Anda di halaman 1dari 409

1

The Cosmic Universe


Collection edited by: Steven Mellema
Content authors: Steven Mellema, OpenStax University Physics, OpenStax, and OpenStax Astronomy
Online: <https://legacy.cnx.org/content/col12169/1.46>
This selection and arrangement of content as a collection is copyrighted by Steven Mellema.
Creative Commons Attribution License 4.0 http://creativecommons.org/licenses/by/4.0/
Collection structure revised: 2017/06/20
PDF Generated: 2017/06/21 03:40:27
For copyright and attribution information for the modules contained in this collection, see the "Attributions"
section at the end of the collection.
2

This OpenStax book is available for free at https://legacy.cnx.org/content/col12169/1.46


Table of Contents
Preface . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1
Preliminaries
Chapter 1: Introducing Physics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5
1.1 Introducing Astrophysics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8
1.2 The Scope and Scale of Physics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11
1.3 Units and Standards . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 18
1.4 Unit Conversion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 23
1.5 Dimensional Analysis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 25
1.6 Estimates and Fermi Calculations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 29
1.7 Significant Figures . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 31
1.8 Solving Problems in Physics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 36
Chapter 2: The Universe at Its Limits . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 47
2.1 Consequences of Light Travel Time . . . . . . . . . . . . . . . . . . . . . . . . . . . . 49
2.2 A Tour of the Universe . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 50
2.3 The Universe on the Large Scale . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 55
2.4 The Universe of the Very Small . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 59
2.5 A Conclusion and a Beginning . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 60
Unit 1: Kinematics
Chapter 3: Straight-Line Motion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 63
3.1 Position, Displacement and Average Velocity . . . . . . . . . . . . . . . . . . . . . . . 64
3.2 Instantaneous Velocity and Speed . . . . . . . . . . . . . . . . . . . . . . . . . . . . 68
3.3 Average and Instantaneous Acceleration . . . . . . . . . . . . . . . . . . . . . . . . . 72
3.4 Motion with Constant Acceleration . . . . . . . . . . . . . . . . . . . . . . . . . . . . 78
3.5 Free Fall . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 90
Chapter 4: Circular Motion as One-Dimensional Motion . . . . . . . . . . . . . . . . . . . . 107
4.1 Rotational Variables . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 108
4.2 Rotation with Constant Angular Acceleration . . . . . . . . . . . . . . . . . . . . . . . 114
4.3 Relating Angular and Translational Quantities . . . . . . . . . . . . . . . . . . . . . . 118
Chapter 5: Relativistic Kinematics in One Dimension . . . . . . . . . . . . . . . . . . . . . . 125
5.1 Invariance of Physical Laws . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 127
5.2 Relativity of Simultaneity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 129
5.3 Time Dilation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 132
5.4 Length Contraction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 142
5.5 The Lorentz Transformation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 147
5.6 Relativistic Velocity Transformation . . . . . . . . . . . . . . . . . . . . . . . . . . . . 155
Chapter 6: Introduction to Vectors . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 169
6.1 Scalars and Vectors . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 170
6.2 Coordinate Systems and Components of a Vector . . . . . . . . . . . . . . . . . . . . 183
6.3 Algebra of Vectors . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 194
Chapter 7: Kinematics in Two and Three Dimensions . . . . . . . . . . . . . . . . . . . . . 213
7.1 Displacement and Velocity Vectors . . . . . . . . . . . . . . . . . . . . . . . . . . . . 214
7.2 Acceleration Vector . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 222
7.3 Projectile Motion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 227
7.4 Uniform Circular Motion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 238
7.5 Relative Motion in One and Two Dimensions . . . . . . . . . . . . . . . . . . . . . . . 245
Dynamics and Solar System I
Chapter 8: Overview of the Solar System . . . . . . . . . . . . . . . . . . . . . . . . . . . . 263
8.1 Overview of Our Planetary System . . . . . . . . . . . . . . . . . . . . . . . . . . . . 264
8.2 Composition and Structure of Planets . . . . . . . . . . . . . . . . . . . . . . . . . . . 273
8.3 Origin of the Solar System . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 277
8.4 Kepler's Laws of Planetary Motion . . . . . . . . . . . . . . . . . . . . . . . . . . . . 279
Chapter 9: Newton's Synthesis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 293
Chapter 10: Newton's Laws for Rotations . . . . . . . . . . . . . . . . . . . . . . . . . . . . 295
Chapter 11: Work and Energy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 297
Chapter 12: Momentum in One Dimension . . . . . . . . . . . . . . . . . . . . . . . . . . . . 299
Chapter 13: Angular Momentum . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 301
Geometric Optics
Chapter 14: Reflection and Refraction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 303
Chapter 15: Dispersion, Luminosity and Apparent Brightness . . . . . . . . . . . . . . . . . 305
Chapter 16: Image Formation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 307
Physical Optics
Chapter 17: Wave Properties of Light . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 309
Chapter 18: Interference and Diffraction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 311
Chapter 19: Spectroscopy and the Doppler Effect . . . . . . . . . . . . . . . . . . . . . . . 313
Chapter 20: Quantum Optics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 315
Thermodynamics and Solar System II
Chapter 21: Temperature . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 317
Chapter 22: The First Law of Thermodynamics . . . . . . . . . . . . . . . . . . . . . . . . . 319
Chapter 23: Kinetic Theory . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 321
Chapter 24: The Nebular Theory of Solar System Formation . . . . . . . . . . . . . . . . . . 323
Chapter 25: Nuclear Physics and Comparative Geology . . . . . . . . . . . . . . . . . . . . 325
Chapter 26: Comparative Atmospheres . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 327
Chapter 27: Exoplanets . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 329
Chapter 28: The Sun . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 331
Stars, Galaxies and Cosmology
Chapter 29: Stellar Properties . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 333
Chapter 30: Stellar Life Cycles . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 335
Chapter 31: The Milky Way as a Prototype Galaxy . . . . . . . . . . . . . . . . . . . . . . . 337
Chapter 32: Survey of Galaxies . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 339
Chapter 33: Galactic Evolution . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 341
Chapter 34: Big Bang Cosmology . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 343
Chapter 35: The Fate of the Universe . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 345
Appendix A: Units . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 347
Appendix B: Unit Conversions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 351
Appendix C: Fundamental Physics Constants . . . . . . . . . . . . . . . . . . . . . . . . . . . . 355
Appendix D: Astronomical Data . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 357
Appendix E: Useful Mathematical Formulas . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 367
Appendix F: Periodic Table of the Elements . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 371
Appendix G: The Greek Alphabet . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 373
Index . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 395

This OpenStax book is available for free at https://legacy.cnx.org/content/col12169/1.46


Preface 1

PREFACE
Welcome to The Cosmic Universe. This textbook was written in collaboration with the OpenStax project, whose purpose
is to increase student access to high-quality learning materials, while maintaining the highest standards of academic rigor
at little to no cost. It was created by re-organizing and editing some material from the OpenStax textbooks University
Physics (http://cnx.org/contents/d50f6e32-0fda-46ef-a362-9bd36ca7c97d@5.22) and Astronomy
(http://cnx.org/contents/2e737be8-ea65-48c3-aa0a-9f35b4c6a966@13.2) , and by the addition of new material.

About OpenStax
OpenStax is a nonprofit based at Rice University, and it’s our mission to improve student access to education. Our first
openly licensed college textbook was published in 2012 and our library has since scaled to over 20 books used by hundreds
of thousands of students across the globe. Our adaptive learning technology, designed to improve learning outcomes through
personalized educational paths, is currently being piloted for K–12 and college. The OpenStax mission is made possible
through the generous support of philanthropic foundations. Through these partnerships and with the help of additional
low-cost resources from our OpenStax partners, OpenStax is breaking down the most common barriers to learning and
empowering students and instructors to succeed.

