CAP1 - Roe Et Al., 2018. Mathematics For Sustainability

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 83

Texts for Quantitative Critical Thinking

John Roe
Russ deForest
Sara Jamshidi

Mathematics
for Sustainability
Foreword by
Francis Edward Su
Texts for Quantitative Critical Thinking

apprecia@ucalgary.ca
Texts for Quantitative Critical Thinking

Series Editors:
Stephen Abbott
Middlebury College, Middlebury, VT, USA
Kristopher Tapp
St. Joseph’s University, Philadelphia, PA, USA

Texts for Quantitative Critical Thinking (TQCT) is a series of undergraduate textbooks, each of
which develops quantitative skills and critical thinking by exploring tools drawn from mathematics
or statistics in the context of real‐world questions. Topics are high sophistication, low prerequisite,
offering students of all disciplines the opportunity to build skills in the understanding, evaluation,
and communication of quantitative information. These are books about mathematical and statistical
thinking, not computation and procedure. Each book explores an application or idea in depth, offering
students in non‐STEM fields a focused, modern context to develop quantitative literacy, while students
with a technical background will gain a broader perspective to bridge beyond procedural proficiency.
Books in Texts for Quantitative Critical Thinking develop writing and communication skills, which
are essential to such cross‐disciplinary discourse and are highly‐transferable for students in any field.
These titles are ideal for use in General Education courses, as well as offering accessible quantitative
enrichment for an independent reader.

More information about this series at http://www.springer.com/series/15949

apprecia@ucalgary.ca
John Roe • Russ deForest • Sara Jamshidi

Mathematics for Sustainability

apprecia@ucalgary.ca
John Roe Russ deForest
Department of Mathematics Department of Mathematics
Pennsylvania State University Pennsylvania State University
University Park, PA, USA University Park, PA, USA

Sara Jamshidi
Department of Mathematics
Pennsylvania State University
University Park, PA, USA

ISSN 2523-8647 ISSN 2523-8655 (electronic)


Texts for Quantitative Critical Thinking
ISBN 978-3-319-76659-1 ISBN 978-3-319-76660-7 (eBook)
https://doi.org/10.1007/978-3-319-76660-7

Library of Congress Control Number: 2018935919

Mathematics Subject Classification (2010): 00A06, 00A69, 00A71, 91A35, 91B76, 97-01, 97F70, 97M99

© Springer International Publishing AG, part of Springer Nature 2018


This work is subject to copyright. All rights are reserved by the Publisher, whether the whole or part of the material is concerned,
specifically the rights of translation, reprinting, reuse of illustrations, recitation, broadcasting, reproduction on microfilms or
in any other physical way, and transmission or information storage and retrieval, electronic adaptation, computer software, or
by similar or dissimilar methodology now known or hereafter developed.
The use of general descriptive names, registered names, trademarks, service marks, etc. in this publication does not imply, even
in the absence of a specific statement, that such names are exempt from the relevant protective laws and regulations and therefore
free for general use.
The publisher, the authors and the editors are safe to assume that the advice and information in this book are believed to be
true and accurate at the date of publication. Neither the publisher nor the authors or the editors give a warranty, express or
implied, with respect to the material contained herein or for any errors or omissions that may have been made. The publisher
remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Cover illustration: Cover image of Lake Erie courtesy of NASA Earth Observatory (Joshua Stevens, using Landsat data from
the U.S. Geological Survey)

Printed on acid-free paper

This Springer imprint is published by the registered company Springer International Publishing AG part of Springer Nature.
The registered company address is: Gewerbestrasse 11, 6330 Cham, Switzerland

apprecia@ucalgary.ca
Contents

Contents i

Foreword iii

Before we begin v
0.1 To the Student . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . v
0.2 To the Instructor . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . xiv
0.3 Acknowledgments . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . xv

I Fundamental Concepts 1

1 Measuring 3
1.1 Units and Measurement . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4
1.2 Scientific Notation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 23
1.3 Estimates, Precision, and Orders of Magnitude . . . . . . . . . . . . . . . . . . . . . 33
1.4 Communicating Quantitative Information . . . . . . . . . . . . . . . . . . . . . . . 44
1.5 Exercises for Chapter 1 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 59

2 Flowing 67
2.1 Stocks, Flows, and Equilibrium . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 69
2.2 Energy Stocks and Flows . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 81
2.3 Calculating Equilibrium States . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 95
2.4 Energy Flows in the Climate System . . . . . . . . . . . . . . . . . . . . . . . . . . 106
2.5 Exercises for Chapter 2 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 118

3 Connecting 127
3.1 Networks and Connections . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 129
3.2 Networks and Behavior . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 140
3.3 Feedback and Dynamic Networks . . . . . . . . . . . . . . . . . . . . . . . . . . . 151
3.4 The Exponential Model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 164
3.5 Exercises for Chapter 3 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 176

4 Changing 183
4.1 Logarithms and Change . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 185
4.2 Logistic Models and the Limits to Growth . . . . . . . . . . . . . . . . . . . . . . . 206

apprecia@ucalgary.ca
vi CONTENTS

4.3 Measuring Feedback Strength . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 221


4.4 Tipping Points . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 239
4.5 Exercises for Chapter 4 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 252

5 Risking 259
5.1 Understanding Data . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 261
5.2 Probabilities and Predictions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 278
5.3 Expectations and Payoffs . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 300
5.4 Assimilating New Information . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 317
5.5 Exercises for Chapter 5 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 337

6 Deciding 345
6.1 Market Perspectives and Large-Scale Change . . . . . . . . . . . . . . . . . . . . . 348
6.2 The Strange Behavior of Rational People . . . . . . . . . . . . . . . . . . . . . . . 360
6.3 The Tragedy of the Commons . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 373
6.4 After Math: Decision-Making and Ethics . . . . . . . . . . . . . . . . . . . . . . . . 388
6.5 Exercises for Chapter 6 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 402

II Case Studies 409

7 Case Studies 411


7.1 Mathematics and Persuasive Writing . . . . . . . . . . . . . . . . . . . . . . . . . . 413
7.2 The Changing Risks of Wildfires . . . . . . . . . . . . . . . . . . . . . . . . . . . . 419
7.3 Is Recycling Worth It, Really? . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 426
7.4 World Population Growth and the Demographic Transition . . . . . . . . . . . . . . 433
7.5 Genetic Engineering and the Future of Food . . . . . . . . . . . . . . . . . . . . . . 443
7.6 Nuclear Power Is a Commitment to the Future . . . . . . . . . . . . . . . . . . . . . 450
7.7 Using Electricity Efficiently at Home . . . . . . . . . . . . . . . . . . . . . . . . . . 458
7.8 Growth and Payback Time for Solar Energy . . . . . . . . . . . . . . . . . . . . . . 466
7.9 Energy Return on Energy Invested . . . . . . . . . . . . . . . . . . . . . . . . . . . 470
7.10 Exercises for Chapter 7 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 475

III Resources 481

8 Resources for Student and Instructor 483


8.1 Resources for Further Reading and Writing . . . . . . . . . . . . . . . . . . . . . . 483
8.2 Useful Numbers for Sustainability Calculations . . . . . . . . . . . . . . . . . . . . 483

List of Figures 491

List of Tables 498

Bibliography 501

Index 519

apprecia@ucalgary.ca
Foreword

I’m excited to introduce this book to you, because it may be different from any math text you’ve read
before. It will change the way you look at the world and enlarge the way you think about mathematics.
No longer will you be just a spectator when people give you quantitative information—you will
become an active participant who can engage and contribute new insights to any discussion. Just
look at the verbs that underlie the chapter titles: measure, flow, connect, change, risk, decide!
Here’s what stands out to me when I read this book: there are many math books that will feed you
knowledge, but it is rare to see a book like this one that will help you cultivate wisdom.
There is a deep difference between knowledge and wisdom. A knowledgeable person may be armed
with facts, but a wise person considers how to act in light of those facts. A knowledgeable person may
think an answer is the end of an investigation, whereas a wise person considers the new questions that
result. And a knowledgeable person might ignore the human element of a problem that a wise person
deems essential to understand. As the authors illustrate, mathematics that pays attention to human
considerations can help you look at the world with a new lens, help you frame important questions,
and help you make wise decisions.
Sustainability asks: how can we be wise stewards of Earth’s resources? One way or another this
question will impinge on some aspect of your life, if it hasn’t already. Sustainability is an economic
concern because resources are limited. Sustainability is a moral concern, because any scarcity of
Earth’s resources will harm the weak and vulnerable first. And sustainability is a scientific concern,
because we may have the power to improve the lives of those who will be affected.
I know that each of the authors shares a deep vocational commitment in bringing this book to you,
and as evidence, I’ll speak personally of the author I have the privilege to know as a friend: John
Roe, a man of deep grace and humility who made this book his highest professional priority while
battling a difficult illness. For him, this project grew out of a conviction and prayerful reflection that
his knowledge as a mathematician and an educator could be channeled into wise action on matters
that will impact us all.
The authors have poured their hearts into this remarkably important and timely book, and I hope
you will engage it with the same fervor. Because it concerns the world you live in, how you will need
to live in it, and the problems that you—yes YOU—can solve so that all of us can live in it well.
Francis Edward Su
Benediktsson-Karwa Professor of Mathematics, Harvey Mudd College
Past President, Mathematical Association of America

vii

apprecia@ucalgary.ca
Before We Begin. . .

0.1 To the Student


A Letter from the Authors
Dear Student,
The world that you are inheriting is full of bright possibilities—and also of big problems. Many
of the problems center on sustainability questions like “can this (key part of our social or economic
system) last?” or to put it in a way that has a little more math in it, “how long can this last?” For
example, modern society is a profligate consumer of energy, most of which is supplied by fossil fuels:
coal, oil, and natural gas. Fossil fuel supplies, though of vast and unknown size, are limited. How
long can they last? What’s more, there is a strong scientific consensus that the carbon dioxide gas
(also known as CO2 ) released by burning fossil fuels is affecting Earth’s climate, making it more
unstable. How much more CO2 can we emit before climate instability becomes dangerous? These are
big problems. Bright possibilities for addressing them include renewable energy sources like wind and
solar. No doubt you have heard all of these things before—as well as many other news stories about
“sustainability,” both positive and negative.
We started developing the “Mathematics for Sustainability” course, and writing this book, because
of three convictions:

• Many of the key choices that humans will have to make in the twenty-first century are rooted in
sustainability questions. These include choices that we must make together, as citizens, as well
as choices related to individual lifestyles.

• In a democracy, as many people as possible need to participate in well-informed discussion of


these sustainability questions. They are too important to be left to “experts.”

• We may engage with sustainability questions from a wide variety of perspectives, including
scientific, technological, political, ethical, and religious. For many of these discussions, we
need some knowledge of mathematics in order to participate in a well-informed way.

The aim of this book is to help you, the student, gain that mathematical knowledge and the ability
to apply it to sustainability questions.
You may not consider yourself a “math person.” Your studies may center on music or English or art
or education or architecture or agriculture.1 But if you want to find out for yourself what “the numbers
say”—not just to choose which “expert” you prefer to listen to—then this book is for you. Together,
we will find out how to model sustainability on local, regional, and global scales. We will learn about
1 Students in all these majors have succeeded in the course on which this book is based.

ix

apprecia@ucalgary.ca
x BEFORE WE BEGIN

measurement, flow, connectivity, change, risk, and decision-making. Some of the topics we discuss
will probably be challenging, perhaps even unsettling. Whatever conclusions you reach, this book will
prepare you to think critically about your own and other people’s arguments and to support them with
careful mathematical reasoning.
As citizens in a democracy, you will ultimately be the ones whose decisions will guide your world
toward a sustainable future. We wish you the very best in your studies, and as you participate in
building the sustainable society of the future.
John Roe, Russ deForest, Sara Jamshidi
August 2017

Sustainability—The Key Idea


In spring 2017, our Earth’s human population surpassed 7 12 billion. Here’s a question. What do you
imagine was the population of Earth two thousand years ago, at the beginning of the Common Era?
Demographers (scientists and historians who study population) obviously don’t know an exact
answer to this question. But they are able to make some good estimates, which are in the range of
200 to 300 million people (Section 8.2). That is to say, the number of people on the whole planet
twenty centuries ago was roughly the same as the number in the United States (U.S.) today. Or, to put
it differently, the population of our Earth has increased by twenty-five times over that period.
That population increase has not been a steady one. Most of the growth has occurred in the last
century. And many other measures of human activity follow a similar pattern. Take a look at the
graphs in Figure 1, which are taken from a book by Will Steffen [308]. These graphs show a pattern
of accelerating increase that mathematicians call exponential growth.2 This is important news: some
good (most societies have regarded large families and increased wealth as good, for example) and
some less so (Steffen’s book includes similar curves about pollution and overuse of resources, which
most would regard as bad). Both “goods” and “bads” have been growing exponentially, especially over
the past two hundred years, since the Industrial Revolution got into gear. Can this pattern continue?

Figure 1: Some measures of the “size” of humanity (from [308]).

Some think so. In July 2015, one presidential candidate declared that his objective for the United
States was “4 percent annual growth as far as the eye can see.” That is about the growth rate for the
2 See Section 3.4 for more about this concept.

apprecia@ucalgary.ca
TO THE STUDENT xi

curves in Figure 1. Others, though, look at similar data and see it differently. “[These] remarkable
charts. . . ,” writes Gus Speth, former dean of the Yale School of Forestry and Environmental Studies,
“reveal the story of humanity’s impact on the natural earth. The pattern is clear: if we could speed up
time, it would seem as if the global economy is crashing against the earth—the Great Collision. And
like the crash of an asteroid, the damage is enormous.” [305]. “Societies are now traveling together
in the midst of unfolding calamity down a path that links two worlds,” he continues. “Behind is the
world we have lost, ahead is the world we are making. . . . The old world, nature’s world, continues, of
course; but we are steadily closing it down, roping it off. It flourishes in our art and literature and in
our imaginations. But it is disappearing.”

The “old world” that Speth describes is a world in which


Earth appears to be huge, teeming with life, abundant, ex-
hilarating and dangerous. Humanity exists on the margin. In
the “old world,” to ask about humanity’s impact on nature
might seem absurd: much more important to worry about
nature’s impact on human beings (diseases? predators? food
shortages?) By contrast, the iconic image of the “new world”
is the Apollo astronaut’s view of Earth: the “blue marble”
(Figure 2), floating in space, gemlike and beautiful, yet cut
down to finite size by human achievement. In this finite world
it makes sense to ask: How long can we keep growing? Have
we already become too big? Can our complex society remain
diverse, active, and productive for an extended period of time?
Or could we overshoot the limits of our resources and then
decline, as many earlier civilizations have done [96]?
This is the sustainability question from which all other such
questions derive. We can put it another way by thinking of
the successive generations of humanity on this planet. Each Figure 2: The “blue marble.”
generation inherits Earth’s resources from its predecessors and
passes them on to its successors. In a lasting or sustainable
society, each generation would leave the world system in as good a shape as it found it; my
generation’s enjoyment of Earth’s bounty would not deprive the next generation (yours) of the
opportunity for similar enjoyment. Thus we arrive at the famous definition3 given in the Brundtland
Report [255] as early as 1987:

Definition 1

Sustainability is the ability of a social, economic, or ecological system to meet the needs
of the present generation without compromising the ability of future generations to meet
their own needs. A process or practice is sustainable to the extent that it contributes to the
sustainability of the social, economic, or ecological systems in that it is embedded.

It is important to recognize that working for sustainability does not mean just trying to keep things
as they are. “Things as they are” include patterns of privilege and inequality that deprive many
members of the present generation of their basic needs. A sustainable world system must continue
to extend the provision of these needs to a growing share of its population—ultimately, to all—even

3 The Bruntland Report in fact defines “sustainable development”; we have slightly modified its language to arrive at a

definition of “sustainability” itself.

apprecia@ucalgary.ca
xii BEFORE WE BEGIN

as it also works to ensure that these needs can be provided in a way that doesn’t ask future generations
to pay the bill. This double challenge, we believe, will define the century that you live in.
We should also recognize that questions about sustainability and about the needs of the present
and future generations are inherently value-laden. Particularly when we approach these questions on
a global scale, we should expect to encounter different value judgments concerning how human well-
being is defined and what are the basic needs whose provision should be sustained.

Think about it. . .

The Brundtland definition of sustainability, Definition 1, was written in 1987. Yet such
ideas appear many years earlier in the thought of indigenous peoples around the world. For
instance, the Iroquois Confederacy’s “Great Law of Peace,” which is older than the U.S.
Constitution, contains a clause that is often paraphrased as, “In every deliberation, we must
consider the impact on the seventh generation. . . even if it requires having skin as thick as the
bark of a pine.” Thus, Westerners’ recent “discovery” of the notion of sustainability might
be more properly described as a “rediscovery” of ideas that are rooted in many traditions
(including some of our own). How do you feel about this? Why might Western society have
lost touch with the sustainability idea for part of its history?

Sustainability and Resilience


The time scale on which we ask sustainability questions is a
long one—many generations. Some changes to our world are not
perceptible unless we take such a long view. The rate at which
Earth’s climate is changing, measured by global average surface
temperature, is no more than a couple of hundredths of a degree
per year: of little consequence from one year to the next but (as we
will see) hugely significant on longer time scales. But change does
not always come in tiny steps: our world also experiences sudden,
extreme events. (Think of Hurricane Harvey, which flooded Houston
in August 2017, doing damage initially estimated at $60 billion.)
Extreme events are unavoidable: the question is not how to avoid
Figure 3: Marine rescue operation in Galveston,
them, but how quickly we can bounce back from them. This “self-
Texas, following Hurricane Harvey, August 31,
2017. healing” capacity is resilience.

Definition 2

Resilience is the ability of a social or economic system to absorb and respond to a sudden
shock without damage to its core functioning.

Sustainability and resilience are not the same, but they are closely related: a system that is already
near the limits of sustainability may be pushed to collapse by a shock that a more resilient system
would easily have survived. In the same way, an infection that a healthy person would shrug off may
be fatal for someone whose immune system is already compromised by malnutrition. Many historical
examples can be found in [96].
One way to think about the relationship between sustainability and resilience is through the idea of
control. It is tempting to believe that the more elements of our socio-ecological system we can bring

apprecia@ucalgary.ca
TO THE STUDENT xiii

under human control, the better we will be able to steer it on a safe path and so the more sustainable
it will be. For example, wildfire management in the Western U.S., and elsewhere, focused for many
years on fire suppression: bringing fires under human control. This seemingly allowed development in
wildfire-threatened areas to be lasting, or “sustainable.” Only recently has it become clear that these
measures, which indeed halted many local fires, also set the stage for huge buildups of fuel that make
the forests less resilient to the possibility of a catastrophic burn. Similar ideas apply to flood control
measures like overflow dams and levees. In Section 4.4 we will develop a mathematical perspective
on tipping points—situations in which a system lacks resilience because a small shock can produce
an overwhelming response.

Ecosystem Services
In Speth’s “new world” there is a tight connection between the economy, environmental quality,
human well-being, and the well-being of the entire community of living and nonliving things that we
call Earth’s ecosystem. This requires a significant change in our thinking. From the perspective of the
“old world,” Earth was simply a source of abundance: nature provided. Our “new world” perspective
requires a deeper understanding of what it is that nature provides, and indeed of the fact that human
activity can no longer be so neatly distinguished from “nature” at all. Economic activity and human
prosperity are embedded in the natural world, and they depend on and are ultimately constrained by
the productive capacity of the whole Earth ecosystem (Figure 4)—which itself is constrained by the
rate at which energy arrives from the Sun.
In Figure 4, we should therefore think of the outer-
most oval, which represents the entire ecosystem of the
Earth, as of more or less unchangeable size. That means
the inner ones cannot grow too much. In other words, Earth’s Ecosystem

the definition of a sustainable society implicitly involves


Human Society
the uncomfortable idea of limits: limits on the “goods”
Sustainability

that are available to each generation and also limits on Economy


the capacity of Earth’s ecosystem to absorb the “bads”
generated by human activities. The recognition that such
limits exist sets the stage for the mathematical tools we
develop throughout this book. We need to be able to
interpret and use quantitative information to assess the
size and nature of such limits and so to reach well- Figure 4: Sustainability and limits. Redrawn from [65].
justified decisions about sustainability questions. If we
lack such tools, we’ll be tempted to believe either that there “are no limits” or that if limits exist, they
are so far off that they need be of no concern to us. That may be true in some cases, but in others we
may find that the limits are startlingly close.
The notion of ecosystem services [342] provides one way to conceptualize this dependence of
human activity on the planet’s ecosystems.

Definition 3

Ecosystem services are benefits that human society obtains from Earth’s ecosystems. These
services range from pollination of food plants and provisioning of freshwater and other
resources to climate regulation, soil production, and recreational opportunities.

Specific ecosystem services are not fixed once and for all: human activity may enhance or degrade
the capacity of Earth’s ecosystems to provide these services. Moreover, many ecosystem services

apprecia@ucalgary.ca
xiv BEFORE WE BEGIN

(such as clean air, clean water, climate regulation) are public goods—they cannot be “owned” in any
practical sense. While some ecosystem services are local or regional (think of the services provided by
a community park or a state forest), many are global in nature (like the pollution-absorbing capacity of
the atmosphere): effective management and stewardship therefore require international cooperation. In
Chapter 6 we develop mathematical tools to study such cooperation. This will allow us to understand
some of the pitfalls and difficulties in reaching mutually beneficial agreements.

Think about it. . .

Gaylord Nelson, former U.S. senator for Wisconsin and founder of Earth Day, wrote in
2002:
The economy is a wholly owned subsidiary of the environment, not the other
way around.
What do you think he means? Do you agree?

