermal energy flows from one object to another because of a
difference in their temperatures.

THE ‘LOST’ ENERGY


The two kinds of changes—work to heat and heat to work—are not
completely symmetric. Carnot showed us that heat can’t be con­
verted completely to work. In a heat engine operating below 100
percent efficiency—in other words, all of them—it seems as if we
might be dealing with, in a sense, two kinds of heat. Some heat is
useful, because we can convert it into work for our desired applica­
tion. The rest is apparently wasted. Then what happens to that
fraction of heat that isn’t converted? Where does it go? Is this a
violation of the First Law of Thermodynamics? (No.)
In any gas, steam for example, the molecules are in constant, rapid
motion in random directions. In a cylinder full of steam, with the
steam held in place by a piston, at any moment some of the molecules
will strike the piston. These molecules will transfer a portion of their
kinetic energy to the piston, moving it and thus operating the engine.
These molecules have done useful work, but they represent only a
portion of the total number of molecules in the cylinder. While doing
work, these molecules lost a portion of their kinetic energy—heat was
converted into work. The remainder of the molecules are also in the
same rapid, apparently chaotic motion, but do not do work, because
they do not help move the piston. The chaotic motion of molecules,
which is not useful energy, is measured by entropy.
Increasing entropy can be considered to be a measure of the
decreasing usefulness of heat. Spontaneous processes or reactions
proceed with an increase in entropy. Provided that they receive any
HEAT, STEAM, AND THERMAL EFFICIENCY 37

necessary energy to start the process, they will take place. Any form
of energy (e.g., gravitational, thermal, electrical) that we can make
use of to do work will dissipate some of the energy as an increase in
entropy. Later in the book we will see cases in which we string
together several energy conversions to get from an initially available
energy source to obtaining work in a form that is useful to us. Every
step along the way proceeds with an increase of entropy, which
means that some useful energy is lost, or, from the other perspective,
the possibility of doing useful work decreases. A practical implication
is that, ideally, we should try to use energy conversions having the
fewest number of steps—preferably one.
A heat engine does work because of the spontaneous flow of
thermal energy from high potential (i.e., high temperature) to low
potential. Operation of a heat engine proceeds with an increase in
entropy. Generalizing, spontaneous processes are those that proceed
with increasing entropy. But spontaneous processes are happening
around us all the time. Also, we want to take advantage of sponta­
neous processes to convert some of the spontaneous flow of energy
into useful work. Does that mean that the entropy of the world—
indeed, of the universe—is steadily increasing? Yes. The German
physicist Rudolf Clausius summed things up as: The energy of the
universe is constant. The entropy of the universe tends to a maximum.

NOTES
1 Nguyen-Kim, Mai Thi. Chemistry for Breakfast. Greystone Books: Van­
couver, 2022, p. 14.
2 Closely related to this is the Zeroth Law of Thermodynamics, which says
that if a system A is in thermal equilibrium (at the same temperature) as
system B, and if system B is in thermal equilibrium with system C, then
systems A and C must also be in thermal equilibrium.
3 Recalling that ‘the hotter, the faster; the slower, the colder,’ if we remove heat
from water, its molecules have less and less kinetic energy until they are very
limited in their movement. Macroscopically, the water freezes to solid ice.
4 Prater, Edward L. Basic Machines and How They Work. Dover Publications:
Garden City, NY, 1997, p. 1–1.
5 Nowadays we take the term to refer to the science and engineering of
automatic control and communication systems in machines and organisms.
6 Quoted in Sen, Paul. Einstein’s Fridge. Scribner: New York, 2021, p. 14; see
also Coopersmith, Jennifer. Energy: The Subtle Concept. Oxford University
Press: Oxford, 2010, p. 219.
7 This argument ignores the fact that all real gases would turn to liquids at
some temperature above –273°C.
8 In Rumford’s era the prevailing concept was that heat was an invisible,
weightless fluid, called caloric. When an object cooled, it was thought to
38 ENERGY: THE BASICS

have lost some of its caloric; if something increased in temperature, it


gained caloric. Rumford’s work helped demolish the caloric theory.
9 Though (or because?) his early scientific work included giving electric
shocks to the family servants and to his brother, Joule became recognized
as one of the great scientists of the nineteenth century. The unit of energy,
the joule, is named in his honor.
10 Rudolf Clausius, quoted in Coopersmith, Jennifer. Energy: The Subtle
Concept. Oxford University Press: Oxford, 2010, p. 292.
5

PETROLEUM FUELS AND THEIR


ENGINES

INTERNAL COMBUSTION
Watt’s engine relied on steam as the working fluid. Expansion of
hot steam causes the piston to move. Movement of the piston
increases the available volume in the cylinder. As volume increases,
temperature drops, giving the spontaneous high-to-low heat flow.
Steam is generated in a separate boiler, heated by the combustion of
a fuel (often coal). Since the fuel is burned outside the engine, it
belongs to the class of devices called external-combustion engines.
In internal-combustion (IC) engines fuel is burned inside the
cylinder. The working fluid is the hot, gaseous products of com­
bustion. Their expansion and cooling, as the piston moves, provides
the spontaneous flow of heat, and the piston movement provides the
work. Piston-in-cylinder IC engines have enormous importance.
They supply the work for the vast majority of road vehicles, many
railway locomotives, ships, mining and construction vehicles, and
small-scale electricity generators. Later we will meet another kind of
IC engine that dominates aviation. Despite different designs and
applications, most IC engines have a common feature: the fuels they
use are derived from petroleum.

PETROLEUM AND ITS DERIVATIVES


Petroleum is one of three major energy resources that collectively
comprise the fossil fuels. The others are coal and natural gas. A
fossil is a remnant of formerly living organisms preserved in the
Earth’s crust; a fuel is a substance burned as a source of energy.
Collectively, petroleum, coal, and natural gas currently represent
the most important fuels that we have for electric power production,
transportation, manufacturing, and use in the home.

DOI: 10.4324/9781032665603-5
40 ENERGY: THE BASICS

Though apparently very diverse, fossil fuels have two things in


common. First, they are the remains of plants and animals. Second,
their major chemical component is carbon. So, their origin can be
understood by considering how carbon is dispersed throughout
nature, represented by the global carbon cycle (Figure 5.1). The key
step, photosynthesis, converts water and carbon dioxide from the
atmosphere into sugars, which plants use to produce the many other
chemicals needed for their life processes. Eventually, these chemicals
will transform into fossil fuels.
The energy source that drives the global carbon cycle, by providing
the energy to drive the reactions of photosynthesis, is sunlight.
Essentially, fossil fuels represent stored solar energy.
Petroleum and natural gas form from the slow chemical transfor­
mation of the accumulated remains of organisms. Petroleum and gas
will migrate from where they originated (called source rocks) through
porous rocks inside the Earth. They migrate until they encounter an
impervious nonporous rock which stops further moment and allows
the oil or gas to collect. This trapped deposit of oil and/or gas is
called a reservoir. If the accumulation is large enough—and if we can
find it—it is worth exploiting to extract these fuels.
Oils range from lightly-colored free-flowing liquids with a mild
inoffensive odor to thick, dark liquids having the consistency of
road tar and an unpleasant odor. As users of petroleum, we expect
products of consistent quality, because all of the items that we have
built to use petroleum products—automobile engines, for exam-
ple—are designed for fuel of consistent properties and behavior.
Otherwise, how could we expect that gasoline would provide the
same engine performance from one tankful to the next? How could

Figure 5.1 A sketch of the global carbon cycle. Fossil fuels arise from an
interruption of the decay process that organisms would normally
experience in the environment
PETROLEUM FUELS AND THE ENGINES 41

we expect a particular grade of gasoline to give the same performance


when purchasing different brands?
The challenge faced by the industry is how to take the highly vari­
able mixture of molecules in the many different kinds of petroleum
and convert it into products that are reliably consistent. To meet this
challenge, petroleum as it comes from the ground, often called crude
oil, is refined into groups of products, each of which has a relatively
narrow range of physical properties and chemical composition.
This separation is done by distillation, which relies on differences
in boiling behavior of the components of a mixture. Distillation is
the most important step in petroleum refining. Gasoline has the
lowest boiling range1 of the liquid products.

GASOLINE
Nearly all IC automobile engines work on the four-stroke principle
invented by the German engineer Nikolaus Otto2 about 150 years
ago. The operation of a simple four-cylinder engine is shown in
Figure 5.2. In the intake stroke, the piston moves downward and a
gasoline–air mixture enters the cylinder. This mixture is compressed
by the upward movement of the piston in the compression stroke.
Then, a spark plug ignites the gasoline–air mixture during the
ignition, or power, stroke. During combustion, the hydrocarbon
molecules in gasoline react with oxygen in the air to form carbon
dioxide and water vapor. More chemical energy is stored in the
carbon–hydrogen bonds in gasoline than is in the carbon–oxygen
and hydrogen–oxygen bonds of the products. As chemical energy
flows ‘downhill,’ the energy released raises the temperature and
pressure in the cylinder. We take advantage of this to do work.
If a gas is confined, as in the space between the cylinder head and
the top of the piston, an increase in temperature will increase its
pressure. This increased pressure pushes against the piston, causing
it to move downward. This allows for the conversion of thermal
energy into mechanical work in the engine.
As the gases expand, they experience a drop in temperature,
allowing for the spontaneous high-to-low temperature flow of heat,
from which we extract work. The chemical potential energy in the
molecular components of the gasoline molecules is converted into
the mechanical work of propelling the car. Finally, in the exhaust
stroke, upward movement of the piston pushes the products of
combustion out of the cylinder, making it ready for the next intake
stroke. Having gone through a complete cycle, we see the feature
42 ENERGY: THE BASICS

that distinguishes an engine from a machine—at the end of per­


forming work, the configuration of the engine is right back where it
started. The engine produces work only in the ignition stroke; in the
intake, compression, and exhaust strokes the piston is being pushed
up or pulled down by the operation of the engine, and not by
energy released from gasoline. Gasoline (also known as petrol) is
the dominant fuel for automobile and light truck engines in many
countries, including the United States.
During the power stroke, it is possible for the unburned portion
of the gasoline–air mixture to be compressed to a point at which it
explodes, rather than burning smoothly. The explosion can be heard
inside the vehicle, and is called engine knock. Knocking is wasteful
of fuel, and, if it persists, can lead to major maintenance issues on
the engine. The knocking tendency of a particular gasoline is
expressed by its octane number. Octane ratings are indicated on the
dispensing pumps at service stations. The higher the octane number,
the less the tendency to knock. Most vehicles run quite well on a
standard grade of gasoline, 87 octane. High-compression engines
might require a premium grade, of about 93 octane.

DIESEL ENGINES AND DIESEL FUEL


In 1804, the French physicist Joseph Mollet produced a device—a fire
piston—that was the forerunner of matches. A sharp blow of the hand
on the end of the piston would compress air in the cylinder, heating it

Figure 5.2 A diagram of the four-stroke Otto cycle for spark-ignition engines
PETROLEUM FUELS AND THE ENGINES 43

enough to ignite a small amount of tinder attached to the piston.


Fire pistons were made obsolete by the development of the match in
the mid-nineteenth century; they became museum pieces. In
Munich, about 1888, a lecturer on heat transfer demonstrated how
to light a cigar by this method; in the audience was a young man
named Rudolf Diesel.
The professor pointed out that steam engines transformed only
6–10 percent of the chemical potential energy of their fuel into
useful work. Diesel wanted to develop an engine that would do
better. By 1893 he had come up with a scheme for an engine that
did indeed operate much more closely to the ideal Carnot efficiency
for a heat engine. In Diesel’s design, ignition of the fuel was
achieved by compression alone, just as in a fire piston.
Diesel developed an engine that not only would be more efficient
than existing steam or gasoline engines, but also could, in principle,
be operated on a variety of fuels: gases, powered coal, coal tar,3 or
higher-boiling petroleum products. All these fuels cannot be used in
the same engine; rather, variants of Diesel’s design can be adapted
to different fuels. The vast majority of diesel engines nowadays
operate on the petroleum product commonly called diesel fuel. Its
name is a testament to the engineer and his creation. No other fuel
is so linked with the inventor and the device intended to use it. We
don’t call gasoline ‘Otto fuel.’
The 1897 version of his engine had a mechanical efficiency of 34
percent, a great advance over all other engines. Originally, the
diesel engine was designed to operate at low speeds, to serve as an
alternative to reciprocating steam engines in ships and in stationary
installations. To this day they do an excellent job in such applications.
By 1909, Diesel had developed a small four-cylinder, high-speed
engine intended as a heavy-duty automobile engine.
Diesel’s engine is, in some ways, a variant of Otto’s four-stroke
cycle. In the diesel engine, the intake and compression strokes
involve only air. The compression ratio is appreciably higher than in
conventional spark-ignition gasoline engines. The rapidly increasing
pressure in the compression stroke means that the temperature also
increases, a consequence of the behavior of all gases (the equations
are in the Appendix). In the diesel engine the goal is to get the air
hot enough to ignite the fuel. Indeed, its temperature rises far above
the ignition point of the fuel. An increase in temperature is a non-
spontaneous change which can occur only if work is put into the
system. In the present context, this means the compression work
done by the engine as it operates.
44 ENERGY: THE BASICS

In the third stroke (the power or combustion stroke) the fuel


injector squirts a fine spray of fuel into the hot compressed air in
the cylinder. Because of the high air temperature, the fuel immedi­
ately begins to burn. The piston begins moving downward. Injec­
tion continues for a substantial part of the power stroke. The
consequent expansion of the gas in the cylinder is not immediately
accompanied by a temperature drop, because heat is still being
released from the burning fuel. Heat from the burning fuel is con­
verted directly to useful work pushing down the piston in the
cylinder. The key difference between Diesel’s engine and all other
internal combustion engines is that heat from the burning fuel is not
‘wasted’ to heat up the air in the cylinder.
A diesel engine is a compression-ignition engine, because the
compression work done by the engine is sufficient to cause the fuel
to ignite. No spark plugs are needed. When the fuel injector shuts
off, the gases in the cylinder can expand further with an accom­
panying drop in temperature and pressure. Here the spontaneous
flow of heat from high to low temperature lets us obtain work. On
the final stroke, an exhaust valve opens and the piston expels the
air and burned fuel. The piston is then back in position for the
intake stroke.
The best diesel engines have efficiencies of up to 42 percent; per­
formances above 30 percent are routinely achievable with any rea­
sonably well-maintained engine. They are some 15–20 percent more
efficient than a comparable gasoline engine.4 Although the ideal
Carnot efficiency of the Otto cycle is 60 percent, the best practical
performance of gasoline engines is only around 32 percent.
When equal masses of gasoline and diesel fuel are burned, the
energy liberated from each is fairly similar. Diesel fuel has a
higher density than gasoline. The ability of a vehicle to carry fuel
is limited by the volume of its fuel tank, not by the weight of the
fuel. Because of its higher density, diesel fuel supplies more energy
than gasoline per unit volume, that is, it has a higher volumetric
energy density. Given two vehicles having the same fuel tank
volume, the diesel-fueled vehicle will be able to travel further than
one using gasoline.
Many diesel engines will run on less-refined fuel than is needed
for spark ignition. A less-refined fuel, though of lower quality, will
also be cheaper. Where petroleum products are not readily avail­
able, diesel engines can be run using such fuels as the oils from
some plants. One of Diesel’s first major demonstrations of his
engine used peanut oil as the fuel. Today there is still interest in
PETROLEUM FUELS AND THE ENGINES 45

using fuels derived from plant oils (usually not the oils themselves)
as ‘biodiesel.’ This will be discussed further in Chapter 10.

ACTIVATION ENERGY
Before introducing a new kind of IC-engine, we digress to consider
a question about Otto- and diesel-cycle engines: why do they need
spark plugs or compression to work in the first place? The reaction
of gasoline or diesel fuel with oxygen (from the air) is spontaneous.
Shouldn’t they catch on fire immediately when mixed with air?
A toy block of rectangular shape is most stable when lying on a
table with its longer side flat on the table. It will stand up on its shorter
side, and will stand like that forever until we give it a bit of a nudge.
Given a nudge, the block will promptly topple over so as to be lying on
its longer side. The block is always stable so long as its center of mass
is over its base.5 The block standing on its shorter side has its center of
mass at slightly higher gravitational potential energy than when lying
on its longer side. The nudge is just enough to displace the center of
mass away from being over its base. Then, gravity takes over, and a
spontaneous change from high to low potential energy occurs. The
nudge that we have supplied was just enough to get the spontaneous
process started. We refer to it as the activation energy.
Similarly, we have to ‘nudge’ molecules in gasoline or diesel fuel
just enough to start their spontaneous reaction with oxygen. That
is, we have to supply activation energy. In this case, activation
energy breaks some chemical bonds in molecules of fuel, forming
highly reactive species. They initiate further reactions that comprise
the spontaneous change from fuel to carbon dioxide plus water
vapor. That activation energy is supplied by the spark plug in a
gasoline engine or by the high temperatures in the diesel engine.

JET ENGINES AND JET FUEL


Let’s now consider an alternative to the piston-in-cylinder concept,
a totally different kind of prime mover. Recall the waterwheel.
Water for turning the wheel comes from outside the wheel, flowing
against the blades. However, suppose that the water was to be
introduced inside the wheel. Such designs, developed in France in
the early nineteenth century, were named turbines. We will see in
Chapter 11 that turbines using water as the working fluid are very
important in electricity generation. Here, we consider turbines using
combustion gases as the working fluid.
46 ENERGY: THE BASICS

A gas turbine engine is an internal-combustion engine in which


the high-temperature, high-pressure combustion products are direc­
ted against the blades of a turbine. The turbine converts the kinetic
energy of a moving fluid—in this case, the products of combus­
tion—into rotary motion. No pistons and no cylinders. We use
rotating motion to run an air compressor, needed to supply air to
the engine. The turbine is directly coupled to the compressor by a
shaft. After leaving the turbine, hot combustion products enter the
tail pipe, which ends in a nozzle. This nozzle converts the energy in
the combustion products into energy of motion. A major applica­
tion of these engines is in aircraft, where they are best known as jet
engines, more formally as aviation gas turbine engines.
Most jet engines nowadays are turbojets. A turbojet engine con­
sists of a metal tube, open at both ends. At the front end is a rotary
air compressor that draws in and compresses air. Its importance is
that fuel needs air to burn, but at high altitudes the atmosphere is
‘thin,’ i.e., air pressure is low. (High-speed long-distance flights have
to be at high altitudes, where air resistance is much less.) Com­
pressed air is passed to combustion chambers, sometimes called
burner cans. Fuel is injected into the stream of compressed air,
through nozzles arrayed around the burner can. The burning fuel
creates very high temperature and pressure.
The great advantage of the jet engine is its high ratio of power to
size, which more than compensates for its noise and high fuel con­
sumption. All large passenger planes now use jet engines. Jet
engines were also adopted for most military uses, because of the
greatly improved performance that they gave in speed and power
relative to piston-engine aircraft.
Consider what happens when you blow up a balloon and let go
of it. The balloon is moved by the pressure of the air inside, not by
the air escaping. Air escaping from the balloon reduces the inter­
ior pressure on the ‘nozzle end’ of the balloon. The high pressure
continues to push out on the other end of the balloon. This
unequal pressure propels the balloon in the direction of the greater
pressure. In the same way, the jet engine moves in the direction of
the greater pressure inside of it. The propulsion is possible because
‘for every action there is an equal and opposite reaction,’ more
formally known as Newton’s third law of motion. In the balloon,
the action is the air rushing out of the neck. In the jet engine, the
action is the hot gases rushing out the rear of the engine at a high
pressure. The reaction in both cases is the unequal pressure inside
pushing forward.
PETROLEUM FUELS AND THE ENGINES 47

Air comes into the engine’s inlet at the speed of the airplane itself,
which could be in excess of 800 km per hour. In the inlet, air slows
down and its pressure increases, but its total energy is unchanged.
Then the compressor blades push the air forward, doing work on it
and increasing its pressure and its total energy. Air arriving at the
combustion chamber has a pressure far above that of the atmo­
sphere. This air will be heated by the release of chemical potential
energy from the burning fuel. Since a given mass of hot air occupies
a greater volume than cold air, the hot exhaust gas takes up more
volume than it would have done before combustion occurred. The
hot exhaust gas exits the combustion chamber traveling faster than
when it entered. Its pressure is still very high. It accelerates out of
the back of the engine, losing its pressure while increasing speed.
On its route out of the engine, the exhaust gas does work on the
turbine and gives up a little of its energy in the process. The turbine
drives the compressor for the incoming air. Overall, the engine has
slowed the air down, then added energy to it, and finally accelerated
it back to high speed. Because the engine has added energy to the
air, air leaves the engine traveling faster than when it arrived. The
jet engine has exerted a force on the air to accelerate it rearward;
the air has pushed back and produced a thrust that propels the
airplane forward.
In the turbojet, the hot gas mixture has two roles. It spins the
turbine, and provides some jet thrust. The thrust from the back of
the engine can be achieved only if the speed of the gases exiting the
back of the engine is greater than the speed of the airplane itself. To
ensure that this will happen, the energy supplied to the compressor
by the turbine is used to raise the pressure, and combustion of the
fuel raises the temperature. The nozzle at the back end of the engine
helps to accelerate the gases as they leave the engine.
The thrust of an engine depends on the mass of air expelled per
second and on the increase in the velocity of the air as it passes
through the engine. If the total mass of air expelled per second can
be increased at the same velocity, greater thrust can be obtained for
a given fuel consumption. In a fan jet engine, common on large
commercial planes flying below the speed of sound, only a fraction
of the air pulled into the engine is sent to the burner cans. The rest
is blown out immediately, adding to the jet of air expelled from the
engine. As in the turbojet, the fan jet engine has slowed the air
down, added energy to it, and returned it to high speed. Air leaves
the engine traveling faster than when it arrived, but the fan jet
engine moves more air than a simple turbojet engine does, giving
48 ENERGY: THE BASICS

that air less energy and, therefore, consuming less fuel in adding
energy to the air. Thus, the fan jet is more economical than a tur­
bojet, but the turbojet gives higher speed.
Turbine engines can be designed and built to operate on a wide
range of fuels, including natural gas, liquid products from petroleum,
synthetic liquid fuels from coal, biofuels, and even specially treated
solid coal. Like the diesel, this does not imply that a given engine will
run on any of these fuels interchangeably; rather, given the fuel of
choice, it is possible to design a turbine engine that will use it. Today
virtually all aviation gas turbine engines are designed to run on a
refined grade of aviation kerosene, more commonly called jet fuel.

WHY PETROLEUM PRODUCTS?


But why petroleum products for all these engines? How did they get
to be the fuels of choice for transportation? There are several reasons.
Given the choices—wood, coal, petroleum liquids, natural gas—the
petroleum products have the highest energy densities. Any transpor­
tation device is going to be limited by the volume of fuel it can
carry.6 Liquids can easily be pumped and handled; somebody’s got to
shovel coal. Gases are easy to regulate and flow, but it is very difficult
to store an adequate amount on a vehicle. Compressed gases come in
heavy metal tanks, creating a substantial weight penalty.
In Chapter 6 we will see that the essence of electricity generation
is to find a way to provide work to turn a generator as cheaply and
reliably as possible. A gas turbine can be used in this application.
Such turbines are similar to those used as aircraft engines (and are
sometimes called ‘aero-derivative’ turbines) but, in electricity gen­
eration, are often fired with natural gas.

WILL WE USE UP ALL THE OIL?


From time to time there are apparent shortages of petroleum, and
attendant price hikes. They may have various causes: the two major
historical events were the 1973 oil embargo by the Organization of
Petroleum Exporting Countries (OPEC), and then the 1979 oil price
shock on the heels of the Iranian revolution. Skyrocketing prices of
petroleum products or gasoline shortages always seem to come
along with the concern that we are running out of oil. Will we ever
use up all the oil in the Earth? No.
It’s sometimes argued that we have yet to explore all the possible
locations of major oil fields, so there’s sure to be more to be found,
PETROLEUM FUELS AND THE ENGINES 49

or that another technological breakthrough like hydraulic fracturing


(fracking) is bound to come along to help us extract more petroleum.
The real reason is different, and depends on the concept of energy
return on energy invested (EROEI). This ratio measures the amount
of any kind of energy that we wish to obtain for the amount of
energy that we have to expend. For example, if there were a deposit
of some fuel that would yield 500 gigajoules when we extracted and
used it, and if it would take us 5 gigajoules of energy to do the
extracting, the EROEI would be 500 ÷ 5, or 100. (That’s a pretty
good EROEI.) As easy-to-obtain energy resources become increas­
ingly depleted, we need to drill deeper, go further offshore, and gen­
erally invest more energy to access that resource. So, EROEI
decreases. When a time comes that EROEI is numerically equal to 1,
we have to use as much energy to obtain a supply of the resource as
we would get back when we used it. That’s a rather ‘iffy’ proposition.
And when EROEI becomes less than 1, we need to expend more
energy than we get back. There is no rational reason for doing that;
we would move on to some different energy source.7 EROEI of pet­
roleum liquids is said to be ‘plummeting,’8 expected to reach about 7.

NOTES
1 A pure chemical compound boils at a specific temperature. It would be too
much to expect that every compound in a mixture would boil at the exact
same temperature, but during distillation we can collect mixtures of which
all of the components have boiling temperatures in a narrow range. We
would call this the boiling range.
2 Who started life as a grocery salesman.
3 A byproduct of converting coal into coke used in metallurgical furnaces,
particularly for the smelting of iron ore.
4 Smil, Vaclav. Numbers Don’t Lie. Penguin Books: New York, 2021, pp. 109ff.
5 The center of mass of any object is a point that moves as if the entire mass
were concentrated at that one point. It also behaves as if all the forces on
the object were concentrated there.
6 There is an extremely important exception to this generalization: trains
directly powered by electricity from overhead wires or an energized rail.
7 An interesting analogy from jurisprudence is the probate case of Jarndyce v.
Jarndyce, which finally ended on the day on which the total lawyers’ fees
and court costs became numerically equal to the value of the estate being
contested. There was no reason to go further. Dickens, Charles. Bleak
House. (Available in many editions.)
8 Misra, Siddharth. ‘Plummeting “energy return on investment of oil” and the
impact on global energy landscape.’ Journal of Petroleum Technology,
March 2023. https://jpt.spe.org/plummeting-energy-return-on-investment-o
f-oil-and-the-impact-on-global-energy-landscape.
6

ELECTRICITY

THE EARLY ELECTRICIANS


Some 2,500 years ago, the Greek philosopher Thales of Miletus
observed that rubbing a piece of amber would cause it to attract
light objects, due to a phenomenon we now call static electricity.
Nothing was done to develop the practical consequences of this
observation until the seventeenth century. Then, substances that
would generate attractive forces when rubbed were said to be ‘elec­
trified.’ The electrified substance was thought to have gained some
sort of electrical fluid that—once actually in the substance—
remained stationary, hence the term static electricity. The scientists
who studied electricity came to be called ‘electricians’—though not
with the same connotation we use today.
In the 1600s, the German inventor Otto von Guericke built a
machine that consisted of a large ball of sulfur rotated at high
speeds by means of a crank. Electricity was produced by friction by
applying a cloth pad, or even his hand. Von Guericke could trans­
mit electricity a meter or more through a moistened string. He had
no way of realizing it at the time, but he had constructed the first
electric generator, and showed that electricity could be transmitted
over distances.
Benjamin Franklin, the first world-class American scientist, pro­
posed in the mid-eighteenth century the concept that electricity is a
fluid. An English ‘electrician,’ Stephen Gray, recognized that the
electric fluid can move through some kinds of objects. Substances
that readily pass the electric fluid we now call conductors; those
that retain the electric fluid we now call insulators. As a rule, good
conductors are metals. Wood, rubber, and plastics are other exam­
ples. The French ‘electrician’ Charles Du Fay observed that there
seemed to be two kinds of electricity: one generated by rubbing

DOI: 10.4324/9781032665603-6
ELECTRICITY 51

glass; the other, from rubbing resin. He showed that the same kinds
of electricity repelled each other, but opposite kinds attracted.
Franklin refined Du Fay’s concept to develop the idea of positive
and negative electricity. In Franklin’s view, matter under ordinary
conditions was not electrified. Franklin had the insight that elec­
tricity was not created by rubbing; it was only being transferred. An
‘un-electrified’ object could either gain electric fluid and become
positive, or could lose electric fluid and become negative. When a
positively charged body is brought into contact with an uncharged
or negatively charged body, excess electric fluid in the positively
charged body would flow—spontaneously—into the uncharged or
negatively charged one.
Franklin’s hypothesis about the flow of electric fluid suggests the
analogy of a flow of water downhill. Something that flows sponta­
neously must do so because of a difference in potential. Electrical
potential energy, or electrical potential, is measured in volts, in
honor of the Italian ‘electrician’ Alessandro Volta, and is given the
symbol V. (Volta’s work will be discussed in more detail later.) In
keeping with the analogy of a waterfall, the low-potential side of an
electrical system is sometimes referred to as the ‘ground.’