About The Cosmic Universe


The Cosmic Universe is based upon the first semester of the four-semester, calculus-based, introductory physics-course
sequence at Gustavus Adolphus College. The text has been developed to meet the scope and sequence of that course. The
entire four-semester sequence at Gustavus, like virtually all university physics courses, provides a foundation for a career
in mathematics, science, or engineering. This book provides a unique combination and ordering of material centered on a
theme of astrophysics. It is distinctive in that it:
intersperses material from both "classical" and "modern" physics;
treats the use of calculus in ways that are largely conceptual; and
serves as a one-semester introduction to astrophysics suitable for physics, astronomy or pre-engineering majors.
Motivation for The Cosmic Universe
Over the years at Gustavus, the successful retention of good students in the physics major has always been a concern.
Certainly, a physics major is one of the most challenging paths through college, and it is definitely not for everyone. As
professors of the liberal arts, we are always happy if a student finds another major that is truly their passion. We also
understand that not every college student has the mathematical aptitude for our subject. And, to be honest, we also accept
the fact that some will leave physics because they do not wish to put in the level of effort required to succeed in such a
rigorous major. Our concerns over the years have focused on a group of hard-working, interested students who seem to
leave the physics major sometime in the first year. Although these particular students have demonstrated both the ability
and the work ethic to succeed in our program, and even though they have yet to find another major interest, they seem to
leave for one of three reasons:
1. They do not find the material in the first-year courses to be particularly interesting or relevant in the world of the
21st century.
2. They are bored by the early repetition of material that they recently covered in high school (especially classical
mechanics).
3. They feel overwhelmed by the immediate use of calculus, especially because they are taking their first semester of
college calculus, while some of their peers in the physics class have had more advanced high-school mathematics
preparation.

Scope, Coverage and Organization


Context for The Cosmic Universe
The four-semester, introductory course sequence at Gustavus Adolphus College consists of:
1. The Cosmic Universe
2. The Mechanical Universe
2 Preface

3. The Electromagnetic Universe


4. The Quantum Universe
The intent of the sequence is to cover 100% of the topical material normally taught in any undergraduate physics program
(and found in any good undergraduate textbook). However, we have re-ordered the topics by using a sequence of themes
(obvious from the course names). In that way, the "story lines" for each course are coherent and facilitate better student
understanding of the underlying physics concepts.

By incorporating the history of ideas, the thematic approaches lead directly to an understanding of the scientific method
- not as some dry set of steps, but as an actual, evolving human experience. As teachers at a liberal arts college, we feel
that every physics major should also understand how history, science, politics, religion, and ethics interact as parts of that
experience.

And, we also feel that it is important for physics majors to encounter, early in their college careers, the major unanswered
questions in physics, many of which are part of the study of astrophysics and cosmology - e.g. dark matter and dark energy.
The Cosmic Universe textbook contains material from various subfields of physics, from classical mechanics to optics to
relativity to quantum mechanics. The choice of topics and of the astrophysics theme were made to provide a first-semester
experience that is:
Interesting, because it deals with a very active field of current study in physics
New to virtually all of the students, because it involves topics not taught in most high-school physics courses
Rigorously mathematical in its approach, at the level of algebra and introductory calculus, more so than a
traditional college textbook in introductory astronomy
Not crucially dependent upon previous fluency in the use of calculus
Presented using a coherent story line
Calculus in The Cosmic Universe
The fact that The Cosmic Universe course is taken mostly by students in their first semester of college strongly influences
our use of calculus in this book. Many introductory calculus-based physics texts will, in an early chapter, derive the
equations of one-dimensional kinematics using integrals. For our student audience, where up to 50% are simultaneously
enrolled in their first semester of college calculus, such an approach can be discouraging and therefore counterproductive.

Knowing that they are studying (first) limits and (next) derivatives and then (perhaps by the end of the semester) integration
influences our use of calculus. We attempt to include calculus, conceptually at first, and hope that its physical significance
(of the derivative in particular) and practical applications can enhance the students' understanding of both the physics and
the math. As we are fond of asking our students, "Why did Newton invent the calculus in the first place?"
The book is organized as follows:
Introduction
Chapter 1: Introducing Physics
Chapter 2: The Universe at Its Limits
Chapter 3: Straight-Line Motion
Chapter 4: Circular Motion as One-Dimensional Motion
Chapter 5: Relativistic Kinematics in One Dimension
Chapter 6: Introduction to Vectors
Chapter 7: Kinematics in Two Dimensions
Chapter 8: Overview of the Solar System and Kepler's Laws
Chapter 9: The Newtonian Synthesis
Chapter 10: Extension of Newton's Laws to Rotations
Chapter 11: Work and Energy
Chapter 12: Momentum in One Dimension
Chapter 13: Angular Momentum
Chapter 14: Reflection and Refraction
Chapter 15: Dispersion, Total Internal Reflection, Luminosity and Apparent Brightness
Chapter 16: Image Formation

This OpenStax book is available for free at https://legacy.cnx.org/content/col12169/1.46


Preface 3

Chapter 17: Wave Properties of Light


Chapter 18: Interference and Diffraction
Chapter 19: Spectroscopy and the Doppler Effect
Chapter 20: Quantum Optics
Chapter 21: Temperature
Chapter 22: First Law of Thermodynamics
Chapter 23: Kinetic Theory
Chapter 24: The Nebular Theory of Solar System Formation
Chapter 25: Nuclear Physics and Comparative Geology
Chapter 26: Comparative Atmospheres
Chapter 27: Exoplanets
Chapter 28: The Sun
Chapter 29: Stellar Properties
Chapter 30: Stellar Life Cycles
Chapter 31: The Milky Way as a Prototype Galaxy
Chapter 32: Survey of Galaxies
Chapter 33: Galactic Evolution
Chapter 34: Big Bang Cosmology
Chapter 35: The Fate of the Universe
Assessments That Reinforce Key Concepts
Many of the assessments that were built into the OpenStax books University Physics and Astronomy have been retained.
In-chapter Examples generally follow a three-part format of Strategy, Solution, and Significance to emphasize how to
approach a problem, how to work with the equations, and how to check and generalize the result. Examples are often
followed by Check Your Understanding questions and answers to help reinforce for students the important ideas of the
examples. Problem-Solving Strategies in each chapter break down methods of approaching various types of problems into
steps students can follow for guidance. The book also includes exercises at the end of each chapter so students can practice
what they’ve learned.
Conceptual questions do not require calculation but test student learning of the key concepts.
Problems categorized by section test student problem-solving skills and the ability to apply ideas to practical
situations.
Additional Problems apply knowledge across the chapter, forcing students to identify what concepts and equations
are appropriate for solving given problems. Randomly located throughout the problems are Unreasonable Results
exercises that ask students to evaluate the answer to a problem and explain why it is not reasonable and what
assumptions made might not be correct.
Challenge Problems extend text ideas to interesting but difficult situations.
For Further Exploration. This section offers a list of suggested articles, websites, and videos so students can delve
into topics of interest, whether for their own learning, for homework, extra credit, or papers.
Answers for selected exercises are available in an Answer Key at the end of the book.

About the Authors


The Cosmic Universe Authors
Steven H. Mellema, Gustavus Adolphus College
Dr. Steve Mellema has taught introductory and advanced physics for over 30 years at Gustavus Adolphus College, where
he is currently Professor of Physics. He began his teaching career as a Peace Corps volunteer in Malaysia, where he taught
both chemistry and physics in secondary schools, worked on curriculum development, and conducted technical training for
science teachers. Dr. Mellema has a PhD in Physics from Ohio University, and he was a post-doctoral research associate in
nuclear physics at both Ohio University and the University of Wisconsin - Madison, before joining the Gustavus faculty. Dr.
Mellema has experience with research in Physics Education, including two Fulbright Professorships in Malaysia, the latest
of which involved bringing active-learning methods into university physics classrooms there.
Charles F. Niederriter, Gustavus Adolphus College
Dr. Chuck Niederriter has taught introductory and advanced physics for over 30 years at Gustavus Adolphus College, where
4 Preface

he is currently Professor of Physics.