About This Book


A few words about the structure of this book. Part I is the longest part of the text. In it, we develop
mathematical tools and apply them to short examples. It is organized into six chapters corresponding
to key concepts that arise in the mathematical study of sustainability: measuring, flowing, connecting,
changing, risking, and deciding. Here is an overview of those concepts:
Chapter 1: Measuring. In this chapter we discuss how to measure, and how to express how big
some quantity is. From the point of view of sustainability, the kind of answer that matters is often not
some absolute number, but a comparison or a level of importance, so we’ll talk about how we might
judge whether some quantity represents something important, and how we might make a decision on
how to respond.
Chapter 2: Flowing. It’s common to talk about the balance of nature. But this image of a “balance”
can suggest something that is static, unchanging. That is not the way natural ecosystems operate.
Individual components of the system are constantly changing, even as the system as a whole maintains
its equilibrium. For example, consider a mountain lake. The water level in the lake may stay the same,
but the actual water in the lake today is not the same as the water that was there yesterday: new water
has arrived through rain and snowmelt, old water has left through runoff and evaporation. It’s the
balance between these various flow processes that keeps the water level—the stock—constant. This
chapter is devoted to exploring these concepts of flow and stock in detail.
Chapter 3: Connecting. Our lives are more interconnected now than at any time in history, and
not just through social media. Rather than most of our food being grown close to where we live, for
example, we have gotten used to obtaining food from supermarkets that are supplied by a production
and transportation network that reaches all over the globe. Energy supply, too, is a vast and complex
network of tanker routes, pipelines, electrical grid connections, and truck deliveries, to mention only
a few. Human-made networks like these are not the only ones: we have also grown much more aware
of the complexity of the natural networks that connect the web of life on earth. In this chapter we will
study the mathematical language that is used to understand these various kinds of networks.
Chapter 4: Changing. In this chapter we will look at examples in which stock-flow systems are
out of equilibrium—that is, how they respond to change. Although we start by studying a simple
model of continuous growth or decay, from the point of view of sustainability the important questions
arise when growth is limited by some external factor (such as the fact that we live on a finite planet).
We’ll study how the strength of feedbacks governs a system’s response to change, and how this leads
to the key idea of a tipping point—a moment when the system “flips” suddenly to a new state. Some

apprecia@ucalgary.ca
TO THE STUDENT xv

scientists are concerned that Earth’s climate system might be approaching one or more tipping points.
We’ll ask how it might be possible to tell.
Chapter 5: Risking. Any kind of realistic thinking about sustainability must consider likelihoods or
risks. Nuclear power generation does not produce any greenhouse gas emissions, and it is constant and
reliable, but what about the possibility of a catastrophic accident? This question asks us to balance the
near-certainty of a steady benefit against a small risk of disaster. It is hard to do so without numbers.
How likely are you to die from a nuclear reactor meltdown? The math involved in working with these
sorts of questions is called probability and statistics. In this chapter, we are going to learn about these
techniques and how they can help us make good decisions when faced with limited knowledge and
uncertain outcomes.
Finally, Chapter 6: Deciding is where the rubber meets the road. In the end, you and your
generation are going to have to make some sustainability decisions. These will range from the
personal, through the local, to the national and global. Human behavior does not always follow the
“rational” principles discussed in Chapter 5. Even if mine always did, I am not the only person
involved. Other people’s decisions interact with mine in a complicated way—we are in fact a
network of decision-makers, with no final authority. This decision-making interaction can be studied
mathematically, which we will do. The chapter concludes, however, with an extended reflection
on how our sustainability decisions can never be purely mathematical, but must also engage our
fundamental personal and ethical commitments.
In Part II (Chapter 7) of the book we provide a collection of Case Studies in which we apply
the mathematical tools developed in Part I to answer particular questions related to sustainability and
to explore extended examples. We believe that it is important that you, the student, learn to write
extended pieces of this sort, and a student writing requirement has been a major part of the course
on which this book is based. Why? This is how you build the ability to assess, develop, and present
quantitative evidence in support of your own ideas. These skills are vital to you as a future leader, as
an engaged citizen, and as an effective advocate for the things you care about. We don’t want you just
to learn mathematical techniques but also to be able to incorporate them in extended and persuasive
written pieces—pieces that might be published in a course blog, in a student or hometown newspaper,
or even on a national platform.
Finally, Part III (Chapter 8) of the book contains reference material: suggestions for further
reading, tables of useful data, and the list of figures, bibliography, and index. We suggest checking
Part III regularly, especially if you need some numerical information (such as the heat capacity of
water, or the amount of sunlight that falls on Earth, or the planet’s estimated coal reserves) to help you
answer one of the exercises or formulate an argument in one of your more extended written pieces.

apprecia@ucalgary.ca
xvi BEFORE WE BEGIN

Online Resources
This book has a website, http://math-for-sustainability.com, as well as an email address for
comments and suggestions, comments@math-for-sustainability.com. On the website you will
find many different kinds of supporting materials:
• Hints and solutions to selected exercises.
• Online calculators and examples. Especially later in the book we will describe many models for
sustainability-related processes that change over time (such models are called dynamic). The
best way to visualize such a model is also dynamic, like a video clip rather than a collection of
snapshots. That can’t be done on paper, but it can be done online, and the website uses several
different modeling tools to help you see the way our dynamic models evolve over time.
• Extended discussions of particular topics (refreshing the case studies in part II of the book,
some of which may become outdated as technology advances or the environment changes).
If you publish a written piece in a newspaper somewhere, please write to us (using the email
address above) and tell us! We’ll be happy to link to your writing from the book’s website.
• Corrections or updates. We’ve tried hard to make this the best book possible. But there are
bound to be some mistakes left. What’s more, some information will simply become outdated.
We’ll keep a list of corrections or updates on the website. Again, if you notice an apparent
mistake, please email us. You’ll be helping many future students by doing so.

Conclusion

We hope that through this book you will gain a clearer under-
standing of the sustainability issues that we humans face and of
some choices that we need to make. But let’s be clear: mathematics
cannot make these choices for us. As remarked above, the choices
human beings make on such fundamental questions reflect their
deepest ethical and personal commitments (compare Figure 5).
What mathematics can do, though, is to inform our choices by
making their likely consequences clearer. It can help us prioritize
issues by ranking them in terms of the relative size of the risk
they pose and the relative severity of their potential outcomes.
Understanding mathematics can help us avoid falling for some
plausible-looking “solutions” that really don’t achieve much or
Figure 5: Pope Francis has said: “These ancient
stories...bear witness to a conviction which we are even harmful. Finally, mathematics carries its own values also,
today share: that everything is interconnected, and like communicating clearly, reasoning logically, and considering
that genuine care for our own lives and our rela- all possibilities. These values, as well as the specific content of
tionships to nature is inseparable from . . . justice
and faithfulness to others” [123].
mathematics, can help us all in the decisions that we will all have
to make together.

apprecia@ucalgary.ca
TO THE STUDENT xvii

Summary of Ideas: To the Student

• Sustainability refers to the ability of a social or economic system to keep functioning


without degrading its environment—to provide for its own needs and also preserve
the ability of future generations to provide for their needs.
• Resilience refers to the ability of a social or economic system to “self-heal”—to
recover from a disruptive event.
• Ecosystem services refer to the benefits that society receives from Earth’s ecosys-
tems.
• Many questions about sustainability and resilience involve measurement, change,
connection, and risk, all of which can be expressed in the language of mathematics.
• This book introduces some of the mathematical ideas that are helpful in making
decisions that involve sustainability.

apprecia@ucalgary.ca
xviii BEFORE WE BEGIN

0.2 To the Instructor


If you want to make a course interesting, then you should study something of interest
[356].

This text supports a course that is aimed at college students—many thousands of them, in our large
universities—who would not describe themselves as “mathematicians” or “scientists” but who need
to take at least one course that supports quantitative literacy as part of their degree requirements.
Often such students have found themselves steered into courses in the precalculus sequence: courses
that may be excellent preparation for future scientists and engineers, but that fail to catch a student’s
attention as their last experience of mathematics. One of us sometimes asks such students, “Would
you rather learn the quadratic formula or would you rather save the world?” This book is for those
students who would like to save the world, or at least take a step in that direction. Similarly, it is for
those instructors who would like to “teach as if life matters” [328], or at least take a step toward using
the mathematics classroom to help students think more clearly about some of the issues that are only
going to become of increasing importance over the twenty-first century.
Each instructor will, of course, use this book in the way that they see fit. However, one of
our primary goals is to advance student skill in quantitative literacy. A required student writing
component has played an essential role in accomplishing this goal and has accounted for over one-
third of the total grade in the course as we have taught it. The Association of American Colleges and
Universities provides the following definition for quantitative literacy [249]:

Quantitative Literacy (QL) is a “habit of mind,” competency, and comfort working


with numerical data. Individuals with strong QL-skills possess the ability to reason
and solve quantitative problems from a wide array of authentic contexts and everyday
life situations. They understand and can create sophisticated arguments supported by
quantitative evidence and they can clearly communicate those arguments in a variety
of formats.
Because of its intended audience, the book does not require any mathematics beyond high school
algebra; the most complicated idea that appears is a fourth root, which shows up in a few places in
Section 2.4. In particular, no calculus is required to read and study this book. If the instructor does
know calculus, though, they will find that it provides background to the chapters on “Flowing” and
“Changing,” and may even find it helpful—for themselves, not for the students—to investigate how
our presentation can be translated into the classical language of differential equations.
Nor does the book require (from you, the instructor) a great deal of sustainability-specific
mathematical background. Naturally, the more you know, the more you will be able to help your
students, but in the end what we’re presenting in Part I of the book is a self-contained set of
mathematical techniques, and if you learn about them from the text, that will be fine. Should you
wish to pursue the material further, the reading suggested in Section 8.1 gives some possible starting
points.
The book’s website is at http://math-for-sustainability.com. As well as the student-
oriented material described in the previous section, this website contains additional resources and
suggestions specifically for instructors. These include a quite specific description of the writing
component of the course as we have taught it. You’re welcome to use this model exactly as it is,
to adapt it, or to do something entirely different. We do believe, however, that this book will be
most effective if it is used in conjunction with a requirement for some student response in the form
of extended writing. Our experience suggests that most students are glad to have the opportunity
to integrate their mathematical learning with social and environmental concerns and to express
themselves in this way.

apprecia@ucalgary.ca
ACKNOWLEDGMENTS xix

0.3 Acknowledgments
A book like this does not come into being without the support and critique of many people. We are
grateful to all the many students who have worked through different versions of the book in their
Math 33 class at Penn State, and for the comments and suggestions they have provided. We’re also
grateful for the advice of those who have read all or part of the manuscript, including Peter Buckland,
David Hunter, Marta Mendoza, Liane Roe, and the anonymous readers who reviewed the manuscript
on behalf of Springer. Numerous colleagues from Penn State and elsewhere have given generously of
their time to make special presentations in one or more Math 33 classes; we are deeply grateful to all
of you. Finally, we especially want to thank Francis Su for the generous and impassioned foreword
that he has contributed to the book.

• John Roe writes: I want to express my thanks to Karl “Baba” Bralich, to Fletcher Harper, to
Byron Smith, to Christopher Uhl, and to Marty Walter, for their various challenges to embrace
this work with my whole self. Marty’s book Mathematics for the Environment [340] helped me
see that something like this volume would be possible. I am deeply grateful to my undergraduate
research assistant for the start of the project, Kaley Weinstein, whose flow of ideas was an
inspiration. And I want to humbly acknowledge the support and love of my whole family, who
now and ever have meant more than I can possibly say.
• Russ deForest writes: Foremost, I would like to express my gratitude to John Roe for inviting
me to be a part of this project, for his kind leadership, and for his deep and personal commitment
to the ideals of general education. I would like to thank Rachel Hoellman for her essential role
as a peer assistant and tutor through multiple iterations and refinements of the accompanying
course, and to Matt Lupold for his current work in this position. Finally, I would like to thank
the many engaged students who make teaching a joyous effort.
• Sara Jamshidi writes: I am inexpressibly grateful to John Roe for giving me an opportunity
to be part of this important project; he has given me an expansive understanding of all that a
mathematician can be and I will carry that with me for the rest of my career and life. I am
thankful for my friends Duane Graysay, Shiv Karunakaran, and Monica Smith Karunakaran,
who have taught me so much about mathematics education. I am also indebted to my partner,
Aleksey Zelenberg, who provided invaluable support and feedback for my work here, and to
my father, Morid Jamshidi, who was the first person to show me that mathematics can be used
in civics and ethics. Finally, I would like to thank you, the reader, for picking up this book
and endeavoring to better understand sustainability through a mathematical lens; I hope this
becomes an empowering tool in your life.

apprecia@ucalgary.ca
Part I

Fundamental Concepts

apprecia@ucalgary.ca
C HAPTER 1
Measuring

Think about this quotation from a BBC news broadcast:

Archaeologists announced Friday that they have discovered human footprints in England
that are between 800,000 and 1 million years old—the most ancient found outside Africa,
and the earliest evidence of human life in northern Europe [188].

We know 1 million years is a long time, but “a long time” is


subjective. Your friend might say that she sat in traffic for “a long
time” because she drove 20 miles in one hour, but one hour does
not compare to 1 million years. Now suppose your friend instead
said it took her one hour to travel 3 million centimeters. How do
we interpret that? Was that a large or small distance to traverse in
that amount of time?
The most basic mathematical question we can ask is: how big?
But there is (or ought to be) an immediate comeback in any real-
world situation: compared to what? Is a million a big number?
A million dollars is a lot of dollars for an ordinary person. But
a million molecules of carbon dioxide is a tiny amount for most
purposes, and even a million dollars is a rounding error if you are
thinking about the total of the U.S. federal budget. What about
a million centimeters? Could you walk that far in a day? In a
week? In a month? Not just the numbers but also the units matter:
a million centimeters might be a manageable distance to walk Figure 1: Measuring instruments.
(we’ll see!), but we can be quite sure that a million miles is not.
In this chapter we will review basic information about numbers,
units, measurements, and comparisons. We will study scientific notation, which is a convenient way to
work with very large and very small numbers, and we will look at some of the most effective ways to
communicate numerical information in human terms. We’ll also learn some of the skills of estimation:
how to get a useful rough idea of the size of some quantity, even when an exact answer is not available.

© Springer International Publishing AG, part of Springer Nature 2018 3


J. Roe et al., Mathematics for Sustainability, Texts for Quantitative Critical Thinking,
https://doi.org/10.1007/978-3-319-76660-7_1
apprecia@ucalgary.ca
4 CHAPTER 1. MEASURING

1.1 Units and Measurement

Objectives

 I can identify the number part and the unit part of a physical measurement.
 I can keep track of units throughout a calculation.
 I can work with pure numbers and their percentage equivalents.
 I am familiar with standard units of time, distance, weight (or mass), and temperature.
 I can convert the same measurement into different units.
 I can communicate the meaning of a measurement by relating it to everyday human
experience.

1.1.1 Number Part and Unit Part


Let’s consider some simple examples of measurements:
(a) The fuel tank of my car holds 13 gallons.
(b) We need 5 pounds of potatoes.
(c) Today’s high temperature is predicted to be 91 degrees Fahrenheit.
(d) The distance by road from New York City to San Francisco is about 2900 miles.
(e) Abraham Lincoln took approximately 2 minutes to deliver the Gettysburg Address.
(f) The Hoover Dam is 726 feet high.
(g) The area of Lake Mead (the lake impounded by the Hoover Dam) is 640 square kilometers.
As you can see from these examples, a physical measurement is made up of two parts:
• the number part—like “13” or “5” or “726”;
• the unit part—“gallons” or “pounds” or “feet.”
It is only the number and the unit part together that make a complete measurement. If my gas tank
has a 13-gallon capacity, and I for some reason decide to measure in teaspoons instead, I could say
that my tank holds about 10,000 teaspoons. Both “13 gallons” and “10,000 teaspoons” are complete
measurements, though the first version is much more useful for most purposes. But it would make no
sense just to say “The capacity of my gas tank is 13.” Without the unit part (gallons), the number part
(13) does not tell us anything.

Critical Thinking

When you read an article that gives numerical information, always ask, “What are the
units?” If no units are given, the information is meaningless.

apprecia@ucalgary.ca
1.1. UNITS AND MEASUREMENT 5

Problem 1: Identify the number part and the unit part in the following
measurements:
(i) The radius of the Earth is approximately 4,000 miles.
(ii) At full takeoff power a Boeing 747 burns over 3 gallons of jet fuel
per second.
(iii) The current U.S. population is about 320 million.
Figure 2: Boeing 747 at takeoff.

Solution: In example (i) it is clear that the number part is 4,000 and the unit part is “miles.”
In example (ii), the number part is 3, but what are the units? The “3” refers to a rate of fuel
consumption, and the units in which this is measured are “gallons per second,” which we may
abbreviate as “ gal/ sec.” Notice that the units are not gallons alone, or seconds alone, but the
combination “gallons per second.” 1
Example (iii) looks at first as though it does not have any units. To see what the units are, try to
expand the statement a bit: ask yourself, “the current U.S. population of what are we describing?”
Clearly the answer is “human beings,” and that tells us the unit part of our measurement: “human
beings” or “people.”

Remark 1: When describing the unit part of a measurement, we don’t distinguish between singular
and plural. Thus, in both “1 inch” and “2 inches,” the units are “inches”; in “1 person” and “300
people” the units are “people” (or “persons” if you prefer). Some units have standard abbreviations:
for example, “2 inches” can also be written “2 in.”

1.1.2 The Unit-Factor Method


Suppose that my car’s gas mileage is 33 miles per gallon. And suppose also that, as in the first example
above, my car’s tank holds 13 gallons. How far can I travel on a full tank?
You probably know that to obtain the answer, we need to multiply 13 by 33. But did you know that
the multiplication can be done with the units as well? Like this:
33 mi
13
gal
× = 13 × 33 mi = 429 mi.
1gal

Using the abbreviations “mi” for miles and “gal” for gallons, the tank capacity is 13 gal, and the fuel
33 mi
consumption is 33 mi/ gal, which we write as a fraction . When we multiply these two together
1 gal
the gallons cancel, leaving the answer (with the correct units) as 429 mi, that is, 429 miles.
This is an example of the unit-factor method.

Definition 1

Using the unit-factor method, whenever we multiply or divide physical quantities, we must
multiply or divide their number parts and their unit parts.

The unit-factor method is valuable in many kinds of problems. It’s especially helpful when we deal
with problems involving unit conversions, like the one below.
1 You may also see this abbreviated as “ gal sec−1 ” in some books; this notation is not wrong, but we’ll try to avoid it.

apprecia@ucalgary.ca
6 CHAPTER 1. MEASURING

Problem 2: There are 1,760 yards in a mile, and there are 36 inches in a yard. Using this information,
express one million inches in miles.

Solution: The key idea is that conversion problems can be expressed in terms of “multiplying by 1”
in different ways—and that multiplying a quantity by 1 does not change that quantity. For example,
the statement “there are 36 inches in a yard,” part of the information given in the problem, can be
re-expressed as
1 yd 36 in
= 1 or = 1. (1)
36 in 1 yd
Both ways of writing the fraction are correct; we have to figure out which way will be more helpful to
us. Now a million inches multiplied by 1 is still a million inches; but if we multiply a million inches
by the left-hand expression for 1 in the display (labeled (1)) above, we get the useful fact that
1 yd 1,000,000 yd
1,000,000 in = 1,000,000
in × = ≈ 27,780 yd.
36in 36


(The symbol “≈” means “is approximately equal to.”) We can use the same idea a second time to
1 mi
convert from yards to miles, using the fraction = 1:
1,760 yd
1 mi 27,780 mi
27,780 yd = 27,780
yd
× = ≈ 15.8 mi.
1,760
yd
 1,760
The answer to our problem, therefore, is approximately 15.8 miles. If we wanted to shorten our work,
we could combine two steps into one and write
1 yd 1 mi 1,000,000
1,000,000 in = 1,000,000
in ×  × = mi ≈ 15.8 mi.
36in 1,760
yd 36 × 1,760


This is also a correct solution.

The unit-factor method is a great help in conversion problems because keeping track of the units
automatically lets us know when to multiply and when to divide. For instance, in the first line of our
solution we multiplied by 1 yd/36 in = 1 to convert our measurement from inches to yards. Only this
form of the conversion factor allows the “inches” to cancel, and that cancellation is the signal that tells
us we are heading in the right direction. Watch how this works in our next example:
Problem 3: Beefy Acres is an imaginary cow/calf farming operation located in the Southeast, used
as an example in a pamphlet published by the Natural Resources Conservation Service (a division of
the U.S. Department of Agriculture) [235]. Beefy Acres has 20 acres of pastureland, and each acre
produces, on average, 9,500 pounds of forage per year. To thrive, each cow requires approximately
17,500 pounds of forage per year. What is the maximum number of cows that can be supported
sustainably on Beefy Acres?

Solution: We have three items of information given to us. Let’s express these in terms of number part
and unit part, where our basic units are “acres” (of pastureland), “cows” (this is a very reasonable
unit for working this problem—units do not have to be just mathy things), “pounds” (of forage; for
historical reasons the abbreviation for pounds is “lb”), and “years.” Then our information is the pasture
lb
area (20 ac), the amount of forage production per acre per year (9,500 ), and the amount of
ac yr
lb
forage consumption per cow per year (17,500 ). We need to combine these quantities using
cow yr
multiplication and division to get an answer whose units are cows; the acres, pounds, and years have
to cancel. There is only one way to do this:

apprecia@ucalgary.ca
1.1. UNITS AND MEASUREMENT 7

1 cowyr 9,500 lb 9,500 × 20


× 20
ac = cow ≈ 11 cows.

 ×
17,500
lb 1
ac 
yr 17,500

  
The answer is approximately 11 cows.
How do we arrive at the correct way to arrange this computation? We start by remembering that our
objective is to arrive at a solution whose units are cows. Only one of the data involves the unit “cow”
lb
at all, and that is the grazing requirement 17,500 . But this piece of information has “cow” in
cow yr
the denominator (the “downstairs” part of the fraction), whereas to get an answer in units of cows we
are going to need “cow” in the numerator (the “upstairs” part). So we consider the reciprocal

1 cow yr
.
17,500 lb
This has “cow” in the numerator as we wanted, but it also involves units of pounds and years, which
we will need to cancel by multiplying by other pieces of information. If we multiply by the yield per
acre, the pounds and years will cancel:

1 cow 
yr 9,500 lb 9,500 cow
 × = .
17,500
lb
 1 ac 
r
y 17,500 ac

This is a meaningful quantity—it tells us how many cows can graze on any given acreage of land,
with units “cows per acre”—but it is not our final answer, whose units must be “cows.” To cancel the
“acre” unit, we multiply by the land area, thus arriving at the final answer

1 cow 
yr 9,500 lb 9,500 × 20
× 20
ac = cow ≈ 11 cows.

 ×
17,500
lb 1
ac 
yr 17,500

  

1.1.3 Standard Units


We already have some knowledge of the standard units for time and distance, area and volume, and
other common physical measurements. Let’s review them.

Units of Time
Each of us is familiar with the idea of time. In the introduction we saw how sustainability questions
can explicitly involve the measurement of time (“how long can this last?”). Table 1 lists some common
units of time.