HOW ELECTRICITY BEHAVES


In a waterfall, water moves from high to low gravitational potential
energy. In a gravitational system, in which anything falls, the ‘thing’
that is moving is matter. Water is just one specific kind of matter.
Both the difference in potential and the amount or quantity of
matter that flows are important. A single drop of water falling the
height of the world’s tallest waterfall will be not nearly as effective
in doing work as thousands of liters of water flowing down just a
few meters over a working waterwheel. The same is true in a ther­
mal system. Pouring a teacup of boiling water onto an icy sidewalk
isn’t going to melt much ice.
In a river or waterfall, the current is the volume of water flowing
in a unit of time, e.g., cubic meters per hour. In an electrical system,
the electric current is the motion of the electrical charge. The cur­
rent is the amount of charge flowing per unit time. The unit of
electric current is the ampere, or, as often called, the amp,1 named
in honor of André Ampère, a French ‘electrician’ who began
studying advanced mathematics at age 12.
Not all materials are equally good insulators or conductors. Alu­
minum is a better conductor than iron, and silver is better than
52 ENERGY: THE BASICS

aluminum. The amount of electric current that will flow depends


not only on the potential difference (the voltage) but also on the
ease with which the current flows. It has become customary to
consider not the ease of flow (i.e., conductivity), but rather resistiv­
ity. All materials have an intrinsic resistivity opposing the flow. In
any actual device, the amount of current will depend not just on the
resistivity, but also on the length and the cross-sectional area
through which the current has to flow. The property that combines
resistivity, length, and area is called the resistance. About a century
later, the German physicist Georg Simon Ohm worked out the
fundamental laws of conductance. He saw the analogy between the
flow of electric current in a wire and the flow of heat through a
material that conducts heat. To develop the analogy, he needed a
concept for electricity that would correspond to the role of tem­
perature in the flow of heat. Ohm conceived a quantity now called
electromotive force, essentially equivalent to the voltage, which
pushed electricity along the wire in the direction of the current. The
amount of heat that flows via conduction depends in part on the
nature of the material it’s flowing through (contrast a metal cup
with a Styrofoam cup, for example). In the same way, the flow of
electric current should vary with the nature of the conductor. From
that analogy, Ohm introduced the idea of resistivity as a property of
the material through which the current is flowing. It has the units of
ohms. Current was directly proportional to voltage and inversely
proportional to the resistance of the conductor. This relationship is
now known as Ohm’s law.
It is often useful to think of the flow of electricity as being similar
to, or even the same as, a flow of water. Voltage is like water pres­
sure, and amperage is like the volume of water.2 In a water hose, the
pressure of water supplied to the hose is constant. This is compar­
able to an electrical system with a constant voltage. Compressing
the hose increases the resistance to the flow of water.3 When resis­
tance increases and pressure stays constant, flow decreases. In the
same way, increasing resistance in an electrical circuit and keeping
voltage (the ‘pressure’) constant, will cause the current to drop.
Connecting a second hose, to double the length, would cause the
flow to decrease. Using a hose of a smaller diameter would also
cause the flow to drop. Electricity flowing through an electrical
system will behave analogously. All samples of copper have the
same resistivity, but a copper wire 10 m long has twice the resis­
tance of a wire of the same diameter but only 5 m long.
ELECTRICITY 53

When a wire breaks, the current stops because air has so high a
resistance (i.e., is so good an insulator) that the voltage is not suffi­
cient to overcome it. Electricity in household circuits stays in the
wires and does not leak out through the walls because the copper or
aluminum in the wires has a much lower resistance (i.e., is a much
better conductor) than the insulation around the wires. Since it is
easier for the current to flow through the low-resistance path—the
wire—than through some route of higher resistance, electricity follows
the circuits we create for it.
A very few materials have no resistivity whatsoever. They are
called superconductors. They have some applications, including one
very important to us, in the magnetic resonance imaging (MRI)
equipment used in medicine. Future applications might include
long-distance transmission of electricity without losses, and high-
speed magnetically levitated (‘maglev’) trains to replace conven­
tional railroads. So far, superconductors show this remarkable
property only at low temperatures. The highest temperature for
superconductivity at ordinary pressures is about –135°C. The need
to keep superconductors insulated and very cold has prevented their
applications in electrical transmission lines, high-speed transport,
and household electrical products. A challenge is to develop mate­
rials that would be superconducting at ordinary temperatures, or at
least at temperatures that could be maintained with common
refrigerating equipment.4 Such a development would be a revolu­
tionary advance in our use of energy.

ELECTRICITY TO CONSUMERS
We previously defined power as the rate of doing work, or the rate
of using energy. Current is the rate of flow of electric charge. Elec­
trical power, which has units of watts, is determined by the product
of current times potential (equations are in the Appendix). A
stream of water flowing through a hose at a pressure of 10 mega-
pascals (MPa) is more powerful than an equal-sized stream at 1
MPa. A high-pressure stream of water can knock a person down, or
do considerable damage to structures. In the same way, a current of
electricity provides more power at the relatively high ‘pressure’ of
100 volts than the same current would at 1 volt.
In many parts of the world, consumers using electricity are
charged a certain cost per kilowatt-hour (kWh). A kilowatt (1,000
watts) is a unit of power. A kilowatt-hour, therefore, is a unit of
(power  time), which is energy. When you pay your electric bill,
54 ENERGY: THE BASICS

you are not purchasing power, you are actually paying for energy.
The local ‘power company’ that produces electricity in ‘power
plants’ is not in the power business at all. It is producing and
selling energy.
Any electric generating facility wants the smallest possible losses
of electricity while it is being distributed to consumers. Though the
wires in electrical distribution systems are excellent conductors,
there is still resistance to the flow of current. Sending electricity
through wires requires doing work to overcome the natural resis­
tance of the wires. When that work is performed, some energy is
inevitably wasted. Much of that wasted energy appears as heat,
sometimes referred to as resistance heating.
For a distribution system having a given resistance, power is
proportional to the current squared. Significant reductions in losses
can be achieved by making the current as small as possible. But if
the current is made very small to reduce losses during transmission,
then the voltage must be very large to maintain a given value of the
power. Consequently, electric generating plants transmit electricity
at extremely high voltages—thousands or even tens of thousands of
volts. The higher the voltage for a given amount of power, the lower
will be the current, and the lower will be the resistance losses.
Voltage is changed in an electrical circuit by using a transformer,
which can be used to increase or decrease voltage. Neglecting small
losses in the transformer itself, the electrical energy going into the
transformer must equal the electrical energy coming out. (Thanks
to the First Law.) For a given period of time, it would also be true
that the electrical power must be the same into and out of the
transformer. For a 220-volt, 100-amp electrical service to a resi­
dence, 22,000 watts of power could be transmitted at 20,000 volts
and 1.1 amps, and then stepped down to 220 volts in an appropriate
transformer.
We can’t really neglect the losses in the transformer. The back of
your laptop will be noticeably warm after using it for a while. This
warmth is due to the fact that no device can operate at 100 percent
efficiency. The charger in your laptop has a transformer to reduce the
voltage of your household electricity, say 110 V, to the 5 V needed to
charge the battery.5 Energy losses in the slightly inefficient function­
ing of the transformer appear as heat. This might not appear to be
too bad for any one of us with our own laptop, but think of how
many laptops (and mobile-phone chargers, which function the same
way) there must be in the world by now. In 2023 alone, about 190
million laptops were sold worldwide.6 Then add the 7.2 billion
ELECTRICITY 55

smartphones in use around the world.7 Not a single one of them


operates with 100 percent efficiency. All produce waste heat.
Electrons flowing through any real material, which will have some
resistance, lose energy. Work must be done in order for electrons to
flow. The electrical energy that is lost is transformed into heat. In a
laptop charger, the small amount of heating is usually considered a
nuisance. However, if we want to, we can force a current through a
material of high resistance, generating considerable heat. This is
taken advantage of in electrical heaters, in the ‘burners’ of electric
stoves, and in the filaments of incandescent light bulbs, as examples.
The importance of electricity comes from its versatility. A steam
engine, for example, can be used for producing reciprocating or
rotary mechanical motion, and that’s about it (disregarding its sig­
nificant noise and vibration). We can use electricity to produce
mechanical motion as well, with electric motors. But we also use
electricity for lighting. Many of us use electricity for domestic
heating and for operating various appliances. Electricity operates
mobile phones, computers, and other devices for communication or
entertainment. Electrically driven transportation is vital. In many
countries, electric locomotives dominate railway traction. Electricity
is easy to regulate, and can be put to use with the flip of a switch or
push of a button. At the places where electricity is used, there is no
pollution and little noise. (This last point is very definitely not true
of the places where it is generated, as we will see in later chapters.)

HOW ELECTRICITY IS GENERATED: THE CONCEPTS


To put electricity to work, we need a continuous flow, to operate the
many electrical devices that form part of everyday life. The first key
observation was made about 250 years ago by an Italian scientist,
Luigi Galvani. While dissecting frogs, Galvani showed that the leg
muscles would twitch when an electric spark was applied to them,
though the tissue was no longer living. When brass or silver hooks
were used to attach the leg muscles to an iron railing, there was a
significant effect on the ‘twitch’ of the leg muscles.
Alessandro Volta, another brilliant Italian physicist of that era,
recognized that what had produced the effect Galvani observed in
his experiments was the two different metals (the brass hooks and
the iron railing) being in contact. Volta reasoned that two different
metals in contact with a solution capable of conducting electricity
would generate an electric current. He produced devices consisting
of two metals, e.g. zinc and copper, in contact with a medium that
56 ENERGY: THE BASICS

would conduct electricity, such as blotting paper soaked in a solu­


tion of salt. (In Galvani’s frogs’ legs, the fluid in the muscle tissues
acted to conduct electricity.) Volta found it possible to draw electricity
continuously from the ends of a stack of alternating zinc and copper
plates, separated with brine-soaked paper (eventually called a Voltaic
pile). The Voltaic pile is the forerunner of our modern battery.
Today the Voltaic pile would be called an electrochemical cell.
Strictly speaking, two or more cells connected together constitute a
battery, though we use the term informally to refer even to a single
cell. Though it was a tremendous advance, the Voltaic pile was still
far from ideal as a steady source of electricity. For one thing, like
modern-day batteries, the pile eventually ‘runs down.’
Its modern descendants, the batteries familiar nowadays, create
an electric potential when needed to do work. In this system, it is
necessary to create a state of high potential energy. To extract work
from a system, we take advantage of a spontaneous change from
high to low potential. In the battery, we must cause a non-sponta­
neous change from low to high potential. To do this, we must put
work into the system. In any system, providing a high potential
energy—whether electricity, thermal, gravitational, or any other
form—is illustrated conceptually by a ‘reverse energy diagram’
(Figure 6.1).
A battery ‘pumps’ electrical charge uphill to establish the high
electrical potential energy (high voltage). Batteries are becoming
extremely important for energy storage, an issue that we will discuss
in detail in Chapter 13.
In 1820 the Danish scientist Hans Christian Ørsted observed that
when a current from a Voltaic pile was flowing through a wire, the

Figure 6.1 A reverse energy diagram for an electrical system. This diagram
shows that a transition from low to high potential energy is not
spontaneous. To cause a non-spontaneous change to occur, we
must supply work to the system, not extract work from it
ELECTRICITY 57

needle of a magnetic compass responded just as if a magnet were


brought near it. Since it is well known that a compass responds to a
magnetic field, Ørsted deduced that a flowing electrical current
generates a magnetic field. Within a week of learning of Ørsted’s
experiment, Ampère published an extensive theoretical treatment of
the subject, showing why an electric current causes a compass
needle to deflect.8
Then, in 1831, the great English chemist and physicist Michael
Faraday asked an extremely profound question: Will Ørsted’s
experiment work backwards? In other words, will a moving mag­
netic field generate an electric current? By passing a magnet in and
out of a loop of wire, Faraday showed that a current is indeed
generated as the magnetic field passes through the conductor (i.e.,
the wire). Faraday’s very simple experiment answered his original
question in the affirmative. A moving magnetic field generates an
electric current. The phenomenon he observed is now called elec­
tromagnetic induction. Faraday realized that electricity and mag­
netism are two manifestations of the same physical force. Kinetic
energy, used in pushing the magnet back and forth, can be con­
verted into electrical energy. This simple observation is responsible
for all electricity generation today, with the exceptions of the small
amounts we obtain from batteries and from the direct conversion of
solar energy into electricity. Electric heating and lighting at home,
our electrical appliances, electric motors and other electric devices
in industries, electrical railway locomotives, streetcars, and subways
all come from this simple experiment. Electricity is of steadily
increasing importance in the use of digital technology, which
nowadays uses about 10 percent of all the world’s electricity.9
As long as the magnetic field was in motion, an electric current
could be generated essentially forever. The reciprocating motion of
Faraday’s original experiment is not desirable for large-scale gen­
eration of electricity. Faraday refined his original approach with two
improvements. First, it’s easier to sustain rotary motion than reci­
procating motion. Second, the key lies simply in having the mag­
netic field and electrical conductor moving relative to each other. It
doesn’t matter which of the two is moving and which is stationary.
Faraday’s improvements to his original experiment led to the design
of the electrical generator or dynamo. (The difference is that a
dynamo produces direct current (DC), in which the current flows
steadily in one direction, while a generator produces alternating
current (AC), which changes direction very frequently.)
58 ENERGY: THE BASICS

Most of the electricity we use directly comes from rotating gen­


erators operated by falling water or by steam.10 A good bit of the
rest comes from battery-operated devices that need to be recharged
periodically from chargers connected to electric mains energized in
turn by rotating generators. So, the essence of electricity generation
on a large scale is to keep the generator turning as cheaply and
reliably as possible. An electrical generator converts mechanical
kinetic energy (almost always rotary motion) into electrical energy.
One cheap, reliable way of spinning the generator is to use the
kinetic energy of falling water, and another is to use wind. All other
approaches make use of steam. Direct conversion of solar energy to
electricity dispenses completely with electromagnetic induction, as
we discuss in Chapter 9.
Work must be supplied to a generator, i.e., the work of spinning
the generator. But, work put into spinning a generator can be phy­
sically separated, even by hundreds of kilometers, from the place
where the electrical energy is used. This gives a way of taking
kinetic energy that might be available in one place, and easily and
conveniently making it available for use, as electricity, someplace
else. Electromagnetic induction made it possible for the first time to
do work free of the geographical constraints of the availability of
water flowing to waterwheels.
As consumers of electricity, we would want the generator to
operate with near-perfect reliability, and to be operated as cheaply
as possible, since sooner or later all of us somehow pay for elec­
tricity. Over about the last half-century, an increasing number of
people in many countries have been concerned with having elec­
tricity generated with minimal impact on the environment. The next
several chapters will discuss some of the principal options for
meeting these goals.

NOTES
1 An ampere is defined as the amount of electric charge in one coulomb
flowing in one second. One coulomb of charge is 6.24x1018 electric charges.
https://www.nist.gov/si-redefinition/ampere-introduction.
2 In physical reality, they are not the same at all. An electric current comes
about from a flow of electrons. A current of water is the flow of water
molecules.
3 For instance, when some fool stands on, or drives over, your garden hose.
4 So far, the record for any superconductor seems to be –23°C, but at
exceptionally high pressures (250 gigapascals) and using a rather unusual
material, lanthanum decahydride.
ELECTRICITY 59

5 There’s also another energy loss. Batteries need to be charged using direct
current (DC) while most mains are alternating current (AC). The conver­
sion of household or office AC to battery-charging DC is effected by a
rectifier. This also has a small inefficiency, another energy loss.
6 https://www.stellarmr.com/report/Laptop-Market/2107.
7 https://explodingtopics.com/blog/smartphone-stats.
8 Ampère likely had been thinking along the same lines before Ørsted pro­
duced the necessary experimental demonstration.
9 Cassoret, Bertrand. Energy Transition. CRC Press: Boca Raton, 2021, p. 38.
10 Smil, Vaclav. Numbers Don’t Lie. Penguin Books: New York, 2021, p. 65.
7

ELECTRICITY FROM STEAM

STEAM TURBINES
Our lights are on thanks to Michael Faraday. Electricity generation
requires that we put mechanical work into a system (by turning the
generator) to pump electrical energy uphill from low to high
potential. Widespread, commercial electricity generation requires a
source of mechanical work to turn the generator as cheaply and
reliably as possible. ‘As long as you can get a shaft to rotate with a
reasonable amount of force, you can hook a generator up to it and
turn that rotation into electric power.’1
In the 1880s the reciprocating steam engine was well developed,
serving to propel ships, pull railway trains, pump water, and operate
many kinds of machinery. It was an obvious choice for providing
the reliable operation of large-scale electricity generators, but not
without disadvantages. The reciprocating motion of the pistons
needs to be converted to the rotary motion of the generator. This
conversion is inefficient and produces significant undesirable vibra­
tion. Reciprocating steam engines are slow. As the nineteenth cen­
tury drew to a close, a major need of the rising electrical industry
was an engine that produced rotary motion directly and could run
at high speeds with smooth, vibration-free operation. Enter the
steam turbine.
A turbine in which the working fluid is steam is an external-
combustion engine. The English engineer Charles Parsons pio­
neered the development of steam turbines that came to be used for
driving generators. Development of this machine for application in
the electricity industry took place quickly, so that steam turbines
came into widespread use in electricity stations by the end of the
nineteenth century. The general style of the Parsons turbine has not
been significantly changed since the 1880s, though there are now

DOI: 10.4324/9781032665603-7
ELECTRICITY FROM STEAM 61

many variants of this type. Remarkable improvements in efficiency


and enormous improvements in power output have been achieved.
How does a steam turbine operate? For a fixed amount and
volume of any gas—such as steam—held at constant temperature,
an increase in temperature will increase the pressure. This is the key
to creating a head of steam. A high-pressure head of steam entering
the turbine will also be at high temperature. As high-pressure steam
goes through the turbine, it increases in volume. If the volume
increases, the temperature will drop. This temperature drop is a
spontaneous flow of heat from high thermal potential energy to low
thermal potential energy. As is always the case, when a system
changes temperature, we can—by inserting the right device into the
system—extract work. The generic device for extracting work from
a spontaneous change in temperature is a heat engine, of which a
steam turbine is one example.
In a thermal system the efficiency will increase the greater the
difference between the high temperature and the low temperature.
It’s generally simpler and less costly to heat something than to cool
it below ordinary temperatures (compare the prices of a stove and a
refrigerator of comparable quality of materials and workmanship).
So, the approach to high efficiencies is to make the high-tempera­
ture side of the heat engine as hot as possible.
In a modern electricity-generating plant, the steam temperature
entering the turbine is about 600°C, or 873 K. The heat eventually
winds up in a body of water such as a river at about 25°C, or 298
K. The Carnot efficiency would be about 65 percent. In reality,
efficiencies of fossil-fuel-fired generating plants are usually less than
40 percent. Nearly two-thirds of the chemical potential energy of
the fuel winds up transferred to the low-temperature repository, or
cold reservoir.
Why put up with this? Why not turn all the heat liberated from
the fuel straight into electricity? It is because we are stuck. Since
electricity production extracts energy from the generator, something
must supply work to the generator to keep it turning. The work
comes from the steam’s thermal energy, converted to rotating
kinetic energy in the shaft that connects the turbine to the gen­
erator. Because heat can’t be converted directly into work at 100
percent efficiency, the steam turbine converts a limited amount of
heat into work as the heat flows from a hotter object—the high-
pressure steam—to a colder object—the external water or air. There
absolutely must be a cold side to the engine (otherwise there is no
heat engine). As Carnot put it, ‘The production of heat alone is not
62 ENERGY: THE BASICS

sufficient to give birth to the impelling power: that there should also
be cold; without it, the heat would be useless.’2 That is, the plant
must dump heat—a large amount of heat, as it turns out—into the
cold reservoir. Turbines achieve much higher efficiencies than were
ever possible with the reciprocating steam engines.
The steam turbine has become the standard technology for operat­
ing the generator in most of the world’s steam-based electricity-gen­
erating plants.3 An alternative turbine-based technology, the gas
turbine, is the major competitor, but does not use steam as the work­
ing fluid. The dominant fuel worldwide for steam turbine plants is
coal, though this statistic is skewed by the Chinese energy economy,
where China is building coal-fired plants at a remarkable pace. Pro­
ducing electricity using steam involves a turbine/generator set, also
called a turbogenerator, in which high-pressure steam is used to oper­
ate the turbine. Earlier we said that the central problem in electricity
generation was finding a cheap and reliable way to turn the generator.
The steam turbine solves that. Taking a ‘step back’ from the generator,
the practical problem to be solved becomes that of finding a cheap and
reliable way to generate high-pressure, high-temperature steam.

LATENT HEAT
Before we discuss actually how to make steam, we need to consider
some heat effects in vaporizing water to steam and condensing
steam back to water. Intuitively, if we add heat to something, its
temperature should rise steadily, as long as we continue to supply
heat. This is usually true. However, when the material being heated
melts or boils something different happens.
In the kitchen it seems to take far longer to get water to boil off
as steam than it did to heat the water to its boiling point in the first
place.4 In the mid-eighteenth century this seemingly odd business
was investigated by the Scottish scientist Joseph Black. He observed
that adding heat to water would indeed raise its temperature stea­
dily, but only until the boiling point (100°C) was reached. Adding
more heat would not raise the temperature until all of the water had
been converted to steam. Then, addition of still more heat would
cause the steam temperature to rise steadily. Comparable observa­
tions can be made with ice. Heating a block of ice raises its tem­
perature, but only up to its melting point (0°C). The temperature
increase stops at that point, until all of the ice has melted to water.
Only then does further addition of heat raise the temperature of
water.
ELECTRICITY FROM STEAM 63

These observations suggest that heat is doing two different things.


Black named the heat that actually increases the temperature as
sensible heat, meaning something that can be perceived by the
senses. But in a way, ‘sensible’ is also consistent with its other
meaning: something reasonable or expected from common sense.
It’s common sense to expect that if we apply heat to something, its
temperature will rise. Black used the term latent heat for the heat
that seemed to go missing at 0°C or at 100°C.
Why is there latent heat? Ice is a solid because moderately strong
interactions between the individual molecules of water hold them in
a rigid configuration in the solid state. Ordinary water is a liquid
because there are also inter-molecular attractive forces holding the
molecules strongly enough to prevent them from vaporizing. To boil
water, we must supply energy to increase the kinetic energy of the
water molecules sufficiently to overcome the inter-molecular forces
holding them in the liquid. The energy that we supply to allow the
molecules to move rapidly enough to escape the liquid phase (i.e.,
to boil) does nothing to increase the temperature. In the ice-liquid
water-steam system, there are two kinds of latent heat. First is the
latent heat of fusion for melting ice to liquid, then the latent heat of
vaporization for converting liquid to steam.
Latent heat seems to go missing, since it doesn’t contribute to the
expected temperature increase. Of course, it really isn’t missing—the
First Law makes sure that doesn’t happen. For boiling, the latent
heat is in the steam. That means that if we were to reverse the pro­
cess, in this case condense steam back to water, the latent heat should
be returned to us. It is. We observe this as the latent heat of con­
densation, for steam to liquid,5 and the latent heat of freezing for
liquid to ice. We’ll deal with the latent heat of condensation later.

PRODUCING STEAM
The short answer to the question of generating large amounts of
high-pressure, high-temperature steam as cheaply and reliably as
possible is this: boil lots of water. In modern electricity-generating
plants, the amount of steam required is often in excess of 500,000
kg (500 tonnes) per hour. Raising the temperature of water, and
then of the steam, requires pumping heat uphill. For this non-
spontaneous change, we have to supply energy, specifically thermal
energy, into the system. The necessary thermal energy is liberated
by burning a fuel; energy in the chemical bonds of fuel molecules is
greater than energy in the bonds of the products of combustion.
64 ENERGY: THE BASICS

Once we supply a bit of activation energy, fuel combustion will be


an energetically downhill spontaneous change. The difference
between chemical bond energies of fuel molecules and of product
molecules will be released from the system, primarily as heat. We
can transfer this heat first to boil water and then to increase the
steam temperature. This sequence of changes gives us the working
fluid for a steam turbine.
The fuels of choice are usually the fossil fuels—coal, petroleum,
and natural gas. Recently-grown plant materials, especially wood,
can also be used. The heat liberated from nuclear fission processes
is another choice. Nuclear plants are sufficiently different, with their
own set of issues, that we will look at them separately, in Chapter 8.
In most countries, coal was the fuel of choice in large-scale genera­
tion from the late nineteenth to the early twenty-first century. In
recent decades, natural gas has taken a greater share of the fossil
fuel market for electricity generation, though in gas turbines, not
steam turbines.
The ability to boil water on the prodigious scale of 500 tonnes
per hour is governed by two factors: the rate at which the
necessary heat can be produced, and the rate at which heat from
the fuel can be transferred into the water or steam. The rate at
which heat is released from the fuel relates directly to the rate at
which it can be burned. Heat transfer depends on many factors,
including the surface area across which heat is transferred. Since
in most steam-generating equipment the other factors are,
roughly, the same, surface area is the key to heat transfer. A
steam-generating system consists of a furnace, where heat is lib­
erated by burning fuel and is transferred to the boiler, where
water is converted to steam. In modern systems, the furnace and
the boiler are physically the same piece of equipment, so the
term ‘boiler’ often means both the place where fuel is burned
and where steam is generated.
In any boiler, the rate of steam generation is limited by the
amount of surface across which heat can be transferred. Maximiz­
ing heat transfer is accomplished by having water or steam flowing
through metal tubes almost completely surrounded by the high-
temperature gases from the burning fuel, a design called a water-
tube boiler. Water or steam circulate in numerous tubes, with the
flames and hot gases acting on the exterior of the tubes. The water-
tube boiler, with dozens or hundreds of water tubes, is about as
good as can be done for increasing the effective heat transfer, at
least in a practical design at reasonable cost.
ELECTRICITY FROM STEAM 65

The heat release rate depends on the fuel of choice and the design
of the burners in which the fuel is used. We will consider the ques­
tion of obtaining a high heat release rate from coal, since coal has
been (and still is, in many countries) the dominant fuel for this
application. Because coal is a solid, the reactions of the chemical
components of coal with oxygen molecules occur where the oxygen
molecules can obtain access to the coal, meaning on the surface of
the particles of coal. As coal is heated when first fed into the boiler,
it gives off some small molecules, collectively called volatiles, which
escape into the surrounding air, ignite, and burn. However, most of
the coal remains behind as a carbon-rich char, which must be con­
sumed in reactions on the char surface. The rate of burning will be
limited by the available surface area. To increase the heat release
rate, the surface available for reaction with oxygen molecules must
be increased. This is done by pulverizing the coal to extremely fine
particle sizes (of diameters between 0.02 and 0.05 mm), an
approach often called pulverized coal firing.
Coal consists mainly of carbon and hydrogen, but also contains
nitrogen and sulfur, as well as a variety of noncombustible mineral
particles. The oxides of sulfur and nitrogen6 are a concern because
of their role in the formation of acidic precipitation, commonly
called acid rain, which has been a serious air-pollution problem in
many parts of the world. Sulfur and nitrogen oxides readily dissolve
in water to produce sulfuric and nitric acids. As these acids fall to
the ground in precipitation, they slowly acidify bodies of water,
such as lakes. The water becomes more acidic, until eventually the
lake becomes biologically dead. Various technological solutions are
available to reduce substantially the amounts of sulfur or nitrogen
oxides emitted into the atmosphere.
During combustion, the mineral particles in coal transform to
various solids that collectively comprise the stuff that we call ash.
Ideally, the ash should fall to the bottom of the boiler, to be col­
lected and removed. Most of the ash does that. However, the
extremely turbulent flame results in some very small ash particles
being carried upward and out of the boiler with the hot combustion
products. Such ash particles are called fly ash. If they escape into
the environment, they become an additional pollution problem
termed particulate emissions.
The worst particulates are those smaller than 2.5 microns in dia­
meter (a micron is one-millionth of a meter), and are now generally
known as PM2.5. About thirty of them together equal the diameter
of a typical human hair. If inhaled, such particles can be trapped in
66 ENERGY: THE BASICS

the fine passages of the respiratory system, adding to pollution-


related respiratory distress. As with sulfur and nitrogen oxides, var­
ious approaches have been developed and are in use to capture most
of the particulates.
When the array of already developed and proven technologies is
applied to a coal-fired plant, large fractions of the sulfur and
nitrogen oxides and of the ash can be captured before they are
emitted. Ideally, the only combustion products that would go ‘up
the stack’ are carbon dioxide and water vapor. It has now become
quite clear that emissions of carbon dioxide are contributing
directly to climate change. This problem is so serious that we will
get back to carbon dioxide in detail in Chapter 14.
Low-pressure, low-temperature steam leaving the turbine is
converted back to water, using a condenser. Cold water from a
natural source, a river, a lake, or the ocean, is brought in to
extract heat from the steam. Loss of heat from the steam causes
the condenser water to warm up, especially as it absorbs heat
given off when the steam condenses back to water, i.e., the latent
heat of condensation. Hot water from the condenser cannot be
discharged back to the environment as-is. The natural waters in
the vicinity of a plant could become warmer than they would
normally be without this additional heat input, a problem some­
times known as thermal pollution. The hot water must be cooled
to near-ambient temperature before it can be discharged. This is
done in a cooling tower. Cooling towers are enormous structures,
often built of concrete with a distinctive ‘narrow waist’ design.
Hot water from the condenser is sprayed inside the cooling
tower, where it is cooled by air flowing through the tower.
Though they look formidable, cooling towers are not much dif­
ferent from a gargantuan hot-water shower bath.
The efficiency of conversion of the chemical potential energy in
coal to electrical potential energy leaving the plant is about 35 per­
cent. Where does the rest of the energy go? The generator efficiency
could be about 99 percent. The boiler will utilize chemical energy in
fuel with an efficiency of about 85–90 percent. The turbine might
have an efficiency of 45–55 percent. Some of the chemical energy in
the coal goes straight up the stack in the hot combustion products.
Some heat is lost as steam moves through the plant, from boiler to
turbine to condenser. More is lost in warming cooling water in the
condensers. A highly efficient plant, by today’s standards, will throw
away the equivalent of 2 megawatts (MW) in wasted power for
every 1 MW put into the distribution system. By the time the
ELECTRICITY FROM STEAM 67

electricity actually reaches the consumer, more losses will have been
encountered in the wires and transformers. Only about 30 percent
of the chemical potential energy in the coal is available as electrical
potential energy at the outlet in the wall.