Senior Contributing Authors - Open Stax University Physics
Samuel J. Ling, Truman State University
Jeff Sanny, Loyola Marymount University
Bill Moebs, PhD
Contributing Authors - OpenStax University Physics
David Anderson, Albion College
Daniel Bowman, Ferrum College
Dedra Demaree, Georgetown University
Gerald Friedman, Santa Fe Community College
Lev Gasparov, University of North Florida
Edw. S. Ginsberg, University of Massachusetts
Alice Kolakowska, University of Memphis
Lee LaRue, Paris Junior College
Mark Lattery, University of Wisconsin
Richard Ludlow, Daniel Webster College
Patrick Motl, Indiana University–Kokomo
Tao Pang, University of Nevada–Las Vegas
Kenneth Podolak, Plattsburgh State University
Takashi Sato, Kwantlen Polytechnic University
David Smith, University of the Virgin Islands
Joseph Trout, Richard Stockton College
Kevin Wheelock, Bellevue College
Senior Contributing Authors - OpenStax Astronomy
Andrew Fraknoi, Foothill College
David Morrison, National Aeronautics and Space Administration
Sidney C. Wolff, National Optical Astronomy Observatories (Emeritus)
Contributing Authors - OpenStax Astronomy
John Beck, Stanford University
Susan D. Benecchi, Planetary Science Institute
John Bochanski, Rider University
Howard Bond, Pennsylvania State University, Emeritus, Space Telescope Science Institute
Jennifer Carson, Occidental College
Bryan Dunne, University of Illinois at Urbana-Champaign
Martin Elvis, Harvard-Smithsonian Center for Astrophysics
Debra Fischer, Yale University
Heidi Hammel, Association of Universities for Research in Astronomy
Tori Hoehler, NASA Ames Research Center
Douglas Ingram, Texas Christian University
Steven Kawaler, Iowa State University
Lloyd Knox, University of California, Davis
Mark Krumholz, Australian National University
James Lowenthal, Smith College
Siobahn Morgan, University of Northern Iowa
Daniel Perley, California Institute of Technology
Claire Raftery, National Solar Observatory
Deborah Scherrer, retired, Stanford University
Phillip Scherrer, Stanford University
Sanjoy Som, Blue Marble Space Institute of Science, NASA Ames Research Center
Wes Tobin, Indiana University East
William H. Waller, retired, Tufts University, Rockport (MA) Public Schools
Todd Young, Wayne State College

This OpenStax book is available for free at https://legacy.cnx.org/content/col12169/1.46


Chapter 1 | Introducing Physics 5

1 | INTRODUCING PHYSICS
Chapter Outline
1.1 Introducing Astrophysics
1.2 The Scope and Scale of Physics
1.3 Units and Standards
1.4 Unit Conversion
1.5 Dimensional Analysis
1.6 Estimates and Fermi Calculations
1.7 Significant Figures
1.8 Solving Problems in Physics

Introduction
You are hopefully reading this book (and taking the course with which it is associated) because you want to get a
mathematically rigorous introduction to the science of physics. To physicists, this means you want to learn to understand
how the world works (and how the things in it work). You likely want to do so because your eventual goal is a career in
physics, engineering, or a related field.
This book is like no other textbook in terms of its approach to introductory physics. The traditional undergraduate physics
course sequence is quite historical. It begins with the 17th and 18th-century development of ideas about motion from
Galileo and Newton, continues with the 19th-century study of energy and heat (by Joule and Carnot), moves on to the
study of electricity and magnetism (by Faraday and Maxwell), and follows through with the 20th-century ideas of Einstein
(especially in relativity) and of Planck, Bohr, Schrodinger and Heisenberg (in quantum physics).
Such an historical approach can have the strength of helping the reader to appreciate the process of science – how each new
discovery is built upon those that came before. A typical textbook’s ordering of the major subdisciplines in physics might
be:
• Classical Mechanics
• Waves and Sound
• Heat and Thermodynamics
• Electricity and Magnetism
• Optics
• Relativity
• Quantum Physics
But, because a rigorous, introductory study of physics can only realistically take place over a protracted period of time
(typically three or four semesters), this traditional, topical approach can sometimes lose track of the big picture. There are
important ideas that cross multiple subdisciplines, have taken many centuries to develop and, in fact, continue to be refined
even in the 21st century.
This book is fashioned around one such big-picture “story line” – our understanding of the Universe. Where did the
Universe (and we) come from? Where are we going? It is the oldest and, to us, the most interesting story in the history of
humankind. It is also multidisciplinary within physics, involving almost all of the traditional subtopics listed above.
Why begin your study of physics with astrophysics? For one thing, in the 21st century, no branch of physics is more active
and dynamic in its pursuit of scientific knowledge. Chances are you may not yet have spent much time in your formal
education studying this subject, but you can surely read almost weekly about new discoveries being made about the nature
and evolution of the Universe. New planets in astonishing numbers have been detected orbiting distant stars. Some are
considered “Earth-like” – are they possibly homes to life? Even some bodies in our own solar system are being revealed
to be very unlike our previous ideas about them. The recent NASA New Horizons mission revealed information about the
dwarf planet Pluto (see Figure 1.1) that has required astronomers to completely rethink the theories of its formation and
6 Chapter 1 | Introducing Physics

evolution.

Figure 1.1 In 2015 NASA's New Horizons spacecraft sent


back this stunning color image of Pluto's surface. Images like
this one have completely overturned astronomers' ideas about
the composition and history of Pluto.

Even more recently, NASA's Juno mission making close approaches to Jupiter has made exquisitely precise images of
magnetic storms there (see Figure 1.2). Those discoveries not only make us re-think our understanding of the planet, but
may in fact lead to new models for the basic formation of planetary magnetic fields.

Figure 1.2 In 2017, NASA's Juno spacecraft sent back this


image of Jupiter's south pole covered in Earth-sized swirling
storms that are densely clustered and rubbing together.

This OpenStax book is available for free at https://legacy.cnx.org/content/col12169/1.46


Chapter 1 | Introducing Physics 7

No questions in science are older than those from astronomy. The recorded history of science begins with astronomy.
Thousands of years ago, from the West (the ancient Greeks) and the East (the Chinese), we find written records of
astronomical data and theories of the Universe. But even before that, it isn’t hard to speculate that human beings’ earliest
questions about their world must have included:
• What causes the regular cycles of the Sun and Moon?
• What are we seeing when we look up in the sky on a clear night?
• Are all those “dots” the same kind of thing? What kind of thing?
• How far away are they?
Indeed, astronomy has been a primary scientific endeavor for well over 2000 years. In the 20th and 21st centuries, the search
for answers to these questions has become “astrophysics”, involving multiple areas of physics. Our story line will include
the study of:
I. Mechanics (the study of motion)
A. Kinematics (a description of motion)
B. Dynamics (an explanation of why things move as they do)
1. Mass, Forces and Newton’s Laws
2. Energy (and its conservation)
a. Kinetic energy
b. Potential energy
3. Momentum (and its conservation)
4. Angular Momentum (and its conservation)
II. Relativity (at high speeds)
A. Time dilation and length contraction
B. Mass-energy conversion: E = mc2
III. Thermodynamics
A. Temperature
B. Heat as transfer of thermal energy
C. Thermal energy as the internal kinetic energy of molecules
D. Phase changes with temperature
1. Liquid-Solid transition: the melting point
2. Gas-Liquid transition: the boiling point
IV. Optics (Everything that we know about what’s “out there” comes from studying the light that reaches us here on
Earth.)
A. Geometric (ray) optics (used to construct lenses, mirrors and telescopes)
B. Physical (wave) optics
C. Electromagnetic waves and the electromagnetic spectrum – from radio waves to gamma rays
D. Types of light (continuous sources vs. line sources)
E. Interference
F. Diffraction
G. The Doppler shift
V. Quantum Physics (After 1900, we learned that the physics that applies at very small scales is not Newtonian.)
A. The ultimate sources of light are molecules, atoms, electrons, and nuclei.
B. Are things in our Universe ultimately particles or waves? (Yes!)
As you can see, if we followed this story line through a typical introductory physics textbook, it would take the whole book
8 Chapter 1 | Introducing Physics

to complete it – insofar as we can ever say it is “complete”. (Perhaps a better way to put it would be “up to its present state
of understanding.”)
It may sound a bit overdramatic when we say that the ultimate goal of this book (and this course) is to understand the
Universe in its entirety. From an astrophysical point of view, we will work from the inside out – beginning with our own
solar system, moving on to consider stars other than our own Sun, and then to study galaxies of stars and what lies beyond,
the large-scale structures that form out of clusters of galaxies.
We will also be faced with the fact the telescopes are time machines, i.e. the farther out in space we look at objects, the
farther back in time we see them. The term cosmology means the study of the history of the Universe, from its beginning to
now and into the future beyond. Our outward journey, then, will help us to paint a picture of that history.
We will study the evidence for the beginning of our Universe in a “Big Bang”, and trace the evolution of stars, solar systems
and galaxies over the roughly 14 billion years of its existence. Through our understanding of the processes that have shaped
the past and the present, we will be able to examine and discuss possible fates (or futures) of the Universe. Does the future
depend upon unseen “stuff” referred to as dark matter? Does it involve an unexplained repulsive force (like an anti-gravity)
called dark energy?
Whether or not the preceding material is enough to convince you that a course in astrophysics is an interesting place to begin
your study of physics, we hope you will leave this chapter with one important take-home thought. We paraphrase the late
astronomer Carl Sagan who, through his ground-breaking television series, Cosmos, did so much to popularize and explain
astrophysics to the general public. Just think about this: “Every atom inside your body, at this instant, once lived inside a
star.”
Does that statement make sense to you? Because it’s absolutely a true, scientific fact. By the end of “The Cosmic Universe”,
you will know both why that is true and how we know it to be so.