Table 1: Units of Time


Unit (abbr) Definition Example
Second (s) – Snapping your fingers
Minute (min) 60 s Singing five verses of “Row, Row, Row Your Boat”
Hour (hr) 60 min About the time you spend during a casual lunch
Day (day) 24 hr Time between one sunrise and the next
Week (wk) 7 day About the time it takes to ship a package by regular mail
from the U.S. to Europe
Year (yr) 365 day The time from one birthday to the next
Century 100 yr About 4 generations in a family (from you to your great-
grandmother)

apprecia@ucalgary.ca
8 CHAPTER 1. MEASURING

Units of Length
Units of length come from one of two systems of measurement: the U.S. system and the metric
system (used in much of the rest of the world). We will review both systems of measurement and
discuss how to go back and forth between the two systems.
First, let us list some U.S. measurements of length. These measurements will be more familiar to
you if you grew up in the United States.

Table 2: Units of Length (U.S.)

Unit (abbr) Definition Example


Inch (in) – About the length of a paper clip
Foot (ft) 12 in About the length of a large human foot
Yard (yd) 3 ft About the width of a single (twin) bed
Mile (mi) 1,760 yd About the distance traveled walking for 20 minutes

If you grew up outside the United States, you will be more familiar with metric units of length.
Some of these are listed in the table below.

Table 3: Units of Length (Metric)

Unit (abbr) Definition Example


1
Micron ( µm) 1,000,000 m Size of a small bacterium
1
Millimeter (mm) 1,000 m About the width of a pencil tip
1
Centimeter (cm) 100 m About the length of a carpenter ant
Meter (m) – A little more than the width of a single (twin) bed
Kilometer (km) 1000 m About the distance traveled walking for 12 minutes

Notice that the metric units all are related by powers of 10, like 10, 100, 1000 and their reciprocals
1 1 1
10 , 100 , 1000 .
Moreover, the names of the units all have a standardized form, made up of a prefix
applied to the basic unit “meter.” You can find a more extensive list of these prefixes, and the powers
of 10 that they represent, in the “Useful Data” section of this book (Table 2 on page 486).
Problem 4: Use the information in the tables to determine how long would it take you to walk 3
million centimeters.

Solution: From the data in Table 3, we know that it takes about 12 minutes to walk 1 kilometer. We
can express this rate of progress as
12 min
.
1 km
We want to use this information to find out how long it takes to travel 3 million centimeters. Since
our rate-of-progress information uses kilometers rather than centimeters, we convert one to the other,
1
using the conversion information 1 cm = 100 m and 1 km = 1, 000 m provided by the table. As we
learned in the solution to Problem 2 on page 6, the way to do this is to re-express the conversions as
1m
fractions equal to 1 (such as = 1, for example). Thus we can obtain the travel time:
100 cm
1m 1
km
 12 min
3,000,000
cm
× × ×  = 360 min.

100
cm
 1000m
 1
km

apprecia@ucalgary.ca
1.1. UNITS AND MEASUREMENT 9

1m

1 ft
1m 1 m2

1 ft 1 ft2

Figure 3: Illustrating the formation of units of area.

Our answer is 360 minutes. That’s a lot of minutes, and it will be easier to understand if we convert it
to hours, using the fact that 60 minutes make 1 hour:

360 × 1 hr = 6 hr.
min
60
min

Based on our work above, then, it would take about 6 hours to walk 3 million centimeters. This is
an example of expressing a measurement in human terms—that is, by reference to familiar human
experience. Instead of the mysterious “3 million centimeters,” we can now communicate the same
information by saying “about six hours’ walk.” This alternative way of expressing our measurement
can be grasped directly, without scientific knowledge.2

Remark 2: We separated out the various conversions when we used the unit-factor method in the
above solution. It would be equally correct, however, to combine them all in one line:
1m 1
km
 12
min
 1 hr
3,000,000
cm
× × × ×  = 6 hr,

100
cm 1,000
 m
 1
km
 60
min
as we did in the solution to Problem 3 on page 6.

Units of Area and Volume


Area is a unit of measurement created from length. Units like “square feet” ( ft2 ) and “square meters”
( m2 ) refer to the area made by multiplying the dimensions. Figure 3 is a diagram demonstrating this
idea. From any unit for length, we can get a corresponding unit for area.
Let’s work some problems involving unit conversions for areas. As we’ll see, we will need to take
the squares of the conversion ratios to make the units come out right. This works because conversions
in the unit-factor method are simply ways of expressing the number 1, and the square of 1 is just 1
again.
Problem 5: What is a square foot in square inches?

12 in
Solution: There are 12 inches in a foot; in terms of the unit-factor method, = 1. So how many
1 ft
2
square inches in a square foot, 1 ft = 1 ft × 1 ft? We can work this out by the unit-factor method
12 in 12 in
   
1 ft × 1 ft = 1 ft × × 1 ft × = 12 in × 12 in = 144 in2 .
1 ft 1 ft
2 We’ll develop this idea further in Section 1.1.5.

apprecia@ucalgary.ca
10 CHAPTER 1. MEASURING

2
12 in

A shorter way to approach this calculation is to take the square of the conversion ratio, =
1 ft
(12 in)2
. This is still equal to 1 (because the square of 1 is still 1), so we can write
(1 ft)2

(12 in)2 144 in2


1 ft2 = 1 ft2 × 2
= 1 2
ft × = 144 in2 ,
(1 ft) 1ft
 2

getting the same answer as before. Look carefully at both approaches to make sure you understand
why they give the same result.

Problem 6: Express 15 square centimeters in square meters.

Solution: Working as in the problem above,

(1 m)2 1 m2
15 cm2 = 15 cm2 × = 15 cm
2
× = 0.0015 m2 .
(100 cm)2 2

10,000
cm

The answer is 0.0015 square meters.


There are a couple of additional units of area that are worth knowing. These are particularly
associated with agriculture.

Table 4: U.S. Units of Area


Unit (abbr) Definition Example
Acre 4840 yd2 Roughly the area of a (U.S.) football field, without the
end zones.
Hectare 104 m2 About 2.5 acres

Volume is also a unit of measurement created from length. This time, though, we are considering
cubes instead of squares (Figure 4). Units like “cubic inches” ( in3 ) and “cubic meters” ( m3 ) refer to
measurement of volume. The technique for carrying out conversions is the same.
Problem 7: What is 1200 cubic inches in cubic feet?

1 ft
Solution: We now need to take the cube of the conversion factor = 1. We obtain
12 in

(1 ft)3 1 ft3
1,200 in3 = 1,200 in3 × = 1,200 in
3
× ≈ 0.7 ft3 .
(12 in)3 3

1,728 in

The answer is approximately 0.7 cubic feet.

Example 1: As well as the “cubic” units for volume, traditional systems of measurement have come
up with many other volume units, often specialized to a particular trade or profession: bushels of
wheat, gallons of milk, hogsheads of ale, and so on. Table 5 on the opposite page gives a few such
volume units worth knowing

Problem 8: How many gallons are in a cubic foot?

apprecia@ucalgary.ca
1.1. UNITS AND MEASUREMENT 11

1m

1 ft

1m 1 m3
1 ft 1 ft3
1 ft
1m

Figure 4: Illustrating the formation of units of volume.

Table 5: Units of Volume


Unit (abbr) Definition Example
Gallon ( gal) 231 in3 Large container of milk
1
Pint ( pt) 8 gal Glass of beer
Barrel, of oil ( bbl) 42 gal Barrel about 3 ft high, 1 12 ft diameter
1
Teaspoon (tsp) 96 pt The smallest spoon in a set of ordinary silverware
Liter ( L) 1000 cm3 or 0.0001 m3 A little over 2 pints

Solution: One gallon is equal to 231 cubic inches (as defined above), and one cubic foot is equal to
123 = 1,728 cubic inches. Therefore,

1 gal
1 ft3 = 1,728 3
in × ≈ 7.5 gal.
3
231in

Problem 9: Approximately how many cubic centimeters make a teaspoon?

Solution: This problem requires us to relate the U.S. and metric systems of measurement for volume.
We can use the unit-factor method and the information in the table above:

1 1 1 L 1,000 cm3
1 tsp = pt ≈ pt × × ≈ 5 cm3 .
96 96 2 pt 1L
Thus, a teaspoon is approximately equal to 5 cubic centimeters.

Units of Weight or Mass

Table 6: Units of Mass (U.S.)


Unit (abbr) Definition Example
Ounce ( oz) – About the mass of an AA battery
Pound ( lb) 16 oz About the mass of a small bottle of soda
Ton ( t) 2,000 lb About the mass of a subcompact car

apprecia@ucalgary.ca
12 CHAPTER 1. MEASURING

In this text, we are going to use the terms “weight” and “mass” interchangeably. Strictly speaking
this is not accurate—the “mass” of an object refers to the amount of “stuff” (matter) that it is made
of, and “weight” refers to the amount of pull that gravity exerts on that mass. But so long as we stay
on the surface of the earth, gravity is pretty much constant, and therefore the two terms “weight” and
“mass” can be taken to refer to the same thing. As earthbound people, then, we will not sweat the
distinction.
Like distances, masses have two unit systems: a U.S. system and a metric system. If you grew up
in the United States, you may be more familiar with the mass units in Table 6 on the previous page.
But if you grew up outside of the United States, you are likely more familiar with the measurements
in Table 7 instead.

Table 7: Units of Mass (Metric)


Unit (abbr) Definition Example
Gram ( g) – About the mass of a paper clip
Kilogram ( kg) 1,000 g About the mass of a hardback book
Tonne ( T) 1,000 kg About the mass of a subcompact car

Notice that a U.S. “ton” and a metric “tonne” are not quite the same. But they are pretty close
(within about 10 percent)—close enough that the difference will not matter for most of our purposes.

Unit Conversions
There are many situations in which we may need to convert a measurement from one unit to another.
These conversions could be within one unit system (for example, expressing 3 million centimeters
as 30 kilometers) or from one unit system to another (for example, expressing 30 kilometers as just
under 19 miles).
Here is a table of some useful conversions. (More extensive and precise tables can be found in
Section 8.2.)

Table 8: Unit Conversions


Metric Unit U.S. Conversion U.S. Unit Metric Conversion
1 cm 0.39 in 1 in 2.54 cm
1m 3.3 ft 1 ft 0.3 m
1 km 0.62 mi 1 mi 1.61 km
1g 0.035 oz 1 oz 28 g
1 kg 2.2 lb 1 lb 0.45 kg
1T 1.1 t 1t 0.91 T
1 gal 3.8 L 1L 0.26 gal

We have already seen how the unit-factor method allows us to handle these conversions efficiently.
Let’s do a couple more examples as a reminder.
Problem 10: The weight limit for a checked bag on Untied Airlines is 50 lb. If my bag weighs 32 kg,
will I be able to check it without paying the overweight charge?

Solution: The fact that 32 is less than 50 does not answer this question! We must convert 32 kg to
pounds. To do this, we use the fact (from Table 8) that 1 kg equals 2.2 lb. Using the unit-factor method
2.2 lb
we express this by saying that the fraction equals 1. Therefore,
1 kg

apprecia@ucalgary.ca
1.1. UNITS AND MEASUREMENT 13

2.2 lb
32 kg = 32
k
g× ≈ 70 lb.
1
kg
It looks as though I definitely will have to pay an overweight charge!

Problem 11: I am pulled over on the freeway for driving at a speed of 100 feet per second. If the
speed limit is 65 miles per hour, was I speeding or not?

Solution: We need to express 100 feet per second in miles per hour. Using Tables 1 and 2 (pages 7–8),
ft ft 1 yd 1 mi 60 sec
 60 min
 100 × 60 × 60 mi mi
100 = 100 ×  × ×  × = ≈ 68 .
sec sec
 3 ft 1,760yd 1min 1 hr 3 × 1,760 hr hr
So I was speeding, though not by very much.
The moral is: be mindful of your units! Forgetting to convert to the appropriate unit system is an easy
mistake that can lead to disastrous results—such as the loss of a spacecraft! According to news reports
from the late 1990s, “NASA lost a $125 million Mars orbiter because a Lockheed Martin engineering
team used English units of measurement while the agency’s team used the more conventional metric
system for a key spacecraft operation” [199].
Here’s a slightly different example
Problem 12: Write the height “5 feet 7 inches” in centimeters.

Solution: “5 feet 7 inches” means an addition: five feet plus seven inches. But when we want to add
two quantities, they must be expressed in the same units. In this example, we need to re-express the
five feet in units of inches before we add the seven inches:
12 in
 
5 ft + 7 in = 5 ft × + 7 in = 60 in + 7 in = 67 in.
1 ft
Now that we have expressed the length in units of inches, we can convert to centimeters:
1 cm
67
in × ≈ 170 cm.
0.39in



Remark 3: A calculator will give 67/0.39 as 171.7048 . . . . So why did we “round off” the answer
above to “approximately 170 centimeters”? To understand this, take a look at the discussion of
precision in Section 1.3.1.

Units of Temperature
Our final example of unit conversions concerns the measurement of temperature. In the United States,
temperatures are measured using the Fahrenheit scale. On this scale, the freezing point of water is
32 ◦ F and the boiling point of water is 212 ◦ F, so that there are 180 Fahrenheit degrees between
freezing and boiling.
In most other countries, the Celsius scale is used. On the Celsius scale, the freezing point of water
is 0 ◦ C and the boiling point of water is 100 ◦ C. Thus there are 100 Celsius degrees between freezing
and boiling.
Remark 4: Converting between temperature scales is more complicated than converting between unit
systems for the other kinds of measurements (length, time, mass, and so on) that we have discussed
so far. The reason is that not only do the units have different sizes (one Celsius degree represents
more of a temperature jump than one Fahrenheit degree) but also the zero points of the two scales are
different. That problem does not come up for other sorts of measurements: zero feet and zero meters
both represent the same length, zero! The differences between Fahrenheit and Celsius are illustrated
in Figure 5.

apprecia@ucalgary.ca
14 CHAPTER 1. MEASURING

What’s more, neither the Celsius nor the Fahrenheit scales have their zero points
rooted in any fundamental physical reality. (According to legend, 0 ◦ F was simply
the coldest temperature that Fahrenheit could reach using the technology available
to him at the time.) Modern physics tells us, though, that heat is the result of the
disordered motion of atoms and molecules, and this means that there is truly a
“coldest possible” temperature, when all atoms and molecules would be at rest.
This absolute zero of temperature is much colder than anything in our ordinary
experience: it is about −273 ◦ C, or −460 ◦ F. In physics calculations the Kelvin
temperature scale is used: Kelvin temperature is simply equal to Celsius temperature
plus 273, so that absolute zero is 0 K, water freezes at 273 K, and water boils at
373 K. Table 9 gives the algebraic formulas for converting between these three
temperature scales.

Table 9: Temperature Scale Conversions


◦C ◦F K
Figure 5: Celsius and
Fahrenheit comparison. ◦C C =C C = 59 (F − 32) C = K − 273
◦F F = 32 + 95 C F =F F = 95 K − 460

K K = C + 273 K = 59 F + 255 K=K

For instance, if you want to convert a temperature from Fahrenheit to Celsius, let the Fahrenheit
temperature be F. Look in the table to the intersection of the ◦ F column and the ◦ C row, where you
will find the algebraic formula C = 59 (F − 32). This tells you how to compute the Celsius temperature
C in terms of the Fahrenheit temperature.
Problem 13: Normal human body temperature is about 98 ◦ F. Express this on the Kelvin scale.

Solution: Looking in the table above, we see the formula K = 59 F + 255 to convert from Fahrenheit
to Kelvin. Plugging in F = 98 ◦ F, we obtain
5 × 98
K= + 255 ≈ 55 + 255 = 320 K
9
for the Kelvin equivalent.

1.1.4 Percentages and Other Pure Numbers


We’ve stressed that physical quantities have both a number part and a unit part, but we sometimes
have a use for pure numbers. These are numerical quantities that have no unit. The most familiar
example is a percentage.
Example 2: In 2015 the population of California was estimated to be 39 million people while the
population of the United States as a whole was 320 million people [261]. Let’s express the California
population as a percentage of the total U.S. population. A percentage is a proportion; we divide the
population of California by the total population,
39,000,000 people
 12
≈ 0.12 = = 12%.
320,000,000  people
 100
The units (people) cancel, leaving us with the pure number 0.12, which we can also write as 12%
(read as 12 percent).

apprecia@ucalgary.ca
1.1. UNITS AND MEASUREMENT 15

Remark 5: Percent comes from the Latin per centum, meaning out of a hundred; 12 out of every 100
people in the U.S. live in California. Notice how you can think of the identity 0.12 = 12% as another
example of the unit-factor method:

0.12 = 0.12 × 100% = 12%,

since 100% is another way of writing 1, and multiplying by 1 makes no difference!


Remark 6: In the previous example, we used the unit “people,” but we could have chosen “million
people” as our unit without changing the result:
39 (
million
((( people
((
320 (million
((( (( ≈ 0.12 = 12%.
people
If a calculation yields a pure number (or percentage) result, then all the unit parts must cancel. That
means we will get the same answer whatever units we use in the numerator and denominator (as long
as we use the same units for both).

Definition 2

Q1
A numerical quantity without a unit is called a pure number. A ratio of two quantities
Q2
Q1 and Q2 having the same units will produce a pure number.

Example 3: The mathematical constant π ≈ 3.14, which is the ratio of a circle’s circumference to
its diameter, is a pure number. It appears (among many other places) in the formula for the area of a
circle, πr2 , where r is the radius.

Example 4: The conversion factors in Section 1.1.3 are also pure numbers. For example,

1 yd 36 in
= 36

=
1 in 1
in

is a pure number.

Example 5: A concentration is a pure number: when we say “the concentration of oxygen in the
atmosphere by volume is about 21%,” we are referring to the ratio
Volume of atmospheric oxygen
.
Volume of entire atmosphere
The units cancel, giving us a pure number, about 0.21 = 21%.
1
A percentage is a multiple of 100 , as we said earlier. Very small (or trace) concentrations may be
1
expressed not as percentages but as parts per million ( ppm, that is, multiples of 1,000,000 ) or even as
1
parts per billion ( ppb, that is, multiples of 1,000,000,000 ). For example, the present concentration of
CO2 in the atmosphere is about 0.04%. We typically express this in parts per million. To do so, use
the unit-factor method again: 1,000,000 ppm = 1, so
1
0.04% = 0.04% × 1,000,000 ppm = 0.04 × × 1,000,000 ppm = 400 ppm,
100
that is, 400 parts per million (by volume).

apprecia@ucalgary.ca
16 CHAPTER 1. MEASURING

Remark 7: In Example 5 on the previous page, we expressed the concentration of atmospheric carbon
dioxide in parts per million by volume. Trace concentrations are also frequently expressed in parts per
million (ppm) or parts per billion (ppb) by weight. In the case of atmospheric CO2 , 400 parts per
million by volume corresponds to about 600 parts per million by weight. The weight of a carbon
dioxide molecule is about 1.5 times the weight of an average air molecule and this accounts for the
difference in these two measures.
For concentrations of CO2 and other trace gasses in the atmosphere ppm is almost always intended
as parts per million by volume, while for concentrations of trace substances in soil and water ppm
is usually intended to mean parts per million by weight. To avoid confusion we will usually use the
abbreviation ppmv when referring to parts per million by volume.

Example 6: Toluene is a byproduct of gasoline production that is used as a solvent in paint thinners
and in industrial applications. Toluene is present in the discharge from petroleum refineries and poses
a concern for groundwater and drinking water supplies. The U.S. Environmental Protection Agency
(EPA) enforces a limit of 1 ppm for toluene in drinking water [20].
1 kilogram (kg) is 1,000 grams and 1 gram is equivalent to 1,000 milligrams (mg). A concentration
of 1 ppm (by weight) is thus equivalent to 1 mg/kg:

1
mg 1 gram
 1k
g 1,000,000 ppm
× 1,000,000 ppm = = 1 ppm
 
×  ×
1
k
g 1, 000mg 1, 000
 gram
 1,000,000

See Exercise 18 on page 60 for examples expressing ppm and ppb in more familiar terms.

Problem 14: Suppose that your electric bill is $70 each month. After following the advice contained
in the EPA’s Energy Saving Tips for Renters [17], you manage to cut your electric bill by 15%. How
much money are you saving each month?

Solution: The amount we save is 15% of $70, that is,

1
15% × $70 = 15 × × $70 = $10.50.
100

Problem 15: Alcohol concentration, ABV, or alcohol by volume, is reported as a percent. Suppose a
particular brand of beer has an ABV of 5%. How much alcohol is in a 12 oz beer of this brand?

Solution: Using the same idea as in the previous solution, we get

1
5% × 12 oz = 5 × × 12 oz = 0.60 oz,
100

or a little more than half an ounce.

There are many circumstances in which a percentage may be the most meaningful way of measuring
some kind of change (for more about why this might be, see Section 4.1.1). For example, changes in
population or the size of the economy are usually reported as percentages.

apprecia@ucalgary.ca
1.1. UNITS AND MEASUREMENT 17

Think about it. . .

According to recent estimates, roughly 10 million tons of plastic makes its way into the
oceans each year to become “marine plastic debris.” Although the U.S. generates more
plastic waste overall than most other countries, it is responsible for only 1% of plastic
flowing into the oceans each year [170], because a much smaller proportion of U.S. plastic
waste is “mismanaged” than plastic waste generated in less well-off countries. How much
responsibility does the U.S. share, in your opinion, for dealing with the problem of plastic
pollution in the world’s oceans and with the global management of plastic waste more
generally?

1.1.5 Measurements in Human Terms


A measurement, as we have seen, has both a number part and a unit part. But we human beings
are not always able to grasp the significance of very large or very small numbers, or of units of
measurement that don’t relate to a familiar scale. If we can express a measurement in terms of more
familiar quantities, it can help us a lot. Let’s make the following definition.

Definition 3

We say that a measurement is expressed in human terms if it is expressed in a way


that allows us to relate it directly to our shared everyday experience, without requiring
specialized scientific knowledge.

Some units of measurement are intrinsically “in human terms.” For example, one foot originally
was just the length of your foot! Here are two simple rules for keeping measurements in human terms:

Rule 1: Human Terms Measurement Rules

To express a measurement in human terms, try, if possible, to follow both of the rules
below:
• Choose units to keep the number part reasonably close to 1; say, between 0.01 and
1,000.
• Use “human scale” units (like feet, inches, hours).

When these rules both apply, it is not hard to express something in human terms. For instance,
the spacing of the studs in standard house framing (in the U.S.) is 16 inches. That is a small number
and a “human scale” unit. It would be foolish to express this spacing as 406,400,000 nanometers or as
0.0002525 miles. Even though both of these conversions are technically correct, they both violate both
of the rules above: the numbers are huge or tiny (violating the first rule) and the units of measurement
(nanometers, miles) are far from human scale (violating the second).

apprecia@ucalgary.ca
18 CHAPTER 1. MEASURING

Problem 16: The mass of a full-grown elephant is about 4 tons. How


should I express this in human terms?