GAS TURBINES
An internal-combustion turbine driving a generator would eliminate
completely the separate boiler, condenser, and cooling tower. These
are large, and therefore expensive, pieces of equipment. Eliminating
them would save on the investment cost of building a plant, simplify
the plant layout, and likely increase overall efficiency. Fortunately,
we can indeed make use of internal-combustion turbines. The fuel
of choice is natural gas. The turbines are conceptually similar to the
jet engines we discussed earlier, so similar in fact that they are
sometimes called aero-derivative gas turbines.
In a gas-fired turbine, combustion temperatures are as high as
1250°C. Expansion of these gases leads to a temperature drop
which provides the thermal energy that is converted to mechanical
energy by the turbine; the mechanical energy is what turns the
generator. By the time the expanded gases leave the turbine they
have cooled to about 600°C, providing efficiencies of over 40 per­
cent.7 A large gas turbine can produce power in excess of 500 MW.
The dominant component of natural gas is methane. It con­
tains only carbon and hydrogen, so its combustion products are
only carbon dioxide and water vapor. Though we no longer
ignore the carbon dioxide (Chapter 14), combustion is relatively
simple and clean.
In general, steam turbines operate at higher pressures and
temperatures than gas turbines. This results in the steam turbines
having higher overall efficiencies. Steam turbines are also built as
larger units, because they need much more supporting hardware
infrastructure—such as boilers, with all the necessary piping and
water- or steam-handling. If we have to build all this stuff, we
might as well make it big. But, the smaller and less hardware-
dependent gas turbines can be built as modules, achieving a
larger total power output from multiple smaller gas turbines. Gas
turbines also tend to require much less maintenance. In terms of
overall plant efficiency, natural gas plants are typically in the
mid-40 percent range, whereas many coal plants tend to be in
the low- to mid-30 percent range.
68 ENERGY: THE BASICS

COMBINED-CYCLE PLANTS
The gases leaving a natural-gas-fired turbine have expanded, dropped
in temperature, and released thermal energy from their ‘downhill’ fall
to the lower temperature. Even so, they are still pretty hot, around
550°C, and far above the boiling point of water. In principle, the
gases leaving the turbine could be passed through a boiler, where they
would be used to produce steam, which could be used it to operate a
steam turbine. This works quite well in practice.
An electricity-generating plant can have two turbines, each operat­
ing its own generator. First, a gas turbine is used to run one generator.
Some of the heat remaining in the gases leaving the turbine can then
be captured in a boiler, producing steam to feed to a steam turbine.
Plants employing this design are called combined-cycle plants.8 The
combined-cycle concept is now about a half-century old. The overall
efficiency of a combined-cycle plant is about 60 percent.

THE REST OF THE HEAT


Plants using steam turbines will have low-grade steam or hot water
to deal with at the very back end of the plant. Cooling towers pro­
vide a common solution. But there’s another approach. The steam
or hot water still has a significant amount of heat, though not at
very high thermal potential energy—not the exceptionally high
temperatures used in the turbines. The idea of using steam, hot
water, or heated air for space heating in both domestic and com­
mercial applications goes back to 100–200 BCE. The available low-
grade steam or heated water provide an opportunity for supplying
heat to buildings in the region around an electric plant, via an
appropriate arrangement of piping and pumps. This concept is
known as district heating.
The concept is very old, the earliest versions dating from ancient
Rome. The United States Naval Academy began a district heating
project in 1853. These, of course, were not tied to electricity gen­
eration, and certainly it is possible to develop district heating sys­
tems that do not depend on an electric plant for the residual heat.
A related concept, co-generation (often known as combined heat
and power (CHP)), uses the hot water from the back end of a plant
for many kinds of industrial applications, such as in the pulp and
paper industry or in petroleum refineries. CHP plants can also
supply heat to district heating systems. Co-generation plants in
which the electricity generation comes from a combined-cycle plant
ELECTRICITY FROM STEAM 69

can reach overall efficiencies in the 80 percent range. The best effi­
ciencies come from having the electricity generation and the heat
user close together, even co-located.
It’s not uncommon to hear the term ‘waste heat’ in reference to
hot air or other gases, low-grade steam, or hot water. That’s an
unfortunate usage, because heat doesn’t ever have to be wasted.
Anytime there is a temperature difference, we should be able to come
up with a way to exploit it, and extract a bit more useful energy.

NOTES
1 Gray, Theodore. Engines. Black Dog & Leventhal Publishers: New York,
2022, p. 164.
2 Sadi Carnot, quoted in Sen, Paul. Einstein’s Fridge. Scribner: New York,
2021, p. 14.
3 Plants relying on falling water, solar energy, and wind energy, all of which
we will discuss in forthcoming chapters, do not use steam.
4 This phenomenon may be the scientific basis of the folk saying that ‘a wat­
ched pot never boils.’
5 This is why being burned by steam at 100°C is much worse than being
burned by water at the same temperature. Of course, any burn is painful and
deserves immediate treatment. Being burned by hot water results in a
transfer of sensible heat from 100°C (in this example) to your skin. But
being burned by steam is a result of the sensible heat of the steam plus the
latent heat of condensation, which could be seriously bad news indeed.
6 At the extremely high temperatures of the combustion reaction, there is also
a reaction between nitrogen molecules in the air and oxygen. This is a
second contribution to nitrogen oxide emissions.
7 Smil, Vaclav. Numbers Don’t Lie. Penguin Books: New York, 2021, pp. 140–141.
8 The name derives from the fact that the two turbines operate somewhat
differently, one a gas turbine and the other a steam turbine.
8

NUCLEAR ENERGY

NUCLEAR FISSION
At 3:45 p.m. on December 2, 1942, the world changed forever. At
that moment, a team led by the Italian-American physicist Enrico
Fermi1 achieved the first controlled release of nuclear energy. The
event paved the way both for nuclear reactors to produce electricity
and for nuclear bombs. It occurred in the rather modest venue of a
squash court underneath the stands of the football stadium at the
University of Chicago. We will first examine what happens in a
nuclear process, and then how we can take advantage of the energy
released to generate electricity.
Henri Becquerel, the third generation of a remarkable family of
French physicists,2 sought a connection between the recently dis­
covered X-rays and the phenomenon of fluorescence at the end of the
nineteenth century. Fluorescent materials emit light when they are
stimulated by light or other electromagnetic rays; we have LED lights
because of fluorescence. Becquerel wondered whether a fluorescent
material that emits visible light also produces X-rays. They were
known to ‘expose’ photographic films—the basis of X-ray examina­
tions by physicians or dentists. Becquerel designed an experiment to
test whether a fluorescent mineral, potassium uranium sulfate, could
give off X-rays, to be detected on a photographic plate. The experi­
ment led, instead, to the discovery of some previously unknown kind
of radiation coming steadily out of the mineral.
This constant emission of radiation was termed radioactivity
by the Polish-French scientist Marie Curie.3 She showed that it
was the uranium in the mineral that was radioactive, and that all
uranium compounds she tested were radioactive. The intensity of
radioactivity was proportional to the uranium content of the
compounds.

DOI: 10.4324/9781032665603-8
NUCLEAR ENERGY 71

It takes a source of energy to expose a photographic plate.


Recalling the energy balance (Appendix), the energy coming from
this system (EOUT) is emitted from the uranium compounds. No
external energy was put into the system (EIN is zero). Therefore,
energy can only be coming out of the samples if a substantial
amount is already stored in them (ESTORED). Unlike the gradual
cooling of a hot object—where the amount of heat given off
becomes less and less until no more is available to exchange with
the surroundings—radiation coming from uranium went on stea­
dily, at the same intensity.
Three kinds of radiation were found. At that time no one under­
stood the nature of the radiation, so they were named using the letters
of the Greek alphabet: α-rays, β-rays, and γ-rays. γ-Rays displayed an
ability to penetrate materials, much like X-rays. β-rays were much less
penetrating, and α-rays were scarcely penetrating at all.
A useful metaphor for thinking about atomic structure is the
solar-system model: a nucleus (the sun) surrounded by one or more
electrons in orbit around it (the planets). Electrons have a negative
charge and are of very light weight. The nucleus consists of protons,
with a positive charge, and electrically neutral neutrons. The mass
of a proton or neutron is about 1,800 times greater than that of an
electron, so the nucleus contains almost the entire mass of the atom,
just as the sun contains most of the mass of our solar system.
Chemical reactions involve gaining, sharing, or losing electrons. In
an electrically neutral atom, the number of protons equals the number
of electrons. Because the number of electrons around an atom might
change in a chemical reaction, the number of protons in the nucleus,
called the atomic number, is a more reliable indicator of the chemical
identity of an element. An atomic number best discriminates among
the chemical elements. A second characteristic, the mass number, is
the sum of the numbers of protons and neutrons in the nucleus. The
generic term nucleons refers to protons and neutrons together.
Since neutrons are electrically neutral, it is not necessary to
restrict the number of neutrons to maintain electrical neutrality in
the atom. Nuclei having the same number of protons could possibly
have different numbers of neutrons. Atoms or nuclei having the
same atomic number but different mass numbers are called isotopes.
In nuclear reactions, we will use a notation that shows the chemical
symbol of the element, the atomic number as a subscript, and the
mass number as a superscript. As an example, uranium, U, has the
atomic number 92. Its most common isotope has the mass number
238. This isotope would be represented as 92U238.
72 ENERGY: THE BASICS

Chemists have many ways to influence the rate or course of a


chemical process. Nothing they can come up with has the slightest
effect on the rate of radioactivity. So, radioactivity cannot be a
chemical process. If radioactivity is not occurring among the elec­
trons, the only other place in the atom in which it can be happening
is in the nucleus.
Marie Curie defined radioactivity as the spontaneous emission of
radiation and showed that radioactive materials release a prodi­
gious amount of energy. Nuclear reactions release about ten million
times as much energy as chemical reactions. This energy keeps
pouring out apparently endlessly. Since a spontaneous change
involves a transition from high potential energy to one of low
potential energy, we can draw an energy diagram (Figure 8.1). To
understand how to tap into this enormous, seemingly limitless
supply of energy, we begin by inquiring where nuclear potential
energy comes from.
The nuclei of all atoms (other than hydrogen) have more than one
proton. Electrical charges of the same sign repel each other. If the
nucleus of an atom of uranium, with 92 protons, could just fly
apart, there would be a tremendous release of energy from the
repulsion of particles having the same electric charge. There must be
a force holding the nucleus together that is stronger than the ten­
dency of like charges to repel each other—as if there was a tiny,
nucleus-sized rubber band around the nucleus holding all the
nucleons in place.
This force is the nuclear binding energy. It overwhelms the electrical
repulsion between protons. Nuclear binding energy acts only over very
small distances, usually when the nucleons are in contact. In a very
large nucleus, some nucleons may find themselves separated by a

Figure 8.1 The energy diagram for a nuclear process, very similar to gravita­
tional, thermal, and other systems
NUCLEAR ENERGY 73

distance at the extreme limit of the binding energy, yet still within
the realm of the repulsive force. In that case, repulsion of similar
charges ‘wins’ and pushes the nucleons apart. This process we
recognize as radioactivity. All nuclei having more than 82 protons
are unstable in this way.
The relationship between the number of nucleons and the average
binding energy per nucleon is shown in the curve of binding energy
(Figure 8.2). The higher the binding energy, the more stable the
nucleus. The highest binding energy— the most stable nucleus—
occurs at mass number 56, an isotope of iron. Nuclear processes
release far more energy than do chemical processes because the
binding force holding nuclei together is much more powerful than is
the electric force that binds electrons to their atoms.
The rising slope of the binding energy curve indicates that the
average binding energy of the nucleons for small nuclei would
increase if they had more nucleons than they already do. At the
crest of the curve, nuclei with 26 protons, the nuclear binding
energy and the repulsive force are well balanced. Such nuclei are
extremely stable. For nuclei with many protons, electrical repulsion
overwhelms binding energy. The decreasing slope of the binding
energy curve in the region of large nuclei signals that they would
become more stable if they had fewer nucleons, not more.
92U
238
emits α-radiation, which consists of the nuclei of helium
atoms—two neutrons and two protons. The loss of two protons and

Figure 8.2 A simplified sketch of the curve of nuclear binding energy. The
higher a nucleus lies on this curve, the greater its binding energy
and the more stable it is
74 ENERGY: THE BASICS

two neutrons from the nucleus of 92U238 creates a product having 90


protons and 144 neutrons. The complete nuclear reaction
92 U
238
! 2 He4 þ90 Th234 . Uranium has been turned into thorium, a
movement slightly to the left on the curve of binding energy.
Because a nucleus of mass number 234 has a higher binding energy
than one of mass number 238, it is more stable. When a system
spontaneously changes from being less stable to being more stable,
falling ‘downhill,’ some energy is released.
A scrupulous account of all the mass on both sides of the equa­
tion would show that a very tiny amount of mass is ‘missing’ after
the reaction. This very tiny amount of missing mass has been con­
verted to energy. The relationship between missing mass and energy
is given by Einstein’s equation, E ¼ mc2, probably the most famous
equation in all of science. Here E is the energy liberated in the
nuclear process, m is the amount of mass that is lost (the missing
mass), and c is the speed of light. Although m may be a very tiny
number, c is very large (and consequently c2 is extremely large).
Thus, the amount of energy liberated, E, can be enormous.
Physicists had learned that bombarding a nucleus with neutrons can
stimulate emission of radiation. When studying the effect of neutrons
on uranium, Fermi found several new species with chemical properties
that did not match those of uranium, nor of other elements in the
range of atomic numbers from 82 to 92, the logical products of known
(at the time) nuclear processes involving uranium.
The German scientists Otto Hahn, Lise Meitner, and Fritz
Strassman4 found that a barium isotope (atomic number of 56)
formed in the process. Lise Meitner and her nephew Otto Frisch5
recognized that the uranium nucleus, activated by neutron bom­
bardment, had split roughly in half. This splitting process seemed
similar to the division of a biological cell, a process called fission;
Frisch suggested the name nuclear fission for the new phenomenon.
The products of nuclear fission are roughly half the size of the
original 92U235 nucleus. When a uranium nucleus undergoes fission,
it does not always divide in exactly the same way. For this reason, a
great variety of isotopes is produced, depending on just how the
division takes place. They are collectively known as fission products.
Some fission products are themselves unstable, and may emit β- or
γ-radiation. We will come back to this problem later.
Transmutation of uranium to thorium by natural radioactivity
results in moving a small distance along the curve of binding
energy, and releases a small amount of energy. The energy release
from fission is prodigious, because fission is a large shift along the
NUCLEAR ENERGY 75

curve of binding energy. Fission of 92U235 releases two million times


more energy as burning an equal weight of coal. A 600 MW elec­
tricity-generating station could run for a year on the 92U235 contained
in 90 tonnes of naturally occurring uranium (which contains only a
small fraction 92U235). The energy equivalent of a full 80-liter tank of
gasoline is a tiny sphere of 92U235 about 0.25 mm diameter.
Each fission event liberates two or three neutrons, while only one
is required to initiate fission. In principle, it should be possible to
use any one of the newly produced neutrons to start a second reac­
tion that will generate three more neutrons. One of the product
neutrons from that second reaction can start a third reaction, and
so on. Any process—whether nuclear or not—in which the product
of one step is used to start the next step is called a chain reaction.
An uncontrolled chain reaction releases immense energy in a short
time. If the three neutrons emitted in each fission step can each
induce further fissions, and if a microsecond elapses between the
emission of a neutron and its subsequent absorption by another
nucleus, a chain reaction starting with a single fission will release
nearly 21 gigajoules of energy in less than one-thousandth of a
second. This is the basis of the atomic bomb. But, a controlled
chain reaction, in which only one neutron released from fission
causes one new fission, can be made to release energy at a constant
output. A controlled nuclear fission chain reaction provides a sus­
tained release of energy that can be used for doing work. The key to
release of useful energy from nuclear fission is to establish a sus­
tained, controlled, chain reaction.
The basic idea of the chain reaction leads to questions about how
to put it into practice. Is it even possible to create and control a
chain reaction in a practical device? That question was answered on
a squash court at the University of Chicago. It then becomes
important to know how much energy is liberated when the chain
reaction takes place. It is also crucial to know how to control the
process while it is going on, and how to stop it when desired.

NUCLEAR POWER PLANTS


Nuclear fission produces enormous amounts of energy, much of
which occurs as heat. The challenge lies in designing and operating
hardware to capture the energy released from fission in a safe, con­
trolled way. There seems to be no way of producing electricity by
directly using the energy released in fission. Heat can be used to
make steam; steam can be used to operate an electrical generator.
76 ENERGY: THE BASICS

Through this sequence, energy from fission is eventually transformed


into electrical energy. Steam turbines can also supply propulsion
energy in submarines and surface ships, leading to applications of
fission reactors in these vessels.
92U
235
more readily undergoes fission than 92U238, but naturally
occurring uranium is about 99.3 percent 92U238, and only 0.7 per­
cent fissionable 92U235. To sustain a nuclear chain reaction based on
uranium, 92U235 must be present in a greater concentration than in
natural uranium. The process for achieving this, uranium enrich­
ment, requires separating isotopes. This is not easy. Isotopes, as
forms of the same chemical element, can’t be separated by processes
that rely on chemistry. Enrichment relies on physical methods
exploiting the slight difference—about 1.3 percent—in mass
between the two isotopes.
To make a suitable fuel, uranium must be enriched to at least 3–4
percent U235, called low-enriched uranium (LEU), or sometimes
reactor-grade or power-grade uranium. Highly enriched uranium
(HEU), containing up to 90 percent U235, has been used for the
construction of nuclear weapons, and is sometimes referred to as
weapons-grade uranium. The difference in U235 concentration
between reactor-grade and weapons-grade uranium removes the
possibility of a nuclear reactor exploding like an atomic bomb,
because the concentration of fissionable U235 isotopes in the reactor
is too low for a nuclear explosion.
LEU is shaped into pellets and packed into long, thin tubes to
make fuel rods, a group of which constitutes a fuel assembly. As the
reactor operates, the concentration of fissionable U235 drops steadily,
since it is being consumed to release energy. Simultaneously, fission
products, some of which are good neutron absorbers, increase.
Eventually a time comes when the loss of U235 and the build-up of
neutron-absorbing fission products leaves a fuel rod basically useless.
The fuel rod is said to be spent, though it still contains 1–2 percent
U235 and a small amount of plutonium (94Pu239).
To control the fission chain reaction, something that will absorb
neutrons without re-emitting them is required. Neutron-absorbing
rods, called control rods, do this. Control rod assemblies are con­
tained within the fuel assemblies. The rate of the chain reaction
depends on the number of control rods inserted into the reactor,
and on how far they are inserted.
During fission, disintegrating 92U235 nuclei emit fast-moving
neutrons having high kinetic energy. Slower-moving neutrons are
more likely to be captured by 92U235 and contribute to sustaining
NUCLEAR ENERGY 77

the chain reaction. A moderator slows neutrons released from fis­


sion to enhance the prospect that U235 nuclei will absorb them and
continue the fission process.
The reactor core contains the fuel assemblies with control rods,
moderator, and a coolant, which removes heat from the core. The
core is inside a thick, strong steel pressure vessel. The pressure
vessel is commonly enclosed in a containment structure, made of
reinforced-concrete at least 1 m thick, which is the outermost bar­
rier between the core and the environment. As its name implies, the
containment structure is intended to contain the reactor contents in
the event of a leak or an accident.
A sustained nuclear chain reaction depends on using a neutron
produced in a fission event to start the next fission. Producing
exactly one neutron per fission event would not be sufficient to sus­
tain a chain reaction. Some neutrons escape from the core, or are
absorbed by non-fissionable nuclei, and would not be available to
interact with U235 nuclei. The minimum amount of fissionable
material needed to balance the production of neutrons with their
loss, so that a chain reaction is self-sustaining, is called the critical
mass. Having not enough fissionable material to sustain a chain
reaction represents a subcritical condition. If more than one neu­
tron produces a further fission, the reaction is said to be super-
critical, and proceeds at an increasing rate. In the worst case, this
could lead to serious problems.
Nuclear plants differ from fossil-fuel plants only by using the heat
from a nuclear reactor to raise steam for their steam turbines. The
steam turbine–electricity generator sets are similar to those of fossil-
fuel-fired stations. Temperatures and pressures of the steam fed to
turbines in nuclear plants are not as high as those of fossil-fuel-fired
systems. This is due partly to the more severe conditions that the
materials in a nuclear plant must withstand. Because the steam
temperatures are lower, resulting thermal efficiencies are also lower
(a 34 percent turbine efficiency is considered good). Since conver­
sion of energy to work is lower, waste energy—in this case, waste
heat output—is necessarily higher.
Most nuclear power stations worldwide use a pressurized water
reactor (PWR). Because heat from the high-pressure water inside
the core is used in a PWR to raise steam in the heat exchanger, no
radioactive material is ever outside the containment building during
normal operation. Of course, this does not mean that there can
never be a problem with operating such a reactor, nor that there are
no potential environmental problems.
78 ENERGY: THE BASICS

Fossil-fuel-fired boilers and nuclear reactors are simply different


ways of boiling water to feed a head of steam to the turbines. All of
the concern about environmental consequences of burning fossil fuels,
and all of the concern, or occasional uproar, over nuclear reactor
safety and nuclear waste, ultimately derive from the question of how
to boil water. In a nuclear plant, the sequence of changes begins with
the release of nuclear binding energy as thermal energy during fission.
Thermal energy is used to increase the temperature of pressurized
water. The thermal energy in high-pressure water is transferred to a
second water loop, converting this water to steam. Thermal energy in
the steam is converted to mechanical work in the turbine-generator,
which produces electrical energy. High-voltage, low-amperage elec­
tricity is transmitted to our communities, where it is converted to low
voltages for use in the home. At that point, that nuclear binding
energy released in the reactor is finally in a form that lets us turn the
lights on and recharge our phones.
Commercial nuclear plants are exceptionally reliable, producing
electricity some 90–95 percent of the time.6 Ordinarily, radiation
from the plant is virtually nil. Radioactive materials are contained
within the reactor system, and only small amounts of short-lived
radioactive elements are allowed to vent into the atmosphere.
Shielding is required by various regulations to protect the health
and safety of workers inside the plant, and, along with the limita­
tions on radioactive discharges, is also designed to meet health
standards for the population outside the plant.
Carbon dioxide emissions are generally low, to the point that
nuclear energy is considered one of the cleanest energy options from
the perspective of carbon emissions. This is the case if our ‘window
of attention’ is the nuclear plant itself. However, there are sig­
nificant carbon emissions associated with the mining, milling, and
processing of uranium ore, and the other steps up until LEU fuel
rods go into the reactor.7

THE NUCLEAR CONTROVERSY


After the Second World War, there were widespread hopes of
extremely cheap energy via nuclear fission. The expression ‘too
cheap to meter’ suggested that electricity could be produced in such
vast quantities in fission reactors, so incredibly cheaply, that it
would not be cost-effective to install electricity meters, nor to
bother mailing out bills for the electricity. It was first used 70 years
ago by Lewis Strauss, chairman of the United States Atomic
NUCLEAR ENERGY 79

Energy Commission.8 Electricity from nuclear plants was not then,


and never will be, too cheap to meter. Massive cost overruns and long
construction delays didn’t help. The situation is complicated further by
the unsolved problem of the disposal and storage of nuclear waste.
The collective impact of concerns about economics, nuclear acci­
dents, and what to do with the radioactive waste have led to enormous
problems for the nuclear power industry in many countries around the
world. Yet another problem is the possibility of subverting an osten­
sibly peaceful civilian nuclear energy program to make weapons-grade
uranium instead of reactor-grade material. This is quite possibly hap­
pening in Iran and North Korea currently.9
We can’t detect radiation with our senses, or feel it affecting us. A
not-unreasonable fear of something we can’t see, smell, taste, or feel
certainly had a negative impact on public perception of nuclear
energy. We are all constantly bathed in natural radiation. Back­
ground radiation comes from cosmic rays from outer space, from
naturally occurring radioactive substances in the environment, and
from medical procedures. Radioactive 90Th232 and 92U238 occur in
some kinds of rocks and soil. Less than 1 percent of our background
radiation comes from sources such as nuclear electricity-generating
plants and their wastes, and from nuclear fuel reprocessing.
During normal operation the radiation level outside the plant
poses no concern to human health or safety. But, when conditions
are more severe than ever anticipated, as in the Fukushima tsunami
of 2011, disaster can happen. Of greatest concern is the prospect
that a nuclear reactor could release substantial amounts of radia­
tion into the environment in the event of a serious accident.
The use of hydroturbines (Chapter 11) is considered to be a very
clean, safe technology. Yet during the twentieth century, far more
people died as a result of the breakage of dams than died in nuclear
reactor accidents.10 Also, France, which has some 60 nuclear plants, has
the second-longest longevity of its population in the European Union.11
This suggests that nuclear plants are not likely a significant health risk.
In commercial fission plants, one-third to one-fourth of the fuel rods
are removed and replaced each year. Fissionable 92U235 remaining in
the rods it could be recovered and recycled. Other reactor products
cause significant concern, especially plutonium (94Pu239), which is
fissionable and can be used to make atomic weapons.
The waste disposal issue remains controversial, due in part to the
NIMBY syndrome (not in my back yard!). Because of the political,
social, and technical problems associated with radioactive waste
disposal, no one wants it stored nearby.
80 ENERGY: THE BASICS

NOTES
1 Fermi was probably the last physicist to master both experimental and
theoretical physics, and is occasionally dubbed ‘the pope of physics.’ Born
in Italy, Fermi had a stellar career there until he and his wife emigrated to
the United States to escape the vile racial policies of the fascists.
2 His grandfather, father, and son were also physicists, and each had their
own productive careers. Grandfather Antoine discovered, among other
things, how to produce metals from their ores by electrolysis. Father
Edmond discovered the photovoltaic effect, the basis of the operation of
solar cells. Son Jean’s researches included studies of the magnetic and
optical properties of crystals; he became the fourth consecutive Becquerel
to hold the chair of physics at the Muséum National d’Histoire Naturelle in
Paris. A pretty good set of genes in that family!
3 This extraordinary woman was one of the great scientists of the twentieth
century: the first woman to win a Nobel Prize, the first member of a mar­
ried couple to win a Nobel, and the first person to receive two Nobel
Prizes. Not to be outdone by the Becquerels, the extended Curie family
among them collected five Nobel Prizes.
4 Hahn was a German chemist who had worked with Rutherford for a year
at McGill University. Strassman, also a German chemist, worked with
Hahn in Berlin. Both opposed the Nazis and their policies. Meitner was an
Austrian physicist who was the first woman to become a full professor of
physics in Germany, so brilliant that Einstein referred to her as ‘the
German Marie Curie.’ But, because she was Jewish, she was forced to flee
Germany, eventually going to Sweden. She was not awarded a share the
Nobel Prize that went to Hahn for the discovery of fission.
5 Lise Meitner’s nephew.
6 Smil, Vaclav. How the World Really Works. Viking: London, 2022, p. 25.
7 For example, Pomponi, F. and Hart, J. ‘The greenhouse gas emissions of
nuclear energy—Life cycle assessment of a European pressurized reactor.’
Applied Energy, Vol. 290, p. 116743, 2021.
8 Strauss was actually speaking of nuclear fusion plants, a much different
technology from fission, and which have not even yet attained sustained
operation for more than a few seconds. The Atomic Energy Commission
was folded into the new Department of Energy when that department was
created in 1977.
9 Other countries include Iraq, Libya, Syria, and Romania. https://tutorials.
nti.org/nuclear-101/uranium-enrichment/.
10 Cassoret, Bertrand. Energy Transition. CRC Press: Boca Raton, 2021, p. 67.
11 Smil, Vaclav. How the World Really Works. Viking: London, 2022, p. 142.
9

ENERGY FROM THE SUN

RENEWABLE ENERGY
Most of the world depends on fossil fuels for both electricity gen­
eration and transportation—for now. Several factors are putting
significant pressure on fossil fuel use: human contributions to global
climate change (Chapter 14), concerns about the useful lifetime of
remaining reserves, and potential geopolitical issues. For the future
we need to consider sources of energy that are not likely to be
depleted. Such sources are called renewables. Examples include
wind, water, solar, and biomass. The term renewables implies that
these sources are continuously being renewed, through natural
events, as in wind or solar energy, or through human intervention,
such as the growing of energy crops. Energy sources that are har­
vested and then regrown are also known as sustainable energy
sources. Renewables do not rely on localized concentrations in
seams or reservoirs, as do fossil fuels, but potentially can be found
and used almost anywhere. Also, the four kinds of renewable energy
mentioned above do not contribute to a net increase in atmospheric
CO2 as they are being used.
The next several chapters discuss the major renewable sources.
They seem quite different among themselves, but share several
features. First, the energy they provide is diffuse—for example,
think of how hot direct sunlight gets, even on a bright, sunny
day, as compared with how hot your oven can get. Second, two
of the renewable resources—wind and water—were exploited for
centuries prior to the burgeoning reliance on fossil fuels that
began in the 1800s. Third, we will see that wind, water, and
biomass all derive in some way from the effects of solar energy.
So, we might as well start with solar.

DOI: 10.4324/9781032665603-9
82 ENERGY: THE BASICS

The sun provides abundant energy. Surely, we ought to be able to


take advantage of it. Solar energy has two major applications: gen­
erating electricity, and providing space heat and hot water for
buildings. Today the primary interest is in using solar energy to
produce electricity. Solar energy is inexhaustible; the remaining
lifetime of the sun is about five billion years. If the sun ever ‘goes
out,’ life on Earth is doomed anyway, so no need to worry about
where we’ll get electricity. The conversion of solar energy to elec­
tricity has no climate-changing emissions. Solar generation of elec­
tricity is the application that we will consider in this chapter
because of its present and steadily growing interest. But first, we will
address the question of where the sun gets its energy.