1.1 | Introducing Astrophysics


Learning Objectives
By the end of this section you will be able to:
• List some of the levels of structure in the Universe that are studied in astrophysics.

Physics + Astronomy = Astrophysics


What's It All About?
What do you make of this picture?

This OpenStax book is available for free at https://legacy.cnx.org/content/col12169/1.46


Chapter 1 | Introducing Physics 9

Figure 1.3 This image might be showing any number of things. It might be a whirlpool in a tank of
water or perhaps a collage of paint and shiny beads done for art class. Without knowing the size of the
object in units we all recognize, such as meters or inches, it is difficult to know what we’re looking at. In
fact, this image shows the Whirlpool Galaxy (and its companion galaxy), which is about 60,000 light-
years in diameter (about 6 × 10 17 km across). (credit: S. Beckwith (STScI) Hubble Heritage Team,
(STScI/AURA), ESA, NASA)

As noted in the figure caption, this image is of the Whirlpool Galaxy. Galaxies are as immense as atoms are small, yet the
same laws of physics describe both, along with all the rest of nature—an indication of the underlying unity in the universe.
The laws of physics are surprisingly few, implying an underlying simplicity to nature’s apparent complexity. In this text,
you learn about the laws of physics. Galaxies and atoms may seem far removed from your daily life, but as you begin to
explore this broad-ranging subject, you may soon come to realize that physics plays a much larger role in your life than you
first thought, no matter your life goals or career choice.

Figure 1.4 These two interacting islands of stars (galaxies) are so far away that their light takes hundreds of millions of years
to reach us on Earth (photographed with the Hubble Space Telescope). (credit: modification of work by NASA, ESA, the Hubble
Heritage (STScl/AURA)-ESA/Hubble Collaboration, and K. Noll (STScl))

We invite you to come along on a series of voyages to explore the universe as astronomers and physicists understand
it today. Beyond Earth are vast and magnificent realms full of objects that have no counterpart on our home planet.
Nevertheless, we hope to show you that the evolution of the universe has been directly responsible for your presence on
10 Chapter 1 | Introducing Physics

Earth today.
Along your journey, you will encounter:
• a canyon system so large that, on Earth, it would stretch from Los Angeles to Washington, DC (Figure 1.5).

Figure 1.5 This image of Mars is centered on the Valles Marineris (Mariner Valley) complex of canyons, which is as long as
the United States is wide. (credit: modification of work by NASA)

• a crater and other evidence on Earth that tell us that the dinosaurs (and many other creatures) died because of a
cosmic collision.
• a tiny moon whose gravity is so weak that one good throw from its surface could put a baseball into orbit.
• a collapsed star so dense that to duplicate its interior we would have to squeeze every human being on Earth into a
single raindrop.
• exploding stars whose violent end could wipe clean all of the life-forms on a planet orbiting a neighboring star
(Figure 1.6).
• a “cannibal galaxy” that has already consumed a number of its smaller galaxy neighbors and is not yet finished
finding new victims.
• a radio echo that is the faint but unmistakable signal of the creation event for our universe.

This OpenStax book is available for free at https://legacy.cnx.org/content/col12169/1.46


Chapter 1 | Introducing Physics 11

Figure 1.6 We observe the remains of a star that was seen to


explode in our skies in 1054 (and was, briefly, bright enough to
be visible during the daytime). Today, the remnant is called the
Crab Nebula and its central region is seen here. Such exploding
stars are crucial to the development of life in the universe.
(credit: NASA, ESA, J. Hester (Arizona State University))

Such discoveries are what make astronomy such an exciting field for scientists and many others—but you will explore much
more than just the objects in our universe and the latest discoveries about them. We will pay equal attention to the process
by which we have come to understand the realms beyond Earth and the tools we use to increase that understanding.
We gather information about the cosmos from the messages the universe sends our way. Because the stars are the
fundamental building blocks of the universe, decoding the message of starlight has been a central challenge and triumph of
modern astronomy. By the time you have finished reading this text, you will know a bit about how to read that message and
how to understand what it is telling us.

1.2 | The Scope and Scale of Physics


Learning Objectives
By the end of this section, you will be able to:
• Describe the scope of physics.
• Calculate the order of magnitude of a quantity.
• Compare measurable length, mass, and timescales quantitatively.
• Describe the relationships among models, theories, and laws.

Physics is devoted to the understanding of all natural phenomena. In physics, we try to understand physical phenomena at all
scales—from the world of subatomic particles to the entire universe. Despite the breadth of the subject, the various subfields
of physics share a common core. The same basic training in physics will prepare you to work in any area of physics and the
related areas of science and engineering. In this section, we investigate the scope of physics; the scales of length, mass, and
time over which the laws of physics have been shown to be applicable; and the process by which science in general, and
physics in particular, operates.

The Scope of Physics


Take another look at the image of the Whirlpool galaxy. That galaxy contains billions of individual stars as well as
12 Chapter 1 | Introducing Physics

huge clouds of gas and dust. Its companion galaxy is also visible to the right. This pair of galaxies lies a staggering billion
trillion miles (1.4 × 10 21 mi) from our own galaxy (which is called the Milky Way). The stars and planets that make up
the Whirlpool Galaxy might seem to be the furthest thing from most people’s everyday lives, but the Whirlpool is a great
starting point to think about the forces that hold the universe together. The forces that cause the Whirlpool Galaxy to act as
it does are thought to be the same forces we contend with here on Earth, whether we are planning to send a rocket into space
or simply planning to raise the walls for a new home. The gravity that causes the stars of the Whirlpool Galaxy to rotate and
revolve is thought to be the same as what causes water to flow over hydroelectric dams here on Earth. When you look up
at the stars, realize the forces out there are the same as the ones here on Earth. Through a study of physics, you may gain a
greater understanding of the interconnectedness of everything we can see and know in this universe.
Think, now, about all the technological devices you use on a regular basis. Computers, smartphones, global positioning
systems (GPSs), MP3 players, and satellite radio might come to mind. Then, think about the most exciting modern
technologies you have heard about in the news, such as trains that levitate above tracks, “invisibility cloaks” that bend light
around them, and microscopic robots that fight cancer cells in our bodies. All these groundbreaking advances, commonplace
or unbelievable, rely on the principles of physics. Aside from playing a significant role in technology, professionals such
as engineers, pilots, physicians, physical therapists, electricians, and computer programmers apply physics concepts in their
daily work. For example, a pilot must understand how wind forces affect a flight path; a physical therapist must understand
how the muscles in the body experience forces as they move and bend. As you will learn in this text, the principles of
physics are propelling new, exciting technologies, and these principles are applied in a wide range of careers.
The underlying order of nature makes science in general, and physics in particular, interesting and enjoyable to study. For
example, what do a bag of chips and a car battery have in common? Both contain energy that can be converted to other
forms. The law of conservation of energy (which says that energy can change form but is never lost) ties together such
topics as food calories, batteries, heat, light, and watch springs. Understanding this law makes it easier to learn about the
various forms energy takes and how they relate to one another. Apparently unrelated topics are connected through broadly
applicable physical laws, permitting an understanding beyond just the memorization of lists of facts.
Science consists of theories and laws that are the general truths of nature, as well as the body of knowledge they encompass.
Scientists are continuously trying to expand this body of knowledge and to perfect the expression of the laws that describe
it. Physics, which comes from the Greek phúsis, meaning “nature,” is concerned with describing the interactions of
energy, matter, space, and time to uncover the fundamental mechanisms that underlie every phenomenon. This concern for
describing the basic phenomena in nature essentially defines the scope of physics.
Physics aims to understand the world around us at the most basic level. It emphasizes the use of a small number of
quantitative laws to do this, which can be useful to other fields pushing the performance boundaries of existing technologies.
Consider a smartphone (Figure 1.7). Physics describes how electricity interacts with the various circuits inside the device.
This knowledge helps engineers select the appropriate materials and circuit layout when building a smartphone. Knowledge
of the physics underlying these devices is required to shrink their size or increase their processing speed. Or, think about a
GPS. Physics describes the relationship between the speed of an object, the distance over which it travels, and the time it
takes to travel that distance. When you use a GPS in a vehicle, it relies on physics equations to determine the travel time
from one location to another.