Solution: There is no single correct answer to a question like this.


We could refer to Table 6 on page 11 to say that the mass of the
elephant is the same as that of 4 compact cars. Or we could look
up the average mass of a U.S. American (about 180 pounds; see
Section 8.2) and say that the elephant has the mass of

2,000 lb 1 American
4t× × ≈ 45 Americans.
1t 180 lb
Figure 6: A grown elephant can weigh 4 tons.
Notice that the notion of “human terms” depends explicitly on whose
shared everyday experience is taken as a reference point. In expressing the mass of the elephant in
terms of a number of compact cars, we are assuming that compact cars are familiar and elephants are
unfamiliar. In a different society, where elephants were abundant but automobiles were rare, one might
reverse the process and express the mass of a compact car in terms of a fraction of an elephant—and
this would be a “human terms” measurement too.

Example 7: Here’s another example from a recent real-life sustainability discussion. In August 2015,
President Obama addressed a meeting of Arctic nations in Anchorage, Alaska. In his remarks [244]
he used the following example to illustrate the shrinking of Arctic ice:

Since 1979, the summer sea ice in the Arctic has decreased by more than 40 percent—a
decrease that has dramatically accelerated over the past two decades. One new study
estimates that Alaska’s glaciers alone lose about 75 gigatons—that’s 75 billion tons—of
ice each year.
To put that in perspective, one scientist described a gigaton of ice as a block the size of the
National Mall in Washington—from Congress all the way to the Lincoln Memorial, four
times as tall as the Washington Monument. Now imagine 75 of those ice blocks. That’s
what Alaska’s glaciers alone lose each year.

For anyone who has visited Washington, DC, the image of a giant ice block “the size of the National
Mall, four times as tall as the Washington Monument” is both striking and accessible, and this speech
is an effective example of putting a very large quantity in human terms. For those who aren’t so
familiar with Washington, though, perhaps not so much. As we said in the previous example, the
notion of “human terms” depends very much on whose experience is our reference point.
More difficult situations arise in which we can’t follow both parts of the Human Terms Measure-
ment Rules at the same time. For example, consider the following data relating to the volume of water
on Earth (Section 8.2). “Fresh surface water” refers to fresh water in lakes, rivers, and streams, as well
as in mountain snow, glaciers, ice caps, and ice sheets, but not underground freshwater (as in aquifers)
nor atmospheric freshwater (as in clouds).

Table 10: Volume of Water on Earth


Type of Water Volume (cubic meters)
All Water 1,400,000,000,000,000,000
Fresh Surface Water 100,000,000,000,000

A cubic meter can perhaps be considered as a “human terms” unit (you can envisage a cube of water
one meter on each side, or just think about the amount of water it might take to fill an average hot tub),

apprecia@ucalgary.ca
1.1. UNITS AND MEASUREMENT 19

so the unit part of these measurements is accessible, but the number part is way out of human scale.
In a situation like this, it is a good idea to ask ourselves just exactly what we are trying to convey in
human terms. Do we want to convey the absolute size of the data—what the total amount of water is,
in some way that a human might be able to grasp? Or are we more interested in the relative size of
the two items—just how much of the Earth’s water is fresh? These are two different questions, and as
we’ll see, they need different kinds of answers.
Example 8: Let’s suppose first that we are trying to convey the absolute size of the measurement
of the total amount of water, 1,400,000,000,000,000,000 cubic meters. This is such a huge number
because the units (cubic meters) are relatively so small. Can we choose a different unit which, though
no longer directly on a human scale, is at least accessible to (some of) our readers’ experience?
Here is one way of doing that. Lake Superior is the largest of the Great Lakes; some readers will
have visited Lake Superior or may even live there. From a reference book, we can find that the volume
of Lake Superior is approximately 12,100,000,000,000 cubic meters. That is,
1 Lake Superior
= 1.
12,100,000,000,000 m3
Using the unit-factor method, the total volume of water on Earth is

3 1 Lake Superior
1,400,000,000,000,000,000
m × ≈ 116,000 Lake Superiors.
12,100,000,000,000
m3

Lake Superior is a pretty huge “unit,” but at least it is somewhat familiar; so here is one answer—the
total amount of water on Earth is a bit over 100,000 Lake Superiors.

Example 9: Another approach to the same problem is one that is useful only for volume (and
sometimes, area) measurements. Imagine all that water rolled up into a single spherical ball. How
big would the ball be?
To use this way of expressing the measurement in human terms, you need to know or remember (or
look up) the formula for the volume V of a sphere in terms of its radius r: that is, V = 43 πr3 . We can
rearrange this using algebra to give r
3 3V
r= .

Thus a sphere of water of volume 1,400,000,000,000,000,000 m3 would have radius
s
3 3 × 1,400,000,000,000,000,000 m3
r= ≈ 690,000 m.
4×π

(Notice, again, how beautifully the unit-factor method works: “cubic meters” ( m3 ) appear under
the cube-root sign; the cube root of m3 is m; so the answer is in meters. Thus, all the water on Earth
would make a ball of radius about 690,000 meters; that is 690 kilometers, or about 430 miles. See
Figure 7 for a graphic illustration of this.

Think about it. . .

A student looking at Figure 7 says, “That can’t possibly be right! After all, we know that
in reality most of the Earth’s surface is covered with water! So how could it all possibly fit
into that tiny droplet?” How would you explain things to this student?

apprecia@ucalgary.ca
20 CHAPTER 1. MEASURING

Example 10: Suppose though that we were more inter-


ested in the second question raised above: not so much
conveying the absolute size of these gigantic amounts
of water, but their relative size: roughly how much
of the Earth’s water is fresh? Here, what we want to
communicate is how several measurements are related
to one another, and we do this in the same way as you
would if you were making a map: by rescaling.
To rescale a group of measurements, you multiply
each one of them by the same number, called the scale
factor. For example, if a map is on a scale of one inch to
one mile, the scale factor is

1 in 1
in
 1mi
 1 y
d 1 ft 1
= × ×  × = .
1 mi 1mi 1,760
yd 3 ft 12in 63,360
Figure 7: All the water on Earth would fit into a “raindrop” 430
miles in radius. A scale factor like this is a pure number (Definition 2 on
page 15), since all the units cancel. The scale factor was
constructed to convert one mile in real life to one inch
on the map. However, now that we know the scale factor, we can see how any distance in real life will
be represented on the map: just multiply by the scale factor. For instance, 1000 feet (or 12,000 inches)
1
in real life will be represented on the map by 12,000 in × ≈ 0.19 in, or just under a fifth of an
63,360
inch.
In our case we have two quantities to consider in our “map” (the volume of all water, and the volume
of fresh surface water). Let’s figure out a scale factor that will convert the volume of all water (in real
life) to ten gallons (on the “map”). This scale factor will be the tiny number

10 gal 10gal
 3.8 L 0.001
m3
= × ×
1,400,000,000,000,000,000 m3 1,400,000,000,000,000,000
m3 1gal
 1L

≈ 0.000000000000000000027.

Now, to work out how the volume of fresh surface water (in real life) will appear on our “map,” we
must multiply by the scale factor:

100,000,000,000,000 m3 × 0.000000000000000000027 ≈ 0.0000027 m3 .

Since a meter equals 100 centimeters, a cubic meter equals 1003 = 1,0000,000 cubic centimeters.
Thus our scaled representation of the volume of fresh surface water is about 0.0000027 × 1,000,000 =
2.7 cm3 , or roughly half a teaspoon (see Problem 9 on page 11). To put it another way, about half a
teaspoon out of every ten gallons of water on earth is (surface) freshwater. That is certainly a striking
“human terms” expression of how relatively rare and valuable fresh water is.
Here is one more example.
Problem 17: A domestic electrical outlet can deliver energy at a maximum rate of perhaps 3
kilowatts—that is, 3,000 joules per second. A gasoline pump delivers gasoline at a rate of 1 gallon
every 10 seconds or so, and each gallon of gasoline contains about 120,000,000 joules of energy (see
Table 5 on page 488). Compare the rate of energy delivery by the electrical outlet and the gas pump.

apprecia@ucalgary.ca
1.1. UNITS AND MEASUREMENT 21

Solution: The energy delivery rate from the gas pump is

120,000,000 J
= 12,000,000 J/sec.
10 sec
The corresponding rate for the electrical outlet is 3,000 J/sec. Thus the ratio of the energy delivery
rates is
12,000,000 J/ sec

= 4,000;
3,000 J/sec

that is, the gas pump delivers energy 4,000 times faster than the domestic electrical outlet. Or to put
the matter in terms of time, the gas pump delivers in one second the same amount of energy that
the electrical outlet takes 4,000 seconds (a little over an hour) to supply. Or again, we can convert
to teaspoons (as in the previous problem) and say that the domestic outlet takes over five seconds to
supply the energy equivalent of a teaspoon of gas. All of these can be thought of as simple rescalings.
A more striking “rescaling” is to express the matter in the following way: if the hose of the gas
pump were shrunk to the diameter of a drinking straw, it would deliver energy at the same rate as the
electrical outlet. The striking image of gasoline dribbling slowly out of a ten-foot hose as thin as a
drinking straw certainly reinforces the contrast between the gas and electric “pumps,” and highlights a
major issue for the acceptance of electric cars: it is extremely difficult to achieve the same “refueling
rates” with electricity as the ones gasoline has gotten us accustomed to. For the details of how the
rescaling is calculated, see Exercise 22 on page 60 at the end of the chapter.

Example 11: The Sagan Planet Walk, a public sculpture in Ithaca, NY,
is a marvelous example of the power of rescaling to convey an important
idea in dramatically visual terms. The Sagan Walk is a scale model,
scale 1 : 5,000,000,000, of our solar system and of the various objects
(the Sun and planets) that are parts of it. It cuts the huge size of the Solar
System down to something that we can appreciate visually.
In Figure 8 you can see the circular opening in the granite obelisk,
almost 11 inches in diameter, that represents the Sun to scale. In each
of the other obelisks in the sculpture, representing the planets, there is a
Sun-sized opening that has a glass window into which the model of the
corresponding planet is embedded. This allows the viewer to experience
directly how much bigger the Sun is than even the largest planets.
You can see the Mercury monolith just above the boy’s hand in
Figure 8, to the right of the Sun monolith. It is just over 10 yards away
from the Sun, and another 10 yards will get you to Venus and another
10 more to our Earth. The first four planets are really close together,
no further than the trees in our picture. What’s more, the models of
the corresponding planets are tiny, the size of a pea or smaller. Then,
suddenly, the scale starts to grow. It will take you ten minutes of brisk Figure 8: Standing by the Sun on the Sagan
walking to get to Neptune from the Sun, passing on the way by Jupiter, Planet Walk.
the largest planet—model size 2.9 cm, greater than an inch but not quite
the size of a Ping-Pong ball. The Planet Walk honors Carl Sagan’s work in science communication by
giving visitors a direct sensory engagement with the size and shape of the solar system.
In 2012 the Planet Walk was expanded. A monolith at the Astronomy Center of the University
of Hawai’i now represents the nearest star to Earth—Alpha Centauri—on the same scale. Imagine
walking the distance from Ithaca to Hilo, Hawai’i (where the Astronomy Center is located), to get
some idea of the difficulties of interstellar travel compared to the journeys of our various spacecraft
within the solar system.

apprecia@ucalgary.ca
22 CHAPTER 1. MEASURING

If you want to learn more about the Planet Walk and the rescaling involved, we encourage you to
read a beautiful article by mathematician Steve Strogatz, titled “Visualizing Vastness” [313].

Summary of Ideas: Units and Measurement

• A measurement, such as “3.2 million centimeters” consists of a number part, 3.2


million, and a unit part, centimeters.
• There are two main systems of units: the U.S. system and the metric system. Some
examples of U.S. units are feet, pounds, gallons. Some examples of metric units are
meters, kilograms, liters.
• We can use the conversion tables in this chapter to convert units within each system
and between them.
• To help us with these conversions, we can use the unit-factor method to make sure
that units cancel out. This tells us which numbers to divide and which to multiply: for
example,
12  × 0.62 mi ≈ 7.4 mi.
km
1km


• The ratio of two quantities having the same units or dimensions is a pure number.
Pure numbers can also be expressed as percentages.
• To communicate clearly, it helps to express measurements or comparisons in human
terms. This means that we express them in a way that is relatable to ordinary human
experience without requiring scientific knowledge.

apprecia@ucalgary.ca
1.2. SCIENTIFIC NOTATION 23

1.2 Scientific Notation

Objectives

 I can explain why scientific notation is used.


 I can convert numbers between decimal and scientific notation.
 I can recognize different ways of writing the same number using scientific notation,
including in standard form.
 I can add, subtract, multiply, and divide numbers written in scientific notation.

1.2.1 Scientific Notation


In the last section we dealt with some very big numbers. When numbers become very large or very
small, the standard way of expressing them (called decimal notation) takes up a lot of space and
becomes hard to read. That is because very large and small numbers are mostly made up of zeros!
For example, two large volumes appeared in Example 10 on page 20:

• 1,400,000,000,000,000,000 cubic meters (the total volume of water on Earth), and

• 100,000,000,000,000 cubic meters (the volume of fresh surface water).

What matters most about these numbers is that the first is in the millions of trillions and the second
is in the hundreds of trillions. To see that, however, you have to carefully count the number of zeros
appearing in each expression. The commas help, but couldn’t we find some more straightforward way
to convey this information?
The key to doing so is to use the idea of powers of 10. Remember that 10 raised to a certain power,
say n, means multiplying 10 by itself that number (n) of times. Thus, for example,

101 = 10,
102 = 10 × 10 = 100,
103 = 10 × 10 × 10 = 100 × 10 = 1,000,
104 = 10 × 10 × 10 × 10 = 1,000 × 10 = 10,000,

Another way of saying this is that 10n can be written as the digit 1 followed by n zeros. So, if we are
given the number 109 , we know our number is 1 followed by 9 zeros; that is a billion. Similarly, 106
is a million, and 1012 is a trillion.
The number 7,000,000,000,000 is written as the digit 7 followed by 12 zeros—7 trillion. We can
write that as 7 times 1 trillion, or 7 × 1012 . This technique of using a power of 10 to keep track of the
zeros is known as scientific notation. Here is a definition.

apprecia@ucalgary.ca
24 CHAPTER 1. MEASURING

Definition 1

A number is written in scientific notation if it is expressed as the product of two parts:


• an ordinary decimal number called the significand (like 7 in the example above), and
• a power of 10 (like 1012 in the example above). The power to which 10 is raised (the
12 in the example) is called the exponent.

Example 1: The expression below represents the number 120,000,000,000,000 in scientific notation.

1.2 × 1014

significand exponent

The significand in this expression is 1.2 and the exponent is 14.

Remark 1: The same number in the example above could be expressed in several different ways. For
instance
1.2 × 1014 = 1.2 × 10 × 1013 = 12 × 1013 ,
because 1.2 × 10 = 12. By shifting the exponent in this way, we can multiply or divide the significand
by 10—that is, we can move the decimal point to the left or right. Usually we choose to move the
decimal point so that the significand is between3 1 and 10. When this has been done, the number is
said to be expressed in standard form. Thus 1.25 × 1014 and 12.5 × 1013 represent the same number,
but the first expression is in standard form and the second is not.

Conversion from Decimal to Scientific Notation


Let’s look at some examples of converting quantities to scientific notation and standard form.
Problem 1: Write 2,000,000,000,000,000,000 in scientific notation, standard form.

Solution: The number 2,000,000,000,000,000,000 has 18 zeros, so we can write it as 2 × 1018 .

Problem 2: Write 150,000,000,000,000,000,000 m3 in scientific notation, standard form.

Solution: This quantity has a number part and a unit part (see Section 1.1.1). In converting to scientific
notation we re-express only the number part; the unit part should remain unchanged. Now in this
example the number part 150,000,000,000,000,000,000 is 15 with 19 zeros. So we could write it as
15 × 1019 . This is a correct expression, but notice that 15 is not between 1 and 10, so this expression
is not in standard form. Instead, we write the number part as 1.5 × 1020 , which is in standard form.
The answer to the complete problem (including the unit part) is therefore

1.5 × 1020 m3 ,

also in standard form.

3 We allow 1 but not 10 for a significand in standard form: the standard form of 1, 000 is 1 × 103 , not 10 × 102 .

apprecia@ucalgary.ca
1.2. SCIENTIFIC NOTATION 25

Remark 2: When we express a physical measurement in scientific notation (as in the previous
problem), it is only the number part of the measurement that we are working with. The unit part
remains unaffected, and must appear in our final solution. Thus, in the last example, the solution is
1.5 × 1020 m3 , not just 1.5 × 1020 .
You can think of these calculations with powers of 10 in terms of “moving the decimal point.” First,
imagine your number written in decimal notation (if it is a whole number, imagine a decimal point to
the right of the ones place, so that you would think of 127 as 127.0). Then moving the decimal point
one step to the left corresponds to dividing by 10 (e.g., 12.7 = 127/10). This gives us a simple rule
for putting a number greater than 1 into standard form .

Rule 1

To express a number greater than 1 in standard form, move the decimal point to the
left until you obtain a number between 1 and 10. What you obtain is the significand
(in standard form), and the number of steps to the left that you moved the decimal
point is the exponent.

For example, to express the number 127 in standard form we move the decimal point 2 steps to the
left to get a significand of 1.27 and an exponent of 2:

127.0
giving us 127 = 1.27 × 102 .
2 steps

Problem 3: The distance from the Earth to the Sun is approximately 93 million miles. Convert this to
inches, and express your answer in scientific notation using standard form.

Solution: Using the unit-factor method we can calculate

 × 1,760
93,000,000
mi
yd 36 in
× ≈ 5,900,000,000,000 in.
1mi
 1
y
d
Moving the decimal point 12 steps to the left gives us 5.9, so the answer in standard form is 5.9 ×
1012 in.
These examples show how to write large numbers using scientific notation. But we can also write
small numbers in the same way. To do this, we need to remember about negative powers of 10. Just as
the positive powers of 10 are obtained by successively multiplying by 10, so the negative powers are
obtained by successively dividing by 10:

100 = 1,
1
10−1 = = 0.1,
10
1
10−2 = = 0.01,
10 × 10
1
10−3 = = 0.001.
10 × 10 × 10

Thus 10−n is

apprecia@ucalgary.ca
26 CHAPTER 1. MEASURING

10−n = 0. 00 . . 00} 1.
| .{z
n − 1 zeros

Suppose for example that we would like to write the number 0.000000567 in standard form. We can
write this as 5.67 × 0.0000001, and from our discussion above, 0.0000001 = 10−7 . Thus the standard
form of our number is 5.67 × 10−7 .
Alternatively, we can again think in terms of “shifting the decimal point.” This time, the rule is the
following.

Rule 2

To express a number less than 1 in standard form, move the decimal point to the
right until you obtain a number between 1 and 10. What you obtain is the significand
(in standard form), and minus the number of steps to the right that you moved the
decimal point is the exponent.

In the example above, we move the decimal point 7 steps to the right to get the number 5.67.
Therefore, the standard form is 5.67 × 10−7 :

0.000000567.
7 steps

Problem 4: Write 0.00000000000000232 in standard form.

Solution: To get 2.32, we need to move the decimal point 15 steps:

0.000000000000002325.
15 steps

That gives us 2.32 × 10−15 as the result in standard form.

Conversion from Scientific to Decimal Notation


Up to now, we have converted numbers from decimal notation to scientific notation. Converting in the
reverse direction, from scientific to decimal, is simply multiplying by a power of 10. Again, we can
think of this in terms of shifting the decimal point.
Problem 5: Convert 9.8 × 107 to decimal notation.

Solution: Because we have a positive exponent (7), we know that this is likely to be a large number.
Our solution is
9.8 × 107 = 98,000,000.
We can think of this as taking the number 9.8 and moving the decimal point 7 steps to the right, filling
in zeros as needed.

Problem 6: Convert 98 × 10−7 to decimal notation.

apprecia@ucalgary.ca
1.2. SCIENTIFIC NOTATION 27

Solution: Because we have a negative exponent (−7), we know that this is likely to be a small number.
Our solution is
98 × 10−7 = 98 × 0.0000001 = 0.0000098.
We can think of this as taking the number 98 = 98.0 and moving the decimal point 7 steps to the left,
filling in zeros as needed.

1.2.2 Calculations with Scientific Notation


Adding, subtracting, multiplying, and dividing numbers in scientific notation is straightforward once
you know the rules. We’ll review these rules and look at some examples.
Remark 3: Most calculators and calculator apps can work with numbers in scientific notation, though
not all offer you a choice about whether to display the final result in scientific or decimal. We’ll show
some examples below. Nevertheless, it’s important to appreciate how these calculations are carried
out, and the best way to do that is to understand the rules for calculating by hand.

Multiplication and division


We’ll begin with the rules for multiplying and dividing with scientific notation. It may seem surprising,
but these are actually simpler than the corresponding rules for adding and subtracting! What’s more,
they are used more often. That’s why we have put them first.
The laws of exponents say that
10a × 10b = 10a+b , 10a ÷ 10b = 10a−b .
For example,
102 × 103 = (10 × 10) × (10 × 10 × 10) = 10 × 10 × 10 × 10 × 10 = 105 .
Thus powers of 10 are multiplied by adding their exponents, and divided by subtracting their
exponents. For numbers in scientific notation this gives us the following rule.

Rule 3

When multiplying (or dividing) two numbers in scientific notation, we multiply (or
divide) their significands and add (or subtract) their exponents.

Example 2: Let’s consider the two numbers 4.23 × 1023 and 9.1 × 106 . First we will multiply them.
What is (4.23 × 1023 ) × (9.1 × 106 )? According to Rule 3, we must multiply the significands and add
the exponents. Therefore, we must do the following calculation:
(4.23 × 1023 ) × (9.1 × 106 ) = (4.23 × 9.1) × 1023+6 = 38.493 × 1029 = 3.8493 × 1030 ≈ 3.85 × 1030 .
Notice how multiplying the significands gives us a number (38.493) that is greater than 10. We shifted
the decimal point to the left in order to get the result in standard form.
Now let’s consider division, (4.23 × 1023 ) ÷ (9.1 × 106 ). According to Rule 3, this time we will
divide the significands and subtract the exponents. Therefore, we do the following calculation:
(4.23 × 1023 ) ÷ (9.1 × 106 ) = (4.23 ÷ 9.1) × 1023−6 ≈ 0.46 × 1017 = 4.6 × 1016 .
Once again we needed to shift the decimal point (this time to the right) in order to get the result in
standard form.

apprecia@ucalgary.ca
28 CHAPTER 1. MEASURING

Remark 4: Whatever notation we use, it is important to remember that a÷b 6= b÷a. So when dividing
numbers in standard form, be careful to carry out the operations in the correct order!
Now, let us look at some more worked problems.
Problem 7: Calculate (9.5 × 108 ) × (4.1 × 108 ).