FUSION
In the late nineteenth and early twentieth centuries, solar astron­
omers learned that hydrogen is the dominant component of the sun,
along with some helium. Both are gases at ordinary temperatures
and pressures. The English astronomer Arthur Eddington1 recog­
nized that the sun could be stable only if its interior temperature
were millions of kelvins. In the late 1930s the German-American
physicist Hans Bethe2 recognized that the conditions inside the sun
would lead to nuclear fusion—a process in which two smaller nuclei
would join together to produce one larger one.
A hypothetical fusion reaction would be 1H2 + 1H2 ! 2He4. In
this case the product nucleus is larger than the reacting nuclei. This
process fits the left-hand side of Figure 8.2, where the curve of
nuclear binding energy is much steeper than it is for the fission
processes. Both processes derive their energy from the Einstein
equation, E = mc2, introduced in Chapter 8, but the mass loss, m, is
much greater in fusion. For example, the fission of a 92U235 nucleus
results in a mass loss of 0.056 percent, whereas fusion of two 1H2
nuclei would be accompanied by a mass loss of 0.63 percent, more
than ten times larger than seen in fission.3
The mass loss associated with the nuclear processes in the sun
amounts to about four million tonnes per second. Since c2 is itself a
huge number, the energy production per second in the sun is so
large that we have no good way of comprehending it in the context
of our normal experience. Fusion of the nuclei in one gram of heavy
hydrogen (1H2) is said to produce energy equivalent to that
obtained by burning over eight tonnes of coal.4
ENERGY FROM THE SUN 83

SOLAR ENERGY
The sun radiates this energy into all directions in space. Since Earth
is a tiny ‘dot’ 150 million km from the sun, the fraction of the sun’s
energy hitting the Earth is only one two-billionth (1/2,000,000,000!) of
the sun’s total output. This extremely tiny fraction of the sun’s energy
would not seem worth bothering with. Yet, almost all the energy that
we use derives in some fashion from the sun. Growing plants require
sunlight for photosynthesis. Food, firewood, and other natural pro­
ducts depend directly on sunlight-driven photosynthesis. Fossil fuels
derive from once-living organisms that required sunlight for their
photosynthesis. Water for waterwheels or hydroelectric turbines is
supplied by the hydrological cycle, driven by the energy in sunlight.
Wind results from differences in the amount of heating of Earth’s
atmosphere by the sun. Only the energy provided by nuclear fission
reactors and the comparatively tiny amount derived from chemical
processes in batteries do not depend, directly or indirectly, on the
energy of the sun. If we could collect completely this tiny fraction of
the sun’s energy that falls on Earth, and convert it into useful work
with 100 percent efficiency, we could provide the world’s annual
energy consumption in only 90 minutes!5
Not all of the tiny fraction of the sun’s energy that comes into the
upper atmosphere actually makes its way to Earth’s surface. One-
third is reflected back into space by clouds, ice caps, and oceans.
The atmosphere absorbs another one-fourth. The remainder that
actually falls on Earth’s surface provides warmth, evaporates water
into the hydrological cycle, drives the winds, and provides the
energy for photosynthesis.
Solar energy can be used by any country or any region. But, not
all parts of Earth’s surface receive the same amounts of solar
radiation, nor does solar radiation at one particular spot remain
constant through the year. The intermittent nature of incoming
solar radiation adds a significant complication. Solar radiation is at
a maximum during daylight hours but is zero at night. In winter the
sun is lower in the sky than in the summer, so the power provided
by the sun is less than in summer; and in winter the available day­
light hours are fewer. Even in a given season at a given location,
available solar radiation can vary from day to day because of
changes in the prevailing cloud cover. The intermittent nature of
solar radiation is complicated further by the fact that neither solar
energy itself nor its two direct products, heat and electricity, can be
stored easily.
84 ENERGY: THE BASICS

From the time of the ancient Greeks, people have developed—


and still use—methods for using solar energy for warming buildings
or heating water. These techniques, if adopted more widely, could
be very helpful in reducing consumption of fossil-fuel or nuclear
energy and in reducing emissions. However, most current interest in
solar energy is due to its role in electricity generation, and particu­
larly the direct conversion of solar energy to electricity, the tech­
nology called photovoltaics.

ELECTRICITY FROM SOLAR ENERGY


We all experience the fact that solar energy provides heat. By
designing the appropriate equipment to collect and concentrate that
heat, we could boil water. If we can boil water, we can raise a head
of steam for a steam turbine/generator set. In this approach, energy
released by nuclear fusion in the sun is used to increase the thermal
energy of water or steam, which in turn provides work to a turbine/
generator, final to producing electrical energy at high potential.
Since the energy in sunlight is not converted directly to electricity,
this approach is sometimes referred to as indirect solar conversion.
All of the principles are the same as those for a fossil-fuel or nuclear
plant. The only significant difference is the source of heat: the
combustion of a fuel, uranium fission, or capturing the heat in
sunlight.
Fossil, nuclear, and indirect solar plants all use large, technologi­
cally-complicated devices to boil water and make electricity. Why
not just throw away all that stuff—boiler, turbine, generator, con­
denser, cooling tower—and convert the energy of sunlight directly
into electricity? This approach is called direct solar electricity gen­
eration, or photovoltaics.
Henri Becquerel’s father Edmond, while working in his father’s
laboratory, coated pieces of platinum with compounds of silver.
Exposing these devices to light generated an electric current. A few
individuals recognized the possibilities of this technology. A century
ago, Alexander Graham Bell, best known as the inventor of the
telephone, advocated solar energy to lessen the consequences of
burning fossil fuels.6 He was well ahead of his time, but it is inter­
esting that much of the early developmental work on silicon-based
solar cells occurred at the eponymous Bell Laboratories.
Currently a large majority of commercial photovoltaic cells start
with silicon. For a substance to conduct an electric current, the
outermost electrons of its constituent atoms must be loosely held, so
ENERGY FROM THE SUN 85

they can move when subjected to an electrical potential difference.


Electrons are tightly held in silicon, so normally it is not a good
conductor of electricity. However, some materials have a few elec­
trons capable of moving under the influence of an electrical poten­
tial, and which, therefore, conduct electricity better than an
insulator, but yet not as good as a conductor. Being somewhere
between a conductor and an insulator, these materials are called
semiconductors.
Adding a small quantity—about one atom in a million—of a
second element to pure silicon is called doping. Suppose we dope
silicon, with four electrons in its outer shell, with arsenic. Arsenic
has five electrons in its outer shell. Since few of the silicon atoms
are replaced, the material retains the crystalline nature of silicon,
but arsenic-doped silicon has some ‘extra’ electrons, thanks to the
fifth electron on atoms of arsenic. These ‘extra’ electrons can move
under the influence of a potential difference. We have created a
semiconductor by doping silicon with arsenic. Because this doping
process has provided the material with ‘extra’ electrons, this mate­
rial is called an n-type semiconductor, where the ‘n’ signals that the
semiconducting properties are due to extra, negatively charged
electrons.
Instead, some silicon atoms could be replaced with atoms of an
element that has only three electrons in its outer shell. Gallium
works well in this application. In this case, gallium-doped silicon
has some ‘missing’ electrons, because each gallium atom contributes
only three, rather than four, electrons. An electron being ‘missing’ is
equivalent to there being a ‘hole’ among the electrons in the mate­
rial. Since the electron has a negative charge, a hole conceptually
would have a positive charge. Gallium-doped silicon is also a semi­
conductor. The shortage of electrons means that the holes, not the
extra electrons, migrate when exposed to an electrical potential.
Since a hole is equivalent to a positive charge, a material such as
gallium-doped silicon is called a p-type semiconductor (‘p’ for
positive).
Putting the two different types of semiconductors together forms
an n–p junction. The separation of the charges creates a difference
in electrical potential energy (voltage) across the junction. Light
striking the photovoltaic material provides extra energy to some of
the electrons. This makes it easier for those excited electrons to
move within the material. An electron moving away from the region
it originally occupied creates a hole that is left behind. We might
expect that the extra electrons ought to flow from the n-type into
86 ENERGY: THE BASICS

the holes in the p-type. The energy needed to excite electrons and
give them mobility turns out to be about the amount of energy
available in sunlight. Joining an n-type and a p-type semiconductor,
connecting it into an electrical circuit, and allowing sunlight to fall
on it, we achieve direct conversion of solar energy to electricity.7
Such a device would be called a solar cell or photovoltaic (PV) cell.
Electrons liberated by the energy in sunlight will move toward the
front surface of the PV cell. This creates a difference in electrical
charges between the front and the back. Because electrons do not
flow as easily in a semiconductor compared with a regular electrical
conductor like copper, this current can be brought outside the cell
and be used to do work. This can be done via metal contacts on the
surfaces of the cell. As long as the cell is exposed to light, the cur­
rent will continue to flow, with the only energy source being the
solar radiation.
Single photovoltaic cells made of crystalline silicon with an n–p
junction have a theoretical maximum efficiency of 29 percent.8
(Good thing the sunlight itself is free!) Present photovoltaic tech­
nology has brought practical cells fairly close to this limit. The
power output of a solar cell varies with the intensity of light. If the
light intensity is cut in half, then the current will also drop by half.
The voltage will decrease by only a few percent. Voltage also
depends on temperature, increasing slightly above 25°C.
Modules of photovoltaic cells are connected to form panels that
are assembled into groups called arrays. A module 1.2 by 0.6 m
would have a power rating of about 250 W. An array of 20 modules
could provide about 5 kW. In typical homes one could run several
appliances with this much power. (But only during the time of
direct sunlight.)
The current and voltage produced by the array depends on how
the modules and panels are wired together. The array is connected
to a battery that has two functions. First, the battery stores elec­
tricity, so that electricity will be available when the sun is not shin­
ing. Second, the battery and its control system are set up to send
electricity into the external circuit at an even rate. This is vital,
because current could fluctuate from one moment to the next as
clouds pass overhead. With today’s technology, an array of an area
less than a square meter could provide some tens of watts of power,
enough for several small electric lights, or for operating portable
radios. With about twice that area, sufficient electricity is produced
to drive a small motor, as, for example, in a water pump.
ENERGY FROM THE SUN 87

For commercial-scale installations, the area of cells becomes a very


significant issue. For example, a nuclear plant rated at 1000 MW requires
an area of about 1 km2, while the Noor solar complex in Morocco pro­
duces half that power (510 MW) from an area of 30 km2. Nonetheless,
solar modules are easy to scale up and easy to connect into arrays to
meet almost any power requirement. A typical photovoltaic installation
for a home would require three main components: the array of solar cell
modules; a set of batteries to provide both a steady output of electricity
and a means of energy storage; and a device to transform the direct
current output of the batteries (called an inverter) into the alternating
current used by most common household appliances. It is a testament to
the progress in solar technology that, in the mid-1950s, a solar array to
operate an average home would have cost 1.5 million dollars.9 (At those
prices, that would be about 17 million dollars today.)
The alternative to building-by-building installations of solar
panels is to set up centralized photovoltaic generating stations.
Many countries now have this technology in place. The Noor com­
plex has been mentioned. A few of the other stations are in the 100–
200 MW range of capacity; many in the 50–100 MW range.
The largest market for photovoltaic cells is currently in providing
on-site generation of electricity to operate equipment or appliances
in communities far from utility grids. Before photovoltaics were
available, these devices needed to use secondary batteries, which
needed to be recharged, or small generators operated by diesel
engines, which needed periodic refueling.
Providing electrical energy for villages in less-developed countries
is very beneficial. In many cases, the cost of extending utility grids to
such locations would be prohibitive. In the most impoverished places,
utility grids may not even exist. On-site photovoltaic systems could
meet much of the need for electricity at a lower cost than trying to
expand utility grids or to build new, centralized generating facilities.
Photovoltaic systems do not use water for cooling or steam gen­
eration, so are well suited to remote or arid regions. They can
operate on any scale. Other than the cost of manufacturing photo­
voltaic cells, there is really only one fundamental problem with
solar energy—the sun doesn’t shine at night.
One answer to the storage question involves use of rechargeable
batteries. Another involves a hybrid energy system, in which a low-
cost fossil-fuel system is added to a photovoltaic system to com­
pensate for variations in sunlight. A combined solar–natural gas
hybrid would be one example. Other possibilities for storage will be
covered in Chapter 13.
88 ENERGY: THE BASICS

The cost of photovoltaic electricity depends on three principal


factors. The first is efficiency, the fraction of energy in the sunlight
striking the cells that is converted to electrical energy. The higher
the efficiency, the fewer the cells required to produce a given
amount of electricity, or the more electricity that can be generated
with a given number of cells. The second factor is the cost of pro­
ducing the cells. The third is the overall expenditure required to
install, operate, and maintain a solar energy facility. The situation
is complicated by the fact that the more efficient cells are also the
more expensive. Thus, it may prove better, for overall cost con­
siderations, to use a large number of cheap, but low-efficiency cells.
The cost of producing electricity dropped from $3 per kilowatt-
hour in 1974 to 30 cents in 1990, and has dropped to about sixteen
cents currently. This represents a tremendous improvement. In
many places, solar (and wind) can now compete with fossil-fuel
facilities, at least in terms of costs per kilowatt. Partly because of
this, solar accounted, worldwide, for over 50 percent of new gen­
erating capacity in 2022.
Solar energy is available anywhere on Earth and can be collected
easily with portable devices. This minimizes the problems of trans­
porting fuel or transmitting electricity from point of production to
point of use. Solar energy is nonpolluting and clean. It produces no
air or water pollution, nor hazardous wastes in use. Another benefit
of solar energy is that it is virtually inexhaustible. The energy source
cannot be owned and monopolized, as fossil fuels have become
because of the unequal geographic distribution of reserves.
Solar energy also has its drawbacks. Sunlight is a dilute energy
source. It’s estimated that generating all of the electricity required in
the United States using solar energy would require about 10,000
km2 of PV cells. That seems like a formidable number, but the
combined areas of all of the parking lots in the United States is
60,000 km2, with another 5,000 km2 in rooftops.10 Creative ways of
using these spaces could generate a lot of PV electricity!
A second disadvantage is the inconsistency of direct sunlight.
Clouds and seasonal changes affect the rate at which direct sunlight
can be collected. Erratic sunlight requires methods of storing the
energy collected. Back-up systems would be needed during pro­
longed periods of bad weather when stored energy might be deple­
ted. Many atmospheric effects can reduce PV-cell efficiency,
including not only thick clouds but also dust storms and smoke.
Severe hailstorms could do serious damage.
ENERGY FROM THE SUN 89

Like all the other energy sources, solar energy has its advantages
and its disadvantages. Given the growth in solar installations in the
recent past, and given its essentially zero carbon dioxide emissions
(once the PV cells have been fabricated and installed), there seems to
be significant potential for continued expansion of solar photovoltaics
in the world’s energy economy.

NOTES
1 Eddington was a very well-known and well-respected astronomer, physicist,
and mathematician of the first half of the twentieth century. He also
developed the Eddington number, E, for bicycling, to express the number of
times a cyclist has gone a certain number of miles. An Eddington number
of 50 would mean that a cyclist has gone 50 miles on 50 occasions.
2 Like Fermi, Hans Bethe emigrated to the United States to escape the fas­
cists. Once dubbed the supreme-problem solver of the twentieth century,
Bethe was publishing original research papers into his nineties.
3 Schobert, Harold H. Energy and Society, 2nd edition. CRC Press: Boca
Raton, 2014, p. 665.
4 Heimler, Charles H. and Price, Jack S. Focus on Physical Science. Charles
E. Merrill Publishing Co.: Columbus, OH, 1984, p. 506.
5 Wagman, D. Renewable Energy. PennWell Books: Tulsa, 2024, p. 43.
6 Sen, Paul. Einstein’s Fridge. Scribner: New York, 2021, p. 244.
7 However, the efficiency is not 100 percent. Some of the sun’s radiation
simply reflects off the surface, some passes all the way through the device,
and only a portion is actually absorbed and converted to electricity.
8 Blakers, Andrew, Zin, Ngwe, McIntosh, Keith, and Fong, Kean. ‘High
efficiency solar cells.’ Energy Procedia, Vol 33, pp. 1–10, 2013.
9 Gertner, Jon. The Idea Factory. The Penguin Press: New York, 2012, p. 172.
10 Magnason, Andri Snær. On Time and Water. Open Letter: Rochester, NY,
2019, pp. 303–304.
10

BIOMASS ENERGY

INTRODUCTION TO BIOMASS
Biomass refers to materials produced by life processes that can be
converted into energy or used directly as energy sources. Biomass
falls into three categories: standing forests (primarily for firewood),
energy crops (such as corn grown to produce ethanol), and wastes
(such as straw, bagasse, and even garbage). Biomass should be
renewable in the literal sense of the word, if we grow a new plant to
replace each one harvested for energy. Biomass might be said to be
carbon-neutral, because removal of CO2 by photosynthesis in the
growing of the ‘replacement’ plant should balance CO2 production
when the same kind of plant is consumed.
However, the carbon-neutral nature of biomass use relates speci­
fically to its use. There are likely CO2 emissions associated with
harvesting biomass, transporting it to its point of use, and prepar­
ing it (e.g., drying or shredding) for use. Then we must consider the
issue of time. It is possible in principle to grow a replacement plant
for each one that was harvested, but such growth could take a long
time. We can indeed grow a new corn plant next year to replace one
we harvested this year, but we can’t do that with an oak tree.
Because of the diffuse nature of renewable energy, it could take
significant land area to grow an energy crop.
For most of human history, biomass was the predominant source
of fuel, well into the mid-nineteenth century. Biomass sources
included wood, dried plants, oils extracted from plant parts, dried
animal dung, and even animal fat, all contributing to domestic
heating, lighting, and cooking needs. The rise of coal as fuel for
stationary steam engines and locomotives in the nineteenth century
and then the importance of petroleum products as transportation
fuels in the twentieth century were responsible for displacing

DOI: 10.4324/9781032665603-10
BIOMASS ENERGY 91

biomass as a major energy source. Those changes meant that a


steadily increasing share of the world’s energy was being derived
from sources that could not be replaced on any timescale mean­
ingful to humans.
Nowadays, biomass supplies about 6 percent of the world’s pri­
mary energy, and about half of the energy supplied from renewable
sources. Possibly two billion people depend almost exclusively on
biomass for energy supplies. Most of these people live in developing
nations where per capita energy consumption is low; their total
energy demand is still relatively small. Biomass supplies a sig­
nificant fraction of the energy—possibly one-third—in developing
nations but could supply even more, with improved production and
with better utilization of municipal, agricultural, or forest wastes.
Biomass fuels utilize the chemical energy fixed by photosynthesis
and stored in molecules within plants. Plants utilize the primary
products of photosynthesis, sugars in a state of relatively high che­
mical potential energy, to biosynthesize the many other compounds
needed for their life processes. Chemical energy in the compounds
created by the plant can be released in various reactions—mainly
combustion—to create heat for domestic cooking and heating, for
industrial processes, or for conversion to other forms: electricity, or
gaseous or liquid fuels. Therefore, biomass is a way of storing solar
energy for later use. Put differently, compounds produced by plants
can be considered to be ‘carbon-based energy carriers.’1
The maximum possible photosynthetic conversion of solar to
chemical energy is only about 7 percent. Plants never achieve this
ideal efficiency. Sugar cane is about 2 percent efficient in converting
available solar energy to biomass, and corn about 1 percent. More
likely, though, less than 1 percent of the solar energy falling on land
is stored as sugars and other compounds in land plants. Plants
growing in water, such as algae, come closer to the ideal maximum.
Biomass is indigenous in most parts of the world. A country that
can grow its own energy need not be fearful of countries owning
large supplies of oil. Biomass is renewable. Agricultural fields can
be used to grow new crops annually, with close attention to soil
quality. Trees can be regrown over longer periods. Biomass provides
convenient energy storage, which is not true of other renewables.
Wood, for example, can be cut and stacked. Biomass captures CO2
during growth. Production of biomass usually does not arouse
much public opposition, though there is on-going debate about the
use of good farm land for growing fuel vs. growing food.
92 ENERGY: THE BASICS

A quarter-century ago, humans used 20 percent of the annual


terrestrial net photosynthetic production (NPP)2 to meet our
demands for food (animals and crops), wood for construction or
other products, and other plant products, such as fibers, for con­
sumer goods—not for energy.3 In those days, there were at most
about 6 billion of us. At the moment this paragraph is in final
editing there are now 8,132,830,674 people.4 Demand for these
applications of biomass is increasing, and will continue to do so
until the population stabilizes. This could impact the proportion of
biomass available for production of useful energy. Raising biomass
for energy production will always compete with land devoted to
agriculture for producing human or animal food.
Using energy plant-based biomass sources, plus human or
animal muscles, to do work represents an ‘organic economy.’5
People relied on the plants growing nearby for their food, their
animals’ food, and fuel. Someone needing additional biomass had
either to obtain more land to grow more, or to adopt a more
intensive approach to the land they did have. It was not until coal
replaced biomass as the major energy source that economic
expansion and increased standards of living were achieved without
intolerable pressure on land use.
A more intensive approach to growing biomass means, ironi­
cally, relying on fossil fuels. Fossil fuels are used in making
agricultural machinery, used in operating such machinery, and
are converted to fertilizers, pesticides, and insecticides. It’s been
pointed out that we no longer eat potatoes made from solar
energy, but eat potatoes partly made from oil,6 thanks to the use
of petroleum-derived compounds to make agrochemicals. Despite
these concerns, interest in biomass is increasing. It currently
supplies about 10 percent of total British energy, and 5 percent
in the United States.7
Chapter 5 introduced the concept of energy returned on energy
invested (EROEI). Certainly, the use of any energy source makes
sense only if its EROEI is greater than 1. Otherwise, why bother?
EROEI calculations for biofuels vary widely, depending on the
specific fuel type and on what assumptions were used in making the
calculation. Some sources suggest that the EROEI for biofuels is
only slightly above 1,8 raising concern about whether, or how, to
incorporate them in a future energy economy.
There are many kinds of biomass energy sources. Here, we con­
sider just a few, as examples of what can be done with biomass.
BIOMASS ENERGY 93

WOOD
Wood was the first fuel to be used on a large scale. Until the nine­
teenth century, it was the predominant fuel in the world. Even
today, cooking with wood still accounts for roughly 60 percent of
all energy consumed in the Sahel (the semiarid region south of the
Sahara).
Trees ultimately store solar energy in the chemical energy of
bonds in the molecules comprising wood. Then, when wood is
burned, the bonds in the wood and in oxygen molecules from the
air are disrupted to form new chemical bonds of lower energy. The
new bonds are mainly in molecules of carbon dioxide and of water.
Because the system is going ‘downhill,’ energy will be released to us,
which we perceive mainly as heat.
Wood does not compare well with respect to natural gas or pet­
roleum products, in terms of convenience of handling and energy
density (i.e., megajoules of energy liberated per kilogram burned),
but for many people, it may be the only choice available. However,
for small-scale applications such as domestic heating and cooking,
wood is much easier—low-tech—to collect and burn, relative to
natural gas and petroleum. Unfortunately, small-scale domestic
appliances are very inefficient. At best, only 5 to 10 percent of the
heat liberated from a traditional open fire is used effectively.
Where lumbering is a major industry, electricity-generating plants
can burn a wood residue called ‘hog fuel’ to produce energy. The
pulp and paper industry is a leader in cogeneration processes
(Chapter 7). In well-designed, well-run combustion equipment,
wood produces lower nitrogen oxide emissions than coal or fuel oil.
Wood has a low sulfur content, so also produces less sulfur oxide
emissions. But, wood combustion at its present scale puts more
particulate matter into the air than all the world’s motor vehicles.9
When wood is heated in the absence of air (so that it won’t catch
on fire), moisture and volatile chemicals are driven off, leaving a
solid product of very high carbon content—charcoal. Charcoal is a
good-quality solid fuel, which can be transported and handled more
economically than an equivalent amount of wood. The energy
density (MJ/kg) of charcoal is at least double that of wood. In some
countries, charcoal is a widely used household fuel in urban areas.
The volatile gaseous and liquid products driven off during heating
are valuable sources of chemical products. If the gaseous and liquid
products are used as well as the charcoal, the energy obtained is
equivalent to an overall efficiency up to 80 percent.
94 ENERGY: THE BASICS

Wood reacted with steam and oxygen produces a gaseous fuel


that is mainly carbon monoxide and hydrogen. This fuel can be
burned in a gas turbine connected to an electricity generator. The
hot gases leaving the turbine can be used to produce hot water used
for heating nearby homes in a district heating scheme. A pioneering
plant in Värnamo, Sweden, produced electricity plus hot water with
80 percent efficiency.
Rapid population increase in developing nations has caused
extensive cutting of forests for fuel. Large-scale wood use also leads
to deforestation. A forest is not just trees or wood—it is an entire
ecosystem. As forests are cut, animal habitats are destroyed along
with the trees. This may reduce the habitat of birds useful in keep­
ing insects in check. Trees help hold soil in place. Loss of trees
increases water runoff and soil erosion. Soil erosion leads to loss of
valuable topsoil and to water pollution from suspended soil washed
into water supplies.
Wood requires both time and energy to be harvested. When
highly mechanized wood-harvesting equipment is used, energy
expended in cutting the wood is likely greater than the energy
available from burning it. The energy consumed in chainsaws, for
operating logging equipment, for transporting wood to its point of
use, and consumed by the workers themselves, is in total likely
greater than the energy liberated when the wood is burned. The
total energy available to society has actually decreased! EROEI is
less than 1. (Not only that, all of these various items of equipment
operate on fossil fuels, adding to carbon dioxide emissions.)
Some electricity-generating stations that burned coal exclusively
now have switched to firing wood pellets, or co-firing wood and
coal. Since the rest of the plant is unchanged, this approach means
that the electricity generated in such biomass-based plants can be
produced and dispatched to consumers just as if it had a come from
a fossil-fuel or nuclear plant. Compared to wind and solar, which
are clearly intermittent energy sources, biomass-generated electricity
can be generated on demand as needed without provisions for
energy storage systems. If nothing else, biomass could serve as a
backup for solar or wind generation.
However, there is the interesting question of where (geo­
graphically) the wood comes from. The majority of the wood used
in converted power stations in Britain is actually imported from the
United States.10 Shipping wood across the Atlantic in vessels pro­
pelled by marine diesel engines, contributes enormous CO2 emis­
sions even before a single pellet of wood was burned.11 Nonetheless,
BIOMASS ENERGY 95

co-firing biomass with coal seems to be coming in countries such as


India and Indonesia.
Though wood is classified as ‘renewable,’ it may take 20 to 100
years to grow a new tree to replace a mature one that has been
harvested. Without strict forest management, we can’t regrow trees
as fast as we cut them. Wood is renewable if we assume that a new
tree is planted for each one cut down and that we measure time in
units of decades. It is obviously not true that wood is a renewable
resource in the short term, because a seedling one or two years old
is not a direct replacement for a mature tree.
There are practices that allow trees to be used and renewed on
relatively short timeframes. Beavers figured this out millennia ago.
They manage the harvesting of aspens and willows from different
places around their ponds, so that regrowth allows them to inhabit
the same pond apparently indefinitely.12 Some species, such as
poplars, can be used as energy crops by the practice of coppicing.
Year-old trees are cut near the ground. The cut stumps sprout about
a half-dozen new shoots, each of which grows rapidly over the next
three to five years. The tree shoots could be harvested every two to
eight years. Since new shoots grow from the cut stumps, replanting
is required only once for every three or four times that the wood is
harvested.

ETHANOL
Liquid fuels are the overwhelming choice as the energy source for
practical internal-combustion engines in vehicles. Solid fuels are not
well suited for Otto or diesel engines. Ash which builds up in the
cylinders erodes metal surfaces, and adds to air pollution. Gases
can be superb fuels for internal combustion engines. But, unless the
gas can be liquefied or compressed to high pressure, it is difficult to
carry enough fuel on a small vehicle to obtain much practical
driving range. On large vehicles that travel short distances, with
access to specialized fueling facilities (e.g., city buses), compressed
natural gas is a fine fuel choice. Liquid fuels provide clean-burning
characteristics similar to gases, but with an energy density that
allows an ample fuel supply to be carried on the vehicle.
Alcohol fuels can be burned in existing spark-ignition engines
with minimal modification. An enormous infrastructure already
exists for the handling, storage, and transporting of liquid fuels (e.
g., gasoline). Conversion from petroleum to alcohol fuels could be
effected with little change in the transportation-fuel infrastructure.
96 ENERGY: THE BASICS

Improvements in the conversion of biomass to alcohols, and possi­


ble introduction of less expensive crops for making alcohols, might
expand the transportation market for these fuels.
A very large family of alcohols exists, but when the word alcohol
is used by itself (as in the term ‘alcoholic beverages,’ for example),
almost always it is ethanol that is meant. Ethanol is also called
ethyl alcohol or grain alcohol. As the name ‘grain alcohol’ implies,
a major source is the fermentation of starches or sugars from
plants, such as corn and other grains.
Ethanol can be derived from renewable biomass, including sur­
plus crops, as well as from waste materials such as sugar beet or
sugar cane pulp and potato peelings. Fermentation converts plant
materials to ethanol, likely the second-oldest chemical process to be
exploited by humans (the oldest being combustion). Virtually any
biomass material that has a high proportion of starches or sugars
could, in principle, be fermented to produce ethanol. During fer­
mentation, starches or simple sugars from biomass are decomposed
by yeasts or other microorganisms to produce ethanol. About 85
percent of the chemical potential energy in the molecule of sugar is
retained in the two molecules of ethanol produced from each
molecule of sugar. Sources of starch are grain crops and some root
plants, such as potatoes. Sugar sources include cane and beet sugar,
and sorghum. The ethanol business in the United States nowadays
is based on corn, whereas other grains and sugar cane are used
elsewhere. The high starch content of corn makes fermentation
relatively easy.
Ethanol can be made in high tonnage by known technologies. It
has a high octane number (Chapter 5), about 110, making it well
suited to high-compression-ratio spark-ignition engines. Ethanol
burns cleanly, with reduced nitrogen oxide emissions compared to
gasoline. A 90:10 blend of gasoline and ethanol can be run in
standard automobile engines with no modifications. In many parts
of the United States service-station pumps dispensing regular-grade
gasoline of 87 octane state that the fuel contains up to 10 percent
ethanol. Pure ethanol has a heating value of 28 MJ/L, about three-
fourths of the value for gasoline. A car using 10 liters per 100 km
with gasoline would use 13 with pure ethanol. Ethanol readily
mixes with water, so care must be taken to ensure that ethanol is
not adulterated by water during handling or storage.
Raising corn for ethanol production is very energy-intensive—
plowing, fertilizing, harvesting, and transportation all require
energy. Raising corn also contributes to the erosion of agricultural
BIOMASS ENERGY 97

land; topsoil washed away eventually contributes to water pollution.