This OpenStax book is available for free at https://legacy.cnx.org/content/col12169/1.46


Chapter 1 | Introducing Physics 13

Figure 1.7 The Apple iPhone is a common smartphone with a


GPS function. Physics describes the way that electricity flows
through the circuits of this device. Engineers use their
knowledge of physics to construct an iPhone with features that
consumers will enjoy. One specific feature of an iPhone is the
GPS function. A GPS uses physics equations to determine the
drive time between two locations on a map.

Knowledge of physics is useful in everyday situations as well as in nonscientific professions. It can help you understand
how microwave ovens work, why metals should not be put into them, and why they might affect pacemakers. Physics allows
you to understand the hazards of radiation and to evaluate these hazards rationally and more easily. Physics also explains the
reason why a black car radiator helps remove heat in a car engine, and it explains why a white roof helps keep the inside of
a house cool. Similarly, the operation of a car’s ignition system as well as the transmission of electrical signals throughout
our body’s nervous system are much easier to understand when you think about them in terms of basic physics.
Physics is a key element of many important disciplines and contributes directly to others. Chemistry, for example—since
it deals with the interactions of atoms and molecules—has close ties to atomic and molecular physics. Most branches of
engineering are concerned with designing new technologies, processes, or structures within the constraints set by the laws
of physics. In architecture, physics is at the heart of structural stability and is involved in the acoustics, heating, lighting, and
cooling of buildings. Parts of geology rely heavily on physics, such as radioactive dating of rocks, earthquake analysis, and
heat transfer within Earth. Some disciplines, such as biophysics and geophysics, are hybrids of physics and other disciplines.
Physics has many applications in the biological sciences. On the microscopic level, it helps describe the properties of
cells and their environments. On the macroscopic level, it explains the heat, work, and power associated with the human
body and its various organ systems. Physics is involved in medical diagnostics, such as radiographs, magnetic resonance
imaging, and ultrasonic blood flow measurements. Medical therapy sometimes involves physics directly; for example,
cancer radiotherapy uses ionizing radiation. Physics also explains sensory phenomena, such as how musical instruments
make sound, how the eye detects color, and how lasers transmit information.
It is not necessary to study all applications of physics formally. What is most useful is knowing the basic laws of physics
and developing skills in the analytical methods for applying them. The study of physics also can improve your problem-
solving skills. Furthermore, physics retains the most basic aspects of science, so it is used by all the sciences, and the study
of physics makes other sciences easier to understand.

The Scale of Physics


From the discussion so far, it should be clear that to accomplish your goals in any of the various fields within the natural
sciences and engineering, a thorough grounding in the laws of physics is necessary. The reason for this is simply that the
laws of physics govern everything in the observable universe at all measurable scales of length, mass, and time. Now, that
14 Chapter 1 | Introducing Physics

is easy enough to say, but to come to grips with what it really means, we need to get a little bit quantitative. So, before
surveying the various scales that physics allows us to explore, let’s first look at the concept of “order of magnitude,” which
we use to come to terms with the vast ranges of length, mass, and time that we consider in this text (Figure 1.8).

Figure 1.8 (a) Using a scanning tunneling microscope, scientists can see the individual atoms (diameters around 10 –10 m) that
compose this sheet of gold. (b) Tiny phytoplankton swim among crystals of ice in the Antarctic Sea. They range from a few
micrometers (1 μm is 10–6 m) to as much as 2 mm (1 mm is 10–3 m) in length. (c) These two colliding galaxies, known as NGC
4676A (right) and NGC 4676B (left), are nicknamed “The Mice” because of the tail of gas emanating from each one. They are
located 300 million light-years from Earth in the constellation Coma Berenices. Eventually, these two galaxies will merge into
one. (credit a: modification of work by Erwinrossen; credit b: modification of work by Prof. Gordon T. Taylor, Stony Brook
University; NOAA Corps Collections; credit c: modification of work by NASA, H. Ford (JHU), G. Illingworth (UCSC/LO), M.
Clampin (STScI), G. Hartig (STScI), the ACS Science Team, and ESA)

Order of magnitude
The order of magnitude of a number is the power of 10 that most closely approximates it. Thus, the order of magnitude
refers to the scale (or size) of a value. Each power of 10 represents a different order of magnitude. For example,
10 1, 10 2, 10 3, and so forth, are all different orders of magnitude, as are 10 0 = 1, 10 −1, 10 −2, and 10 −3. To find the
order of magnitude of a number, take the base-10 logarithm of the number and round it to the nearest integer, then the order
of magnitude of the number is simply the resulting power of 10. For example, the order of magnitude of 800 is 103 because
log 10 800 ≈ 2.903, which rounds to 3. Similarly, the order of magnitude of 450 is 103 because log 10 450 ≈ 2.653,
which rounds to 3 as well. Thus, we say the numbers 800 and 450 are of the same order of magnitude: 103. However, the
order of magnitude of 250 is 102 because log 10 250 ≈ 2.397, which rounds to 2.

An equivalent but quicker way to find the order of magnitude of a number is first to write it in scientific notation and then
check to see whether the first factor is greater than or less than 10 = 10 0.5 ≈ 3. The idea is that 10 = 10 0.5 is halfway
between 1 = 10 0 and 10 = 10 1 on a log base-10 scale. Thus, if the first factor is less than 10, then we round it down
to 1 and the order of magnitude is simply whatever power of 10 is required to write the number in scientific notation. On
the other hand, if the first factor is greater than 10, then we round it up to 10 and the order of magnitude is one power of
10 higher than the power needed to write the number in scientific notation. For example, the number 800 can be written in
scientific notation as 8 × 10 2. Because 8 is bigger than 10 ≈ 3, we say the order of magnitude of 800 is 10 2 + 1 = 10 3.
The number 450 can be written as 4.5 × 10 2, so its order of magnitude is also 103 because 4.5 is greater than 3. However,
250 written in scientific notation is 2.5 × 10 2 and 2.5 is less than 3, so its order of magnitude is 10 2.
The order of magnitude of a number is designed to be a ballpark estimate for the scale (or size) of its value. It is simply a
way of rounding numbers consistently to the nearest power of 10. This makes doing rough mental math with very big and
very small numbers easier. For example, the diameter of a hydrogen atom is on the order of 10−10 m, whereas the diameter
of the Sun is on the order of 109 m, so it would take roughly 10 9 /10 −10 = 10 19 hydrogen atoms to stretch across the
diameter of the Sun. This is much easier to do in your head than using the more precise values of 1.06 × 10 −10 m for a
hydrogen atom diameter and 1.39 × 10 9 m for the Sun’s diameter, to find that it would take 1.31 × 10 19 hydrogen atoms
to stretch across the Sun’s diameter. In addition to being easier, the rough estimate is also nearly as informative as the precise
calculation.