Solution:

(9.5 × 108 ) × (4.1 × 108 ) = (9.5 × 4.1) × 108+8 = 38.95 × 1016 ≈ 3.9 × 1017 .

Problem 8: Calculate (3.2 × 104 ) ÷ (4.5 × 108 ).

Solution:

(3.2 × 104 ) ÷ (4.5 × 108 ) = (3.2 ÷ 4.5) × 104−8 ) ≈ 0.71 × 10−4 = 7.1 × 10−5 .

If units are involved, we multiply the number parts according to Rule 3 on the previous page, and
the unit parts according to the unit-factor method (Section 1.1.2).
Problem 9: According to [136], a cow emits about 0.3 kilograms of methane per day.4 There are
thought to be about 1.2 × 109 cows on Earth. Compute the total mass of methane emitted by cows
over one year, expressing your answer in scientific notation.
The mass of the entire atmosphere is about 5 × 1018 kg (see Section 8.2). Methane lasts about 8
years in the atmosphere before breaking down to other gases. What proportion (by mass) of the whole
atmosphere is made up of bovine methane emissions?

Solution: We solve the first problem using the unit-factor method:

kg days kg kg
× (1.2 × 109 ) ≈ 130 × 109 = 1.3 × 1011 .

0.3 cows
 × 365

cows
 
 days
 yr yr yr

Bovine methane emissions are about 1.3 × 1011 kg/ yr.


For the second problem, the atmosphere contains about 8 years worth of bovine methane. That is,

8 yr × (1.3 × 1011 ) kg/ yr ≈ 10 × 1011 kg = 1012 kg.

Since the mass of the whole atmosphere is about 5×1018 kg, the proportion of bovine methane (which
is a pure number) is
1012
kg 1
= × 10−6 = 0.2 × 10−6 = 2 × 10−7 ,
5 × 1018kg
 5
or 0.2 parts per million. Though this is a small number, it is in fact a substantial fraction of the total
amount of methane in the atmosphere (which is about 1 part per million by weight).

Addition and subtraction


Now we will think about adding and subtracting numbers in scientific notation. There is one situation
in which we can do this directly:
4 Often referred to as “cow farts” online, though [136] showed that the great majority of the methane is burped rather than

farted.

apprecia@ucalgary.ca
1.2. SCIENTIFIC NOTATION 29

Rule 4

In order to directly add (or subtract) two numbers in scientific notation, they need
to have the same exponent. Once two numbers have the same exponent, we add (or
subtract) their significands.

The two numbers below can be added automatically because they have the same exponent, 8.

(2.32 × 108 ) + (3.1 × 108 ) = (2.32 + 3.1) × 108 = 5.42 × 108 .

But you won’t always be adding two numbers with the same exponent. Suppose, for example, that
you need to compute the following sum:

(2 × 109 ) + (4.3 × 1010 ).

These numbers do not have the same exponent and therefore cannot be added together directly. We
must rewrite one number to have the same exponent as the other. Let’s rewrite 2 × 109 to have an
exponent of 10. We’ve already seen how to do this in the previous section; let’s express it by rules.

Rule 5

• For every increase in the exponent, we have to move the decimal point in the
significand to the left.
• For every decrease in the exponent, we have to move the decimal point in the
significand to the right.

This means that 2 × 109 = 0.2 × 1010 . Now we can add using Rule 4:

(0.2 × 1010 ) + (4.3 × 1010 ) = 4.5 × 1010 .

Alternatively, we could have rewritten 4.3 × 1010 , moving the decimal point in the other direction to
write 4.3 × 1010 as 43 × 109 . If we add using this approach, we get another situation in which we can
apply Rule 4:
(2 × 109 ) + (43 × 109 ) = 45 × 109 .
This is equal to 4.5 × 1010 , the same answer as before. Notice that depending on how we choose to
perform the calculation, we may need to re-express our final answer to make sure that it is in standard
form.

Critical Thinking

What does the method we’ve illustrated here have in common with the method we used to
solve Problem 12 on page 13?

Here are some more worked examples.

apprecia@ucalgary.ca
30 CHAPTER 1. MEASURING

Problem 10: Calculate 9.5 × 108 + 4.1 × 108 .

Solution:
9.5 × 108 + 4.1 × 108 = 13.6 × 108 = 1.36 × 109 .

Problem 11: Calculate 8.1 × 107 − 3 × 106 .

Solution:

8.1 × 107 − 3 × 106 = 8.1 × 107 + 0.3 × 107 = (8.1 − 3) × 107 = 7.8 × 107 .

Problem 12: Calculate 5.75 × 1027 + 2 × 1025 .

Solution:

5.75 × 1027 + 2 × 1025 = 5.75 × 1027 + 0.02 × 1027 = (5.75 + 0.02) × 1027 = 5.77 × 1027 .

Problem 13: Calculate


4.23 × 1023 − 9.1 × 106 .

Solution:
4.23 × 1023 − 9.1 × 106 ≈ 4.23 × 1023 .
Although we might consider 9.1 × 106 = 9,100,000 to be a large number, it is tiny compared to
4.23 × 1023 , which is
423,000,000,000,000,000,000,000
A value in the millions has little impact on a number that is in the thousands of sextillions. So the
subtraction is a negligible change.

Think about it. . .

A student suggests that the rule for adding two numbers in scientific notation should be
“add the significands and add the exponents.” How will you convince this student that he is
wrong? Avoid appeal to authority (“The textbook says that you should do it this other way”)
and try to find an approach, maybe by means of examples, that helps the student understand
why the correct rule makes sense, whereas his proposal does not.

apprecia@ucalgary.ca
1.2. SCIENTIFIC NOTATION 31

Using a Calculator

Most calculators or calculator apps allow you to enter numbers in scientific notation. Typically, you
should look for a button labeled “E” or “EE.” Some calculators may have a button labeled “Exp”
instead.5 You can enter a number in scientific notation like this:

significand EE exponent.

For example, Figure 9 shows the result of entering 1.56 × 109 into a smart phone calculator app. In
this particular app, rotating the phone to the vertical position changes the display to scientific notation,
“1.56e9,” as shown in the same figure. It also changes the buttons available on the calculator.6

(a) The result of entering “1.56 EE 9” (b) The result of rotating


the phone to the vertical
position

Figure 9: A smart phone calculator app. Turning the phone on its side changes the display and the set of buttons available.

Another smart phone app, shown in Figure 10 on the next page, displays both what is entered
(below the white line) and the result (above the white line). We’ve entered the number 97.5 × 1015
using “97.5 E 15,” which displays as 97500000000000000 (shown on the left). We can switch to
scientific notation by pressing “FIXED” in the display, which changes the mode from “FIXED” to
“FLOAT” (shown in the center).7
Once you know how to enter numbers in scientific notation, you can do arithmetic on them
with your calculator in the ordinary way. We’ve used the previous calculator app to work out
5.7 × 10−8 × (2.6 × 1012 ); the result is shown in Figure 10 on the next page (on the right).


5 The calculator included with Windows 10 uses “Exp” for this key. However, this is a terrible choice, since many other

calculators, as well as advanced mathematics textbooks, use “exp” for the exponential function y = ex where e ≈ 2.718 is an
important mathematical constant. If you’re going to use Windows 10’s calculator, you will have to get used to this (and the
exponential function won’t appear anywhere in this book, which reduces the possibility of confusion), but nevertheless, watch
out if you switch to another calculator app or program.
6 This is the case with many calculators available on smart phones. If you are using your phone as a calculator, try turning

the phone on its side and see whether there are more or different buttons available.
7 “Fixed” is short for fixed-point notation. The decimal point remains in its usual place. “Float” is short for floating-point

notation. We use powers of 10 to move the decimal point to a place that makes the number easier to read (putting the number
into standard form).

apprecia@ucalgary.ca
32 CHAPTER 1. MEASURING

Figure 10: A smart phone calculator app. To enter numbers in scientific notation, use the E button. Results can be displayed
in the usual way (“FIXED” mode, shown on the left) or using scientific notation by choosing the “FLOAT” mode (center and
right).

Summary of Ideas: Scientific Notation

• Scientific notation is a concise way of representing quantities that are very large or
very small.
• A number is represented in scientific notation as the product of two numbers, a
significand (between 1 and 10 if the number is in standard form) and a power of
10, written 10n for some integer n called the exponent. For example, 2.32 × 10107
is an expression in scientific notation, standard form; the significand is 2.32 and the
exponent is 107.
• To write a quantity that has both a number part and a unit part in scientific notation,
re-express only the number part; the unit part remains unaffected.
• For every increase (decrease) in the exponent, we have to move the decimal point in
the significand to the left (right). For example, 4.3×103 can be rewritten as 0.43×104
or as 43 × 102 .
• When multiplying (dividing) two numbers in scientific notation, you must multiply
(divide) the significands and add (subtract) the exponent. For example, (6 × 102 ) ÷
(3 × 107 ) = 2 × 10−5 , since 6 ÷ 3 = 2 and 2 − 7 = −5.
• When adding (subtracting) two numbers in scientific notation, you must first make
their exponents the same and then add (subtract) their significands.

apprecia@ucalgary.ca
1.3. ESTIMATES, PRECISION, AND ORDERS OF MAGNITUDE 33

1.3 Estimates, Precision, and Orders of Magnitude

Objectives

 I know what is meant by the precision of a measurement claim.


 I can understand and use significant figures as a measure of precision.
 I can express the result of a calculation with appropriate precision, taking into account
the precision of the data used in the calculation.
 I can say when two quantities have the same the order of magnitude.
 I can make order-of-magnitude estimates of quantities that are relevant to sustain-
ability.

1.3.1 Precision

Example 1: An old story [345] tells of a museum guard whose


post is next to a dinosaur skeleton. One day, a kid asks the guard
how old the skeleton is. He responds, “It’s 70,000,003 years old.”
Puzzled about the “3,” the kid asks the guard how he knows.
“Well,” he replies, “when I started work I was told that the skeleton
was 70 million years old. . . and that was three years ago.”
The 70-million-year figure was some expert’s best estimate.
Perhaps what the expert meant was “between 65 and 75 million
years,” in which case we’d say that the precision of her estimate
was plus or minus 5 million years. Maybe the estimate was more
specific than that—perhaps it was only plus or minus 500,000 Figure 11: Dinosaur skeleton.
years. But surely it was not specific enough for an additional 3 years to make any significant
difference! The guard’s mistake arose from a misunderstanding of precision: by using all the digits of
the number 70,000,003, he gave an answer that seemed much more precise than was in fact possible.

Example 2: Here is another example that will help us come to grips with this notion of precision.
Consider the following statements about the Earth’s current human population.
(a) The population of Earth is 7,188,895,672.
(b) The population of Earth is about 7 billion, plus or minus a billion.
(c) The population of Earth is between 109 and 1010 .
(d) The population of Earth is 6.4 × 107 , to 2 significant figures.8

Each of statements (a)–(d) makes a claim that Earth’s human population is within a certain range of
values. In statements (b) and (c) the range of values is given explicitly, while in statements (a) and (d)
the range of values is implied. A statement of this kind, about a physical or environmental quantity, is
what we call a measurement claim.
8 Significant figures will be explained in more detail later in this chapter.

apprecia@ucalgary.ca
34 CHAPTER 1. MEASURING

Definition 1

A measurement claim, about some physical or environmental quantity, is the claim that
the true value of that quantity lies within a certain range.

When someone makes a measurement claim we often pay most attention to the center of the range
of values (“Polls show Clinton ahead by three percentage points”). Often the width of the range
receives less attention (Footnote: “The margin of error is plus or minus five percentage points”), but
it is equally important. That width is what we call the precision of the measurement claim.

Definition 2

The precision of a measurement claim, about some physical or environmental quantity,


describes the width of the range of values within which the true value of that quantity is
claimed to lie. This width can be described explicitly, or can be expressed implicitly using
significant figures or in other ways.

Thus, of the four statements (a)–(d) above, statement (a) is the most precise, and statement (c) is
the most imprecise. The others are intermediate in their level of precision, with (a) being more precise
than (d), which is more precise than (b), which is more precise than (c).
Notice that precision is still a claim! A highly precise measurement claim does not for that reason
have to be true (nor does a highly imprecise claim have to be false). In fact, of the statements (a)–(d)
above, which are actually true? At the time of writing, (b) and (c) are true, while (a) and (d) are false.
The highly precise statement (a) may have been true for an instant sometime in 2014, but it is now
false; the moderately precise statement (d) has not been true for thousands of years.

Remark 1: Notice that statements (a)–(d) also provide examples of several different ways of
describing the precision of a measurement claim. You could explicitly give a range of values for
the measurement (example (c)); you could specify a single value with a possible error “plus or minus”
(example (b), and also the language we imagined our expert using in the dinosaur example above);
or you could use the notion of significant figures (example (d)), which we’ll discuss in a moment.
Example (a) does not include a specific statement of precision, but since it specifies a whole number
of discrete objects without mentioning possible error, its implied claim is to complete precision—“this
exact number is the population of the Earth.” It is difficult to imagine how anyone could accumulate
sufficiently precise data to make such a claim, even for single instant.

Remark 2: You may read, in some texts, about a distinction between the words precision and
accuracy. For instance, consider again the examples (a)–(d) above. Even though both (a) and (d)
are false, we might want to say that (a) is more accurate (that is, closer to the truth) than (d).
For another example, consider the Volkswagen emissions scandal [72]. Many models of VW diesel
automobiles were equipped with special software that detected when EPA-mandated emissions tests
were being run, and if so, altered the engine’s settings to make it behave more “virtuously” than
normal. Measurements from these tests could then have been highly precise, but they would not have
been accurate. In this text, we will speak mostly of precision, and we won’t use the word “accuracy”
in this specific, technical sense.

apprecia@ucalgary.ca
1.3. ESTIMATES, PRECISION, AND ORDERS OF MAGNITUDE 35

Critical Thinking

Open a newspaper today and find some examples of measurement claims. Which (if any) of
them include an explicit description of the level of precision that they claim? For those
that don’t, what seems to be the implied level of precision? Do you think that these
measurements are, in fact, as accurate as the claims made for them? Why or why not?

1.3.2 Significant Figures


One way to describe the precision of a measurement claim is to use the concept of significant figures.
Consider again the example of the dinosaur skeleton. If we know that its age is 70 million years
“plus or minus five million,” that tells us that the true age is nearer to 70,000,000 than to 60,000,000
or to 80,000,000 or to any other number given by one digit followed by a string of zeros. We would
say that the 70-million-year estimate is accurate to one significant figure.
Or if we learn from census data that the population of Pennsylvania, on April 1, 2010, was
12,702,884 plus or minus 40,000, then we know that the true population was closer to 12,700,000
than to any other number given by three digits followed by a string of zeros. We would say that the
count of 12,700,000 is accurate to three significant figures.
These examples involved integers (whole numbers). To understand the concept of “significant
figures” when decimals are involved, it helps to use our work with scientific notation. Our examples
can be written
70,000,000 = 7 × 107 , 12,700,000 = 127 × 105 .
Notice that we have written both numbers so that the significand is an integer. When this is done, the
number of digits in that integer is the number of significant figures.

Definition 3

An approximation to a number x, using n significant figures, is the n-significant-figure


number that is closest to x. (We also can say that the approximation is obtained by rounding
off x to n significant figures.)

Problem 1: Write an approximation to 83,333,333,334.4 in standard form, using one significant


figure.

Solution: The leading digit in the given number is 8. Because we are told to use only one significant
figure, the subsequent digits (the 3’s and the 4’s) are simply place holders—the only information that
they provide is how far to the left of the decimal point the 8 occurs. In fact, all these subsequent digits
may as well be zeros. In other words, we can write
83,333,333,334.4 ≈ 80,000,000,000 = 8 × 1010 .
If we were asked to write an approximation to the same number using two significant figures, we
would keep the first two digits (the remainder being place holders) and write
83,333,333,334.4 ≈ 83,000,000,000 = 8.3 × 1010 .
Look carefully at the next example to see how we may sometimes need to take the size of the first
“placeholder” digit into account also.

apprecia@ucalgary.ca
36 CHAPTER 1. MEASURING

Problem 2: According to the 2010 census, the U.S. population on April 1, 2010, was 308,758,105.
Find an approximation to this, using three significant figures.

Solution: According to our rules, a number with three significant figures can be expressed in scientific
notation with a significand that is a three-digit integer. So we write the given number in scientific
notation with a significand that has three digits before the decimal point:

308,758,105 = 308.758105 × 106 .

Then we choose the nearest integer to the significand. Here, since 0.758105 is greater than 0.5, we
“round up”: the nearest integer significand is 309, and our answer to a precision of 3 significant figures
is 309 × 106 (it would also be correct to express this in standard form as 3.09 × 108 , or to write it out
in long form as 309,000,000).
We can express the technique that we used in the previous solution as a rule.

Rule 1

To approximate a datum x to n significant figures, first express x in scientific notation


in such a way that the significand has n digits before the decimal point, then replace
the significand by the nearest integer to it.

When you make a measurement claim, you should approximate it to a number of significant figures
that reflects the precision of your claim. What does this mean? I just measured the depth of the
computer keyboard on which I’m typing this manuscript, and got an answer of 16.4 centimeters.
Then I asked Google what that is in inches, and got the answer 6.456693 inches. Does that mean
that I should quote 6.456693 inches as the depth of my keyboard? No! In the first place, I am not
the world’s most careful measurer. Let’s say that my original measurement of 16.4 cm was within
0.5 cm of the true value. That means it has two significant figures. Then I should not claim a greater
degree of precision in the measurement after I have converted it to inches than it had before. Thus, the
appropriate way to express my measurement claim in inches would be “the depth of my keyboard is
6.5 inches (to two significant figures).”
Sometimes, people make measurement claims without specifically stating the number of significant
figures involved. What to do in this situation? We try to interpret the claim as “economically” as
possible, which is to say that we remove all the trailing zeros (zeros to the right that are followed only
by more zeros) and leading zeros (zeros to the left that are preceded only by more zeros). The number
of digits that are left is the implied number of significant figures.

Rule 2

If a measurement claim is made without specifically stating a number of significant


figures, the default assumption is that as few as possible of the digits are significant:
that is, the number of digits left after leading and trailing zeros have been removed.

The following examples show how this works out in practice:

apprecia@ucalgary.ca
1.3. ESTIMATES, PRECISION, AND ORDERS OF MAGNITUDE 37

Problem 3: Recent annual wheat yields in Bangladesh are stated to be 1,370 pounds per acre. How
many significant figures are claimed by this statement?

Solution: The final 0 is a trailing zero; the other three digits are significant. Thus the default
assumption is that our original claim was made to three significant figures.

Problem 4: The manufacturer claims that a new hybrid automobile obtains 61.07 miles per gallon.
How many significant figures are implied by this claim? Do you believe that the manufacturer can, in
fact, measure gas mileage to this level of precision?

Solution: There are no leading or trailing zeros, so the manufacturer’s original claim was expressed
with four significant figures. Notice that the zero between the 1 and the 7 is significant.
It is highly unlikely that the manufacturer, or anyone else, can measure gas mileage to the implied
precision of one part in ten thousand. Even if such a precise measurement could be carried out under
highly controlled conditions, such a level of precision would be irrelevant to highway driving. The
manufacturer’s statement is probably just trading on the idea that a more complicated-looking number
is somehow more impressive or “scientific.”

Remark 3: You might have noticed what seems like an ambiguity in our definitions. For instance, how
many significant figures has the number 200? We can write it as 2×102 (one significant figure). But we
could also write it as 20 × 101 (looks like 2 significant figures), or even as 200 × 100 or 2,000 × 10−1
(three or four significant figures)! Rule 2 tells us that we take the least number of significant figures
that makes sense: one significant figure in this example. If we want to express greater precision than
that, we need to add extra explanation. For instance, “there are 200 teaspoons in a liter, to 2 significant
figures,” means that there are between 195 and 205 teaspoons in a liter. (The exact figure is 202.884)

Significant Figures and Calculations


An important fact about calculations is this: you cannot get a more accurate result out of a calculation
than the accuracy of the data you put into it. (As the computer scientists say: “Garbage in, garbage
out”!). In terms of significant figures this translates to the following rule of thumb:

Rule 3

• In multiplying or dividing two or more numbers, the number of significant figures in


the result is the smallest number of significant figures in any piece of the input data.
• The result of a calculation should be approximated to the appropriate number of
significant figures (given in the previous bullet point).
• If a calculation involves intermediate steps, don’t approximate the intermediate steps.
Only approximate when you have reached the final result.

Thus, for example, if we multiply one number (a) that is given to five-figure accuracy by another
number (b) that is given only to two-figure accuracy, the result (a × b) will have only two-figure
accuracy, and we should round it off to express this. (There is also a rule for how to deal with
significant figures when adding or subtracting, but it is more complicated—just as, on pages 27–
29 in the previous section, Rules 4 and 5 are more complicated than Rule 3—and we are not going to
worry about it.)

apprecia@ucalgary.ca
38 CHAPTER 1. MEASURING

Remark 4: Using a calculator can easily tempt one to express something with inappropriate precision.
For example, suppose that next time I fill my gas tank, the odometer indicates that I have driven 316.2
miles since my last fill-up, and the gas pump shows that I used 9.12 gallons. Using my calculator, I
obtain
316.2
= 34.6710526
9.12
(and I can compute more decimal places if I like). But it would be ridiculous to quote my gas mileage
in this form, which suggests that I measured all the quantities involved to a precision of 1 part in a
billion! The input data claim 4-figure precision (the odometer reading) and 3-figure precision (the gas
pump measurement), so it would be appropriate to express the result to 3-figure precision at most,
34.7 miles per gallon. In fact, you might prefer to use the nearest whole number and say, “My car gets
about 35 miles to the gallon.”

Definition 4

Mathematicians use the symbol ≈ to signify that two quantities are approximately equal—
equal within the level of precision appropriate to the context. Thus, using this symbol, we
could write the mileage calculation of Remark 4 as
316.2
≈ 35,
9.12
adding the phrase “to 2 significant figures” if we needed to make it clear what level of
precision we were using.