The distillation step required to separate ethanol from the water in
which it was produced requires large amounts of heat, generally
supplied from petroleum or natural gas. Fertilizers used in corn
production are made from petroleum-derived chemicals. A break­
through in chemical engineering that would replace distillation by
some other operation for recovering ethanol from the fermentation
mixture would be helpful, as would further progress in alternative
processes that don’t even require distillation. The EROEI for etha­
nol derived from corn is just slightly greater than 1.
Growing crops for energy removes land, or the crops grown on it,
from food production. With a large percentage of the world’s popula­
tion malnourished or even starving, it may seem not to make sense to
divert lots of agricultural land to the production of ‘energy crops.’
Plants make the sugar called glucose as the primary product of photo­
synthesis. Glucose is stored in two ways inside the plant: as starch, for
long-term energy storage, and as cellulose, to provide mechanical
strength to the plant. We can digest starch, but not cellulose. Using
cellulose as the starting material for ethanol production would have a
major impact on the argument of using crops such as corn for food vs.
for fuel. Ethanol produced from cellulose is sometimes called cellulosic
ethanol. It is no different chemically from ethanol from any other
source. The adjective cellulosic signals that the original biomass was a
cellulose source, and not starch. A challenge remains to find easy, quick
chemical treatments to convert cellulose back to glucose. A break­
through in that area, permitting large-scale production of cellulosic
ethanol instead of starch-based ethanol, would be a very significant
boost for ethanol. At present cellulosic ethanol appears to be uneco­
nomical in comparison with gasoline and ethanol made from grains.
By 2011, Brazil was making nearly 21 billion liters of ethanol per
year from sugar cane, the most successful program for biofuels
anywhere in the world. The Brazilian car industry has developed
flex-fuel vehicles, capable of operating on gasoline, E22, and pure
ethanol (E100) without modification. The great success of the Bra­
zilian ethanol program—demand remains strong—comes from
numerous factors. The feedstock is sugar cane, in which the juice
already contains simple sugars—and so no conversion of starch or
cellulose is needed. Sugar cane grows readily, in some places two
crops per year, with relatively ‘low-tech’ agriculture. Bagasse, the resi­
due from squeezing the juice out of cane, is burned on-site as the fuel
source for the process heat needed in fermentation and distillation. In
a sense, this energy source is free.
98 ENERGY: THE BASICS

WASTES
We will never run short of waste products. Many materials now
routinely thrown away are combustible and could be used as fuels.
This aggregate of waste material has been given various names, such
as municipal solid waste (MSW), eco-fuel, or refuse-derived fuel.
In the United States, burning such wastes—in waste-to-energy
(WTE) plants—provides about 14,000 GWh of electrical energy
annually, from about 60 plants.13 Such plants can count on a never-
ending source of fuel, which is usually inexpensive. About 90 per­
cent of the plant output is electricity. The rest is mostly steam that
could be used for district heating.
Many biological materials, such as animal manure, produce
methane. It occurs naturally in a landfill, or by intent in equipment
designed for the purpose. The reaction is facilitated by anaerobic
bacteria that do not use oxygen for their life processes. The fuel
produced by the anaerobic digestion of wastes is sometimes called
biogas; natural conversion of organic wastes in a landfill produces
landfill gas. The important component is methane, an excellent fuel
that is also the main component of natural gas.

THE TIMESCALE OF RENEWABLES


The biofuels discussed here, along with many others, are renewable
energy sources, though one must be careful to define the timescale
on which they are renewable. Trees are not renewable from one year
to the next. On the other hand, as long as there are humans, there is
certain to be garbage and waste that could go to biogas digestors.
All of the biofuels can be used in combustion equipment, but, as
cautioned with other fuels, the same device will not work equally
well on different kinds of biofuels. Each biofuel has its own advan­
tages and disadvantages. Some could be sufficiently low-tech as to
be attractive for places that lack the financial resources for expen­
sive, precision-engineered equipment or for building extensive
energy distribution infrastructures.

NOTES
1 Moore, Andrew. The Decarbonization Delusion. Oxford University Press:
New York, 2023, p. 176.
2 This is the difference between all of the carbon that’s incorporated into
plants minus that portion that is consumed by the plant for its own needs.
BIOMASS ENERGY 99

3 Wyn Jones, R. Gareth. Energy: The Great Driver. University of Wales


Press: Cardiff, 2019, p. 123.
4 https://www.theworldcounts.com/challenges/planet-earth/state-of-the-pla
net/world-population-clock-live. Accessed September 25, 2024.
5 Wrigley, E.A. Energy and the English Industrial Revolution. Cambridge
University Press: Cambridge, 2010, p. 111.
6 Odum, Howard T. Environment, Power, and Society. Wiley-Interscience:
New York, 1971, p. 116.
7 Wagman, David. Renewable Energy. PennWell Books: Tulsa, 2024, p. 102.
8 Cassoret, Bertrand. Energy Transition. CRC Press: Boca Raton, 2021, p. 79.
9 Wohlleben, Peter. The Heartbeat of Trees. Greystone Books: Vancouver,
2021, p. 159.
10 Wagman, David. Renewable Energy. PennWell Books: Tulsa, 2024, p. 102.
11 The largest marine diesel engines, as used in container ships, consume
immense quantities of fuel, on the order of 4 liters per second. Gray, Theodore.
Engines. Black Dog & Leventhal Publishers: New York, 2022, p. 67.
12 Pastor, John. What Should a Clever Moose Eat? Island Press: Washington,
2016, pp. 50–51.
13 Wagman, David. Renewable Energy. PennWell Books: Tulsa, 2024, p. 105.
11

ELECTRICITY FROM FALLING


WATER

WATER TURBINES
Michael Faraday’s discovery of electromagnetic induction (Chapter
6) provides the essence of large-scale electricity generation: find a
good way of getting a shaft to rotate reliably at low cost, and con­
nect a generator to it. This requires a source of reliable and inex­
pensive energy. Waterwheels have been known since antiquity as a
source of cheap, reliable rotary motion, and could be used to oper­
ate a generator. In such a system, water flows spontaneously from
high to low gravitational potential energy producing work that
allows the generator to effect a non-spontaneous movement of
electrical charge from low to high electrical potential energy—from
low voltage to high voltage.
Medieval waterwheels were not very efficient, because water
leaving the wheel still had significant kinetic energy. The evolution
of more efficient waterwheels began with Leonardo da Vinci, who
recognized that water flowing through curved blades, rather than
the flat ones normally used in waterwheels, bears against those
blades with more force or impulse. Such a device nowadays is gen­
erally known as a turbine. The principle is as with the turbines we
have met previously: a turbine converts the kinetic energy of a
moving fluid into rotating kinetic energy. In the context of this
chapter, the working fluid is water. When the kinetic energy of water
is used to drive the turbine/generator set, the technology is referred
to as hydroelectricity, hydropower, or often just ‘hydro.’ The term
‘hydroelectricity’ does not imply that the electricity is somehow dif­
ferent from electricity generated by other means. The prefix ‘hydro-’
simply reminds us that water is the working fluid in the turbine.
In water turbines, water flows through pipes or ducts over curved
vanes. The turbine must be enclosed in a casing to provide the

DOI: 10.4324/9781032665603-11
ELECTRICITY FROM FALLING WATER 101

impulse to the vanes. The higher the pressure of water supplied to


the turbine—the higher the column of water, or its ‘head’— the
faster the turbine will spin. The rotational speed of a traditional
waterwheel is limited by the speed of the water. This is not so with
turbines, which attain high rotational speeds with appropriate
design. The development of the turbine was a significant advance
for providing a source of work for an electrical generator.
The essential feature of the turbine is the flow of the water over
curved vanes. Energy is captured by the impulses—forces acting
over very brief times—between the fast-flowing water and the pas­
sages created by the curvature of the vanes. These designs will work
only if water is confined, so the vanes are enclosed in a closely fit­
ting casing. Such a turbine can convert more than 90 percent of the
kinetic energy of the water into rotary motion, at high rates of
revolution. The power output of a plant is determined by the
quantity of water flowing and by its potential drop, or ‘head.’ (This
is similar to the dependence of electrical power on the amount of
electricity flowing—the current—and on its potential—voltage.)
Water coming into a hydro plant does not need to drop exactly
vertically, only from higher to lower gravitational potential energy.
Where does the water come from? It results from the hydrologic
cycle, described in Chapter 3. It has operated for eons, driven by the
energy of sunlight providing enough energy to allow water mole­
cules to leave the liquid state and enter the atmosphere as vapor.
Because the hydrologic cycle is driven by solar energy, our ability to
put hydro to work for us derives from the sun. Of course, the
available amount of water varies from place to place on Earth.
Even in one spot, the amount could be affected by normal seasonal
changes through the year, or by longer-term climate changes, such
as droughts.

HYDROELECTRIC PLANTS
It’s aptly said that hydro is ‘an old form of renewable energy.’1 In
the United States, commercial hydro plants opened in 1882. The
Whiting plant in Portage County, Wisconsin, first began operation
in 1891 and is still running. The scale of operation has changed by
four orders of magnitude since then—the Whiting plant generates 5
MW; the immense Three Gorges plant on the Yangtze River in
Hubei Province, China, is rated at 22,500 MW. However, the fun­
damental principles have not changed. A good engineer from the
1890s, somehow returned to life and wandering into a modern
102 ENERGY: THE BASICS

hydro plant, would be stunned by the electronic sensing, control,


and computing systems, but would immediately comprehend the
way in which the plant transmutes the kinetic energy of water to
electrical energy going to customers.
The hydroelectric plant at Fully, Switzerland, illustrates some of the
features of a modern installation. Water is led from a lake through a
pipe almost 5 km long that drops nearly 2 km elevation. This very
steep gradient results in a water pressure, at the lower end of the pipe,
of 14 MPa. Water leaves the nozzle at a speed of nearly 650 km per
hour.2 The turbine wheels are 3.5 m in diameter; given the water
pressure and velocity, they turn at 500 revolutions per minute.
The world’s largest hydroelectric plant, the Three Gorges plant, is
one of the largest engineering projects of any kind ever undertaken
anywhere, if not the largest. It has an intended generating capacity
of 22,500 MW. The output is some 20 times larger than that of a
typical large coal-fired or nuclear electrical plant.
But regardless of size, at the heart of any hydro plant is the same
principle: the conversion of the kinetic energy of water to electricity.
That principle provides the possibility of taking advantage of the
water flow in numberless small streams, or even in water flowing
through public water systems. Small-scale hydroelectric plants have
various names: less than 10 MW are called small; less than 1 MW,
mini-hydro; and less than 100 kW, micro-hydro. Small hydro plants
feeding electricity into a regional or national grid represent a source
of relatively inexpensive, renewable energy. Installation of mini-
hydro plants could be an economically attractive alternative to
expanding nationwide electricity grids that connect a few large,
expensive electricity-generating stations.
Micro-hydro projects provide an output sufficient for a small
community or small commercial installation. The smallest micro-
hydro units would serve one, or a few, homes. Micro-hydro units
can be set up to generate electricity in regions that have no con­
nections to an electric grid, or used by those who choose to dis­
connect from the electric grid. Vendors in several countries now
offer virtually complete, commercial packages, many of which can
be installed as a do-it-yourself project. Although the output of each
individual microturbine is small, they can be deployed in many
thousands of locations, collectively providing a significant amount
of electricity.
Worldwide, hydropower provides about 16 percent of electricity
demand, equivalent to about 4 percent of total energy demand.
China leads the world in hydropower capacity, followed by Brazil,
ELECTRICITY FROM FALLING WATER 103

the United States, and Canada. Hydro has been the main con­
tributor to electricity generation in Canada for over a century.
Hydro, along with tidal energy (discussed later), provide about 60
percent of Canadian electricity. In 2022, Paraguay’s electricity was
generated completely from renewables, mainly from hydro.3 Elec­
tricity in Ecuador and in Norway is also mainly derived from
hydro. At least 60 countries obtain more than 50 percent of their
electricity from hydropower.

WHY GO HYDRO?
Hydropower has a number of advantages. Hydropower is renew­
able; that is, probably not going to be depleted on any human
timescale. There has seemed to be little likelihood that rivers will
dry up, although nowadays that is not so certain, given the effects
of climate change combined with insatiable demands for water.
Parts of such major rivers as the Yangtze, the Rhine, and the Col­
orado are becoming extremely low. Hydropower offers high effi­
ciency, between 80 and 90 percent, as compared, for example, with
30–35 percent for electricity generation in coal-fired plants. Unlike
fossil-fuel or nuclear plants, there are no waste products and vir­
tually no pollution, including no carbon dioxide emissions (from
the operation). The energy source, water, is free—there is no recur­
ring cost to buy fuel.4 From the plant operator’s perspective, an
important advantage of hydropower is that it can easily be ramped
up and down, or even switched on and off entirely (by diverting
water onto or away from the turbines), making it an excellent
choice for providing extra electricity at times of peak demand or in
energy emergencies.
In most places, hydropower has always been one of the least
expensive sources of electricity. Though the initial capital invest­
ment required to build a new hydropower installation can be very
high, once the plant is built and running, the recurring costs for
operations and for maintenance are usually quite low. The only
moving parts are the turbine and the generator.
Electricity generation is also judged on the basis of the capacity
factor, which indicates the fraction of the time that a plant is really
producing electricity. We will see this again in Chapter 12. Hydro
has a capacity factor (about 38 percent) slightly greater than wind,
and better than solar PV.5 However, fossil-fuel-fired plants often are
around 50 percent, and nuclear plants can get up to 90 percent.
104 ENERGY: THE BASICS

Only a fraction, perhaps as little as one-tenth, of potential


hydropower capacity has been developed.6 While this may sound
very attractive and promise a great future for hydro, many potential
sites for new development are places that are inaccessible, far from
potential consumers and markets for electricity, and often both. In
many developed countries, there is growing public concern about
the environmental effects of constructing dams to provide hydro­
electric sites. In some nations, construction of dams and their large
reservoirs could cause many people to have to be displaced from
their homes. (As discussed below, dams can have negative impacts
on some species of endangered fish and other wildlife.) Considering
these issues, it is likely that no more than about half of the world’s
potential hydro capacity will ever be developed.
Actually, the total hydro capacity could drop in the future. There
is already increasingly strenuous competition for water resources,
with needs for drinking water and agricultural irrigation contesting
with its use in hydro plants. Such competition will only intensify in
the future. This makes a fierce, potentially explosive, issue where a
major river flows through more than one country, and questions of
water rights turn into international squabbles. Some serious obser­
vers believe that wars over water will replace wars about oil as a
major source of conflict in the world—water will be the ‘oil’ of the
twenty-first century. As global climate change results in climate
shifts to reduce rainfall, water resources could become even scarcer.
A decrease in hydro capacity due to climate change would be a
sadly ironic situation, because hydroelectricity generation is one of
the technologies that does not produce emissions thought to con­
tribute to climate change in the first place.

THE DAM PROBLEM


Not many rivers have the fast currents or significant waterfalls needed
to provide an adequate head for a turbine. If so, we can create an
artificial head by building a dam. Water in the reservoir upstream
flows over the dam to create a useful head. Dams provide several
benefits in addition to creating a source for hydropower generation.
They control seasonal fluctuations in water flow, helping navigation
and downstream flood control. The newly created artificial lake can be
used as for recreation, or the water in it can be used for agricultural
irrigation. However, barring major changes in public perception and
public policy, it remains questionable whether hydropower will expand
into a significantly larger share of the world’s electricity generation.
ELECTRICITY FROM FALLING WATER 105

Decades ago, there was little opposition to the construction of


new dams. Hundreds were built in the early part of the twentieth
century in many countries of the world. They were, in many cases,
symbols of technological progress and of economic prosperity. In
the 1930s and 1940s dam-building projects were seen as a way of
bringing electricity to many rural communities and farms, and the
availability of cheap electricity would foster the growth of industries
and create jobs. This was especially important during the Great
Depression of the 1930s. During those decades, coal was the domi­
nant energy source for electricity generation; hydro became known
as ‘white coal.’7
However, there are numerous counterarguments. The artificial lake
floods valuable farmland, or pristine wilderness. Damming of a river
or stream has a major impact on the local ecosystem. Dams create a
reservoir where a flowing river used to exist. Aquatic plants and
animals best adapted to free-flowing water may not flourish in the
stagnant conditions of the reservoir. New species, adapted to still or
sluggish water, may move in. The reservoir can serve as a breeding
ground for mosquitoes. Dams can block fish migration and destroy
spawning grounds, affecting commercial and sport fishing. In some
places there is a growing effort to dismantle hydroelectric plants to
restore trout and salmon to the streams. (It’s said that, ‘Many a tur­
bine turns expensive hatchery-raised fish into sushi the moment they
start their journey to the sea.’8) Depending on where the dam is
built, reservoirs can flood large areas of forest and wildlife habitat,
even flood valuable farmland or towns. Sediments that may have
vital nutrients accumulate in the reservoir instead of being carried
downstream, impacting downstream ecosystems. That situation is
further aggravated because the water flowing out of the hydro plant
may be warmer, have lower oxygen content, or both, compared with
normal river conditions.9 The reservoir usually raises the water table
behind the dam (as a result of seepage) and lowers it below the dam.
Evaporation from the reservoir increases the amount of dissolved salt
and minerals in the water.
Dams do not last forever. Silt can accumulate behind the dam,
eventually filling in the reservoir. If the dam fails, there is great
potential for a catastrophic accident with large loss of life and
property damage. For example, a dam failure in the Hunan pro­
vince of China in 1975 cost the lives of 20,000 people. Dams can be
vulnerable to acts of war or terrorism. The most spectacular exam­
ple was the 1943 attack of the Royal Air Force on three dams on
the Ruhr River, in the heartland of Germany. Two of the dams were
106 ENERGY: THE BASICS

breached, with an estimated death toll of about 1,600. It’s been


estimated that ‘several tens of thousands of people’ have been killed
in the past century by dam failures, very much more than the
number of deaths caused by nuclear accidents.10
The Three Gorges Dam required the relocation of at least
1,300,000 people. The dam led to a loss of antiquities, and a need
to salvage some of the priceless heritage of ancient China. A parti­
cular concern is the fate of endangered species such as the Chinese
river dolphin and the Siberian crane. The dam is built in a seismi­
cally active region. If a major earthquake should strike, especially
when the reservoir is holding a huge weight of water pressing
against the dam, perhaps—we can only hope not—the dam would
rupture. The resulting flood could be a catastrophe of unimaginable
proportion.
Since most environmental problems associated with dam building
derive from the creation of the reservoir, ideally the amount of
electricity generated should be maximized while minimizing the size
of the reservoir. This is measured in terms of an indicator number,
the power produced per hectare (ha) of reservoir surface. The larger
this number, the smaller the likelihood of a serious impact on the
environment; the smaller the number, the worse the problems. The
Balbina hydroelectric plant in Brazil is a major environmental dis­
aster; only about 1 kW is generated per hectare of reservoir surface.
Another plant in the same country has an indicator value of 588
kW/ha, a substantial generation of electricity for a relatively small
area covered by the reservoir.

TIDAL ENERGY
Rivers make their way into oceans, and there we have another pos­
sible source of hydroelectricity, though scarcely exploited at present.
Anyone who has spent time by the seaside has experienced the fact
that tides rise and fall. From a different perspective, they’re simply a
movement of water, and we could extract useful work from that
form of moving water.
With the changing tide, water will flow into, or out of, bays or
rivers that empty into the sea. If the river or bay could be closed by
a dam built near its mouth, water flowing in and out could be used
to generate electricity. Also, ocean currents and waves, which result
from wind and from temperature gradients in the ocean, represent
another enormous source of moving water. Waves in particular
arise from wind blowing over the surface of the ocean (we will see
ELECTRICITY FROM FALLING WATER 107

in Chapter 12 where the wind gets it energy). Waves and currents could
be used for generating electricity. An estimate made some 60 years ago
suggests that the ‘energy of the tides in continuously being dissipated at
a rate [i.e., power] whose order of magnitude is a billion horsepower,’11
i.e., about 750 GW. This amount of power is generated by the com­
bined output of about one-third of the world’s coal-fired plants. A more
recent view is that the billion-horsepower figure is low.12
The fact that the ocean is in constant motion means that we
ought to be able to tap the associated kinetic energy to make
renewable electricity. The job won’t be easy. Consider some
common observations: rivers tend to flow one way, oceans don’t;
they surge, heave, and swirl. Capturing the kinetic energy will not
be as simple as doing it from a river. Oceans are salt water. For
some kinds of metals, salt water is far more corrosive than fresh
water. Oceans can transport finely suspended sand, which can
wreak havoc passing through mechanisms.
Compared to many free-flowing rivers or dams, oceans have a
relatively low head. Since power generation depends on head and
volume, a low head means that large volumes of water have to be
processed. (Of course, oceans are large volumes of water.) Often,
the head is inconsistent. Tides are generally available for only a
portion of the day. A tidal power installation would need to be
designed to extract work from the rising and falling tides, which
flow in different directions. Despite these complications, compared
to terrestrial water sources ‘there are no dry years for a tidal
installation.’13

SOME BENEFITS AND COMPLICATIONS


Let’s consider three additional points. Those of us fortunate enough
to live in countries with abundant, never-failing (almost) electricity
likely overlook the extraordinary freedom that electricity gives us.
The early days of exploiting the energy of flowing water—up to the
late nineteenth century—gave rise to some remarkable waterwheels
and ways to connect them to the devices they were intended to
drive. But, any water-driven facility necessarily had to be located
directly on the river. Mechanical work provided by a waterwheel
had to be used on the spot. Electricity freed us from this. The
hydroturbines needed water, but the users of the electricity—towns
or factories for example—could be many kilometers away. Elec­
tricity broke the geographical linkage between the waterwheel and
the machine it drove.
108 ENERGY: THE BASICS

We are seeing nowadays an increasing concern for reducing


carbon dioxide emissions from human sources, including elec­
tricity generation. Some people even advocate a complete ‘dec­
arbonization’ of the energy sector. Hydropower could be a great
candidate for contributing to such goals. But in addition to the
concerns raised earlier, there’s another: time. A dam can’t be
built quickly. The first tidal power installation, on the La Rance
estuary in France, took six years to build.14 These days, figure
on a decade. The Three Gorges Dam took 13 years from the
start of work until construction was completed. The Grand
Coulee Dam, the largest hydropower source in the United States,
took nine. The impact on carbon emissions is going to be a long
time coming.
Third, dams are made of concrete. The Grand Coulee Dam
required about 22 million tonnes. The cement necessary for concrete
is made by first heating limestone (calcium carbonate) to make lime
(calcium oxide). The breakdown of the limestone during heating
liberates more CO2, about 28 percent of the weight of the limestone.
Heating is usually done using fossil fuels, producing abundant
carbon dioxide from their combustion.
There are no simple answers.

NOTES
1 Smil, Vaclav. How the World Really Works. Viking: London, 2022, p. 216.
2 Water has extraordinary properties under these conditions; a water jet
coming from a 7.5-cm pipe at a pressure of one-quarter of the value in the
Fully plant—3.5 MPa—cannot be cut through or deflected even by hitting
it with a steel bar.
3 David Wagman. Renewable Energy. PennWell Books: Tulsa, 2024, p. 85.
4 Compare free ‘fuel’ with the situation of a coal-fired plant burning, say,
10,000 tonnes of coal per day at 50 dollars per tonne.
5 Rhodes, Richard. Energy. Simon and Schuster: New York, 2018, pp. 330–331.
6 Heimler, Charles H. and Price, Jack S. Focus on Physical Science. Charles
E. Merrill Publishing Co.: Columbus, OH, 1984, p. 500.
7 Hawks, Ellison. Popular Science Mechanical Encyclopedia. Popular Science
Publishing Co.: New York, 1941, p. 144.
8 Peter Wohlleben. The Secret Network of Nature. Vintage: London, 2017, p. 29.
9 Why? First, no process is 100 percent efficient; the energy not converted to
work usually appears as heat. Second, the solubility of gases in liquids (e.g.,
oxygen in water) decreases as the liquid temperature increases.
10 Cassoret, Bertrand. Energy Transition. CRC Press: Boca Raton, 2021, p. 67.
11 Clancy, Edward P. The Tides: Pulse of the Earth. Anchor Books: Garden
City, NY, 1968, p. 133.
ELECTRICITY FROM FALLING WATER 109

12 Nahin, Paul J. In Praise of Simple Physics. Princeton University Press:


Princeton, 2016, p. 51.
13 Clancy, Edward P. The Tides: Pulse of the Earth. Anchor Books: Garden
City, NY, 1968, p. 147.
14 Clancy, Edward P. The Tides: Pulse of the Earth. Anchor Books: Garden
City, NY, 1968, pp. 140, 142.
12

ELECTRICITY FROM WIND

WHERE WIND COMES FROM


Winds are generated by uneven heating of the Earth and its atmo­
sphere by the sun. The Earth is generally warmest at the equator. As
the air there is heated, it expands. (This is a universal property of
gases—increasing the temperature while the pressure is constant
increases volume.) Because the same mass of air now occupies a larger
volume, its density has decreased. Because the density has decreased,
the air will rise. This expansion and consequent reduction of density is
the science behind the bit of folk wisdom that ‘hot air rises.’
Two effects follow. First, cooler air from elsewhere will flow into
the region, to replace the heated air that has moved higher in the
atmosphere. This flow of air is felt as a wind. Second, the heated air
will travel at high altitudes until it cools. Then, its volume decreases
and it becomes denser, sinking to lower elevations. These effects
establish a gigantic ‘conveyor belt’ of air movement. These large
circulation patterns owe their origin to the differential heating of
Earth by the sun. Wind energy is an indirect form of solar energy.
Here again we have a matter (in this case, air) in motion, which
gives us the opportunity to extract useful energy as work.
On a local scale, this simplistic model becomes much more com­
plex. Each component of the local landscape—soil, rocks, plants,
and water—has different abilities to absorb solar energy or to
reradiate it back into the atmosphere. Each experiences a different
degree of heating. Different landscapes will have different features
of topography: some may be flat, others hilly or mountainous. Some
may incorporate large bodies of water. Urban landscapes may be
marked by numerous tall buildings. As a result, the frequency,
direction, and speed of winds vary greatly from one location to
another.

DOI: 10.4324/9781032665603-12
ELECTRICITY FROM WIND 111

The rise of the steam engine diminished the industrially impor­


tant role of wind, but wind never entirely stopped contributing to
energy use. Windmills remained in use well into the twentieth cen­
tury, especially in rural regions throughout the world. Windmills
provided energy for chores such as pumping water for livestock or
for irrigation. By the late nineteenth century, the technology of
generating electricity was understood: the need to rotate the shaft of
a generator cheaply and reliably. Using windmills to produce elec­
tricity dates from the late 1880s. Applications of wind energy for
local or on-site electricity generation were effectively ended by the
spread of electrical grids bringing electricity from central generating
plants, possibly tens or hundreds of kilometers away. As with hydro,
electricity from the grid breaks the geographic constraint requiring
the user of energy to be very close to the generation of the energy.
Denmark has been a leader in applications of wind technology.
The easy availability of cheap oil after the Second World War
eliminated European interest in using wind to produce electricity,
but the oil price shocks of the 1970s promptly revived it. The
Danish response provided the impetus for the revitalization of the
wind industry. The Danish government actions included a 30 per­
cent subsidy on the purchase of wind machines, and a requirement
that utilities purchase the electricity that these devices produced.