This OpenStax book is available for free at https://legacy.cnx.org/content/col12169/1.46


Chapter 1 | Introducing Physics 15

Known ranges of length, mass, and time


The vastness of the universe and the breadth over which physics applies are illustrated by the wide range of examples of
known lengths, masses, and times (given as orders of magnitude) in Figure 1.9. Examining this table will give you a
feeling for the range of possible topics in physics and numerical values. A good way to appreciate the vastness of the ranges
of values in Figure 1.9 is to try to answer some simple comparative questions, such as the following:
• How many hydrogen atoms does it take to stretch across the diameter of the Sun?
(Answer: 109 m/10–10 m = 1019 hydrogen atoms)
• How many protons are there in a bacterium?
(Answer: 10–15 kg/10–27 kg = 1012 protons)
• How many floating-point operations can a supercomputer do in 1 day?
(Answer: 105 s/10–17 s = 1022 floating-point operations)
In studying Figure 1.9, take some time to come up with similar questions that interest you and then try answering them.
Doing this can breathe some life into almost any table of numbers.

Figure 1.9 This table shows the orders of magnitude of length, mass, and time.
16 Chapter 1 | Introducing Physics

Visit this site (https://openstaxcollege.org/l/21scaleuniv) to explore interactively the vast range of length
scales in our universe. Scroll down and up the scale to view hundreds of organisms and objects, and click on the
individual objects to learn more about each one.

Building Models
How did we come to know the laws governing natural phenomena? What we refer to as the laws of nature are concise
descriptions of the universe around us. They are human statements of the underlying laws or rules that all natural processes
follow. Such laws are intrinsic to the universe; humans did not create them and cannot change them. We can only discover
and understand them. Their discovery is a very human endeavor, with all the elements of mystery, imagination, struggle,
triumph, and disappointment inherent in any creative effort (Figure 1.10). The cornerstone of discovering natural laws is
observation; scientists must describe the universe as it is, not as we imagine it to be.

Figure 1.10 (a) Enrico Fermi (1901–1954) was born in Italy. On accepting the Nobel Prize in
Stockholm in 1938 for his work on artificial radioactivity produced by neutrons, he took his
family to America rather than return home to the government in power at the time. He became an
American citizen and was a leading participant in the Manhattan Project. (b) Marie Curie
(1867–1934) sacrificed monetary assets to help finance her early research and damaged her
physical well-being with radiation exposure. She is the only person to win Nobel prizes in both
physics and chemistry. One of her daughters also won a Nobel Prize. (credit a: United States
Department of Energy)

A model is a representation of something that is often too difficult (or impossible) to display directly. Although a model
is justified by experimental tests, it is only accurate in describing certain aspects of a physical system. An example is the
Bohr model of single-electron atoms, in which the electron is pictured as orbiting the nucleus, analogous to the way planets
orbit the Sun (Figure 1.11). We cannot observe electron orbits directly, but the mental image helps explain some of the
observations we can make, such as the emission of light from hot gases (atomic spectra). However, other observations
show that the picture in the Bohr model is not really what atoms look like. The model is “wrong,” but is still useful for
some purposes. Physicists use models for a variety of purposes. For example, models can help physicists analyze a scenario
and perform a calculation or models can be used to represent a situation in the form of a computer simulation. Ultimately,
however, the results of these calculations and simulations need to be double-checked by other means—namely, observation
and experimentation.

This OpenStax book is available for free at https://legacy.cnx.org/content/col12169/1.46


Chapter 1 | Introducing Physics 17

Figure 1.11 What is a model? The Bohr model of a single-


electron atom shows the electron orbiting the nucleus in one of
several possible circular orbits. Like all models, it captures
some, but not all, aspects of the physical system.

The word theory means something different to scientists than what is often meant when the word is used in everyday
conversation. In particular, to a scientist a theory is not the same as a “guess” or an “idea” or even a “hypothesis.” The
phrase “it’s just a theory” seems meaningless and silly to scientists because science is founded on the notion of theories. To
a scientist, a theory is a testable explanation for patterns in nature supported by scientific evidence and verified multiple
times by various groups of researchers. Some theories include models to help visualize phenomena whereas others do not.
Newton’s theory of gravity, for example, does not require a model or mental image, because we can observe the objects
directly with our own senses. The kinetic theory of gases, on the other hand, is a model in which a gas is viewed as being
composed of atoms and molecules. Atoms and molecules are too small to be observed directly with our senses—thus, we
picture them mentally to understand what the instruments tell us about the behavior of gases. Although models are meant
only to describe certain aspects of a physical system accurately, a theory should describe all aspects of any system that falls
within its domain of applicability. In particular, any experimentally testable implication of a theory should be verified. If an
experiment ever shows an implication of a theory to be false, then the theory is either thrown out or modified suitably (for
example, by limiting its domain of applicability).
A law uses concise language to describe a generalized pattern in nature supported by scientific evidence and repeated
experiments. Often, a law can be expressed in the form of a single mathematical equation. Laws and theories are similar
in that they are both scientific statements that result from a tested hypothesis and are supported by scientific evidence.
However, the designation law is usually reserved for a concise and very general statement that describes phenomena in
nature, such as the law that energy is conserved during any process, or Newton’s second law of motion, which relates force
(F), mass (m), and acceleration (a) by the simple equation F = ma. A theory, in contrast, is a less concise statement of
observed behavior. For example, the theory of evolution and the theory of relativity cannot be expressed concisely enough to
be considered laws. The biggest difference between a law and a theory is that a theory is much more complex and dynamic.
A law describes a single action whereas a theory explains an entire group of related phenomena. Less broadly applicable
statements are usually called principles (such as Pascal’s principle, which is applicable only in fluids), but the distinction
between laws and principles often is not made carefully.
The models, theories, and laws we devise sometimes imply the existence of objects or phenomena that are as yet
unobserved. These predictions are remarkable triumphs and tributes to the power of science. It is the underlying order in
the universe that enables scientists to make such spectacular predictions. However, if experimentation does not verify our
predictions, then the theory or law is wrong, no matter how elegant or convenient it is. Laws can never be known with
absolute certainty because it is impossible to perform every imaginable experiment to confirm a law for every possible
scenario. Physicists operate under the assumption that all scientific laws and theories are valid until a counterexample is
observed. If a good-quality, verifiable experiment contradicts a well-established law or theory, then the law or theory must
be modified or overthrown completely.
The study of science in general, and physics in particular, is an adventure much like the exploration of an uncharted ocean.
Discoveries are made; models, theories, and laws are formulated; and the beauty of the physical universe is made more
sublime for the insights gained.
18 Chapter 1 | Introducing Physics

1.3 | Units and Standards


Learning Objectives
By the end of this section, you will be able to:
• Describe how SI base units are defined.
• Describe how derived units are created from base units.
• Express quantities given in SI units using metric prefixes.

As we saw previously, the range of objects and phenomena studied in physics is immense. From the incredibly short lifetime
of a nucleus to the age of Earth, from the tiny sizes of subnuclear particles to the vast distance to the edges of the known
universe, from the force exerted by a jumping flea to the force between Earth and the Sun, there are enough factors of 10
to challenge the imagination of even the most experienced scientist. Giving numerical values for physical quantities and
equations for physical principles allows us to understand nature much more deeply than qualitative descriptions alone. To
comprehend these vast ranges, we must also have accepted units in which to express them. We shall find that even in the
potentially mundane discussion of meters, kilograms, and seconds, a profound simplicity of nature appears: all physical
quantities can be expressed as combinations of only seven base physical quantities.
We define a physical quantity either by specifying how it is measured or by stating how it is calculated from other
measurements. For example, we might define distance and time by specifying methods for measuring them, such as using
a meter stick and a stopwatch. Then, we could define average speed by stating that it is calculated as the total distance
traveled divided by time of travel.
Measurements of physical quantities are expressed in terms of units, which are standardized values. For example, the length
of a race, which is a physical quantity, can be expressed in units of meters (for sprinters) or kilometers (for distance runners).
Without standardized units, it would be extremely difficult for scientists to express and compare measured values in a
meaningful way (Figure 1.12).

Figure 1.12 Distances given in unknown units are


maddeningly useless.

Two major systems of units are used in the world: SI units (for the French Système International d’Unités), also known
as the metric system, and English units (also known as the customary or imperial system). English units were historically
used in nations once ruled by the British Empire and are still widely used in the United States. English units may also be
referred to as the foot–pound–second (fps) system, as opposed to the centimeter–gram–second (cgs) system. You may also
encounter the term SAE units, named after the Society of Automotive Engineers. Products such as fasteners and automotive
tools (for example, wrenches) that are measured in inches rather than metric units are referred to as SAE fasteners or SAE
wrenches.
Virtually every other country in the world (except the United States) now uses SI units as the standard. The metric system
is also the standard system agreed on by scientists and mathematicians.