Problem 5: A local lake has a surface area of 1.273×106 m2 and an average depth of 5.1 m. A polluter
dumps a truckload (7,300 gallons) of contaminated wastewater into the (previously unpolluted) lake;
each gallon of the wastewater contains about 2 grams of arsenic. Assuming that the contaminated
wastewater is uniformly dispersed throughout the lake, find the resulting concentration of arsenic
in the lake water, expressed9 in micrograms per liter (µg/ L). Be sure to express your answer with
appropriate precision.

Solution: The concentration of arsenic in the lake will be the total weight of arsenic deposited, divided
by the total volume of lake water. Using the unit-factor method, we calculate

2g 106 µg
Weight of arsenic = 7,300
gal
× × = 1.46 × 1010 µg.
1gal
 1
g
Because this is an intermediate result, we do not approximate it yet. Similarly,
1,000 L
Volume of lake = (1.273 × 106 )
m2 × (5.1
m) × = 6.4923 × 109 L.
1m3
 

Again, this is an intermediate result. Our final answer will be obtained by dividing the weight of
arsenic by the volume of the lake, and then approximating to the appropriate number of significant
figures. How many significant figures are appropriate? The data have 4 significant figures (lake area),
2 significant figures (lake depth), 2 significant figures (truckload volume), and 1 significant figure
(level of contamination of the wastewater). According to Rule 3 on the previous page, we should
approximate the result to the smallest of these, i.e., 1 significant figure. Thus we write
9 This “concentration” is a weight divided by a volume, so it is not a pure number. Compare Example 5 on page 15;

where is the important difference?

apprecia@ucalgary.ca
1.3. ESTIMATES, PRECISION, AND ORDERS OF MAGNITUDE 39

1.46 × 1010 µg
Arsenic concentration = = 2.2466 . . . µg/ L ≈ 2 µg/ L.
6.4923 × 109 L
The answer is 2 micrograms per liter, to the appropriate level of precision, which is 1 significant
figure. (For comparison, in 2001 the EPA set a maximum level of 10 micrograms per liter for arsenic
in domestic drinking water [12].)

1.3.3 Orders of Magnitude


One or two significant figures give a useful level of precision. In many of the calculations that we will
do in this course, one or two figure precision will be quite sufficient. Quite often, indeed, the quantities
that we are trying to understand cannot be measured with greater precision than this.
But there are other situations in which even one significant figure represents a greater level of
precision than we can attain—or than it is useful to work with. Is there a notion of precision still more
wide-ranging than that expressed by “one significant figure”? Indeed there is: in fact we have seen an
example in Statement (c) of Example 2 on page 33. The level of precision expressed here is called an
order of magnitude. Let’s make a definition:

Definition 5

Let A and B be two numbers, or two physical quantities measured in the same units. We
say that A and B have the same order of magnitude if when we divide them, the result is
between 10−1 and 10,
1 A
< < 10.
10 B
It does not matter which quantity we pick as A and which as B: the definition gives the same
result either way.

Problem 6: Do the numbers 7 and 28 have the same order of magnitude?

Solution: Yes. Dividing the larger number (28) by the smaller (7) yields 4, which is between 10−1
and 10.

Problem 7: Does 1 million feet have the same order of magnitude as 5 kilometers?

Solution: In order to answer this question, we need to express both quantities in terms of the same
unit. We choose to use kilometers as the common unit, so we must express 1 million feet in kilometers.
Using the conversion factors from Tables 2 on page 8 and 8 on page 12, we get
1 yd 1mi 1.61 km
106 ft ×  ×

3 ft 1760
yd
×
1 ≈ 305 km.
mi
Thus the ratio
106 ft 305 km

≈ ≈ 61
5 km 5
km

is greater than 10, so the answer is no: the two lengths are not of the same order of magnitude.

Problem 8: According to Vaclav Smil, a professor at the University of Manitoba, the total mass of
all the termites on Earth is at least 500 million tonnes. In terms of order of magnitude, how does this
compare with the mass of all the humans on Earth?

apprecia@ucalgary.ca
40 CHAPTER 1. MEASURING

Solution: Let’s try to estimate the mass of all humans on Earth. We’ll assume 100 kg for the mass
of a human, which is something of an overestimate but gives an easy calculation. The total mass of
humans is equal to the number of humans times the mass of each one. Let’s include the conversion
factor from kilograms to tonnes as well, so as to get a final answer in tonnes:

7 × 109   × 100
humans
k
g 1T
= 7 × 108 T.
 × 103

1human kg


Divide this by 5 × 108 T of termites, and you’ll see that we get a number between 10−1 and 10. That
means the two masses have the same order of magnitude. When all insect species are taken into
consideration, the total mass of insects is thought to be of greater order of magnitude than that of
humans—in fact, several tonnes of insects per human!
Orders of magnitude are closely related to scientific notation,
the topic that we have studied in Section 1.2. When we represent a
number in standard form, the most important information about that
number is the power of 10, or the exponent. In fact, if we completely
neglect the significand and simply take the “nearest” power of 10,
we will get a number that has the same order of magnitude as the
original one. Thus, “order of magnitude” calculations simply involve
manipulating powers of 10.
Remark 5: You might ask which power of 10 is “nearest” to a given
significand. The rule we use is that significands (in standard form)
beginning with 1, 2, 3 are “nearest” to 1, while significands beginning
Figure 12: Planes at San Francisco Airport. with 4 and greater are “nearest” to 10. So, for instance, the “nearest”
power of 10 to Earth’s human population of 7 billion is 1010 , not 109 ,
whereas the “nearest” power of 10 to the mass of a flea (estimated at
2.2 × 10−4 g) is 10−4 g.

Estimating Orders of Magnitude


The order of magnitude of a quantity is a very crude approximation to that quantity. You might be
surprised to hear, though, that their very crudeness makes orders of magnitude extremely useful. There
are two reasons for that:

• Orders of magnitude are easy to estimate. We already observed this in Problem 8 on the
previous page where we used the easy order-of-magnitude estimate 100 kg for the mass of an
average human being. We know that this number (about 220 pounds) is a bit of an overestimate,
but it’s certainly closer than 10 kg or 1000 kg, so it has the right order of magnitude.

• If two quantities have different orders of magnitude, then that difference is pretty significant. If
the quantities are related to sustainability questions, an order of magnitude comparison can give
us important information on which to prioritize. For example, an article in the San Francisco
Chronicle [269] reported that the installation of solar panels at San Francisco International
Airport’s Terminal 3 “is expected to reduce energy use at the airport by 15 percent” within a few
years (Figure 12). That’s true—if you neglect the energy used by the actual airplanes arriving
and departing from the airport, which are, after all, the reason for the airport’s existence! The
amount of energy used by these airplanes is about three orders of magnitude greater than the
savings from installing solar panels.

apprecia@ucalgary.ca
1.3. ESTIMATES, PRECISION, AND ORDERS OF MAGNITUDE 41

Think about it. . .

When you read that energy use by airplanes at San Francisco Airport is much greater than
the savings coming from installing solar panels at the airport, what is your reaction?
(a) Installing solar panels at an airport is simply a distraction from the much larger
question of the fossil fuel consumption (and greenhouse gas emissions) created by
our desire for air travel.
(b) Why are you so negative about a well-intentioned effort? Every little bit helps.
(c) The comparison is not appropriate. The demand for air travel comes, through the
market, from thousands of consumers. The San Francisco Airports Authority is one
public utility making its contribution to save energy and fight climate change.
(d) Something else.
Whatever alternative best describes your reaction, try to imagine the best arguments that
someone else could make for the other alternatives. How would you respond to them?

What matters most in understanding a quantity is its order of magnitude. And, as we said
above, orders of magnitude are relatively easy to estimate. In their book Guesstimation [345],
authors Lawrence Weinstein and John Adam give two simple ideas for generating order-of-magnitude
estimates. Here they are.

Rule 4: Order-of-Magnitude Estimation

Idea 1: Write down the answer, straightaway. In other words, come up with a “reasonably
close” solution. Remember we are only trying to find the nearest power of 10.
Idea 2: If you can’t estimate the answer straightaway (in other words, if Idea 1 fails), break
the problem into smaller pieces and estimate the answer for each one. You only need
to estimate each answer to within a factor of ten. How hard can that be?

Here are some examples.


Problem 9: Find an order-of-magnitude estimate for the number of clowns who can fit inside a VW
Beetle.

Solution: This is a problem we can solve using our imagination directly (Idea 1). Clearly, 4 clowns
will fit into a Beetle, and 40 clowns will not. Since our answer must be a power of 10, bigger than 4,
and smaller than 40, our order-of-magnitude answer is 101 .

Problem 10: Find an order-of-magnitude estimate of the volume of gasoline burned in a year by a
typical U.S. driver.

Solution: This is probably not something we can imagine directly. But we can break the problem
down into two simpler ones (Idea 2) that we can solve using our imagination and experience:
• How many miles are driven per year by the typical driver?

apprecia@ucalgary.ca
42 CHAPTER 1. MEASURING

• What is the typical gas mileage?


How many miles a year does the average American drive? Is it 12,000? 15,000? 30,000? Notice that
all of our guesses have the same order of magnitude. While we might not agree what number is a
good guess, we can agree that the order of magnitude should be 104 . Similarly, what about the typical
gas mileage? You may know cars that get 20 miles per gallon while others get 30 miles per gallon.
Regardless, the order of magnitude is 101 . Therefore, our order-of-magnitude answer is

 × 1 gal = 103 gal.


104
mi
101mi

Our estimate is 1,000 gallons of gasoline per driver per year. In this example we can compare our
back-of-the-envelope estimate to a more accurate figure. According to the U.S. Energy Information
Administration [6], the U.S. consumed about 3.3 × 109 barrels of highway gasoline in 2015
(remember, a barrel is 42 gallons). According to the Department of Transportation [111], we may
estimate10 approximately 2.2 × 108 licensed drivers in that same year. Dividing, we get

3.3 × 109 × 42 gal ÷ (2.2 × 108 drivers) = 630 gal


bbl
1bbl
 driver
which is fairly close to our order-of-magnitude number.
Problems that require you to estimate orders of magnitude are sometimes called Fermi problems,
after the physicist Enrico Fermi, who was famous for the skill and speed with which he could
make estimates of this sort (see the video [220]). Such estimates are a great way to understand the
world around you from a simple yet quantitative perspective, whether you use that skill to work on
sustainability or on other issues. According to the authors of [345],
You can also use these tools to further your career. Many top companies use estimation
questions in job interviews to judge the intelligence and flexibility of their applicants.
Leading software firms, management consultants, and investment banks. . . ask questions
such as What’s the size of the market for disposable diapers in China? How many golf
balls does it take to fill a 747? and How many piano tuners are there in the world?
Companies use these questions as an excellent test of the applicants’ abilities to think on
their feet and to apply their mathematical skills to real-world problems.
We are not necessarily encouraging you to become an investment banker, but this is a good moment
to remember that the skills we are learning through this book have a very wide range of applications!
You can get much more practice in making these estimates at the Fermi Questions website [217]. Here
is one more example from [345]:
Problem 11: Find an order-of-magnitude estimate for how many people are airborne (as commercial
airline passengers) over the United States right now.

Solution: Again, our imagination probably doesn’t give us any direct access to this number. But what
we can do without too much trouble is to think about our own experience. How many flights did you
take in the last year? How long did they last, in total? Then we argue as follows (Idea 2): if your
flying is representative of the average American, then the fraction of your time that you spent in the
air is the same as the fraction of everyone else’s time that they spend in the air, which is in turn
equal to the fraction of the U.S. population that is in the air right now. Now we know that the U.S.
population is about 3 × 108 , and I might guess that I spend 12 hours (half a day) in the air each year.
Imagining myself as representative of the average American, I estimate that the average American
10 The most recent year for which the OHP provides data is 2009; this estimate assumes that the growth in the number of

drivers continued for the next years at the same rate as in the previous decade.

apprecia@ucalgary.ca
1.3. ESTIMATES, PRECISION, AND ORDERS OF MAGNITUDE 43

spends 21 /365 ≈ 10−3 of his/her time airborne. Therefore, I estimate that the number of airborne
Americans right now is
(3 × 108 ) × 10−3 ≈ 3 × 105 ,
or, to the nearest power of 10, just 105 . That is a bigger number than I might have expected!

Summary of Ideas: Estimates, Precision, and Orders of Magnitude

• The precision of a measurement claim expresses how close the measurement is


claimed to be to the true value of the quantity measured.
• One way of describing precision is in terms of the number of significant figures.
• When we multiply or divide quantities, the number of significant figures in the result
is the least number of significant figures among the quantities we multiplied or
divided.
• Another (lesser) level of precision is described by order of magnitude. Two
quantities (with the same units) A and B have the same order of magnitude if
1 A
< < 10.
10 B

• An Order-of-magnitude estimate for a quantity refers to the nearest power of 10 to


that quantity.
• Order-of-magnitude estimates can often be obtained easily by following Rule 4 on
page 41.

apprecia@ucalgary.ca
44 CHAPTER 1. MEASURING

1.4 Communicating Quantitative Information

Objectives

 I can understand when and why I might want to communicate quantitative informa-
tion to nonmathematical readers or listeners.
 I can distinguish between verbal, tabular, and graphical ways to communicate
quantitative information, and I know when each is effective.
 I can design tables to be clear and readable.
 I can identify when it is appropriate to use a line graph, a bar chart, a scatter plot,
and a pie chart to represent a set of data.
 I can design each of the above four types of graphical presentation to be clear and
readable.
 I am aware of some of the ways graphical presentations can be used to misrepresent
data. I will avoid them in my own work and be alert for them in the work of others.

By the time you have finished this book, you will be able to carry out quite sophisticated calculations
about growth, pollution, planetary climate, risk and risk management, decisions, and payoffs. The day-
to-day work of a professional scientist in one of these areas—an atmospheric physicist or an ecological
economist or a statistician—will involve still more complicated calculations of the same general kind.
In the end, though, all the analysis and calculation ought to lead a democratic society to some
decisions. Do we build this pipeline? Do we sign this treaty? Do we subsidize this innovative industry
(and if so, by how much)? Do we build a nationwide, interconnected energy grid, or would a
more decentralized system be safer or cheaper? And as society tries to make these decisions, the
scientists and mathematicians need to communicate quantitative information in a way that makes its
implications clear and easy to grasp. That is what this section is about.
We will think about three ways of communicating such information: verbal (using words),
numerical (using numbers in a text or table), and graphical (using charts or graphs).

1.4.1 Verbal Communication


Sometimes, the quantitative data to be presented are so straightforward that a summary in words
is enough to vividly convey the underlying information. This is often the case when what we wish
to convey is a single “data point” rather than a comparison (between multiple data points) or a trend
(information about how something varies over time). We saw a striking example of this in Example 10
on page 20. After some calculation, we reached the conclusion that “out of every ten gallons of water
on Earth, just under a teaspoon is fresh.” No numbers or charts are needed to explain this; it stands
by itself as a reminder of how rare a resource fresh water is, even though water itself is abundant on
Earth and indeed covers well over half the planet’s surface.
Here is another, similar example.
Problem 1: Geologists tell us that our present-day oil reserves were formed from the bodies of tiny
marine organisms that became trapped in ocean-floor sediments during the Mesozoic period, roughly
100 million years ago. In roughly 250 years of industrial civilization, humanity has used about one-
quarter to one-half of these reserves. Find a vivid verbal way to contrast the length of time that it took
our oil reserves to be formed and the speed at which we are burning them up.

apprecia@ucalgary.ca
1.4. COMMUNICATING QUANTITATIVE INFORMATION 45

Solution: If we assume (perhaps optimistically) that humanity has consumed only 41 of the Earth’s oil
reserves in 250 years, then the length of time to consume the entirety of those reserves would be
250 yr ÷ 14 = 250 yr × 4 = 1,000 yr.
Since the time to form the reserves is 100,000,000 yr, and
100,000,000 yr
= 100,000,
1,000 yr
we could say “humanity is burning up Earth’s oil reserves a hundred thousand times faster than they
were formed.” That is pretty striking! But let’s look for a way to illustrate it in still more direct terms—
in a way that does not involve the big number 100,000 (which is too large for many of us to wrap our
heads around).
Remember, there is no one right answer here. But a way of expressing the information that may
be very evocative involves rescaling (see Example 10 on page 20 for this concept) the times involved
by a factor of 100,000,000. Thus the 100,000,000-year period to form the oil reserves rescales to 1
1
year, and the 1,000-year period to burn them up rescales to 100,000,000 of 1,000 years. We’ll convert
the answer to minutes:
1 365day
 24 hr 60 min
× 1,000yr × ≈ 5 min.

× ×
100,000,000 1
yr 1
day 1hr
  

Thus, if the time to form all our oil reserves is represented by a year, the time during which humanity
is expected to consume all of them is represented by five minutes.
Every Fourth of July, State College, PA (the home of Penn
State), holds the largest all-volunteer fireworks display in the
United States. Planning for the Central Pennsylvania 4th Fest
begins on July 5th the year before, so it is not inaccurate to
say that the Fest takes a year to set up. It usually takes 30–45
minutes to shoot off the 12,000 or so firework shells that make
up the display. That makes an impressive, indeed overwhelming,
amount of noise, light, and smoke.
Imagine now that, perhaps owing to a software error, all the
shells are shot off in five minutes11 instead of 45. You are
basically looking at a nonstop explosion for five minutes. But
you are also looking at an analogy for the extraordinary rate at
Figure 13: Fireworks.
which humanity is consuming its fossil fuel reserves. A year to
set up, five minutes to burn. This picture of a crazy, out-of-control fireworks display provides a vivid
way of representing the numerical data about oil consumption. And it invites the right question: No-
one thinks of a fireworks display as anything other than a brief, spectacular excitement—it lights up
the night for a moment, but darkness rolls back again soon enough. What might make us think, then,
that by our discovery of oil and other fossil fuels we have driven darkness away forever?

Human Terms Again


Did you recognize what we were doing in the previous example? We were simply following the
methods laid out in Definition 3 on page 17 and the subsequent discussion to express the age of our oil
reserves in human terms. Specifically, we were using the rescaling method to translate the comparison
between two large quantities—the age of the oil reserves and the length of time it may take to consume
them—into another comparison that it is easier for us to grasp: the comparison between the one-year
setup time for the fireworks show and the five minutes it takes to shoot off all the shells at once.
Here is one more example, which relates to plastic pollution in our oceans.
11 Something like this actually happened in San Diego in 2012.

apprecia@ucalgary.ca
46 CHAPTER 1. MEASURING

Example 1: Consider the Great Pacific Garbage


Patch (GPGP). If you know a little bit about ocean
currents, then you probably know that they move in
circular patterns called gyres. (There’s a bunch of
reasons for this—the basin topography (the shape of
the land under the ocean), the rotation of the Earth,
the differences in temperature and saltiness of the
different parts of the ocean—are three of the most
important.) When garbage (especially plastic) makes
its way to the ocean, these gyres pool it in particular
locations. The biggest of the gyres is the North Pacific
Gyre, and the pool of our garbage in this gyre has
become infamous as the Great Pacific Garbage Patch
Figure 14: Ocean currents and garbage.
(see Figure 14, and [167] for more information). It
has been estimated that the area of the GPGP is
between 700,000 and 15,000,000 square kilometers.
How can we express these measurements in human terms? Because the areas involved are so large,
we can’t express them with “human size” units and small numbers—that is, we can’t follow both
parts of Rule 1 on page 17. But we can try to make the number part more humanly accessible by
choosing a measurement that will be more familiar to the reader and is of somewhat similar size to
the measurement we are trying to express. That is what we did by choosing the “volume of Lake
Superior” as a “unit” in Example 8 on page 19.
How might this work? Let’s take as our “unit” the area of the commonwealth of Pennsylvania.
(Pennsylvania happens to be our state, so it is reasonable for us to hope that at least some students
have traveled the state and have gained a feeling for the amount of land that it contains. But if you
are using this book in a different state, of course you should substitute that state for Pennsylvania and
change the numbers accordingly. The point is to have a large area that is a little familiar at least.)
Let’s do some calculations. We find online that the area of Pennsylvania is 119,283 square
kilometers. How does that compare to the size of the GPGP? We’ll consider both the “low end”
and “high end” estimates of its area.

Solution: We will start with the area of the GPGP and estimate how many Pennsylvanias we could fit
in that area. We will first compute the low estimate and then the high one:

1 Pennsylvania
700,000 km2 × ≈ 5.87 Pennsylvanias;
119,283 km2
1 Pennsylvania
15,000,000 km2 × ≈ 126 Pennsylvanias.
119,283 km2
From our two calculations, we see that the area of the GPGP is roughly between 6 Pennsylvanias
and 126 Pennsylvanias. That’s a very large range of large areas! But by expressing the area in terms
of “Pennsylvanias,” we have made the huge size at least a bit easier to grasp. It would be completely
safe to say, “An area of the Pacific Ocean many times larger than the state of Pennsylvania—perhaps
even a hundred times larger—is covered with plastic trash. What do you think about that?”

1.4.2 Tables
Simple verbal communication, as we have discussed in the previous section, can be used effectively
to convey numerical data, provided that only one or two numbers are involved. For example:

apprecia@ucalgary.ca
1.4. COMMUNICATING QUANTITATIVE INFORMATION 47

In a survey, 27 percent of students identifying as male said that they worked out regularly
(twice or more a week), compared to 34 percent of students identifying as female.

Or this:

At the end of 2015 the U.S. had an installed solar photovoltaic generating capacity of
approximately 25 gigawatts (GW) peak. The installed wind power generating capacity at
the same time was 75 GW peak.

But if three or more numbers are involved, a text including


them all will feel cluttered and will not be easy to understand. Source Peak Capacity (gigawatts)
For instance, suppose we also wanted to show the installed Natural Gas 504
power generating capacities sourced by coal, natural gas, Coal 305
Nuclear 104
hydropower, and nuclear fission. Writing all these down in a
Hydro 79
single sentence would be unwieldy and unhelpful. It is better Wind 74
to display the data in a table, as in Table 11. Solar PV 22

Remark 1: Notice that Table 11 contains only a couple of Table 11: U.S. installed electric generating capacity by
dividing lines—one beneath the first (title) row and one sepa- source in 2015 (gigawatts) [7].
rating the columns. Especially if you often use a spreadsheet
program like Excel, it is tempting to put each piece of data
into its own “box.” Don’t do this! The extra row and column dividing lines will actually make your
table harder, rather than easier, to read. If you want to help the reader’s eye move across the rows, an
alternating, gentle shading is usually more effective than heavy grid lines. For example, see Table 12,
which shows the steady increase in U.S. gas output over the past ten years.12

Year Q1 Q2 Q3 Q4
2005 4595 4652 4432 4370
2006 4474 4591 4693 4743
2007 4644 4784 4856 4980
2008 4999 5059 5001 5098
2009 5192 5162 5148 5120
2010 5115 5240 5404 5555
2011 5415 5665 5789 6031
2012 5937 5898 6091 6105
2013 5879 5999 6158 6168
2014 6077 6345 6585 6719
2015 6606 6736 6885 6804
Table 12: U.S. quarterly natural gas production (billions of cubic feet) [8].