WIND TURBINES
Since wind represents a fluid (air) in motion, and has kinetic energy,
it can also be used in a turbine. We use the term wind turbine to
refer to machines that use the energy of wind specifically to supply
work to an electricity generator. The term windmill is usually
restricted to devices that use the wind for other applications of
mechanical work, such as pumping water or making flour.
Wind turbines can be configured such that the axis of the wheel is
either vertical or horizontal. The horizontal-axis design is the more
common. In a horizontal axis machine, the turbine blades look
somewhat like the propeller on an airplane. The shaft on which the
turbine blades are mounted enters a housing called a nacelle. (This
term generically refers to a streamlined enclosure that is intended to
protect something, such as aircraft landing gear.) The nacelle pro­
tects the generator, any necessary gears, and the ‘back end’ of the
turbine shaft. Conversion of the kinetic energy of the wind into
rotary mechanical work of the shaft results from a difference in air
pressure across the two sides of the blade. This pressure difference
112 ENERGY: THE BASICS

creates lift and drag forces, somewhat like an airplane wing. Since
we prefer that the turbine not fly, stronger lift force causes the rotor
to spin, and it turns the generator. The electricity that is generated
is conducted down to ground level by a cable.
The scale on which the wind turbine-generator can be built varies
greatly, depending on the intended use. The general size ranges are
small machines for individual farms or households not connected to
the grid, and larger machines generating electricity for distribution
into the grid by an electric utility. Smaller units have generating
capacities typically in the range 0.5 to about 50 kW. The larger
machines now range up to about 15 MW. A wind turbine 30 m in
diameter, operating in a wind of 30 km/h (this is strong enough to
make large branches or small trees sway) would generate about 200
kW, an amount of power sufficient for about eight to ten modern
homes with their usual array of electrical goods.1 The larger the
power rating of an individual unit, the fewer would be needed to
achieve a desired total generation.
Current designs of wind turbines can operate over a range of
wind speeds, from 10 km/h to about 80 km/h. The theoretical
maximum efficiency of converting wind energy to work, the Betz
limit, is 59 percent. (It has been pointed out that this is the Betz we
can do.2) Just as real efficiencies of actual heat engines are lower
than the Carnot efficiency, the efficiency of actual wind turbines
doesn’t reach the Betz limit, but is about 50 percent.3 Achieving a
100 percent efficiency would require that the airflow downstream of
the turbine would have zero velocity, obviously impossible because
some of the air has to keep moving. Conversion of mechanical
work to electrical energy in the generator is typically above 90
percent.
The power in wind is related to the cube of its speed. For exam­
ple, if the wind speed doubles from a moderate breeze of 25 km/h,
which is enough to move small branches, to a very strong breeze of
50 km/h, that will keep even large branches in continual motion, the
power of that wind increases by a factor of eight.4
Generally, wind speeds increase somewhat with increasing height
above the ground. At higher altitudes, there are fewer obstacles to
air flow, such as tall trees or buildings, or even hills. So, the evolu­
tion of wind turbine designs is to go taller and taller, now about 120
m. We’ve reached a point where it can be said that ‘a windmill is
basically a generator on a stick.’5
The potential of wind energy is extraordinary. Worldwide, total
wind flow is ample to supply several hundred times the world’s total
ELECTRICITY FROM WIND 113

annual energy consumption. In the United States, the total amount


of kinetic energy dissipated by winds across the country is 40 times
larger than the annual energy consumption. It could—in theory—
be possible to satisfy all of the energy needs of the United States by
extracting just one-fortieth of the energy from the wind. Forty years
ago, it was predicted that wind energy could ‘someday’ provide
more than 13 percent of the United States’ electricity.6 In 2023
production almost got there, hitting 12 percent.7
The capacity factor, introduced in Chapter 11, is the amount of
electricity actually generated to the amount that could have been
produced running continuously at full power, compared over a spe­
cified time period. The ratio is expressed as a percentage. For
example, plant that generated 100 million kWh in a year, but had a
maximum capacity of 200 million kWh, would have a 50 percent
capacity. The capacity factors for wind farms are about 35 percent.8
In comparison, nuclear plants have capacity factors above 90 per­
cent and coal about 48 percent. So, to replace a 500 MW nuclear
plant a utility would need to install over 1250 MW of wind capa­
city. It could be argued that this issue is counterbalanced by the fact
that the ‘fuel’—the wind—for the wind farm is free, certainly not
true of a nuclear or fossil-fuel plant.
It is not conceivable that every possible site for wind-to-electricity
will be developed. Land use may be among the most contentious
issues for wind energy development, especially in highly populated
urban areas. Some of the best locations are over oceans, but near to
shore, providing the possibility of offshore generation from ocean
winds. Offshore generation might eliminate, or at least reduce, poten­
tially contentious issues of transmission rights-of-way and alternative
uses for the ‘land’ (or, in this case, ocean) surface. Offshore wind
energy will be more costly, because of the added challenges of working
in the oceans, and because of connections needed between the elec­
trical gear on the turbines and the shore. Nonetheless, it seems rea­
sonable to expect that over the next several decades about 10 percent
of total worldwide electricity production could be supplied by wind.
Commercial generation of electricity from wind nowadays relies
on wind farms. A wind farm contains large numbers—dozens to
hundreds—of wind turbines. Wind farms are now in operation in
many countries around the world, though there remains a market
for smaller machines used individually to supply electricity for a
single user. The costs of generating capacity for wind energy can
now, in many places, compete with the lower end of costs for fossil
fuel generation. (This is true of solar as well.)
114 ENERGY: THE BASICS

ARGUMENTS PRO AND CON


Wind is very environmentally friendly. It is essentially pollution-free
at the point of generation. There is no air or water pollution, no
radioactive waste, or ash. There are no emissions of greenhouse
gases from a turbine in operation, so there is no impact on global
climate change. About the only imaginable effect on the environ­
ment is a very slight reduction in the strength of the wind as it
passes through a wind farm. A wind farm poses no significant
threat to the safety of the public. In principle, wind turbines can be
installed on any site where there is reasonably consistent wind.
Each wind turbine requires a relatively short time to build. They
can be built on an assembly line. It would be possible to produce
enough wind turbines to generate 1000 MW in a year. In compar­
ison, the construction time to build a 1000 MW fossil-fuel plant
would be several years and, for a nuclear plant, about ten years.
Wind turbines are modular. Once a site has been selected, new tur­
bines can be added quickly and relatively inexpensively as electricity
demand increases. In comparison, fossil-fuel or nuclear plants
would usually be built in increments of several hundred megawatts
at a time, at a cost well into the hundreds of millions of dollars.
Although the total area occupied by a wind farm may be large,
the only actual use of the ground surface is the ‘footprint’ of each
wind turbine. Over the area of an entire wind farm, the turbines
themselves use no more than about 5 percent of the total. The
remaining land can be put to other uses. Thus, wind energy devel­
opment is ideally suited to farming areas. Crops can be planted, or
livestock can graze, right up to the bases of the turbine towers.
Wind energy has the advantage, relative to solar energy, that
wind can blow day or night, and can blow on cloudy days when
solar would not be at its best. Winds are often at their strongest and
most reliable during the dark nights of winter, when most of us have
our greatest demand for electricity. This suggests that a combina­
tion of solar and wind facilities might balance the effects of sunny
vs. cloudy, day vs. night, and calm vs. windy. Nonetheless, both
solar and wind are considered to be intermittent sources, for which
energy storage is needed. We will discuss this issue in Chapter 13.
An obvious disadvantage of wind energy is that turbines only
work in areas that are windy. Even in suitably windy sites, in most
places the wind does not blow all the time, nor does it usually blow
steadily. Nuclear, fossil fuel, and hydro plants can operate 24 hours
per day, in all but the most extreme of weather, and can generate
ELECTRICITY FROM WIND 115

electricity on demand. Wind farms are necessarily dependent on


available wind. Capacity factors can be over one-third of rated
capacity, but in unfavorable cases produce an average output of
one-fifth of the rated capacity. It appears that some wind farms can
affect local temperatures, with near-surface temperatures a bit
higher at night and a bit cooler during the day.9
The low-frequency noise made by the rotating blades can be
considered to be a form of noise pollution. The rotating blades can
kill birds and bats. This problem has not gone away (and probably
never will), but the modern designs of turbines operating at much
slower rates of rotation, have a reduced the likelihood of killing
birds. Bird kills are not unique to wind energy systems. By far the
most effective killer of birds is glass—windowpanes.
Some people object to the visual impact of a large group of tur­
bines clumped in a large wind farm or arrayed along an otherwise
scenic ridgeline. This is an issue of personal aesthetic choices that
cannot be addressed or fixed by science and engineering. To some,
wind turbines are elegant symbols of modern, environmentally-
friendly technology. To others, they are quite ugly. Though wind
farms produce no tangible pollution of air or water, others object to
the very sight of tens or hundreds of wind turbines, and their asso­
ciated transmission lines, on the landscape. To help assure that tur­
bines capture as much wind as possible, it is good to locate them on
a hilltop. A situation develops in which people for miles around find
themselves looking at dozens and dozens of turbines.
Wind turbines are virtually emission-free at the point of elec­
tricity generation. However, their overall impact depends on how
big we make the ‘window of attention’. Those emission-free wind
turbines rest on pads made of reinforced concrete. Many of the
components, particularly the tower, are made of steel. The esti­
mated steel requirement is 200 tonnes per MW of generating capa­
city.10 Both the cement and steel industries make significant
contributions to carbon emissions, about 5 and 7 percent, respec­
tively. It is a good thing that may of the components can be made in
factories, even on assembly lines, but then how do they get to the
wind farm? On very large trucks or possibly partway by rail.
Transportation adds further carbon emissions.
Some blades are made from carbon-fiber reinforced fiberglass
composites. Using carbon fibers is a promising technology, because
they can be used to fabricate lighter, thinner, and stiffer blades.
Carbon fibers can be made from materials derived from coal. We
will discuss in Chapter 14 some important reasons why we need to
116 ENERGY: THE BASICS

reduce significantly the combustion of fossil fuels. I have argued


elsewhere11 that it makes no sense at all to burn coal—but this need
not be the death knell of the coal industry, but rather offer an
opportunity to help build the infrastructure for a renewable-energy
industry from coal.
Cost is always of concern for any energy system. At least two
measures of cost must be considered. The first is the cost to build
and install the generating system. The second is the actual cost of
the electricity that is eventually delivered to the consumers. Cost
estimates must always be treated with caution, because they are
usually made on the basis of numerous assumptions, and unless
those are clearly known, it may not be obvious how a final esti­
mated cost was developed. Further, it might be that different kinds
of assumptions were made for, say, a wind farm and a nuclear plant,
leading to the proverbial ‘apples vs. oranges’ comparison.
The cost for the capital investment to build and install an
onshore wind farm is second-lowest among the likely sources of
large-scale electricity generation. Currently the lowest capital cost
would be for a natural-gas fired plant. The major costs in the wind
power industry are in capital equipment, since the turbines do not
need ‘fuel.’
The same caution about capital costs also applies to comparing
the cost of electricity from different sources. Also, the cost of gen­
eration might have been increased by virtue of various taxes, tariffs,
subsidies, surcharges, or other costs. These additional costs vary
from one country to another and from one state or region within a
country. Generally, in terms of cents per kilowatt-hour, electricity
from wind is about the cheapest, though hydro and nuclear are
close behind. Electricity from wind is less than one-third the cost as
from solar.
On balance, wind is an essentially pollution-free method of gen­
erating electricity that appears now to be very cost-competitive with
conventional sources. It seems reasonable, therefore, that wind is the
renewable energy resource with fewest technical, economic, or envir­
onmental hurdles to making a significant contribution to our elec­
tricity supply.
Wind most certainly qualifies as a renewable resource, along with
hydro, at least some forms of biomass, and solar. Wind energy
therefore shares the common problem of renewables, that they are
mostly diffuse, scattered energy sources. We must invest time,
energy, and ingenuity to gather and concentrate these resources. We
must be cognizant of the EROEI. We must recognize that an energy
ELECTRICITY FROM WIND 117

source that is low- or even near-zero emitting at its point of use


might have significant issues in infrastructure and transportation.
And, we should also recognize that what we’re doing is shifting
toward relying on energy sources that may go back to antiquity, or
at least to the pre-industrial era.

NOTES
1 Based on data from Nahin, Paul J. In Praise of Simple Physics. Princeton
University Press: Princeton, 2016, p. 25.
2 Nahin, Paul J. In Praise of Simple Physics. Princeton University Press:
Princeton, 2016; p. 25.
3 Wagman, David. Renewable Energy. PennWell Books: Tulsa, 2024, p. 73.
4 A useful guide to estimating what one might see or feel in winds of various
strengths is available from the U.S. National Weather Service at: https://
www.weather.gov/pqr/wind.
5 Gray, Theodore. Engines. Black Dog & Leventhal Publishers: New York,
2022, p. 164.
6 Heimler, Charles H. and Price, Jack S. Focus on Physical Science. Charles
E. Merrill Publishing Co.: Columbus, OH, 1984, pp. 500–501.
7 Wagman, David. Renewable Energy. PennWell Books: Tulsa, 2024, p. 66.
8 All capacity factor data are from Miesner, Thomas O. and Gallo, A.
Andrew. The Electric Power Industry. PennWell Books: Tulsa, 2022, p. 96.
9 Roy, Somnath Baidya and Traiteur, Justin J. ‘Impacts of wind farms on
surface air temperatures.’ Proceedings of the National Academy of Sciences,
Vol. 107, pp. 17899–17904, 2010.
10 Smil, Vaclav. How the World Really Works. Viking: London, 2022, p. 101.
11 Schobert, Harold H. Rethinking Coal. Oxford University Press: New York,
2023, Chapters 16–17.
13

ENERGY STORAGE

THE NEED FOR ENERGY STORAGE


Of the all the energy forms, of all the ways of obtaining useful
work, electricity is in many ways the best, when we consider it at its
point of use. There’s miniscule effort in using it. Pushing a button or
flipping a switch certainly beats shoveling coal or splitting firewood.
When we do use electricity, it’s clean, emission-free, and quiet.
There’s no need to devote space to oil tanks, wood piles, or coal
bins. Electricity can be converted into light, heat, or mechanical
work with equal ease. The efficiency of conversion of electricity
when we use it is generally very good.
Most of us have become accustomed to having electricity avail­
able around the clock, every day of the year. As we try to accom­
plish a transition from fossil fuels, with their attendant CO2
emissions, to renewables, we come up against a problem: some of
the major sources of renewable energy—notably solar and wind—
are intermittent. A further complication is that our electricity
demand is not constant during the day. It tends to be higher during
daylight hours and lower late at night. It also has peaks in demand
at times, such as the early evening when most people are coming
home from school or work, preparing meals, and turning on heat or
air conditioning. Even hydro, which works well for providing steady
output, might not be able to cope with the daily swings in demand.
These considerations affect electricity used in stationary locations.
In addition though, more and more we need to use electricity in
mobile applications, with telephones, laptops, and electric vehicles
(EVs) leading the way. For these applications we need ‘portable
electricity.’
Concerns about global warming have led to a desire to achieve
near-zero net carbon emissions by 2050.1 To hit that target will

DOI: 10.4324/9781032665603-13
ENERGY STORAGE 119

require a massive shift to renewables for electricity generation. But


infrastructure takes time to build.2 Is it possible, starting from 2025,
to build enough renewables generating capacity in 25 years? And,
would we have the energy storage technology and capacity adequate
to back up a global renewable-energy system?
It’s possible to envision electric utility systems that combine sev­
eral generating technologies, perhaps linked by very-long-distance
grid systems, that might smooth out issues of intermittent genera­
tion and daily demand. They don’t help us with the mobility issue.
Users of electricity would expect to use the GPS on their smart-
phones while driving their EVs. This leads us to consider ways of
storing electrical energy.
At least three strategies are available. The most familiar is the use
of batteries. A more convoluted approach is to use excess electricity
to produce a fuel that could be stored and then later converted back
to electricity. Specifically for hydro, a third option is pumped storage.

BATTERIES
Batteries3 are over 200 years old. They date from the Voltaic pile
(Chapter 6). It contained all the essentials of the modern battery:
two electrodes made of some conducting material (copper and zinc
in Volta’s case) to allow electricity to leave or enter the battery, and
some sort of medium, usually a liquid (saltwater for Volta), that
allows ions to move between the electrodes.
A flow of electricity from a battery represents a spontaneous
change from high to low electrical potential energy. To obtain the
high electrical potential energy, work must be supplied to the system.
In essence, we must pump electrical energy uphill to have it at high
potential. The work necessary to do that non-spontaneous change
comes from reactions among the chemical components of the battery.
This last point brings us to the two major kinds of batteries.
First, when the chemicals in the battery have reacted to the point
at which they are unable to pump electrical energy uphill, we can
no longer get a flow of electricity from the battery. We commonly
say that the battery has died, run down, or gone dead. The battery
is now useless, and the great majority of us throw it away at that
point, sometimes in a fury. Batteries of this kind are called primary
cells. The near-ubiquitous AA and AAA batteries are of this kind.
Second, in a different kind of battery, we begin by supplying elec­
trical energy into the battery. The electrical energy that we supply to
the battery does the work of pumping the battery components up
120 ENERGY: THE BASICS

the ‘chemical hill’ to high potential. Then, when we use the battery
to supply electricity, the change from high to low chemical potential
drives the change from low to high electrical potential, and we then
are able to make use of the electricity. Batteries of this type are
called secondary cells, and are also known by the apt name of sto­
rage batteries. The battery has stored (as chemical energy) the elec­
tricity that we originally supplied to it. Batteries in automobiles and
in electronic devices are examples of storage batteries.
When a secondary cell has run down, we can supply more electricity
to it from some external source. In other words, the battery can be
recharged. The two kinds of secondary cells of by far the greatest
importance to us are the lithium-ion batteries in various devices such
as laptops and cell phones, and the lead-acid batteries used in internal-
combustion-engine vehicles. Both are rechargeable. If it were not for
the Second Law of Thermodynamics, we could in principle run a bat­
tery through an unlimited number of discharge-recharge cycles. One
good battery would last us a lifetime. However, since no process can
run with 100 percent efficiency, a very small fraction of its capacity is
lost. Battery technology nowadays is quite good, so that batteries can
run through thousands of cycles before the capacity is reduced to the
point at which the battery is no longer useful, and must be replaced.
Storage batteries should allow us to store electricity when there is
excess generating capacity, and then to pull electricity out when the
sun isn’t shining or the wind isn’t blowing. But, it’s not quite that
simple. As with every other technology we’ve discussed, there are
issues to be addressed. They include questions of energy density,
capacity, and the availability of materials to make the batteries.
We’ve seen previously (Chapter 5) that energy density—the
amount of energy that can be supplied per unit of mass—can be an
important characteristic of energy sources, especially for vehicular
or mobile applications. Energy density can be appreciated in one
simple experiment—pick up, or try to pick up, a lead-acid storage
battery in an automobile-parts store. They’re heavy! This might be
considered an unfair test, since lead is one of the densest of
common materials. The type of battery at the forefront of today’s
energy storage technology is the lithium-ion battery. Lithium, in
marked contrast to lead, is less dense than water.4 Even so, the
energy density of the best rechargeable batteries presently available
is only about one-sixtieth (not one-sixth) of the energy density of
gasoline.5 Graham Hoyland has pointed out the unfortunate fact
that a ham and cheese sandwich has thirty times better energy
density than the best lithium-ion batteries.6
ENERGY STORAGE 121

These issues aside, lithium-ion batteries have become the energy


storage of choice for many stationary and mobile applications.
Their popularity derives from providing a high energy density
without adding undue weight, a reasonably quick recharge time,
and a long useful lifetime in a phone or laptop.
The essential components of a lithium-ion battery are exactly the
same as those of a Voltaic pile: a pair of electrodes and an electro­
lyte. The electrodes are known as the cathode, which is positively
charged, and the anode, negatively charged. The cathode is typi­
cally a lithium cobalt oxide; the anode is generally made of gra­
phite. The electrolyte is a mixture of an organic compound (such as
dimethyl carbonate) and a lithium salt; other formulations are used
as well. The lithium salt dissolves in the organic component to form
a conductive solution.
Before it can be used, the battery has to be charged from an
external source of electricity. We are pushing electrical energy uphill
to achieve a high electrical potential energy that, when we need it,
we can put to work for us as it spontaneously flows back downhill.
For all of today’s amazing electronic devices, the external source of
electricity still comes from a prime mover (i.e., a device that con­
verts energy into mechanical work) driving a generator to achieve
the electromagnetic induction discovered by Michael Faraday
nearly two centuries ago.7
When we want to operate an electrical device—that is, we are
now discharging the battery—the chemical reactions that took place
during the charging cycle are reversed. Electrons leave the anode,
pass through the device and provide the energy needed to operate it,
and ultimately flow into the cathode. The electric circuit is com­
pleted as positively charged ions move through the electrolyte and
pick up electrons at the anode. Let’s consider the specific case of a
lithium-ion battery.

LITHIUM-ION BATTERIES
When a lithium-ion battery is being charged—that is, we are sup­
plying energy to it—the cathode releases lithium ions. They migrate
through the electrolyte to the anode. At the anode, lithium ions
combine with electrons (which we are supplying) to produce lithium
metal. It remains there during the charging cycle. Then, when we
put the battery into use for us, lithium metal at the anode gives up
its electrons, and the resulting lithium ions move back through the
electrolyte toward the cathode. Electrons flow in the opposite
122 ENERGY: THE BASICS

direction through the outer circuit (the outer circuit may include,
for example, your phone).
In terms of energy density, lithium-ion batteries are substantially
superior to the much older lead-acid battery technology. Even so,
for very large-scale uses, and even for possible electric vehicle
applications, the energy that can be supplied by the biggest battery
systems is not much.
Presently the largest lithium-ion battery system will deliver 400
megawatt-hours (MWh) of electricity,8 the annual electricity con­
sumption for about 40 average houses in the United States.9 Even
so, the prospects of large-capacity battery storage, with recharging
from solar and/or wind installations, offer enormous benefits to
people in developing nations having no access to an electric grid.
Having enough electricity to provide refrigeration for safe, long-
term storage of food and medicines and adequate lighting for per­
sonal safety would be an inestimable boon for such people.
In electric vehicles, a 300–400-volt, 30-amp battery pack will
drive a car for about 160 km.10 So will 4 liters of gasoline. A really
good lithium-ion battery will deliver about 300 watt-hours per
kilogram of mass (Wh/kg). Petroleum-derived liquid fuel in a jet
engine can deliver 12,000 Wh/kg.11 Especially for transportation,
batteries need continued reduction of size and weight, improved
storage for large amounts of electrical energy, and the ability to
deliver that stored energy rapidly.
Electric motors used in transportation have the advantage of
being virtually emission-free. In addition, electric motors are smal­
ler, lighter, and quieter than internal-combustion engines. They can
run continuously for very long times, often years, with little or no
maintenance. However, using battery-operated electric motors—as
in electric cars, for example—as a strategy for reducing carbon
emissions, requires that the electricity used for recharging the bat­
teries comes from carbon-free generation. Getting enough elec­
tricity becomes a major concern.
The energy density and energy storage capacity are issues that
have affected batteries from the first Voltaic pile, around 1800, to
today. Even Thomas Edison couldn’t come up with a solution.
About 125 years ago, he put a great deal of effort into the devel­
opment of a battery that could offer an energy density comparable
to that of gasoline. We still don’t have one.
Despite these concerns, it is undeniable that lithium-ion batteries
are greatly superior to older storage-battery designs, and that they
have already made great impacts on the daily lives of enormous
ENERGY STORAGE 123

numbers of people. Further, there are vigorous research and devel­


opment efforts in battery science and technology going on in many
countries of the world. But, the burgeoning popularity of these
batteries brings up another question: what are we going to make
them from?
The cathode is a compound of lithium with oxygen and (usually)
one other metallic element, such as cobalt. The anode is commonly
made of graphite. Major producers of lithium are Australia, Chile,
and China. Currently, cobalt is produced mainly by the Democratic
Republic of Congo, China, and Zambia. Neither of these elements
occurs in high average concentrations in Earth’s crust, so con­
centrated deposits—ores—are very important. Graphite, a form of
pure carbon, occurs naturally in about a dozen countries, of which
China and Brazil hold the largest reserves. Graphite can also be
produced synthetically, usually starting with a coal of high carbon
content, or a coke made from the heavy fractions of petroleum.12

PUMPED STORAGE
We’ve discussed how the water supply that is at high gravitational
potential energy got there through the natural operation of the
hydrologic cycle, which is driven by solar energy. There is an alter­
native way of assuring a supply of water at high potential: we can
pump it uphill ourselves. Doing so is the basis of an energy storage
strategy known as pumped storage.
During periods of high demand for electricity, a hydro plant
operates exactly as described in Chapter 11. Water flowing through
turbines drives a generator that produces electricity to be fed into
the electrical grid. Generally, demand for electricity fluctuates
through a 24-hour period, which means that there can be periods in
which demand is less than the generating capacity of the plant. At
such times, part of the plant can be run backwards, so that elec­
tricity is supplied to some of the turbines, in which case they will
operate as pumps. Excess electricity from the plant can be used to
move water back uphill (literally, in this case) from low to high
gravitational potential energy.
So, when electricity demand is high, water flows from high to low
potential, work is extracted to operate a generator, which converts
mechanical work into electricity at high electrical potential. At
times of low demand, we supply electricity to a turbine which does
work pumping water from low to high potential. The whole concept
has been likened to a ‘giant, complicated rechargeable battery.’13
124 ENERGY: THE BASICS

The pumped storage concept has the advantage that it can pro­
vide fairly large storage capacity. Conceptually, pumped storage
could be used as a way to store excess electricity from wind or solar
installations, getting the electricity back, so to speak, by running
the water through hydro turbines. In the United States, pumped
storage is the largest contributor to energy storage, accounting for
about 93 percent of energy storage.14
Since no process can operate at 100 percent efficiency, some of
the ‘extra’ electricity not going into the grid at night is going to be
consumed in running the pumps to get water back uphill. We can’t
use pumped storage as an enormous version of a perpetual-motion
machine. Also, any pumped storage installation is going to require
two reservoirs, one uphill and one downhill. Each is going to
occupy a significant area of land; setting aside that much may be
questionable, especially near large urban areas. The local topo­
graphy has to be able to provide useful elevation differences
between the two reservoirs.

HYDROGEN
Energy can be stored in chemical bonds. One possibility is to use
any excess electricity production to split water molecules (electro­
lysis). The products of electrolyzing water are hydrogen and oxygen.
Hydrogen can be produced in many different ways, but water elec­
trolysis is likely the direct route and—if the electricity is made from
renewables—the cleanest route.
Hydrogen has many fervent advocates. Since about 70 percent of
Earth’s surface is water, there’s no shortage of the raw material.
When hydrogen is consumed to produce energy, the product is
water. In principle, one might think we could run the water-to­
hydrogen-to-water cycle forever, but the Second Law reminds us
that none of the steps will work with 100 percent efficiency. There is
already a large infrastructure in many countries for hydrogen pro­
duction, transport, and storage. Other routes to hydrogen are
available, so that hydrogen could conceivably be used for energy
storage even in very arid regions.
The most direct way of recovering the stored energy used to make
hydrogen is in fuel cells. Primary batteries eventually stop working
because their chemical components are used up, to the point where
the battery is no longer able to generate adequate potential. We
could get around that problem if we could continually provide a
supply of the necessary chemicals. In that case, the battery could
ENERGY STORAGE 125

run for very long times. Devices into which we supply a stream of
chemicals and from which we extract electrical energy are called
fuel cells. A fuel cell operating on hydrogen converts the chemical
energy in the bond between hydrogen atoms to electrical energy.
Like many new and exciting ideas, fuel cells are actually quite
old. The original developmental work was done in the 1830s. The
idea lay around for a century, until, in the 1930s, the British engi­
neer Francis Thomas Bacon15 breathed new life into it—we might
say breathed hydrogen into it. Hydrogen fuel cells were first used on
spacecraft some 60 years ago. A stand-alone fuel cell doesn’t store
energy, but, if we combine a fuel cell with an electrolyzer and with
hydrogen storage, we then have an energy storage system.
Stationary fuel cells are already in use both as backup energy
systems and for a primary source of energy. Fuel cells are being
used—or are in trials—in various kinds of vehicles, including cars,
buses, trucks, and the even the humble forklift. None of these are
designed for energy storage, but rather would require a replenish­
ment of hydrogen.
Like other energy-storage strategies, making hydrogen will always
use more energy that we will get back by consuming the hydrogen
later. If we didn’t need to store energy, we would certainly be better
off to use solar or wind electricity directly. Thanks to the immutable
Second Law, making hydrogen from water and converting it back to
water in a fuel cell is not going to be some kind of perpetual energy
machine.

E-FUELS
Solar energy has been stored in carbon compounds at least since the
evolution of photosynthesis over four billion years ago. It’s well said
that ‘carbon-based energy carriers are the most natural thing in the
world.’16 Nuclear fusion reactions in the sun release energy that is
captured by photosynthesis and stored in numerous kinds of carbon
compounds. Prodigious quantities of energy can be stored this way.
The German forester Peter Wohlleben points out that, if only our
human digestive systems had evolved in a way to allow us to eat
wood, a single mature beech tree could provide the necessary daily
energy intake for an adult for 40 years.17
Electro-fuels (or e-fuels) provide another approach. Electro­
chemical processes are able to convert mixtures of carbon dioxide
and hydrogen into hydrocarbons. We already make both of these
raw materials on huge scales. To make an e-fuel process that is
126 ENERGY: THE BASICS

carbon dioxide neutral, there are some limitations. Hydrogen would


have to be produced from electrolysis of water using an electricity
source which does not release CO2 itself, e.g., solar. The carbon
dioxide would, ideally, be obtained from CO2-capture processes
such as those applied to fossil-fuel electricity generation.
Several kinds of e-fuels could be produced: synthetic diesel fuel;
methanol, once touted as a replacement for gasoline; or such gaseous
fuels as methane or butane. A virtue of the e-fuels concept is that we
already know how to handle, store, transport, and use these fuels.
E-fuels could represent a strategy for energy storage, if electricity
from excess capacity were used to drive the water electrolysis and
the e-fuel synthesis processes. The eventual combustion of the e-fuel
would return the original raw materials, carbon dioxide and water,
as the products. This represents another potentially cyclic process,
though, just like the others, subject to Second-Law limitations.
However, producing carbon dioxide might not be such a good idea,
as we will see in the next chapter.