This OpenStax book is available for free at https://legacy.cnx.org/content/col12169/1.46


Chapter 1 | Introducing Physics 19

SI Units: Base and Derived Units


In any system of units, the units for some physical quantities must be defined through a measurement process. These are
called the base quantities for that system and their units are the system’s base units. All other physical quantities can
then be expressed as algebraic combinations of the base quantities. Each of these physical quantities is then known as a
derived quantity and each unit is called a derived unit. The choice of base quantities is somewhat arbitrary, as long as they
are independent of each other and all other quantities can be derived from them. Typically, the goal is to choose physical
quantities that can be measured accurately to a high precision as the base quantities. The reason for this is simple. Since the
derived units can be expressed as algebraic combinations of the base units, they can only be as accurate and precise as the
base units from which they are derived.
Based on such considerations, the International Standards Organization recommends using seven base quantities, which
form the International System of Quantities (ISQ). These are the base quantities used to define the SI base units. Table 1.1
lists these seven ISQ base quantities and the corresponding SI base units.

ISQ Base Quantity SI Base Unit


Length meter (m)
Mass kilogram (kg)
Time second (s)
Electrical current ampere (A)
Thermodynamic temperature kelvin (K)
Amount of substance mole (mol)
Luminous intensity candela (cd)

Table 1.1 ISQ Base Quantities and Their SI Units

You are probably already familiar with some derived quantities that can be formed from the base quantities in Table 1.1.
For example, the geometric concept of area is always calculated as the product of two lengths. Thus, area is a derived
quantity that can be expressed in terms of SI base units using square meters (m × m = m 2). Similarly, volume is a derived
quantity that can be expressed in cubic meters (m 3). Speed is length per time; so in terms of SI base units, we could
measure it in meters per second (m/s). Volume mass density (or just density) is mass per volume, which is expressed in
terms of SI base units such as kilograms per cubic meter (kg/m3). Angles can also be thought of as derived quantities
because they can be defined as the ratio of the arc length subtended by two radii of a circle to the radius of the circle. This
is how the radian is defined. Depending on your background and interests, you may be able to come up with other derived
quantities, such as the mass flow rate (kg/s) or volume flow rate (m3/s) of a fluid, electric charge (A · s), mass flux density
[kg/(m 2 · s)], and so on. We will see many more examples throughout this text. For now, the point is that every physical
quantity can be derived from the seven base quantities in Table 1.1, and the units of every physical quantity can be derived
from the seven SI base units.
For the most part, we use SI units in this text. Non-SI units are used in a few applications in which they are in very common
use, such as the measurement of temperature in degrees Celsius (°C), the measurement of fluid volume in liters (L), and
the measurement of energies of elementary particles in electron-volts (eV). Whenever non-SI units are discussed, they are
tied to SI units through conversions. For example, 1 L is 10 −3 m 3.
Check out a comprehensive source of information on SI units (https://openstaxcollege.org/l/21SIUnits) at
the National Institute of Standards and Technology (NIST) Reference on Constants, Units, and Uncertainty.

Units of Time, Length, and Mass: The Second, Meter, and Kilogram
The initial chapters in this textbook are concerned with mechanics, fluids, and waves. In these subjects all pertinent physical
quantities can be expressed in terms of the base units of length, mass, and time. Therefore, we now turn to a discussion of
these three base units, leaving discussion of the others until they are needed later.
20 Chapter 1 | Introducing Physics

The second
The SI unit for time, the second (abbreviated s), has a long history. For many years it was defined as 1/86,400 of a mean
solar day. More recently, a new standard was adopted to gain greater accuracy and to define the second in terms of a
nonvarying or constant physical phenomenon (because the solar day is getting longer as a result of the very gradual slowing
of Earth’s rotation). Cesium atoms can be made to vibrate in a very steady way, and these vibrations can be readily observed
and counted. In 1967, the second was redefined as the time required for 9,192,631,770 of these vibrations to occur (Figure
1.13). Note that this may seem like more precision than you would ever need, but it isn’t—GPSs rely on the precision of
atomic clocks to be able to give you turn-by-turn directions on the surface of Earth, far from the satellites broadcasting their
location.

Figure 1.13 An atomic clock such as this one uses the


vibrations of cesium atoms to keep time to a precision of better
than a microsecond per year. The fundamental unit of time, the
second, is based on such clocks. This image looks down from
the top of an atomic fountain nearly 30 feet tall. (credit: Steve
Jurvetson)

The meter
The SI unit for length is the meter (abbreviated m); its definition has also changed over time to become more precise. The
meter was first defined in 1791 as 1/10,000,000 of the distance from the equator to the North Pole. This measurement was
improved in 1889 by redefining the meter to be the distance between two engraved lines on a platinum–iridium bar now
kept near Paris. By 1960, it had become possible to define the meter even more accurately in terms of the wavelength of
light, so it was again redefined as 1,650,763.73 wavelengths of orange light emitted by krypton atoms. In 1983, the meter
was given its current definition (in part for greater accuracy) as the distance light travels in a vacuum in 1/299,792,458 of a
second (Figure 1.14). This change came after knowing the speed of light to be exactly 299,792,458 m/s. The length of the
meter will change if the speed of light is someday measured with greater accuracy.

Figure 1.14 The meter is defined to be the distance light travels in 1/299,792,458 of a second in a vacuum. Distance
traveled is speed multiplied by time.

The kilogram
The SI unit for mass is the kilogram (abbreviated kg); it is defined to be the mass of a platinum–iridium cylinder
kept with the old meter standard at the International Bureau of Weights and Measures near Paris. Exact replicas of the
standard kilogram are also kept at the U.S. National Institute of Standards and Technology (NIST), located in Gaithersburg,
Maryland, outside of Washington, DC, and at other locations around the world. Scientists at NIST are currently investigating
two complementary methods of redefining the kilogram (see Figure 1.15). The determination of all other masses can be

This OpenStax book is available for free at https://legacy.cnx.org/content/col12169/1.46


Chapter 1 | Introducing Physics 21

traced ultimately to a comparison with the standard mass.


There is currently an effort to redefine the SI unit of mass in terms of more fundamental processes by 2018. You
can explore the history of mass standards and the contenders in the quest to devise a new one at the website
(https://openstaxcollege.org/l/21redefkilo) of the Physical Measurement Laboratory.

Figure 1.15 Redefining the SI unit of mass. Complementary methods are being investigated for use in an upcoming
redefinition of the SI unit of mass. (a) The U.S. National Institute of Standards and Technology’s watt balance is a machine that
balances the weight of a test mass against the current and voltage (the “watt”) produced by a strong system of magnets. (b) The
International Avogadro Project is working to redefine the kilogram based on the dimensions, mass, and other known properties of
a silicon sphere. (credit a and credit b: National Institute of Standards and Technology)

Metric Prefixes
SI units are part of the metric system, which is convenient for scientific and engineering calculations because the units are
categorized by factors of 10. Table 1.2 lists the metric prefixes and symbols used to denote various factors of 10 in SI units.
For example, a centimeter is one-hundredth of a meter (in symbols, 1 cm = 10–2 m) and a kilometer is a thousand meters (1
km = 103 m). Similarly, a megagram is a million grams (1 Mg = 106 g), a nanosecond is a billionth of a second (1 ns = 10–9
s), and a terameter is a trillion meters (1 Tm = 1012 m).

Prefix Symbol Meaning Prefix Symbol Meaning


yotta- Y 1024 yocto- y 10–24
zetta- Z 1021 zepto- z 10–21
exa- E 1018 atto- a 10–18
peta- P 1015 femto- f 10–15
tera- T 1012 pico- p 10–12
giga- G 109 nano- n 10–9
mega- M 106 micro- μ 10–6

kilo- k 103 milli- m 10–3

Table 1.2 Metric Prefixes for Powers of 10 and Their Symbols


22 Chapter 1 | Introducing Physics

Prefix Symbol Meaning Prefix Symbol Meaning


hecto- h 102 centi- c 10–2
deka- da 101 deci- d 10–1

Table 1.2 Metric Prefixes for Powers of 10 and Their Symbols

The only rule when using metric prefixes is that you cannot “double them up.” For example, if you have measurements in
petameters (1 Pm = 1015 m), it is not proper to talk about megagigameters, although 10 6 × 10 9 = 10 15. In practice, the
only time this becomes a bit confusing is when discussing masses. As we have seen, the base SI unit of mass is the kilogram
(kg), but metric prefixes need to be applied to the gram (g), because we are not allowed to “double-up” prefixes. Thus, a
thousand kilograms (103 kg) is written as a megagram (1 Mg) since
10 3 kg = 10 3 × 10 3 g = 10 6 g = 1 Mg.