We will summarize “best practice” for using tables in a series of rules:

12 This is mostly the result of hydraulic fracturing (“fracking”), a new technology that has enabled us to extract gas from

deposits that were previously thought to be inaccessible.

apprecia@ucalgary.ca
48 CHAPTER 1. MEASURING

Rule 1: Rules for Tables

• When you want to present three or more pieces of numerical information, consider
using a table rather than a verbal presentation.
• Presenting information in a table is most helpful when the reader will need to look
up individual values. (If trends or overall comparisons are of more interest, consider
a graphical presentation instead.)
• Since most people find it easier to see patterns in numerical data by reading down
columns rather than across rows, it is a good idea to plan your table so that the most
important patterns appear in the columns.
• The level of precision used for data presented in a table should not exceed the
minimum needed to communicate effectively with your readers. See the discussion
in Section 1.3.1, as well as Remark 2.

Remark 2: When we present numerical information in the form of a table, the question of precision
is bound to arise (see Section 1.3.1). Remember, the level of precision claimed by a numerical
measurement describes the width of the range (around the claimed value) within which the true value
of the measured quantity should lie. Thus, if I say (as in the last line of Table 12 on the previous page)
“U.S. natural gas production in the first quarter of 2015 was 6606 billion cubic feet,” I am implicitly
claiming a level of precision of plus or minus a billion cubic feet (compare Rule 2)—that is, I am
saying that the actual production level was somewhere in between 6605 and 6607 billion cubic feet.
In Section 1.3.1 we talked about how important it is, especially in environmental calculations, not to
claim a greater level of precision than we can in fact justify.

1.4.3 Graphical Communication


Along with verbal and numerical (or tabular) ways of communicating mathematical information, we
need also to be familiar with graphical communication. Using graphical communication, we encode
numerical values as visual objects, often presented relative to one or two axes. A graph, therefore, is
a visual display of quantitative information. There are many different kinds of graphs and charts. We
will consider four in this section: a bar chart, a pie chart, a scatter plot, and a line graph. Let’s think
about the two kinds of charts (bar charts and pie charts) first.
These charts are used when we are dealing with categorical information—data that is presented
in a variety of distinct categories. The information in Figure 11 on the previous page, for example,
is categorical—the energy source is categorized into coal, natural gas, nuclear and so on. We can
represent this data in the form of a bar chart by drawing six horizontal bars, of the same width but of
different lengths in proportion to the various measures of generating capacity. Thus we obtain a bar
chart that might look something like Figure 15 on the opposite page.

Remark 3: One can make bar charts with either horizontal or vertical “bars”; use whichever approach
is most convenient. A bar chart with vertical “bars” is sometimes called a column chart or “column
bar chart.”
In a bar chart, the lengths (or heights in the vertical case) of the bars represent the size of the
numerical data being conveyed. An alternative way of representing similar data is by a pie chart: in
this case, the numerical values are represented by the angles of different sectors rather than by the
lengths of different bars. Figure 16 on the opposite page represents the same data in pie chart form.

apprecia@ucalgary.ca
1.4. COMMUNICATING QUANTITATIVE INFORMATION 49

Solar 22
Wind 74
Hydro 79
Nuclear 104
Coal 305
Natural Gas 504

0 50 100 150 200 250 300 350 400 450 500 550
Peak Capacity (Gigawatts)

Figure 15: Bar chart representation of the data in Table 11 on page 47 on U.S. peak generating capacity from various sources
(gigawatts).

Natural Gas

504

22
Solar
74
305
Wind
79
Coal 104
Hydro

Nuclear

Figure 16: Pie chart representation of the data in Table 11 on page 47 on U.S. peak generating capacity from various sources
(gigawatts).

In both bar and pie chart representations, the areas of the bars or “pie pieces” represent the size of
the relevant numerical data. Because in a pie chart, the total area of all the pieces has to add up to one
whole pie, this kind of chart is particularly helpful in giving a visual impression of what proportion
of the total a particular category makes up. For instance, from the pie chart above you can quickly
see that coal, nuclear and hydro together account for about half of U.S. (peak) electricity generating
capacity.
Remark 4: It is helpful to use software such as Excel to prepare charts of this sort—of course
you can draw them by hand, but if a computer can draw them quickly and accurately instead, why
bother? One problem with using software, though, is that many programs offer a wide range of purely
cosmetic options—making the bars or pies “3-dimensional,” adding images or shading or shadows,
and so on. Most of these “effects” do not improve communication, but rather the reverse; this is
particularly so for 3-dimensional effects, which can confuse our understanding badly, for reasons
we will see in a moment. Therefore, even if you have access to software that enables you to add
“effects” to your charts, we encourage you not to do so. Remember, our desire is simply to achieve
clear communication; nothing more than that.

Bar and pie charts are used when we are relating categorical data (like the different types of
electricity generation) to numerical data. By contrast, scatter plots and line graphs are used when

apprecia@ucalgary.ca
50 CHAPTER 1. MEASURING

we are relating two different sets of numerical data. For example, we could consider the information
in Table 12 on page 47 as representing the relationship between two different sets of numerical data—
one set representing the year and quarter, and the other set representing the natural gas production in
that quarter. Figure 17 is a scatter plot showing the data from this point of view.
Remember, this plot shows exactly the same data as in
Table 12. Compared to the table, it has some obvious advantages
6,000 and disadvantages. If you want to know the exact amount of
natural gas produced in the third quarter of 2011, the table is
more useful to you. But if you want to estimate, overall, how
Production

4,000
fast U.S. natural gas production has been increasing over the past
decade, the graph will be much more helpful. It will even allow
2,000 you to “eyeball” the answer to a question like “is the increase in
natural gas production leveling off?”—it clearly looks from the
0 graph as though the answer is “no” (at least over the range of
2004 2006 2008 2010 2012 2014 2016
dates that it covers), whereas it would take some thinking even
Year
to figure out how to express the question in terms of data in the
Figure 17: Scatter plot of data from Table 12 on table, let alone how to answer it!
page 47 on U.S. natural gas production (billions of
cubic feet per quarter). The final kind of graphical representation that we mentioned
is a line graph. You can think of this, if you like, as a special
kind of scatter plot in which the points are so close together that
they make up a smooth curve. The kinds of “graphs” you learned about in high school algebra (“Draw
the graph of y = 2x − 7”) are line graphs. Here is another example, the so-called Keeling curve. This
has been called one of the most important geophysical records ever made [224].

Figure 18: The Keeling curve, which shows atmospheric CO2 concentrations in parts per million by volume (ppmv).

The Keeling curve summarizes measurements of the carbon dioxide concentration (see Example 5
on page 15) of Earth’s atmosphere, made continuously since 1958 at the Mauna Loa Observatory
in Hawai‘i under the direction of C.D. Keeling. The red curve shows the actual measurements,
which fluctuate according to seasonal changes in global vegetation; the black curve is an average
that smooths these seasonal variations out. This graphical presentation makes it vividly clear that
carbon dioxide levels in our atmosphere are rising quite rapidly, and it can be shown that this rise
corresponds very accurately with the increasing rate at which humans are releasing carbon dioxide by
burning fossil fuels. The problems that this may give rise to, and the possible ways these problems
might be addressed, will occupy our attention in several later sections of this book. For the moment, we
simply want you to observe how effective graphical presentation of this data is—much more effective
than a corresponding table.

apprecia@ucalgary.ca
1.4. COMMUNICATING QUANTITATIVE INFORMATION 51

Think about it. . .

Notice that the y-axis of the Keeling curve does not start at zero, but at about 320 parts per
million by volume—the level of atmospheric CO2 when Keeling began his measurements.
How would the appearance of the curve differ if it was plotted using axes where the y-axis
did begin at zero parts per million? Which possible graphic (the version in Figure 18, or
the replotted version with y-axis starting at zero) seems more informative to you? More
accurate? Read ahead to the next section for more discussion of these issues.

Remark 5: To learn more about the skillful use of graphical presentation, and about how such
presentations can mislead (the subject of our next section!) we recommend Edward Tufte’s classic
book [325].

1.4.4 How Presentations Can Mislead


Visual, graphical communication is incredibly powerful. A well-constructed graphic can allow the
human eye to take in, in an instant, the significance of a mountain of data that might be impossible
to comprehend in numerical form. Software companies and consultants thrive in the area of big data
visualization, creating graphical tools that allow companies and researchers to get a visual picture of
the huge amounts of information that their efforts have generated.
But just because graphics can be so powerful, they can also be powerfully misleading. In this section
we’ll review a couple of the more common ways in which graphics can mislead, whether by accident
or design. We want to help you recognize misleading graphics when you encounter them, and also
to help you avoid creating graphics that fail in their purpose of communication by misleading other
people. Here are some key problems to look out for.
Example 2: Questions about the choice of origin
Go back and look again at Figure 17, showing the change in
7,000
U.S. natural gas production over time. In that figure, we started the
y-axis at zero. What happens if we instead start the y-axis about at
the lowest production figure (as we saw was done with the Keeling 6,000
Production

curve; see Question 1.4.3)? We get the result in Figure 19.


Both Figures 17 and 19 show the exact same data, with the
same accuracy, and yet there is no question that the second figure 5,000
gives a much more dramatic impression of growth than the first
one does. Is this a misleading impression? At the very least, it has
the potential to mislead. The percentage increase in gas production 2004 2006 2008 2010 2012 2014 2016
over the period covered is a little over 50% . It seems that Figure 17 Year
represents this accurately, whereas Figure 19 suggests a much
Figure 19: Scatter plot of the data from Table 12 on
more dramatic increase. page 47 on U.S. natural gas production (billions of
Political news releases often give us much more egregious cubic feet per quarter).
example of the same kind of thing. Figures 20 and 21 on the next
page are fairly recent ones. In both cases, the changes involved are
made to look much more dramatic than they really are by choice of origin on the y-axis. The first
figure doesn’t even specify the origin, though by examining the other data we can deduce that it is
at about 70% . In the second figure, the origin is at 94 million. Sure, in both cases the numbers have
increased over time, but we can be pretty confident that those who created the graphics have chosen
the y-axis origin with intent to make the increase look as dramatic as possible.

apprecia@ucalgary.ca
52 CHAPTER 1. MEASURING

Figure 20: Obama administration (2015) graphic of high school graduation rates.

Figure 21: Senate Budget Committee (2012) graphic of welfare rolls.

104
How can we avoid this kind of problem? Some people propose a rule that the
y-axis should always start at zero. But this is too general. There are plenty of
Temperature (degrees F)

102 examples for which the zero point is not relevant, or even is arbitrary, whereas the
changes in the y-value are extremely significant.
100
Consider, for instance, a chart of someone’s body temperature over time, such
98 as might be recorded in a hospital. If my temperature chart over a couple of days
0 10 20 30 40 50
Hour looks like Figure 22, we can definitely say that I am very sick; I have a fever
Figure 22: Body temperature
that is spiking rapidly and has reached the level (over 104 ◦ F) that is considered
plot. extremely dangerous. But if we plotted the same data with a baseline of zero, this
important information would be obscured (Figure 23).
In this case, starting the y-axis at zero has led to a decrease in the effectiveness
of communication. Zero, in this case, does not reflect anything relevant to the
100
Temperature (degrees F)

80
question at hand. In fact, we have already seen that zero degrees Fahrenheit is
60
simply the coldest that Fahrenheit could achieve (see Remark 4 on page 13) using
40
the technology of the early eighteenth century in Europe.
20
A similar point, by the way, applies to the Keeling curve. We asked what the
0
0 10 20
Hour
30 40 50
significance might be of the fact that the Keeling curve is not based at zero. Here
again, the answer is that what is significant is not how the CO2 measurements
Figure 23: Body temperature
plot starting at zero.
are related to zero but how they are related to a “normal” level; we want to see
whether the planet is “running a fever” as measured by carbon dioxide levels that
are far above normal. The usual presentation of the Keeling curve takes the levels
when Keeling started his measurements as “normal.” (In fact, human-generated emissions had already
had a measurable effect by that point, but much smaller than the effect recorded by Keeling between
the 1960s and the present day.)

apprecia@ucalgary.ca
1.4. COMMUNICATING QUANTITATIVE INFORMATION 53

Figure 24: Prevalence of West Nile virus. Avoid the use of three-dimensional charts.

Even though starting the y-axis at zero should not be considered an absolute rule, it is a good policy
in cases in which the zero level is significant, as it is, for example, in Figures 20 and 21.

Example 3: Problems with three-dimensional plots


We said above that three-dimensional plots are particularly liable to be perceived wrongly, whether
by accident or by intent on the part of the plot designer. Let’s look at some examples.
As you can see, Figure 24 is built on a map of the continental United States and purports to track the
incidence of the West Nile virus in different states for one year. But what is it that actually represents
the number of West Nile cases? The height of the box—or its volume? The eye is naturally drawn to
the volume of the box, so that a “prism” of given height above Texas looks much more significant than
one of the same height above Pennsylvania, but it is not clear from the captions whether that is in fact
how the chart is to be read. (These kinds of confusion between height and volume are very common
when people read “3-dimensional” charts.) Perspective makes the areas of the more northern states
look smaller in any case—another opportunity for confusion. Finally, some “prisms” are partly hidden
behind other prisms. All in all, not a good representation of an interesting set of data. The author is
really trying to convey a relationship between two numerical quantities (distance from the Mississippi
and West Nile prevalence), so a clearly labeled scatter plot would be much more informative.

Critical Thinking

We have cited 3-dimensional plots as one example in which sophisticated data visualiza-
tions, made possible by technology, can obscure the actual information being presented—
whether accidentally or malevolently. Can you think of other examples in which new ways
to communicate information sometimes work to obscure it?

Perspective errors, which we mentioned above, are front and center in the next chart (Figure 25
on the next page). The two pie charts show the exact same data on U.S. energy annual consumption
by each source as a percentage of the total. But in the left-hand pie chart, the “pie” has been made
three-dimensional, then flattened out and displayed in perspective. The effect is to make the “pie
segments” nearest to the viewer, especially the red segment corresponding to natural gas, look bigger
relative to the others. Without the reported percentages, the viewer, who naturally perceives areas,
could be left with the misleading impression that natural gas accounts for the largest share of our

apprecia@ucalgary.ca
54 CHAPTER 1. MEASURING

Petroleum Natural Gas Coal Renewables Nuclear

8.6%
8.6%
36.9% 10.5% 36.9%
10.5%
14.6%
29.3%
14.6%

29.3%

Bad Good
Figure 25: U.S. annual energy consumption by fuel source as a percentage of total (2016). Avoid using 3-dimensional charts.

energy consumption. The right-hand chart, which is an ordinary two-dimensional pie chart, correctly
shows that the largest share is made up by petroleum, and by a sizable margin.

Example 4: “Cherry-picking” (showing only a subset of the data) You will sometimes encounter
graphical presentations that mislead by ignoring or excluding data that contradicts a desired message
or conclusion. When you present data demonstrating a trend or correlation it is essential to include all
relevant data, not just a carefully selected subset of the data that is most favorable to a predetermined
conclusion.
As an extreme example, consider the following graph showing average temperatures in Pennsylva-
nia over time.
90

80
Temperature (degrees F)

70

60

50

0 1 2 3 4 5
Time interval

Figure 26: Average temperatures in PA plotted against time.

Wow! So much for global warming, huh? What these figures demonstrate is a rapid and definitive
cooling trend. But no doubt you have guessed the trick: the “time intervals” on the x-axis are months,
and the plot runs from July to December. No great surprise to learn that it is colder in winter than
in summer, and no long-term conclusions about global warming can possibly be drawn. You might
say that the long-term trend, whatever it is, has been “contaminated” by the short-term cycle of the
seasons.
Most attempts to refute global warming by cherry-picking data are not as foolish as this (though
Senator Inhofe’s stunt presenting a snowball on the Senate floor, Figure 28 on the opposite page,

apprecia@ucalgary.ca
1.4. COMMUNICATING QUANTITATIVE INFORMATION 55

perhaps comes close). But more sophisticated cherry-picking is surprisingly common. In 2012
Britain’s Daily Mail newspaper published a front-page article entitled “Global warming stopped 16
years ago,” [275] and accompanied it by a chart similar to that shown in Figure 27.
Global Temperature Anomaly 1998–2014

0.8

Temp Anomaly ◦ C
0.6

0.4

0.2
98

00

02

04

08

10

12

14
06
19

20

20

20

20

20

20

20
20
year

Figure 27: Global monthly mean temp anomaly1997–2014, data from [119].

Certainly, this chart tells us that one data set showed a very small
change in surface temperatures over the period indicated. But as
it happens, there was some cherry-picking going on here, some
inadvertently. The starting point of the chart was an unusually hot
year (because of the El Niño phenomenon, a periodic weather
cycle in the Pacific Ocean). And it turns out, though of course
the report’s authors could not have known this at the time, that
the end point of the chart was unusually cool (the next El Niño
was about to get going, and the successive years 2013–2016 have
each been notably warmer than the previous one). We can put this
figure in context by looking at a larger data set. Figure 29 on the
next page plots the global mean monthly temperature anomaly
from the year 1880 through 2016. The small data set included
in Figure 27 is indicated by the yellow line. In this larger data
set, the so-called “global warming pause” is barely visible. We’ve
highlighted a longer “pause” from 1945 to 1980 in orange. We are Figure 28: Sen. J. Inhofe carries a snowball into the
Senate to “refute” global warming.
reporting the anomaly in average surface temperatures, but most
of the warming that Earth experiences is absorbed by the oceans.
The reasons for these apparent pauses have to do with cycles of ocean warming and cooling, such as
the El Niño phenomenon; it turns out that even a 35-year time period may not be long enough to avoid
“contamination” of the long-term trend by other, shorter-term variable phenomena.
Question 1: What do you mean by the word anomaly in the previous paragraph and on the y-axes of
Figures 27 and 29 on the next page?

Answer: The anomaly in some quantity is the amount by which it differs from its normal or expected
value. The global temperature anomaly therefore measures how much the whole planet is overheating
or “running a fever”.

apprecia@ucalgary.ca
56 CHAPTER 1. MEASURING

Global Temperature Anomaly 1880–2016


1.5

1
Temp Anomaly ◦ C
0.5

−0.5
00

20

40
0

80

00

20
60
8

19

19

19
18

19

20

20
19
year

Figure 29: Global monthly mean temp anomaly 1880–2016, data from NASA [119].

Critical Thinking

Cherry-picking data is very tempting, especially if you think that you “know” what the right
or virtuous answer has to be. To avoid it, try to cultivate the ethics of a scientist, whose
job is to tell the story “as it is” in the most accurate way possible; not to manipulate the
presentation of information so as to steer people to a predefined conclusion. This viewpoint
is sometimes called “skeptical” or “critical.” At the time of writing, the website Skeptical
Science [336] is a useful source of this kind of “skepticism” in matters related to global
warming. (But treat its claims skeptically, of course!)

Example 5: Fallacies of multiplication


Our final example of misleading communication is a purely verbal one, which comes about when
we put together a great many small quantities to obtain a large one. For example, you may have heard
the term energy vampires used to refer to appliances that continue to draw electrical power even when
they are not obviously being used. Microwave ovens, network routers, TVs, stereos, wireless phones,
and so on can all be energy vampires.13 Often the cell phone charger is singled out as a particularly
notorious example.
Problem 2: A public service announcement says, “Unplug your cell phone charger when it is not in
use! If everyone did this, we could save enough electricity to power half a million homes!” Assuming
that this is true, what percentage of your overall electricity use is represented by your cell phone
charger?

13 For definitive information, see the measurements carried out by the Lawrence Berkeley Laboratory at [190].

apprecia@ucalgary.ca
1.4. COMMUNICATING QUANTITATIVE INFORMATION 57

Solution: Half a million homes certainly sounds like a lot, but then there are a lot of homes in the
United States. Let’s unpack the numbers a bit. According to the Census Bureau there are something
like 130 million households in the USA. Half a million is about 0.4 percent of all U.S. households
500,000
(that is, 130,000,000 × 100%). So an equivalent way to express the claim made by the advertisement
is to say that cell phone chargers represent less than half of one percent of U.S. household electrical
consumption—or that (if you are “average” in this respect) your cell phone charger represents less
than half of one percent of your electrical energy consumption. It is good to save energy, but one half
of one percent is not a large saving.
In other words, we’ve obtained what sounds like an impressive total saving by multiplying a small
saving (from unplugging your cell phone charger) by a large number (the number of households in
the USA). But we have forgotten that the total energy consumption of each household must also be
multiplied by that same large number. Once we have done that, we realize that we have not made such
a big difference, relative to total energy consumption, as we might have thought.

Think about it. . .