NOTES
1 International Energy Agency. Net Zero by 2050. 2021. https://www.iea.org/
reports/net-zero-by-2050.
2 Not only does it take time, it also takes large amounts of steel, cement, and
polymers, all of which come from industries that themselves currently have
significant carbon emissions.
3 Strictly speaking, one of the devices that we commonly call a ‘battery’ is
actually a cell. However, almost no one pays attention to that bit of ped-
antry, and we won’t either.
4 The comparison isn’t quite fair, since neither battery is made only of lead
or lithium. Nevertheless, for batteries of equal size, the lead-acid battery
will be substantially heavier.
5 Theodore Gray. Engines. Black Dog & Leventhal Publishers: New York,
2022, p. 118.
6 Hoyland, Graham. Merlin. William Collins: London, 2020, p. 54.
7 There is one exception: electricity generated from solar photovoltaic systems.
8 Smil, Vaclav. Numbers Don’t Lie. Penguin Books: New York, 2021,
pp. 162–165.
9 Of course, this figure is wildly variable, since energy consumption depends
on many factors, such as the local climate, the quality of construction of
the house, and the lifestyles of the occupants.
10 Nahin, Paul J. In Praise of Simple Physics. Princeton University Press:
Princeton, 2016, p. 30.
11 Smil, Vaclav. How the World Really Works. Viking: London, 2022, p. 41.
12 The reason for going to the trouble and expense of synthesizing a material
that occurs in nature is that we can often tailor the properties of the
ENERGY STORAGE 127

synthetic material, whereas with the natural material we take the properties
we can get.
13 Gray, Theodore. Engines. Black Dog & Leventhal Publishers: New York,
2022, p. 165.
14 Wagman, David. Renewable Energy. PennWell Books: Tulsa, 2024, p. 100.
15 Francis Thomas Bacon is a descendant of the family of the Francis Bacon,
the sixteenth-century English philosopher who conceived of, or at least
contributed to, the development of the scientific method. ‘Our’ Francis
Bacon served his engineering apprenticeship in the firm established by
Charles Parsons, of steam-turbine fame.
16 Moore, Andrew. The Decarbonization Delusion. Oxford University Press:
New York, 2023, pp. 175–176.
17 Wohlleben, Peter. The Secret Network of Nature. Vintage: London, 2017, p. 103.
14

ENERGY AND CLIMATE

EARTH’S HEAT BALANCE


Growing plants are warmed by sunlight. If something has been
warmed, its temperature has increased, which means that energy
has been absorbed (in this case the energy is supplied by the sun).
At night, the plant will cool, because the air surrounding the plant
has cooled, and heat flows spontaneously from the warm plants, to
the cooler air. The plants lose heat by convection and by radiation.
The usual energy balance applies: EIN  EOUT ¼ ESTORED : EIN is
the energy provided by sunlight, and EOUT is the energy lost by
convection and radiation as the plants cool.
Similarly, plants in a greenhouse1 will also be warmed. Energy in
the sun’s light helps to warm the plants and other contents of the
greenhouse. When daylight hours are over, the plants will cool, just
as any object hotter than its surroundings will lose heat sponta­
neously to the surroundings. Objects only slightly warmer than their
local environment radiate in the lower-energy (longer-wavelength)
infrared region of the spectrum. While ordinary glass is transparent
in the visible, it is nearly opaque in the infrared. Radiated heat is
trapped inside the greenhouse. (In a real greenhouse, convection
also helps retain heat inside because the roof keeps currents of
warm air from rising out of the greenhouse; but we’ll focus on heat
transfer by radiation.) The energy balance equation applies just as well
0 0 0
to the system of plant-plus-greenhouse: EIN  EOUT ¼ ESTORED .
(Here the ‘prime’ signs (´) signal that we’re writing the equation for a
different system.)
The energy coming in is the same in both systems; it’s the energy
0
in sunlight. But, EOUT is less than EOUT, because some of the heat
has been trapped in the greenhouse by the glass being opaque to
0
infrared radiation (and by its roof). Then ESTORED must be larger

DOI: 10.4324/9781032665603-14
ENERGY AND CLIMATE 129

than ESTORED. That is, energy stored in the greenhouse is larger


than the heat stored in a plant outdoors.
The observable effect of the larger value of ESTORED is that the
greenhouse is warmer. The ability of the greenhouse to maintain an
artificially warm temperature inside is why we build them in the first
place—to provide an environment for growing fresh flowers, fruits,
or vegetables in the winter, or for growing plants that normally
require much warmer climates.
The sun’s energy is absorbed by the Earth, warming the con­
tinents and oceans. The average temperature of the planet, about
15°C, is much higher than the –270°C of outer space. Consequently,
the Earth will transfer heat spontaneously outward to the frigid
surroundings of outer space.
Our atmosphere, like glass, is transparent to visible radiation. At
night, objects that warmed during the day will cool by radiating
heat to outer space, in the form of infrared radiation. Our planet
achieves its energy balance, in which energy stored as heat main­
tains the average temperature. The principal components of our
atmosphere, nitrogen and oxygen, are essentially transparent to
infrared radiation. However, the atmosphere also contains small
amounts of other gases—carbon dioxide and water vapor being
examples—that absorb some infrared radiation. Without these
gases and their ability to reduce heat loss, the average temperature
of Earth would be substantially lower. The observed average tem­
perature of the Earth results, in part, from the effect of some of the
atmosphere’s components on loss of infrared radiation from the
atmosphere. Because such infrared-absorbing gases act somewhat
like the glass of a greenhouse, they are called greenhouse gases. By
extension, this warming of the Earth by trapping infrared radiation
is commonly called the greenhouse effect. Though these terms have
now become common in our language, the analogy is not exact,
because of the very important role of preventing convection losses
in an actual greenhouse.
Infrared radiation does not become trapped permanently in
molecules of greenhouse gases. A molecule of a greenhouse gas gets
pumped ‘uphill’ by absorbing infrared radiation, increasing its che­
mical potential energy. At some point the molecule will release
energy by reverting to its preferred ‘downhill’ state.2 The energy that
is released is radiated equally in all directions. Some of that energy
is radiated right back to the Earth’s surface, and does not escape
into outer space. The net effect is exactly the same: the EOUT term
becomes smaller, and ESTORED increases.
130 ENERGY: THE BASICS

Without a greenhouse effect, Earth would have an average tem­


perature of –18°C. Most of the water on the planet would be frozen
year-round. It is very questionable whether life as we know it could
have evolved in such an ‘ice world.’ But even if it had, certainly our
lives would be much different if we had to cope with such tem­
peratures. The warming of the planet by some 33°C as a result of
carbon dioxide, water vapor, and other gases that occur naturally in
the atmosphere is the natural greenhouse effect. It exists and occurs
without human intervention. As has been pointed out, ‘earth with­
out the greenhouse effect would be only marginally more tropical
than Mars.’3
The composition of the atmosphere plays the crucial role in the
Earth’s enjoying an average temperature much warmer than expec­
ted by balancing solar radiation with infrared losses. Two centuries
ago, the French mathematician Joseph Fourier stated that carbon
dioxide, water vapor, and methane in the atmosphere helped con­
tribute to Earth’s warmth. Experimental measurements on the
absorption of heat by CO2 and water vapor were published in 1856
by the American scientist Eunice Newton Foote, her contribution
being doubly noteworthy not only for its content but also for its
being the first scientific paper in physics in America published by a
woman. Shortly thereafter the Irish physicist John Tyndall said that
the contribution of water vapor in the atmosphere was ‘a blanket
more necessary to the vegetable life of England than clothing is to
man.’ He continued, ‘Remove it for a single summer night…and the
warmth of our fields and gardens would pour itself unrequited into
space, and the sun would rise upon an island held fast in the iron
grip of frost.’4
Increased atmospheric concentrations of greenhouse gases
increase absorption of infrared radiation and augment the warming
of the planet beyond that provided by the natural greenhouse effect.
This additional heating by increased concentrations of greenhouse
gases is called the enhanced greenhouse effect. Since much of the
increase results from human activities, it is also called the anthro­
pogenic greenhouse effect.
In addition to its ability to absorb infrared radiation, the effec­
tiveness of any gas in contributing to global warming depends on its
lifetime in the atmosphere, and on its interactions with other gases
and water vapor. These factors are combined into an indicator
called global warming potential (GWP).
Methane has a GWP of 11. Its concentration in the atmosphere is
about 2 parts per million (ppm), about three times higher than it
ENERGY AND CLIMATE 131

was in 1750. Human use of natural gas has likely led to increased
methane emissions from leaks in gas wells, pipelines, and storage.
Methane is also released by decaying plant matter. Anthropogenic
sources include landfills, the decaying residue of cleared forests, and
rice paddies. The breakdown of organic matter forming methane
also occurs in the digestive systems of cattle and sheep, where bac­
teria decompose the cellulose of plant material. A healthy adult cow
can add about 500 liters of methane to the air every day. Fodder,
taken in at one end of the cow, is readily converted to methane,
which is emitted from the other end. Worldwide, about one-third of
our anthropogenic methane emissions come from livestock. Ter­
mites also convert cellulose to methane, and may be responsible for
about 10 percent of total methane emissions.
Small amounts of nitrous oxide, with a GWP of 265, are released
from fossil-fuel combustion. Its concentration in the atmosphere is
only about 320 parts per billion, but, molecule for molecule, it is over
300 times better than CO2 at trapping infrared radiation. Major
anthropogenic sources include artificial nitrogen fertilizers, discharge
of sewage, and the burning of plant material such as forests.
Other atmospheric effects can act to cool the planet. In addition to
being a greenhouse gas, water vapor also forms clouds, which help
cool the Earth by reflecting sunlight. Aerosols from volcanoes also
absorb and reflect radiation. When a major volcanic eruption occurs,
for example Mount Pinatubo in the Philippines in 1991, unusually
cool summers and harsher winters can persist for several years.
These gases are all important, and all have a role in the enhanced
greenhouse effect. However, the main concern is on carbon dioxide.
Atmospheric CO2 concentrations have been increasing steadily for
over a century. A significant fraction of the increased CO2 results
directly from human use of energy, and especially from the com­
bustion of fossil fuels. In 2023 China led the world in CO2 emis­
sions, with the United States in second place.
Processes that add CO2 to the atmosphere are called sources.
They include the decay of organisms; fossil-fuel combustion; decom­
position of carbonate rocks, either by natural processes or in the
manufacture of cement; and CO2 dissolved in the ocean returning to
the atmosphere. Processes that remove CO2 from the atmosphere are
sinks, and include photosynthesis, the dissolving of CO2 into the
oceans, or the fixation of CO2 into various carbonate forms.
The enhanced greenhouse effect is, and will continue to be, the
greatest issue to impact future energy use in the world, and the
greatest issue to impact the sources from which we obtain it. Many
132 ENERGY: THE BASICS

complex issues are involved: scientific and technical issues related to


the utilization of energy sources and the reduction of greenhouse
gas emissions, and societal issues of choices relating to energy use
and energy conservation. Resolving these issues is made all the
more complicated by the injection of politics. In some places it is a
tenet of parties on the extreme right that global warming is not
occurring, and that those who claim that it is take part in a massive
fraud. Even four centuries ago, a king of Scotland had the perfect
response to such statements, ‘It is a tale told by an idiot, full of
sound and fury, signifying nothing.’5

EVIDENCE FOR GLOBAL CHANGE


In 1938 the British engineer Guy Callendar found a very good cor­
relation between ground temperatures and atmospheric CO2 in 50
years’ meteorological data. Callendar apparently did all of his data
compilations and calculations by hand; nonetheless later investiga­
tors found his work to be extremely accurate.
Based on analyses of ice cores, we know that for the last 18,000
years atmospheric CO2 concentrations fluctuated around 280 ppm,
and that, in the relatively recent past, there has been an increase to
a concentration of 422 ppm,6 with a current rate of increase of 2.3
ppm per year. Concentrations of CO2 are now higher by at least 30
percent than the maximum concentrations found prior to the rapid
growth in the past 150 years. The increase in atmospheric con­
centrations of CO2 measured at the observatory at Mauna Loa,
Hawaii, has been tracked since 1958 (Figure 14.1). Not only is the
absolute amount of CO2 increasing, but the rate at which the con­
centration is increasing is also growing.
CO2 is an inevitable product of fossil-fuel combustion. The rise in
CO2 at ever-increasing rates coincides with the onset of extensive
use of fossil fuels, also at ever-increasing rates. Direct evidence
linking increased CO2 to fossil-fuel combustion derives from carbon
isotopes. Natural carbon has three isotopes: 6C12, 6C13, and 6C14.
There is no C14 in fossil fuels; this isotope is radioactive with a half-
life of about 5,700 years, and is long gone from materials tens or
hundreds of million years old. Burning fossil fuels produces no
C14O2. The CO2 in the atmosphere does contain some C14O2.
Mixing billions of tonnes of fossil-fuel derived CO2 into the atmo­
sphere should cause a decrease in C14 in atmospheric CO2. Also,
the ratio of C12 to C13 is slightly higher in fossil fuels than in its
average value in the world.7 Mixing billions of tonnes of fossil-fuel
ENERGY AND CLIMATE 133

Figure 14.1 Data for the increase in atmospheric carbon dioxide concentration,
taken at Muana Loa, Hawaii
Source: ©Nils Simon.

derived CO2 into the atmosphere should cause an increase in the


C12 to C13 ratio of atmospheric CO2. So, if fossil-fuel combustion is
contributing significantly to the rise in CO2 in the atmosphere, we
should expect that the amount of C14 should drop, and the C12 to
C13 ratio should go up. That’s exactly what happens. These obser­
vations are named the Suess effect, for the Austrian-American iso­
tope scientist Hans Suess. If present-day rates of fuel consumption
continue unchanged, atmospheric concentration of CO2 could
double the 1860 level sometime between 2030 and 2050.
Earth’s population is increasing rapidly. The additional people
will, not unreasonably, aspire to something more than a sub­
sistence-level existence. Increased energy use provides an improved
standard of living, as shown in Figure 1.1: readily available trans­
portation, electricity, education and health care, and the other
amenities that those of us in the developed nations often take for
granted. It’s estimated that between the years 1800 and 2000, fossil-
fuel combustion increased the amount of carbon put into the
atmosphere by a factor of 650, though the world population
increased by only a factor of 6.8
134 ENERGY: THE BASICS

Deforestation to create more agricultural land has a significant


impact. Burning the trees that have been felled causes an immediate
CO2 release. Loss of the trees means the immediate loss of an
important CO2 sink. When the trees are cleared, the humus (the
organic matter in the soil) in the forest floor, which contains CO2
locked up for decades or centuries, also decomposes, releasing all of
that stored carbon.

POSSIBLE CONSEQUENCES OF GLOBAL WARMING


Sea levels have risen 10 cm over the past 30 years. Extreme global
warming could result in a further sea level rise of about 5 m.
The ice cap at the North Pole is melting, which has already led to
squabbles among the countries that border the Arctic Ocean,
regarding who can lay claim to energy and mineral resources on,
and below, the ocean floor, and in what locations. In the Antarctic,
if the Thwaites Glacier melted suddenly, the sea level would rise by
about 3 m, drowning many major seaports.
Higher concentrations of CO2 result in ocean water becoming
more acidic. Many ocean-dwelling organisms would find it increas­
ingly difficult to secrete or maintain their shells or skeletons; they
might not be able to survive. Coral reefs could be damaged and
weakened to a point where they could not survive, bringing
immense changes to the marine ecosystems around them.
Climate zones ideal for farming will shift toward the poles, in some
cases crossing national borders, with possibly enormous impacts on
global geopolitics. Areas of several Mediterranean countries, parts of
the United States, and some former Soviet countries could become
deserts.
Global warming could exacerbate the release of even more CO2
or other greenhouse gases.

THE GLOBAL CARBON CYCLE


The global carbon cycle (Figure 5.1) helps track the processes that
can change CO2 concentration in the atmosphere. Changes in the
concentration of CO2 in the atmosphere result from the relative
balance between sources and sinks. Since the concentration of CO2
in the atmosphere has been increasing steadily for many years, this
is an indication that more CO2 is being produced from sources than
can be absorbed into sinks.
ENERGY AND CLIMATE 135

The dominant anthropogenic CO2 source is the combustion of


fossil fuels and biofuels. Current annual CO2 emissions from
worldwide fossil-fuel combustion are about 35 gigatonnes, four
times the rate of release in 1950 and ten times the rate in 1900.

POSSIBLE POLICY OPTIONS


The global carbon cycle shows two strategies for slowing the build­
up of carbon dioxide in the atmosphere: either to reduce the rate
coming from sources, or to increase the rate going into sinks. Rea­
listically, both need to be employed.
The only routes to removing substantial quantities of CO2 from
the atmosphere on a human timescale—as opposed to a geological
scale—would be to encourage and enhance photosynthetic con­
sumption of CO2 by plants, and dissolving it into the ocean. The
latter brings with it its own family of consequences.
The concept of ‘decarbonizing’ the economy seems at first to be very
attractive. Basically, we just convert from using fossil fuels to using
renewable fuels. Decarbonizing would help reduce the source side of
the carbon cycle. The deployment of the emerging technology of
‘direct air capture’ to remove CO2 that’s presently in the air would help
on the sink side. There’s only one major problem. We can’t do it yet.
Several issues stand in the way. It takes time, often a lot of time,
to build new energy-production facilities, or even to retrofit existing
ones. The building of large-scale facilities requires years, even after
land acquisition, permissions, and regulatory compliances are fin­
ished. We can’t put the shutters on a coal-fired plant next week, and
expect to have a wind farm running in its place the following week.
Nor can we switch the plant from burning coal this week to burning
aspen wood next week. Boilers, along with their ancillary equip­
ment, are meticulously designed around a selected fuel. A radical
change will require lots of downtime for a major retrofit. For the
most part all of the necessary equipment is made of steel, concrete,
and plastics. Producing and transporting these three families of
materials relies on fossil fuels. Even if current efforts to ‘dec­
arbonize’ steel and concrete are technically and economically suc­
cessful, a long time is going to be needed for the complete
conversion of industries worldwide. And what about us? Perhaps by
now some 80 percent of the world’s population is alive thanks to
food produced using synthetic ammonia or ammonia-based chemi­
cals as fertilizers.9 Production of synthetic ammonia produces about
3 tonnes of CO2 per tonne of ammonia.
136 ENERGY: THE BASICS

By far the most effective, and least costly, approach to reducing


CO2 coming from sources is to increase the efficiency of processes
that consume fuel, thus increasing the amount of useful energy
obtained per unit carbon dioxide emitted. Examples include
switching to cars that are more fuel efficient, increasing the effi­
ciency of electricity-generating plant operation, and practicing
energy conservation at home.
Fuel consumption could be shifted to those fuels that produce
more usable energy per unit weight of carbon dioxide emitted. In
most cases, this comes down to using less coal and more natural
gas. Some sincere individuals have argued that we must stop using
fossil fuels, right now. That’s impossible. The International Energy
Agency has said that ‘no more than one-third of proven reserves of
fossil fuels can be consumed prior to 2050.’10 Yes, doing so would
probably limit the increase in average global temperature to about
2°C, which should probably stave off the worst outcomes of global
warming. But there’s no way to do this quickly, given our depen­
dence on fossil fuels for energy and for crucial materials.
Third, the use of ‘non-fossil’ or ‘non-carbon’ sources of energy can
be increased, though, once again, a switch to large-scale use cannot
be made quickly. Such sources include hydro, nuclear, biofuels, wind,
and solar. Each of these options has its advantages and dis­
advantages, some of which have been discussed in previous chapters.
The second major strategy involves enhancing the uptake of CO2
into sinks. The first step that could be made is to slow the rate of
destruction of the sinks we already have. Stopping the destruction
of rainforests and similar natural sinks entirely is probably hopeless.
The ability of other natural ways to absorb CO2 can be enhanced,
to mimic the sinks in the global carbon cycle. Carbon dioxide fed to
these processes could be kept out of the atmosphere, ideally for
millennia. Such processes are broadly known as carbon capture and
storage (CCS), or as CO2 sequestration, and include its being
incorporated into carbonate rocks, dissolved into the oceans, or
used by living plants in photosynthesis.
Mineral carbonation involves reacting carbon dioxide (as in the
flue gases of a boiler, for example) with a mineral using conditions
that force uptake of CO2 in a relatively short time, rather than the
very long timescales of natural processes. Minerals rich in calcium
or magnesium would form stable calcium or magnesium carbo­
nates. They have industrial uses, so might be sold to offset some
costs of the process, but also could be buried and locked away in
the Earth’s crust for millennia. Alternatively, natural brines, which
ENERGY AND CLIMATE 137

occur in such places as oil or natural gas reservoirs, can be used.11


Carbon dioxide dissolves in the brine much as it would in the ocean.
The current challenge for CCS is mainly one of the mismatch of
scale between the prodigious volumes of CO2 produced each year,
and our ability to capture and store it. In 2020 it was estimated that
the global capacity for CO2 capture was about 40 million tonnes
per year, with at least that much more capacity under development.
The challenge comes from the fact that global emissions are a
thousand times greater than our capacity for capturing CO2.
The global warming problem is enormous, highly complicated,
and both literally and figuratively global. The fact that it is global is
one of the things that makes it hard. Action is important. ‘Screaming
headlines about impending catastrophe and images of orphaned
polar bear cubs have not yet solved many problems.’12 Despite the
magnitude of the problems and the difficulties to be overcome, it’s
still fair to say that ‘major reductions in carbon emissions—resulting
from the combination of continued efficiency gains, better system
designs, and moderated consumption—are possible, and a deter­
mined pursuit of these goals would limit the eventual rate of global
warming.’13 The consequences of making the wrong decisions could
be catastrophic. We—all of humankind—are in the process of doing
an experiment on ourselves. We’ll see how it turns out.

NOTES
1 In various parts of the world, greenhouses are called glass houses, or hot
houses. Both terms are apt.
2 Continuing an analogy with waterfalls, the lowest possible ‘downhill’
energy state of a molecule is called its ground state.
3 Streever, Bill. Cold. Little, Brown & Company: New York, 2009, p. 164.
4 Quoted by Sara Wheeler. The Magnetic North. Jonathan Cape: London,
2009, p. 226.
5 Shakespeare, W. The Tragedy of Macbeth. Act 5, scene 5, 1606. Always in
print in a wide variety of editions.
6 Lan, X., Tans, P. and Thoning, K.W. ‘Trends in globally-averaged CO2
determined from NOAA Global Monitoring Laboratory measurements.’
Accessed 05-Nov-2024 13:17:54 MST: https://doi.org/10.15138/9N0H-ZH07.
7 This is due to the kinetic isotope effect, a well-known and well-understood
phenomenon in chemistry that pertains to isotopes of any element in any
reaction.
8 Smil, Vaclav. Numbers Don’t Lie. Penguin Books: New York, 2021, p. 304.
9 This story is admirably documented by Smil, Vaclav. Enriching the Earth,
MIT Press: Cambridge, 2001.
10 Quoted in Wilson, Edward O. Half-Earth. Liveright Publishing Corp.: New
York, 2016, pp. 65–67.
138 ENERGY: THE BASICS

11 Some natural brines contain enough dissolved lithium to make them


attractive potential lithium sources for production of lithium-ion batteries.
Perhaps someone might develop a process to extract lithium and then use
the residual brine for CO2 capture.
12 Wheeler, Sara. The Magnetic North. Jonathan Cape: London, 2009,
pp. 162–163.
13 Smil, Vaclav. How the World Really Works. Viking: London, 2022, p. 200.
15

THE 18-TERAWATT CHALLENGE

THE CHALLENGE
We have welcomed the eighth-billionth person to our planet, and
might greet the nine-billionth by 2050. Collectively, we use enough
energy in the course of a year such that, converted to an annual rate
of using energy, i.e., our global power consumption, it would be
17.7 terawatts (TW), or 17,700,000,000 kilowatts.1 In comparison, a
human metabolism operates at around 80 watts. A superbly trained
athlete at peak condition performing at his or her optimum might
be able to produce about 500 watts. The energy used for all the
interior processes inside the Earth, including the motion of portions
of continents, is about 40 TW. Right now, our annual consumption
is more than one-third of the power required for all of the geologi­
cal processes inside our planet. There are more powerful activities:
a volcanic eruption can release about 50 TW and an enormous
earthquake even more. The 2004 earthquake in the Indian Ocean,
which produced a devastating tsunami, may have released 2000 TW.
However, earthquakes and volcanoes are very short-term events on
our human timescale, and are a mere moment on a geological
timescale. Our consumption of nearly 18 TW goes on all the time,
all day every day.
At least one billion people live in conditions so dreadful that they
lack access to clean water and dependable food, adequate housing
and sanitation, basic health care, and enough education for lit­
eracy.2 Possibly a third to a half of the world’s population lacks
access to at least one of these services. Suppose we were to try to lift
all of the world’s population to a level at which any person, any
place, could reasonably expect at least to have safe and warm shel­
ter, drinking water, enough food, and access to a school and a
clinic. This does not mean providing everyone with the resources to

DOI: 10.4324/9781032665603-15
140 ENERGY: THE BASICS

use energy at a profligate scale, such as the energy consumption in


the United States, but to a level of a rather modest standard of
living just to assure health, nutrition, and safety. In terms of energy
rather than power, by 2050 we will need twice as much energy as we
consume now.3 Doing only this job will require about 18 more
terawatts. One way of formulating the challenge is this: Where are
we going to get the ‘extra’ 18 terawatts?
Here we are confronted with a super-wicked problem, character­
ized by the lack of a central authority that is dedicated to finding a
solution, in which the people seeking solutions are also contributing
to the problem, in which there is a significant time deadline (such as
reaching zero net carbon emissions by 2050), and in which there
may even be policies that irrationally impede progress toward a
solution.4 Richard Rhodes says that, ‘The great challenge of the
twenty-first century will be limiting global warming while simulta­
neously providing energy for a world population not only increasing
in number but also advancing from subsistence to prosperity.’5 Jon
Gertner adds this perspective:

Moving the global economy from one that runs on fossil fuels to one
that runs on renewables is almost certainly the most difficult chal­
lenge—the wickedest problem—of the twenty-first century. Whether
we succeed or fail at these efforts, which are only now in their
infancy, will determine how dramatically our climate will change in the
coming century. It will also determine the political and economic
strength of nations that now rely on the production or consumption
of oil, gas, and coal.6

During my career, I have been fortunate to meet executives and


senior managers from many energy industries and government
agencies. When the topic comes up of addressing future energy
needs, all of them, regardless of personal professional interests, have
said almost exactly the same thing: We’re going to need everything.
Every energy resource we have available will have to be mobilized
to meet the future energy needs, and expectations, of the ever-
growing world population.7
In mobilizing all the world’s energy resources, we need to keep three
things in mind. First, every energy source has some technical advan­
tages, and some technical disadvantages. Second, every energy source
has an impact on the environment, and may have some advantages or
benefits for the environment as well. And third, every energy source
has some economic advantages, and some economic disincentives.
THE 18-TERAWATT CHALLENGE 141

These facts mean that all of us, not just professional workers in
energy science and engineering, have to understand how to navigate
between the technical advantages and disadvantages, the environ­
mental impacts and benefits, and the economic incentives and dis­
incentives, to arrive at a mixture of energy sources for wherever we
happen to live that provide a reasonable balance among reliability,
affordability, and impact on the environment. It may help us to
remember these comments: ‘It’s not necessarily the best technology
that prevails, but rather the one that works reliably, and that can be
purchased at an attainable price point.’8; and, ‘Physics is indis­
putable, but economics rules.’9 There is no ‘one size fits all’ solution
to energy needs.
Further, addressing the many aspects of the 18-terawatt challenge
is going to require the efforts not only of scientists and engineers,
but also sociologists, political scientists, economists, historians,
anthropologists, and people from many other disciplines or areas of
expertise. In other words, not only are we going to need everything,
we’re going to need everybody.
Improving standards of living, at least in terms of GDP, require
improving per capita energy consumption (recall Figure 1.1). In
2020, the average annual per capita energy used by about 3 billion
people (some 40 percent of world population) was comparable to
that enjoyed by the people of Germany and France—but in 1860!10
As developing countries continue to improve their standards of
living, their energy use will inevitably increase. If everyone on the
planet enjoyed a standard of living comparable to that in many
European countries, the collective power used would be somewhere
around 40 TW. This scenario is highly unlikely, because raising
everyone’s standard of living to that of Europe would require an
international humanitarian effort on a scale and on a level of
international cooperation unprecedented in history. But, as the
population expands, more people will necessarily consume more
energy. Continued population growth coupled with improved stan­
dards of living means that it’s not unlikely that a time will come
when we humans generate as much power as does the planet itself
(i.e., 40 TW).
Some years ago, I used to teach a course on Energy Conversion
Processes. Toward the end of the term, I presented this 18-terawatt
challenge for discussion. (It was 12 TW in those days.) One of the
students said nothing, but merely pointed to the sky. The energy
source that is certain to last beyond the lifetime of the human spe­
cies is solar energy. Almost everything we do in terms of energy is,
142 ENERGY: THE BASICS

in some way, solar. In addition to direct conversion of solar energy,


solar drives the wind and water cycles, grows the plants, and, eons
ago, grew other plants that have given us fossil fuels. The solar
energy flow to our planet is so large that, if it could all be captured
and transformed to useful work, we could satisfy our current annual
energy needs with incoming solar energy in an hour.11 Perhaps
continued improvements in photovoltaic energy conversion, com­
bined with the development of improved energy storage technolo­
gies (including reductions in the cost of both), would allow
deployment of decentralized, small-scale systems that combine
photovoltaics with batteries almost anywhere in the world. These
systems would allow any impoverished society to bypass steam tur­
bines, large central electricity-generating stations, the electricity
grid, and all their assorted paraphernalia.
If a person living in a remote, impoverished spot in Asia or
Africa had the money, he or she could, right now, buy a smart-
phone as a first piece of communication technology. There is no
requirement that such a person first buy a telegraph, and then a
landline telephone connected to a central switchboard, a two-way
radio, then a dial telephone and, later, a dial-up computer
modem. That person goes from absolutely nothing to a smart-
phone in one step. We should push for the analogous situation in
energy development. Those living in impoverished countries
should be able to aspire to move directly to the most modern,
efficient, low-emission energy technologies without themselves
having to start with Watt & Boulton steam engines, and then
wait for their own versions of Michael Faraday and Charles
Parsons to come along.
Small, simple, decentralized, inexpensive electricity generation
offers advantages that could provide immediate improvements in
quality of life: electric lights and electric motors. Lighting provides
safety for working or being outdoors after dark. It provides a way
to read after dark, which means that even people who must work
full time could study during the evening, toward improving them­
selves and possibly moving into a better job. Electric motors can
operate water pumps, for improved watering of livestock and irri­
gation of crops. Electric motors can operate refrigerators and free­
zers. In addition to food preservation, another vital application of a
refrigerator is storing the many medicines and vaccines that need to
be refrigerated. Being able to do this in remote rural villages can
have an immediate impact on public health.
THE 18-TERAWATT CHALLENGE 143

At the same time, we cannot forget the vital role to be played by


energy conservation. Conserving, or curtailing, energy use is certainly
the least expensive way to make energy available. For the average energy
consumer, it’s a money-maker—buy less gasoline, use less electricity.
John Hofmeister, former president of Shell Oil Company and
head of the United States-based organization Citizens for Afford­
able Energy, described the solutions to future energy needs in terms
of the ‘four mores.’12 These are: more energy from all sources; more
efficiency in producing and using energy; more environmental pro­
tection; and more infrastructure—the technologies to transport or
transmit energy from wherever it is produced to the places where it
is needed to be used.
Without doubt we will need more energy. We can start getting
more energy right now, by increasing the efficiency of how we use
the energy we already have, and conserving it in use. We will need
to increase the contribution from energy sources that today are
relatively small contributors in many places—wind, water, and bio­
fuels, as examples. We need new ideas in energy science and new
breakthroughs in energy technology. And perhaps the ultimate
challenge is that, we need to find these additional 18 terawatts with­
out destroying our environment.
Finally, we can all profit from the advice of one of the great fig­
ures of our time, Nelson Mandela: ‘It always seems impossible until
it’s done.’13

NOTES
1 theworldcounts.com/stories/current-world-energy-consumption.
2 Wilson, Edward O. The Social Conquest of Earth. Liveright Publishing:
New York, 2012, p. 104.
3 Cassoret, Bertrand. Energy Transition. CRC Press: Boca Raton, 2021, p. 126.
4 Levin, Kelly, Cashore, Benjamin, Bernstein, Steven, and Auld, Graeme
‘Overcoming the tragedy of super wicked problems: constraining our
future selves to ameliorate global climate change’. Policy Sciences, Vol.
45, pp. 123–152, 2012.
5 Rhodes, Richard. Energy. Simon and Schuster: New York, 2018, p. 332.
6 Gertner, Jon. The Idea Factory. Penguin Books: New York, 2012, p. 355.
7 There is another alternative, which is utterly apocalyptic: to shrink the
world’s population. Short of all-out nuclear war or a pandemic so ferocious
as to make the 1918 influenza outbreak seem like a mere speed bump, there
is no rational, humane way to accomplish this on any reasonable time scale.
8 Clavey, William. ‘Solid state batteries have moved one step closer to rea­
lity.’ TopSpeed: https://www.msn.com/en-us/autos/news/solid-state-batterie
s-have-moved-one-step-closer-to-reality/.
9 Smil, Vaclav. Numbers Don’t Lie. Penguin Books: New York, 2021, p. 297.
144 ENERGY: THE BASICS

10 Smil, Vaclav. How the World Really Works. Viking: London, 2022, p. 5.
11 ucdavis.edu/climate/definitions/how-is-solar-power-generated.
12 Hofmeister, John. ‘The promise and reality of energy independence.’
Hydrocarbon Processing, February 2012, pp. 42–44.
13 https://www.socratic-method.com/quote-meanings/nelson-mandela-it-always­
seems-impossible-until-its-done.
APPENDIX

The Rules of the Game

THE LAWS OF THERMODYNAMICS


The Zeroth Law: When a body or system A is in thermal equilibrium with
another body or system B, and at the same time A is in thermal equili­
brium with C, then B and C will be in thermal equilibrium with each other.

The First Law: Energy can never be created or destroyed. (But it


can change its form or location.) A machine of greater than 100
percent efficiency is impossible.

The Second Law: Heat cannot flow spontaneously from cold to hot,
unless work is also done on the system. All energy conversions result
in the dissipation of low-temperature heat that cannot be used to do
work. A machine of exactly 100 percent efficiency is impossible.

The Third Law: It is impossible to cool a body to absolute zero in a


finite number of steps. (Basically, it is impossible to attain a tem­
perature of absolute zero.)

The First Law says that you can’t win; the Second Law says that
you can’t even break even.

THE ENERGY BALANCE


The difference between the energy put into a system and the energy
taken out (or emitted) is somehow stored in the system.
EIN  EOUT ¼ ESTORED
146 APPENDIX

EFFICIENCY
Efficiency is commonly expressed as a percent. The equation is:

P ercent efficiency ¼ work output=work input x 100

For heat engines, this can be expressed mathematically as:


Efficiency ¼ ðThigh  Tlow Þ=Thigh

where Thigh and Tlow represent the high and low temperatures,
respectively, expressed in kelvins. Using this equation, the numerical
value of efficiency will be a fraction having a value between 0 and 1.
Often it is more convenient or easier to express efficiency as a percentage,
rather than as a fraction, which is readily done by multiplying by 100.

THE GAS LAWS


Avogadro’s Law: For a gas held at constant pressure and constant
temperature, the volume is directly proportional to the quantity of gas.
V /n

where n is the quantity expressed in moles (i.e., the actual weight


divided by the molecular weight).
Boyle’s Law: For a fixed quantity of gas held at a constant tem­
perature, the product of pressure and volume is a constant.

PV ¼ k

where P is pressure and V, volume.


Boyle’s Law is also known, especially in France, as Mariotte’s Law.
Charles’s Law: For a fixed quantity of gas held at a constant pressure,
volume is directly proportional to temperature expressed in kelvins.
V /T

or
V ¼ kT

where k is a constant of proportionality, T is the absolute temperature,


and V is the volume. All gases increase or decrease in volume by the
same amount when the temperature is changed.
APPENDIX 147

Charles’s Law is also known as Gay-Lussac’s Law.


The Ideal Gas Law: Combining the equations above, the pressure,
volume, and temperature of a fixed quantity of gas are related by

P V ¼ nRT

Here R is the ideal-gas constant. The units of R depend on the units


used for the other variables. R can be expressed as, for example,
8.314 liters · kilopascals / kelvins · moles.

ELECTRICITY
Ohm’s Law: Current is equal to the potential difference divided by
the resistance:
Current ¼ Potential  Resistance
or, equivalently,
Potential ¼ Current  Resistance
Volts ¼ Amps  Ohms

Power is expressed as the product of current and potential:

Power ¼ ðCharge  TimeÞ  Potential


Power ¼ Current  Potential
Watts ¼ Amps  Volts
The relationship among power, current, and resistance is:
Power ¼ Resistance  ðCurrentÞ2
Watts ¼ Ohms  ðAmpsÞ2

Then for a transformer:


ðPowerÞin ¼ ðPowerÞout
ðVoltage=CurrentÞin ¼ ðVoltage=CurrentÞout

MASS-ENERGY EQUIVALENCE

E ¼ mc2
where E is energy, m is mass, and c is the speed of light.
SUGGESTED FURTHER READING

Collectively, these sources emphasize various topics, and represent various view­
points. Most are aimed at interested readers who are not specialists.
Amrith, Sunil. The Burning Earth. W.W. Norton & Company: New York, 2024.
Cassoret, Bertrand. Energy Transition. CRC Press: Boca Raton, 2021.
Clancy, Edward P. The Tides. Anchor Books: Garden City, NY, 1969. Chapter
VIII discusses using the energy of tides to generate electricity, related to
Chapter 12 here.
Conklin, Jeff. Dialogue Mapping. John Wiley & Sons: Chichester, 2006. A
good introduction to the concept of wicked problems.
Coopersmith, Jennifer. Energy: The Subtle Concept. Oxford University Press:
Oxford, 2010.
Edison Tech Center. Generators and dynamos. edisontechcenter.org/generators.
html.
Environmental Protection Agency (U.S.). Overview of greenhouse gases. https://
www.epa.gov/ghgemissions/overview-greenhouse-gases. Washington, D.C., 2024.
Goodman, Ruth. The Domestic Revolution. Liveright Publishing Corp.: New
York, 2020. Discusses the remarkable societal effects of switching from wood
to coal for domestic heating and cooking in Britain.
Gray, Theodore. Engines. Black Dog & Leventhal Publishers: New York, 2022.
Jorgensen, Timothy J. Strange Glow. Princeton University Press: Princeton, 2017.
Jorgensen, Timothy J. Spark. Princeton University Press: Princeton, 2021.
Moore, Andrew. The Decarbonization Delusion. Oxford University Press:
New York, 2023.
Nahin, Paul J. In Praise of Simple Physics. Princeton University Press: Princeton,
2016.
Odum, Howard T. Environment, Power, and Society. Wiley-Interscience: New
York, 1971.
Rhodes, Richard. Energy. Simon and Schuster: New York, 2018.
Rogers, Tom. Insultingly Stupid Movie Physics. Sourcebooks Hysteria: Naperville,
IL, 2007.
Sen, Paul. Einstein’s Fridge. Scribner: New York, 2021.
Smil, Vaclav. Creating the Twentieth Century. Oxford University Press: New
York, 2005.
Smil, Vaclav. Energy and Civilization. MIT Press: Cambridge, 2017.
SUGGESTED FURTHER READING 149

Smil, Vaclav. How the World Really Works. Viking: London, 2022.
Smil, Vaclav. Numbers Don’t Lie. Penguin: New York, 2022.
Solomon, Susan. Solvable. University of Chicago Press: Chicago, 2024.
Streever, Bill. Cold. Little, Brown & Company: New York, 2009.
Streever, Bill. Heat. Little, Brown & Company: New York, 2013.
Streever, Bill. And Soon I Heard a Roaring Wind. Little, Brown & Company:
New York, 2016.
Violett, Pierre-Louis. ‘From the waterwheel to turbines and hydroelectricity.
Technological evolution and revolutions.’ Comptes Rendus Mecanique, Vol.
345, pp. 570–580, 2017.
Wohlleben, Peter. The Heartbeat of Trees. Greystone Books: Vancouver, 2021.
Wyn Jones, R. Gareth. Energy: The Great Driver. University of Wales Press:
Cardiff, 2019.
INDEX

Note: Locators in italics refer to figures. Locators followed by “n” refer to footnotes.
18-terawatt challenge, 139, 141; energy use, 143; wicked problem, 140

absolute zero, 145; Carnot efficiency, atoms, 11, 28, 35, 71, 73; hydrogen,
33; quest for, 32 125; kinetic energy of, 34; nuclei
acidic precipitation, 65 of, 72
acid rain, 65 automobile engines, 40–41, 96
activation energy, 37, 45, 64 Avogadro’s Law, 146
‘aero-derivative’ turbines, 48
air, 12, 42–43, 47, 53, 65, 93, 110, background radiation, 79
131, 135; compressor, 46; flows, Bacon, Francis Thomas, 125
25; mass of, 24, 47, 110; Balbina hydroelectric plant in
moisture-laden warm, 23 Brazil, 106
alcohol, 96; fuels, 95; grain, 96 batteries, 56–57, 59n5, 119–121, 142;
α-rays, 71 battery-operated electric motors,
alternating current (AC), 57, 122; chemical processes in, 83;
59n5, 87 lithium-ion, 121–123; primary, 124;
ampere (amp), 51, 57, 58n1 rechargeable, 87; secondary, 87
Ampère, André, 51, 57 Becquerel, Henri, 70, 84
anode, 121, 123 β-rays, 71
anthropogenic CO2 source, 135 Bethe, Hans, 82
anthropogenic greenhouse effect, biofuels, 48, 92, 97–98, 135–136, 143
130, 131 biogas, 98
ash, 65–66, 95, 114 biomass, 81, 90–92; energy sources,
atmosphere, 11, 30, 40, 46–47, 65, 78, 92; ethanol, 95–97; forms of, 116;
83, 101, 110, 131; CO2 in, timescale of renewables, 98;
131–135; composition of, 17, 130; wastes, 98; wood, 93–95
of Earth, 24; loss of infrared Black, Joseph, 62–63
radiation from, 129; pressure boiler, 39, 64–66, 68
differences regions in, 25; principal Boyle’s Law, 146
components of, 129; sun’s energy Brazilian ethanol program, 97
absorption by, 83; volcanic
eruption affects, 131; water vapor Callendar, Guy, 132
in, 130 caloric, 37–38n8
Atomic Energy Commission, 80n8 Calories (unit), 11
atomic number, 71 calorimeter, 11, 13
INDEX 151

carbon, 12, 40, 67; carbon-neutral copper, 52–53, 55–56, 86


nature of biomass, 90; carbon-rich crude oil, 41
char, 65; emissions, 78, 108, 115; Cunene River in Angola, 21
fibers, 115–116; global carbon Curie, Marie, 70, 72
cycle, 40, 134–135; natural, 132 current, 51–54, 56–57, 147;
carbon capture and storage (CCS), alternating current, 57; direct
136–137 current, 57; electric, 51–52,
carbon dioxide (CO2), 11–13, 30, 40, 55, 57
66–67, 130, 132–133, 136–137; cybernetics, 31
concentration in atmosphere, 134,
135; sequestration, 136 da Vinci, Leonardo, 25, 100
carbon dioxide emissions, 78, 90; in De architectura (Vitruvius), 21
China, 131; in United States, 131; ‘decarbonizing’ concept, 135
wood and, 94–95; zero, 89 deforestation, 94, 134
Carnot efficiency, 33, 43–44, 61, 112 diesel engines, 42–45, 87, 94–95
Carnot, Sadi, 32, 36, 61 diesel fuel, 42–45
cathode, 121, 123 Diesel, Rudolf, 43
cellulose, 97, 131 digestion, 13–14
cellulosic ethanol, 97 direct current (DC), 57, 59n5, 87
chain reaction, 75; controlled nuclear direct solar electricity generation, 84
fission, 75; fission, 76; nuclear, distillation, 41, 97
76; sustained nuclear, 77; district heating, 68, 94, 98
uncontrolled, 75 Domesday Book, 22
charcoal, 12, 93 doping process, 85
Charles’s Law, 146 Du Fay, Charles, 50–51
chemical energy, 30, 41, 66, 91,
93, 125 earthquakes, 106, 139
Citizens for Affordable Energy, 143 eco-fuel, 98
Clausius, Rudolf, 37 Ecuador, electricity in, 103
climate, energy and: consequences of Eddington, Arthur, 82
global warming, 134; Earth’s heat Edison, Thomas, 122
balance, 128–132; evidence for efficiency, 31–32, 145–146
global change, 132–134; global Einstein equation, 74, 82
carbon cycle, 134–135; policy electrical generator, 57–58, 75, 101
options, 135–137 electrical potential, 51, 85, 120, 123
coal, 2, 12, 13, 32, 39, 48, 62, 64–65, electrical potential energy, 51, 56,
90, 105, 136; burned, 94; chemical 66–67, 85, 100, 119, 121
potential energy in, 66; co-firing electricity, 1–3, 8, 50, 55, 57, 107,
biomass with, 95; tar, 43 118, 147; to consumers, 53–55;
co-generation, 68–69 early electricians, 50–51; in
combined-cycle plants, 68 Ecuador and in Norway, 103;
combined heat and power (CHP), 68 in Ecuador and Norway, 103;
compressed natural gas, 95 electrical generator, 58; generation,
compression ratio, 43 45, 60; generation from hydro,
conductance, fundamental laws of, 52 103; matter, 51; megawatt-hours
conduction, 35–36, 52 of, 122; metals, 55–56; from
conservation of energy, 13 nuclear plants, 79; resistance, 52;
control rods, 76–77 reverse energy diagram for elec­
convection, 35–36, 128–129 trical system, 56, 56; from solar
cooling towers, 66, 68 energy, 84–89; solar generation of,
152 INDEX

81; Voltaic pile, 56–57; from wind, ethanol, 90, 95–97


110–117 ethyl alcohol see ethanol
electric motors, 55, 57, 122, 142 exhaust gas, 47
electric vehicles (EVs), 118, 122
electrified substance, 50 Faraday, Michael, 57, 60, 100, 121, 142
electrochemical cell, 56 farming, 1–2, 134
electro-fuels (e-fuels), 125–126 fermentation, 96–97
electrolyte, 121 Fermi, Enrico, 70, 74
electromagnetic induction, 57–58, fertilizers, 1, 92, 97
100, 121 Feynman, Richard, 7
electromotive force, 52 fire pistons, 43
electrons, 55, 71–73, 84–86, 121 firewood, 2, 83
energy, 1, 6; and climate, 128–137; First Law of Thermodynamics, 17n6,
conservation, 132, 136, 143; crops, 35–36, 145
81, 90, 95, 97; density, 44, 93, 95, fission, 74, 77, 82
120–122; diagram for gravitational fission products, 74, 76
system, 22; gravitational potential, fly ash, 65
21–23, 32, 45, 51, 100–101, 123; foods, 1–2, 11, 13; composition of,
and nations, 2–4; needs, 1–2; and 12; energy balance, 15
power, 7–8; resources, 4, 39, 49, Foote, Eunice Newton, 130
140; science and technology, 4–5; fossil fuels, 39–40, 64, 83, 92, 132,
of wind and water, 2; and work, 7, 135, 140, 142; fossil-fuel-fired boi­
8–9; see also nuclear energy lers, 78; fossil-fuel combustion,
energy balance, 14, 24, 71, 128–129, 133; see also coal; petroleum;
145; equation, 128; food, 15; gasoline
greenhouse effect, 17; waste four-stroke Otto cycle, 41, 42
heat, 16 four-stroke principle, 41
energy consumption: growth in, 3; in Fourier, Joseph, 130
terms of GDP, 141; in United Franklin, Benjamin, 50–51
States, 140 fuel(s), 12; alcohol, 95; cells, 124–125;
energy return on energy invested combustion, 64; consumption,
(EROEI), 49, 92, 94; for biofuels, 46–47, 133, 136; e-fuels, 125–126;
92; for ethanol, 97 fossil, 39–40, 64, 83, 92, 132, 135,
energy storage: batteries, 119–121; 140, 142; hog, 93; liquid, 91, 95;
e-fuels, 125–126; hydrogen, petroleum fuels and engines, 39
124–125; lithium-ion batteries, furnace, 13, 64
121–123; need for, 118–119; fusion, 82; latent heat of, 63;
pumped storage, 123–124; nuclear, 84
technologies, 142
engine, 30, 32–33, 42, 60; Diesel’s Galvani, Luigi, 55
design, 43–44; heat, 29, 36; internal γ-rays, 71
combustion, 3; jet, 46, 122; knock, gas(es), 33, 43, 95; laws, 146–147;
42; Newcomen’s design, 31; steam, natural, 39–40, 64
2, 26, 31, 111; turbojet, 46; Watt’s gasoline, 2, 14, 41–42, 96, 122;
design, 31, 39 gasoline–air mixture, 42;
engineering, 25, 141; calculations, 5; shortages, 2; see also petroleum
chemical, 97; studies in, 10 gas turbines, 64, 67; aero-derivative,
enhanced greenhouse effect, 130, 67; engine, 46
131–132 Gay-Lussac’s Law see Charles’s Law
entropy, 36–37 Gertner, Jon, 140
INDEX 153

glass houses, 137n1 hydrologic cycle, 23–24, 101, 123


global carbon cycle, 40, 40, 134–136 hydropower, 100, 102–104, 108
global climate change, 17, 81, 104, 114 hydroturbines, 79, 107
global warming, 130, 134, 137;
concerns, 118; problem, 137 Ideal Gas Law, 147
global warming potential (GWP), 130 ignition stroke, 42
glucose, 97 inclined plane, 8
Graham Bell, Alexander, 84 Indian Ocean earthquake (2004), 139
grain alcohol see ethanol indirect solar conversion, 84
Grand Coulee Dam, 108 infrared radiation, 128–131
graphite, 121, 123 internal combustion engine (IC
gravitational potential energy, 21–23, engine), 3, 39, 44, 46, 95, 122
32, 45, 51, 100–101, 123 International Energy Agency, 136
gravitational system, 28, 34 inverter, 87
Gray, Stephen, 50 isotopes, 71, 74, 76, 132
greenhouse effect, 5, 17, 129–131
greenhouse gases, 114, 129–130, 134 jet: engines, 45–48, 67; fuel, 45–48;
gross domestic product (GDP), 3; thrust, 47
growth, 4; relationship with per joule, 38n9
capita energy use, 3 Joule, James Prescott, 34
ground state, 137n2
Kelvin, Lord 11
Hahn, Otto, 74 Kelvin scale, 33
heat, 28–29, 68–69, 75, 145; flows, kilojoules (kJ), 11
32, 35–36, 61, 128; release rate, 65; kilowatt-hour (kWh), 53, 116
to work, 34, 36 kinetic energy, 19–21, 24, 27n3, 28,
heat engine, 29, 36–37, 61, 146; 57, 113; conversion, 102, 111; of
efficiency, 33; temperature water molecules, 30, 100
difference within, 32 kinetic isotope effect, 137n7
heat transfer, 64; convection, 35;
radiation, 36 landfill gas, 98
higher-boiling petroleum products, 43 land yacht, 27n9
highly enriched uranium (HEU), 76 latent heat, 62–63
high-pressure steam, 61–62 Lavoisier, Antoine, 12
high temperature, 28, 46, 61, 63 Laws of thermodynamics, 145
high-to-low pressure flow, 25 lead-acid batteries, 120
Hofmeister, John, 143 levers, 8
hog fuel, 93 lighting, 1, 55, 57, 90, 122, 142
hominids, 8 liquid fuels, 91, 95; see also ethanol;
hot houses see glass houses gasoline; diesel fuel; jet fuel
Hoyland, Graham, 120 lithium-ion batteries, 120–123
human and animal muscles, 2 ‘lost’ energy, 36–37
human energy, 10; conservation of, low-enriched uranium (LEU), 76
13; digestion, 13–14; fuels, 12; low temperature, 28, 32–33, 35, 44,
released from foods, 11; see also 53, 61, 145
energy lumbering, 93
hybrid energy system, 87
hydroelectricity, 100, 104, 106 magnetic resonance imaging (MRI), 53
hydroelectric plants, 101–103, 105 magnetism, 57
hydrogen, 12, 65, 67, 82, 124–126 Mandela, Nelson, 143
154 INDEX

mass-energy equivalence, 147 73; nuclear reactions, 72;


mass number, 71, 73–74 radioactivity, 70, 72
matter, 28–29, 33–34, 36, 51, 110 nuclear fusion, 82, 125
Medieval waterwheels, 100 nuclear plants, 64, 77, 103, 113–114;
Meitner, Lise, 74 commercial, 78; electricity from, 79
metals, 50, 55–56, 107; see also nuclear power plants, 75; isotopes,
copper; uranium 76; PWR, 77; thermal energy, 78
methane, 67, 98, 126, 130–131 nuclear reactions, 71–72
micro-hydro projects, 102 nuclear reactors, 70, 78
mineral carbonation, 136 nuclei/nucleus, 71–75, 82
molecules, 11, 28, 34, 36, 41; gaso­ nucleons, 71–73
line, 41; of greenhouse gases, 129; nutritional calories, 11
kinetic energy of water, 30; oxygen,
13, 65; water, 124 octane ratings, 42
Mollet, Joseph, 42 Ohm, Georg Simon, 52
money, 4, 7, 14–15, 142 Ohm’s law, 52, 147
moving water, 19, 21, 24, 106 oil(s), 40, 44, 90; 1973 oil embargo by
municipal solid waste (MSW), 98 OPEC, 48; crude, 41; peanut, 44;
see also petroleum
nacelle, 111 on-site photovoltaic systems, 87
natural brines, 136–137, 138n11 organic economy, 92
natural carbon, 132 Organization of Petroleum Exporting
natural gas, 2, 39–40, 48, 64, 67, 93, Countries (OPEC), 48
95, 97, 131, 136 Ørsted, Hans Christian, 56–57
natural greenhouse effect, 130 Otto, Nikolaus, 41
near-zero net carbon emissions,
118–119 Papin, Denis, 30
negative electricity, 51 Parsons, Charles, 60, 142
net photosynthetic production Parsons turbine, 60–61
(NPP), 92 particulate matter (PM2.5), 65
neutrons, 71, 73–77 peanut oil, 44
Newcomen engine, 31 petrol see gasoline
Newcomen, Thomas, 30 petroleum, 39–40, 64; activation
Newton’s third law of motion, 46 energy, 45; and derivatives, 39–41;
nitrogen oxides, 65–66 diesel engines and diesel fuel,
‘non-carbon’ sources of energy, 136 42–45; EROEI, 49; fuels and
‘non-fossil’ sources of energy, 136 engines, 39; gasoline, 41–42;
Norway, electricity in, 103 internal combustion, 39; jet
not in my back yard! syndrome engines and jet fuel, 45–48; OPEC,
(NIMBY syndrome), 79 48; petroleum-derived liquid fuel,
n-type semiconductor, 85, 86 122; products, 48
nuclear binding energy, 72–73, 78, 82 photosynthesis, 11, 40, 83, 90–91, 97,
nuclear energy, 70; nuclear 125, 131, 136
controversy, 78–79; nuclear photovoltaic(s), 84; cost of
fission, 70–75; nuclear power photovoltaic electricity, 88;
plants, 75–78; see also energy effect, 80n2; installation, 87;
nuclear fission, 70, 74; chain reaction, technology, 86
75; Einstein’s equation, 74; energy photovoltaic cell (PV cell) see solar cell
diagram for nuclear process, 72; piston, 30–31, 36, 39, 41–42; fire, 43;
nuclear binding energy, curve of, piston-in-cylinder IC engines, 39
INDEX 155

plastics, 50, 135 soil erosion, 94


plutonium (94Pu239), 79 solar cell, 86–87
positive electricity, 51 solar energy, 11–12, 24, 82–84, 88,
post-industrial economy, 4 114, 125; direct conversion of, 142;
post-industrial society, 4 electricity from, 84–89; see also
power, 6–8, 30, 147; failure, 1; thermal energy
nuclear power plants, 75–78 solar generation of electricity, 81
pressurized water reactor (PWR), 77 solar radiation, 24, 83, 86, 130
primary batteries, 124 solid fuels, 9 see also coal,
primary cells, 119 refuse-derived fuel; wood
protons, 71–74 source rocks, 40
p-type semiconductor, 85, 86 static electricity, 50
pulleys, 8 stationary fuel cells, 125
pulverized coal firing, 65 steam, 30; combined-cycle plants, 68;
pumped storage, 119, 123–124 electricity from, 60; engines,
31–32, 43, 55; gas turbines, 67;
radiation, 35–36, 70, 78; α-radiation, heat, 68–69; latent heat, 62–63;
73; β-or γ-radiation, 74; infrared, production, 63–67; turbines,
129; solar, 24, 83, 86, 130; 60–62; work from, 30–31
spontaneous emission of, 72; Stevin, Simon, 25–26
of sun, 24 storage batteries, 120
radioactive materials, 72, 78 Strassman, Fritz, 74
radioactivity, 70, 72–74 Strauss, Lewis, 78
reciprocating steam engine, 60 Suess effect, 133
refuse-derived fuel, 98 Suess, Hans, 133
renewable energy, 26, 81–82, 98, 118; sugar cane, 91, 96–97
see also ethanol; hydropower; sulfur oxides, 65–66
refuse-derived fuel; solar energy; sun, 24, 83, 130; energy absorption
wind; wood by Earth, 129; energy release by
reservoir, 16, 40, 61–62, 104–106 nuclear fusion in, 84
resistance, 6, 19, 46, 52–55, 147 sunlight, 11, 25, 40, 81, 83, 88; direct,
resistivity, 52–53 88; energy in, 84, 128
respiration, 12 superconductors, 53, 58n4
reverse energy diagram for electrical sustainable energy sources, 81
system, 56, 56 synthetic ammonia, 135
Rhodes, Richard, 140 system, definition of, 10, 20
rubber, 50, 72
Rumford, Count, 34 temperature, 28–29, 62; difference,
Rumford’s system, 34 32, 69; scales, 34
Thales of Miletus, 50
sailing ship, 25 thermal energy, 30, 32, 68; flow of,
secondary batteries, 87 28; in high-pressure water, 78;
secondary cells, 120 point of zero, 29; of water or
Second Law of Thermodynamics, 35, steam, 84
120, 145 thermal equilibrium, 37n2
semiconductors, 85t thermal systems, 28, 29
shaft, 19–20, 31, 60–61, 100, 111 Third Law of thermodynamics, 145
shielding, 78 Thompson, Benjamin, 34
silicon, 85 Thomson, William, see Kelvin, Lord
single photovoltaic cells, 86 Three Gorges plant, 101–102, 106, 108
156 INDEX

tidal energy, 103, 106–107 30; cycles, 23–24; dam problem,


tower mill, 25 104–106; electricity from, 100;
turbine engines, 48 hydroelectric plants, 101–103;
turbines, 45–46, 48, 62, 67–68, 100 hydropower, 103–104; tidal energy,
turbogenerator, 62 106–107; turbines, 100–101;
turbojet engine, 46–47 vapour, 67; wind and, 26
turret mill, 25 waterfall, 19–20, 20, 51, 104
Tyndall, John, 130 water-tube boiler, 64
waterwheels, 8, 20, 20–22, 26;
United States (US), 8; commercial advantage of, 23; designs of, 22
hydro plants in, 101; energy Watt & Boulton steam engines, 142
consumption in, 140; ethanol Watt, James, 31
business in, 96; using solar Whiting plant in Portage County
energy, 88; waste-to-energy plants (Wisconsin), 101
in, 98 wicked problem, 140
uranium, 70–71, 74, 76 wind, 110–111; arguments pro and
con, 114–117; electricity from,
Vitruvius, 21 110; energy, 110; turbines,
volcanic eruption, 131, 139 111–113; and water, 26; work
volcanoes, 131, 139 from, 24–26
Volta, Alessandro, 51, 55 windmills, 25–26, 111
voltage, 52–54, 86 Wohlleben, Peter, 125
Voltaic pile, 56–57, 119, 121–122 wood, 50, 93–95, 125
von Guericke, Otto, 50 work, 6–9, 10, 28, 55, 58; to heat, 36;
von Mayer, Julius, 13 from steam, 30–31; from wind,
24–26
waste(s), 79, 90; biomass, 98; MSW,
98; waste heat, 16, 55, 69, 77 zero net carbon emissions, 140
waste-to-energy plants (WTE zero point of thermal energy, 29
plants), 98 Zeroth Law of thermodynamics,
water, 11, 19, 30; benefits and 37n2, 145
complications, 107–108; boiling, zinc, 55–56

You might also like