David MacKay [202] writes about claims like the one about cell phone chargers:
“But surely, if everyone does a little, it’ll add up to a lot?” No. This if-everyone
multiplying machine is just a way of making something small sound big. The if-
everyone multiplying machine churns out inspirational statements of the form
“if everyone did X, then it would provide enough energy/water/gas to do Y,”
where Y sounds impressive. Is it surprising that Y sounds big? Of course not.
We got Y by multiplying X by a big number—the number of people involved.
If everyone does a little, we’ll achieve only a little. We must do a lot.
Do you agree with MacKay? Why or why not?

apprecia@ucalgary.ca
58 CHAPTER 1. MEASURING

Summary of Ideas: Communicating Quantitative Information

• Three ways of communicating quantitative information are verbal, tabular, and


graphical.
• Verbal communication is most effective when only one or two pieces of information
need to be shared. “Human terms” comparisons enhance its effectiveness.
• Tabular communication is effective when three or more data values are to be presented
and where it is important that the reader be accurately aware of the individual values.
• Principles for designing effective tables can be found in Rule 1 on page 48.
• Graphical communication is effective to demonstrate trends and comparisons in a
way that can be grasped visually.
• Bar charts and pie charts are used to share categorical data, that is, data that is divided
into different types or “categories.”
• Scatter plots and line graphs display the relation between two types of numerical data.
• Graphical (and other) means of communication can mislead. Avoid misplaced
origins, abuse of three-dimensional effects, cherry-picking data, and fallacies of
multiplication.
• All these means of communication are intended to enhance citizen understanding and
thus the effectiveness of democratic decision-making.

apprecia@ucalgary.ca
1.5. EXERCISES FOR CHAPTER 1 59

1.5 Exercises for Chapter 1

 8. Alaska Wildfires. In 2015, a below normal winter


1.1. UNITS AND MEASUREMENT snowpack and relatively dry summer contributed to an
above average summer wildfire season in Alaska in
 1. Identify the number part and the unit part in which 5.1 million acres burned. During the first week
the measurements below. Which of the measurements of August 2015, 246 active fires were increasing the
would you say are expressed in “human terms”? area burned by 86 thousand acres per day.
Convert the total acres burned to square kilometers
(a) The distance around the equator is about 24,000
miles. and convert the additional acres burned per day to
square kilometers per day. Put these into human terms
(b) The lower 48 states receive, in total, about 1,400
by comparing with other geographical areas.
cubic miles of rainfall per year.
(c) Each year, the lower 48 states receive enough
 9. In early May of 2016, the city of Fort McMurray,
rain to cover them to a depth of about 2 12 feet.
Alberta was evacuated because of an encroaching wild-
(d) Penn State (University Park) produced about fire. On the morning of May 5, the fire was described
15,000 tons of solid waste in 2012. as being 25% larger than the area of Manhattan Island,
(e) The average amount of solid waste produced at which has an area of 88 square kilometers. Estimate
University Park each day weighs as much as ten the size of the fire and compare this with a familiar ge-
elephants. ographical area (such as your campus, city, or county—
(f) My height is approximately 0.0011 nautical that is, if Manhattan Island isn’t already familiar!).
miles.
(g) My height is approximately 6 feet.  10. What is 4,000,000 millimeters in miles? Round
(h) Worldwide consumption of jet fuel is roughly your answer to the nearest whole number.
5,000,000 barrels each day.
 11. Which is larger, 3 cubic feet or 2 square meters?
(Does the question even make sense?)
 2. In what other ways could you express 3 million
centimeters in human terms?
 12. If we wish to convert cubic inches into cubic
centimeters, what fraction should we multiply by?
 3. The total rainfall in the State College, PA, area
averages 39 inches per year. If my house has a roof area
 13. Finn wishes to convert 300 ◦ F into Celsius. He
of 650 square feet, how much rainwater (in gallons) has
does the following calculation:
fallen on it since it was built, 18 years ago?
5
(300 ◦ F − 32 ◦ F) = 167 − 32 = 135 ◦ C.
9
 4. Use the unit-factor method to express 1 million
seconds in days. What, if anything, is wrong with Finn’s calculation?

 5. Make a table of definitions and examples for units  14. Express yesterday’s minimum and maximum
of area, like the one in Table 2 on page 8 for units of temperatures for your region in the Fahrenheit, Celsius,
length. Try to come up with your own examples of each and Kelvin scales
area.
 15. The daytime surface of Mercury reaches a tem-
 6. An acre is defined to be 4840 square yards. How perature of 700 kelvin. Express this temperature in both
many acres in a square mile? Celsius and Fahrenheit.

 7. About how many cubic yards are in 3 cubic  16. The American Dental Association recommends
meters? that you brush your teeth for two minutes twice a
day [302]. Calculate the amount of time you spend

apprecia@ucalgary.ca
60 CHAPTER 1. MEASURING

brushing your teeth as a percentage of your total time, under fixed conditions (constant length of pipe,
assuming you follow this recommendation. constant pressure difference) is proportional to
the fourth power of the pipe’s diameter. Using
this fact, figure out by what factor one should
 17. In 2017 the concentration of carbon dioxide in
change the diameter of a pipe in order to change
the atmosphere was around 405 parts per million by the flow rate by a factor of 4,000.
volume (ppmv). Describe what this means in your own
(b) A standard gas pump hose has an internal diam-
words, and express 405 ppmv as a percentage. eter of about 34 inch. Using the result of (a), find
the diameter of a hose of the same length that
 18. In Exercise 4 on the previous page you used the would deliver gas (and, therefore, energy) 4,000
unit-factor method to show that one million seconds times more slowly.
is about 11.5 days. Thus, we can think of 1 part per
Your answer should be approximately the diameter of
million (1 ppm) as corresponding to 1 second out
a drinking straw.
of 11.5 days. Use this idea to express each of the
quantities listed below as 1 part per million (1 ppm)
and 1 part per billion (1 ppb) of quantities that are in  23. An electrician might be able to install a specialist
familiar terms. electrical outlet in your home that would deliver power
at the rate of 20 kilowatts: see the picture below. Redo
(a) 1 second,
the calculations of Exercise 22 using this outlet instead.
(b) 1 minute, How much difference does it make?
(c) 1 inch,
(d) 1 teaspoon (tsp) (1 gallon is 768 teaspoons).

 19. The diameter of the Earth is about 8,000 miles. In


a scale model, the Earth is represented by a sphere the
size of a basketball, diameter 9 12 inches. How far away Figure 30: Five-pronged industrial AC plug.
is the Moon in the scale model? (The actual distance
from Earth to the Moon is about a quarter of a million
miles.) What size of sphere represents the Moon? (The 1.2. SCIENTIFIC NOTATION
actual diameter of the Moon is about 2,200 miles.)
 24. Convert the following to scientific notation in
Does this correspond to the ball used in any game that
standard form:
you know?
(a) 75,000, (e) 0.0000002319,
 20. The International Space Station (ISS) orbits (b) 2,100,000,000, (f) 23.189,
about 250 miles above Earth. Since the year 2000, no (c) 0.007, (g) 0.1056,
humans have traveled much farther from Earth than the (d) 17, (h) 0.0156.
ISS. Where would the ISS be situated in the model
described in the previous exercise?
 25. Convert the following to decimal notation:
 21. In 2015, Penn State’s University Park campus (a) 2 × 1010 , (d) 885 × 1011 ,
diverted 9000 tons of waste from landfills through (b) 3.1 × 10−9 , (e) 4.783 × 10−1 ,
recycling and composting programs. Convert this to
(c) 6.66 × 102 , (f) 1.01 × 10−2 .
tons per day and express the result in human terms,
using a unit that is easy to visualize.
 26. Let a = 1.27 × 10−6 and b = 8.9 × 10−7 . Calcu-
 22. Look back at Problem 17 on page 20. In its late the following and express your answers in scien-
solution, we calculated that the ratio of the energy tific notation:
delivery rates of the gas pump and the electrical outlet (a) a + b, (c) a × b,
was about 4,000.
(b) a − b, (d) a ÷ b.
(a) It is a law of physics (Poiseuille’s law) that the
rate of flow of a fluid through a circular pipe

apprecia@ucalgary.ca
1.5. EXERCISES FOR CHAPTER 1 61

 27. Let x = 7.55 × 1018 and y = 2.1 × 1019 . Calculate make tea each year? Use scientific notation to express
the following and express your answers in scientific your answer.
notation:

(a) x + y, (c) x × y,  33. The Moon orbits the Earth once every 27.3 days.
How many orbits has the Moon made during your
(b) x − y, (d) x ÷ y.
lifetime? Suppose that the orbit of the Moon is circular,
with radius r = 250,000 miles. Then, during each
orbit, the Moon travels 2πr miles (the circumference
 28. Assume that there are 250 million cars on the of the circle) in its journey around Earth. How far has
road in the U.S., that on average each car gets 25 miles the Moon traveled in this way during your lifetime?
to the gallon and drives 10,000 miles per year, and that Express your answer using scientific notation.
each gallon of gas burned generates 20 pounds of CO2
(carbon dioxide). How many tons of carbon dioxide are
emitted by all these vehicles over the course of a year? 1.3. ESTIMATES, PRECISION, AND
(Use scientific notation.) ORDERS OF MAGNITUDE

 34. Without consulting any references, estimate the


 29. In 2015, in a scandal affecting half a million
distance (in a straight line) from your present location
vehicles in the United States, Volkswagen was caught
to Des Moines, IA. What do you believe is an appro-
using software to cheat on emissions testing for diesel
priate level of precision for your estimate? Now use an
vehicles. VW diesels were found to emit about 1 gram
online mapping tool to check your answer. Was your
of nitrous oxide per kilometer driven (about 40 times
estimate correct to within the precision you specified?
the legal limit). In a vehicle with the software, the
emission control systems were fully activated only
when the vehicle was being tested. If each car is driven  35. A human being needs approximately 12 kilo-
20,000 kilometers per year (on average), estimate the calories of food energy per pound of bodyweight in
mass of the nitrous oxide pollution that is emitted order to sustain life for a day. The planetwide average
in one year from these 500,000 “dirty diesels.” (Use bodyweight of a human is about 140 pounds (Table 6
scientific notation). on page 489). The total human population of the planet
can be estimated as 7.5 × 109 people. Estimate the
number of kilocalories of food energy required each
 30. A U.S. one dollar bill is 0.0043 inches in thick-
day by the entire human population, being sure to
ness. Convert this to meters, and express the result in
express your answer with appropriate precision.
scientific notation.

 36. The meltdown at the Fukushima Daiichi nuclear


 31. The salary of a member of Congress is $174,000 power plant, triggered by the Tohoku earthquake and
per year, a figure that (at the time of writing) has tsunami of March 11, 2011, was one of the most severe
been constant since 2009. Suppose that a member of nuclear accidents of modern times. In Iitate, Fukushima
Congress was paid at this rate from January 1, 2009, prefecture, outdoor radiation levels of approximately
until the day that you work this problem, and that the 30 microsieverts per hour were measured immediately
total salary payment was converted into dollar bills. following the accident. Suppose that (contrary to gov-
Using the result of the previous exercise, estimate the ernment advice) a resident of Iitate stood in the open
air continuously for 3 days following the accident.
thickness (in meters) of the resulting pile of dollar
Estimate the total amount of radiation this person
bills.
would receive.
A CT (computed tomography) scan of the abdomen
 32. The current population of Britain is about 65 and pelvis can deliver a dose of approximately 20 mil-
million people. Assume that on average, each Briton lisieverts of radiation (1 millisievert = 1,000 microsiev-
drinks 3.5 cups of tea per day, each containing 0.2 erts). Compare the order of magnitude of radiation
kilograms of water that must be heated from supply received by the patient undergoing the CT scan and the
temperature (about 15 ◦ C) to boiling point (100 ◦ C). It hypothetical resident of Iitate whose radiation dose you
takes about 4200 joules of energy to heat 1 kilogram of calculated above.
water through 1 ◦ C. How many joules do Britons use to

apprecia@ucalgary.ca
62 CHAPTER 1. MEASURING

 37. “In an attempt to run a marathon in under 2 hours, is correct, how much did your fingernails grow during
Eliud Kipchoge ran a 26.2 mile marathon course in the time it took to complete this problem?
2 hours and 25 seconds, corresponding to an average
speed of 13.0547001903 miles per hour.” 3.56 × 106
What, if anything, is wrong with the preceding  42. Compute . Express the result in stan-
4.64 × 1011
statement? dard form with the appropriate number of significant
figures.
 38. An aging underground natural gas storage facility
in Porter Ranch, near Los Angeles, leaked 1.1 × 105  43. Particulate matter is a form of air pollution con-
pounds of methane per hour between October 2015 sisting of small particles that are harmful when inhaled.
and February 2016, when the leak was finally plugged. PM2.5 is the abbreviation for fine particulate matter,
Convert this leakage rate to cubic feet of methane per smaller than 2.5 microns in size. Use scientific notation
year. You may use the fact that 1 cubic foot of methane to express 2.5 microns in meters to 2 significant figures.
weighs approximately 4.5 × 10−2 pounds at standard Follow-up questions: As of February 2017, there are 20
conditions for pressure and temperature. counties in the U.S. that have failed to meet EPA air
quality standards for PM2.5 pollution. These are called
 39. Approximate the following quantities to the num- nonattainment areas. Do you live or go to school in
ber of significant figures specified. Write your answer a nonattainment area? What are the health effects of
in scientific notation, standard form: PM2.5 exposure?
(a) 467.967 to 2 significant figures;
 44. Look in the “Useful Numbers” section (Sec-
(b) 299,792,458 to 3 significant figures. (This is the
tion 8.2) to find the area of the Earth’s oceans, and the
speed of light in meters per second.)
volume of the Greenland ice cap.
(c) 0.0000000059055118 to 2 significant figures. If the entire Greenland ice cap were to melt into the
(d) 15,157,486,080 to 1 significant figure. (This ocean, estimate the resulting sea level rise.
number was found online as the result of some-
one’s calculation of the distance, in inches, from  45. In 2016, the United State generated 35.5 terawatt-
the Earth to the Moon. It is a good example hours (TWh) of electrical energy from large-scale solar
of an excessively precise measurement claim.
power plants. Use scientific notation to express this
According to NASA, the distance from the Earth
in kilowatt-hours (the standard unit used in measuring
to the Moon varies by almost 10 percent over the
course of a month, because the Moon’s orbit is residential electrical energy consumption). 1 kilowatt-
not a perfect circle. So most of those impressive- hour is 1,000 watt-hours. 1 terawatt-hour is 1012 watt-
looking figures are. . . not significant.) hours. Express your answer to 2 significant figures.

 46. Use scientific notation to express the following


 40. Approximate the following quantities to the num- quantities in the specified units with the specified
ber of significant figures specified (you may need to number of significant figures:
search for the information online). Write your answer
in scientific notation, standard form: (a) Width of a human hair (100 microns), in inches,
to 1 significant figure.
(a) The current population of the United States to 2
(b) Diameter of tobacco smoke particles (0.01 mi-
significant figures.
crons), in meters, to 1 significant figure.
(b) The global population to two significant figures.
(c) Average length of an E. coli bacterium (0.002
(c) The U.S. federal debt to 2 significant figures. millimeters), in meters, to 1 significant figure.
(d) Last year’s U.S. federal budget to 3 significant
figures.
 47. Is the average length of an E. coli bacterium of
the same order of magnitude as the width of a human
 41. It is estimated that your fingernails grow at a rate hair? (See Exercise 46)
of 1.2 nanometers per second. Express this quantity
in units of meters per second, using two significant  48. A car is lowered (carefully) into an empty
figures and scientific notation. Assuming this estimate backyard swimming pool. If the pool is then filled with

apprecia@ucalgary.ca
1.5. EXERCISES FOR CHAPTER 1 63

all the gasoline that the car will burn over its working  55. Give an order-of-magnitude estimate of the num-
lifetime, will the gasoline cover the car? (If you try this ber of Rubik’s Cubes that could fit in an Olympic size
at home, be sure not to strike a match.) swimming pool.

 49. Penn State’s Beaver Stadium has a capacity of  56. The ambient level of atmospheric mercury is 2
approximately 106,000 spectators. Estimate the num- nanograms per cubic meter. Estimate the amount of
ber of Porta-Potties needed to cater to all the tailgaters mercury inside your classroom (assuming the ambient
on a day when Beaver Stadium is filled to capacity. level of atmospheric mercury).

 57. Estimate the number of unused AA batteries in


 50. Penn State’s University Park campus serves
all homes in the United States right now.
roughly 50,000 students, faculty, and staff. Estimate the
number of plastic water bottles that are discarded on the
University Park campus in one year. (You can look this  58. Give an order-of-magnitude estimate for the total
figure up, but the point of this problem is to make a amount of water used to flush toilets in the United
reasonable estimate based on your own experience and States during one day.
things that you already know.)
 59. Your local news station is on the scene of a crash
on I-95 involving a tractor trailer hauling unminted
 51. Total U.S. federal expenditures for fiscal year pennies to the U.S. Mint in Philadelphia. The news
2014 were 3.50 trillion dollars (to 3 significant figures). desk urgently needs an estimate for the total number
Suppose that for fiscal year 2015, federal expenditures of pennies scattered over the northbound lanes of I-95.
increased by 396 billion dollars over the 2014 baseline. The legal weight limit for a tractor trailer on I-95 is
Using the appropriate precision, give your best estimate 40 tons (you can assume the truck was at this limit);
for 2015 total federal expenditures. one penny weighs 2.5 grams.
(This describes a real event that occurred on Septem-
ber 8, 2016. A newscaster reported that there were a
 52. The total area of the U.S. is 3.8 × 106 square “gajillion pennies” on I-95.)
miles, and its population (at the time of writing) is
3.19 × 108 people. Calculate the average number of
 60. Estimate the total amount of blood (in gallons)
people per square mile, expressing your answer, with
donated in the U.S. each year. After making your
appropriate precision, in ordinary (not scientific) nota-
estimate, research statistics on blood donation. How
tion.
close is your estimate? What assumptions did you
make about blood donations and how did these differ
 53. Suppose that the average home in Centre County, from what you found in your research?
PA, has a roof area (viewed from above) of 1600 square
feet. Centre County receives about 40 inches of rain
or equivalent precipitation (snow in the winter) per 1.4. COMMUNICATING QUANTITATIVE
year. Estimate the total amount of precipitation, in INFORMATION
gallons, that falls on a Centre County roof during a
year. According to the EPA Partnership Program Water  61. In this section we have talked about verbal, tab-
Service, the average American family of four uses ular, and graphical ways to communicate quantitative
1.5×105 gallons of water per year in their home. Is this information. Do you think these are the only possibil-
figure of the same order of magnitude as the volume of ities? Try to name some other ways of communicating
rain you calculated above? this kind of information, especially some that modern
digital technology has made possible.
 54. Estimate the height (in miles) of a stack of $100
bills equivalent to the net worth of the wealthiest person  62. It is estimated that U.S. consumers used, and
on Earth. The thickness of a $100 bill is approximately threw away, 16 billion disposable paper coffee cups
4 × 10−3 inches. in 2006, and that this number had increased to 23
billion by 2010. Each cup weighs approximately 0.7

apprecia@ucalgary.ca
64 CHAPTER 1. MEASURING

ounces. What do you think is the most effective way to temperature. How does this relate to the discussions in
communicate this information? Prepare a chart, table, Section 1.4.4 and Example 2 on page 51?
or verbal presentation that you believe will bring home
to the viewer the amount of waste involved here.  67. Pick up a copy (or visit the website) of your
favorite major newspaper and search for a story that
 63. It is estimated that the preindustrial level of CO2 , includes quantitative information (if visiting a website,
that is, the amount before the Industrial Revolution you might try a topic-based search on “Environment”
started the large-scale consumption of fossil fuels, may or “Energy” or “Science”). Discuss how the quan-
have been about 270 parts per million by volume titative information is presented. Is the information
(ppmv). Replot the Keeling curve on a scale that begins presented clearly? Does it connect with and support the
with 270 ppmv on the y-axis. Does the result convey a overall story?
significantly different message?
 68. In 2016, the United States produced 315 million
 64. The following table (see [79]) shows the annual tons of bituminous coal, 338 million tons of sub-
sales of hybrid (gasoline-electric) vehicles in the U.S. bituminous coal, 73 million tons of lignite coal, and
for years from 2000 to 2013.
1.6 million tons of anthracite coal. Present this data as
Year Sales a table and in graphical form.
2000 9,350
2001 20,282  69. Think of at least three questions related to
2002 36,035 Exercise 68 that would require additional (or different)
2003 47,600 data, and then try to answer them. For example, you
2004 84,199
might ask, “How does the coal produced in the U.S. get
2005 209,711
used?” The U.S. Energy Information Administration is
2006 252,636
2007 352,271 an excellent source of data for answering your ques-
2008 312,386 tions. If you look for data on coal at the EIA website,
2009 290,271 take note of the various ways they present quantitative
2010 274,210 information.
2011 268,752
2012 434,498
 70. The EPA estimates that 258 million tons of
2013 495,771
municipal solid waste were generated in 2014 [15]
Represent this information in graphical form, using and 34.8% of this waste was recycled or composted.
what you feel is the most appropriate kind of chart Figure 10 on page 426 shows the makeup of mu-
or graph. What do you notice? Can you suggest any nicipal solid waste that was recovered (recycled or
explanation for this? composted), and Figure 15 on page 431 shows the
makeup of waste that was discarded (landfilled). Use
the information in these figures to discuss the potential
 65. In a poll of voters in Florida carried out some
for increasing the recovery rate of municipal solid
time before the 2016 election, 45 percent supported
waste.
the Democratic candidate, 40 percent supported the Re-
publican candidate, 5 percent supported the Libertarian
candidate, 2 percent supported the Green Party can-  71. According to the EPA, 34.8% of the municipal
didate, and the remainder supported other candidates solid waste generated in 2014 was recovered (com-
or were undecided. Represent this information in a bar posted or recycled). In 1960, the recovery rate was
chart and in a pie chart. Which representation is better under 6%. Assuming you could access similar data for
at communicating the information here? all years in the time-period 1960–2014, what way(s) do
you think would be most effective for conveying how
 66. If you search for information on the Earth’s global this rate has changed over time?
surface temperature, you may notice that scientists
often present the temperature anomaly (the departure  72. The EPA presents information on composting and
from the long-term average) in place of the actual recycling in both percentage terms and in terms of total

apprecia@ucalgary.ca
1.5. EXERCISES FOR CHAPTER 1 65

weight. Discuss situations in which each method of  75. On a webpage about water use [315], the U.S.
reporting might be more effective than the other. Geological Survey presents the following table on wa-
ter use of the top states as a percentage of the national
total. Discuss the usefulness of presenting information
 73. Average household water use for indoor fixtures
about water use in this way.
by category is shown in the table below (as reported in
the Water Research Foundation’s 2016 Residential End State Percentage of Total
Use Study [93]). Present this data in a graphical form. Domestic Use
California 15%
Fixture Gallons per day Texas 8%
Bath 3.6 Florida 6%
Clothes Washer 22.7 New York 6%
Dishwasher 1.6 Illinois 4%
Faucets 26.3
Leaks 17.0  76. During President Obama’s first term, gasoline
Other 5.3
prices increased from a national average of $1.83 per
Shower 28.1
gallon in January 2009 to a national average of $3.96
Toilet 33.1
in May 2011, an increase of over 115%. This statement
is factually correct but misleading. Why might this be
so? Historical data on gasoline prices is available from
 74. The study on residential water use referenced in
the U.S. Energy Information Administration.
Exercise 73 updates a previous study done in 1999.
Discuss how you might present information comparing
household water use in 1999 with household water use  77. Refer back to Figure 29 on page 56. Can you
in 2016. You can then refer to the study to see how this identify other intervals of time where, if viewed in
was done. isolation, it would appear that global temperatures are
not changing over time?

apprecia@ucalgary.ca

You might also like