Incidentally, 103 kg is also called a metric ton, abbreviated t. This is one of the units outside the SI system considered
acceptable for use with SI units.
As we see in the next section, metric systems have the advantage that conversions of units involve only powers of 10. There
are 100 cm in 1 m, 1000 m in 1 km, and so on. In nonmetric systems, such as the English system of units, the relationships
are not as simple—there are 12 in. in 1 ft, 5280 ft in 1 mi, and so on.
Another advantage of metric systems is that the same unit can be used over extremely large ranges of values simply by
scaling it with an appropriate metric prefix. The prefix is chosen by the order of magnitude of physical quantities commonly
found in the task at hand. For example, distances in meters are suitable in construction, whereas distances in kilometers are
appropriate for air travel, and nanometers are convenient in optical design. With the metric system there is no need to invent
new units for particular applications. Instead, we rescale the units with which we are already familiar.

Example 1.1

Using Metric Prefixes


Restate the mass 1.93 × 10 13 kg using a metric prefix such that the resulting numerical value is bigger than one
but less than 1000.
Strategy
Since we are not allowed to “double-up” prefixes, we first need to restate the mass in grams by replacing the
prefix symbol k with a factor of 103 (see Table 1.2). Then, we should see which two prefixes in Table 1.2 are
closest to the resulting power of 10 when the number is written in scientific notation. We use whichever of these
two prefixes gives us a number between one and 1000.
Solution
Replacing the k in kilogram with a factor of 103, we find that
1.93 × 10 13 kg = 1.93 × 10 13 × 10 3 g = 1.93 × 10 16 g.

From Table 1.2, we see that 1016 is between “peta-” (1015) and “exa-” (1018). If we use the “peta-” prefix, then
we find that 1.93 × 10 16 g = 1.93 × 10 1 Pg, since 16 = 1 + 15. Alternatively, if we use the “exa-” prefix we
find that 1.93 × 10 16 g = 1.93 × 10 −2 Eg, since 16 = −2 + 18. Because the problem asks for the numerical
value between one and 1000, we use the “peta-” prefix and the answer is 19.3 Pg.
Significance
It is easy to make silly arithmetic errors when switching from one prefix to another, so it is always a good idea to
check that our final answer matches the number we started with. An easy way to do this is to put both numbers in
scientific notation and count powers of 10, including the ones hidden in prefixes. If we did not make a mistake,
the powers of 10 should match up. In this problem, we started with 1.93 × 10 13 kg, so we have 13 + 3 = 16
powers of 10. Our final answer in scientific notation is 1.93 × 10 1 Pg, so we have 1 + 15 = 16 powers of 10. So,

This OpenStax book is available for free at https://legacy.cnx.org/content/col12169/1.46


Chapter 1 | Introducing Physics 23

everything checks out.


If this mass arose from a calculation, we would also want to check to determine whether a mass this large makes
any sense in the context of the problem. For this, Figure 1.9 might be helpful.

1.1 Check Your Understanding Restate 4.79 × 10 5 kg using a metric prefix such that the resulting number
is bigger than one but less than 1000.

1.4 | Unit Conversion


Learning Objectives
By the end of this section, you will be able to:
• Use conversion factors to express the value of a given quantity in different units.

It is often necessary to convert from one unit to another. For example, if you are reading a European cookbook, some
quantities may be expressed in units of liters and you need to convert them to cups. Or perhaps you are reading walking
directions from one location to another and you are interested in how many miles you will be walking. In this case, you may
need to convert units of feet or meters to miles.
The Power of 1
Even though it will be expressed in new units, we do NOT want to change the actual value of the quantity. In algebra, there
is a safe way to keep the value of a quantity unchanged: multiply it by 1. So, if all we ever do is multiply our quantity by 1,
we are assured that we keep the same value.
The secret is in a clever use of the many ways there are in which to write the quantity 1. In particular, any fraction whose
numerator and denominator are equal does in fact have the value 1. The particular fractions we will choose are called
conversion factors.
Let’s consider a simple example of how to convert units. Suppose we want to convert 80 m to kilometers. The first thing to
do is to list the units you have and the units to which you want to convert. In this case, we have units in meters and we want
to convert to kilometers. Next, we need to determine a conversion factor relating meters to kilometers. A conversion factor
is a ratio that expresses how many of one unit are equal to another unit. For example, there are 12 in. in 1 ft, 1609 m in 1
mi, 100 cm in 1 m, 60 s in 1 min, and so on. Refer to Appendix B for a more complete list of conversion factors. In this
case, we know that there are 1000 m in 1 km. Now we can set up our unit conversion. We write the units we have and then
multiply them by the conversion factor so the units cancel out, as shown:

80 m × 1 km = 0.080 km.
1000 m
Why did the actual quantity (the distance involved) not change? Because all we did, mathematically, was to multiply it by
1. Our conversion factor is a fraction, the value of whose numerator (1 km) is equal to the value of its denominator (1000
m). So, it is just another way to write 1.
Note that the unwanted meter unit cancels, leaving only the desired kilometer unit. You can use this method to convert
between any type of unit. Of course, the conversion of 80 m to kilometers is simply the use of a metric prefix, as we saw in
the preceding section, so we can get the same answer just as easily by noting that
80 m = 8.0 × 10 1 m = 8.0 × 10 −2 km = 0.080 km,
since “kilo-” means 103 (see Table 1.2) and 1 = −2 + 3. However, using conversion factors is handy when converting
between units that are not metric or when converting between derived units, as the following examples illustrate.
24 Chapter 1 | Introducing Physics

Example 1.2

Converting Nonmetric Units to Metric


The distance from the university to home is 10 mi and it usually takes 20 min to drive this distance. Calculate the
average speed in meters per second (m/s). (Note: Average speed is distance traveled divided by time of travel.)
Strategy
First we calculate the average speed using the given units, then we can get the average speed into the desired
units by picking the correct conversion factors and multiplying by them. The correct conversion factors are those
that cancel the unwanted units and leave the desired units in their place. In this case, we want to convert miles to
meters, so we need to know the fact that there are 1609 m in 1 mi. We also want to convert minutes to seconds,
so we use the conversion of 60 s in 1 min.
Solution
1. Calculate average speed. Average speed is distance traveled divided by time of travel. (Take this definition
as a given for now. Average speed and other motion concepts are covered in later chapters.) In equation
form,

Average speed = Distance .


Time
2. Substitute the given values for distance and time:

Average speed = 10 mi = 0.50 mi .


20 min min
3. Convert miles per minute to meters per second by multiplying by the conversion factor that cancels miles
and leave meters, and also by the conversion factor that cancels minutes and leave seconds:
(0.50)(1609)
0.50 mile × 1609 m × 1 min = m/s = 13 m/s.
min 1 mile 60 s 60

Significance
Check the answer in the following ways:
1. Be sure that each conversion factor is a fraction whose numerator and denominator are equal. This ensures
that all you ever do is multiply your quantity by 1 (sometimes repeatedly).
2. Be sure the units in the unit conversion cancel correctly. If the unit conversion factor was written upside
down, the units do not cancel correctly in the equation. We see the “miles” in the numerator in 0.50
mi/min cancels the “mile” in the denominator in the first conversion factor. Also, the “min” in the
denominator in 0.50 mi/min cancels the “min” in the numerator in the second conversion factor.
3. Check that the units of the final answer are the desired units. The problem asked us to solve for average
speed in units of meters per second and, after the cancellations, the only units left are a meter (m) in the
numerator and a second (s) in the denominator, so we have indeed obtained these units.

1.2 Check Your Understanding Light travels about 9 Pm in a year. Given that a year is about 3 × 10 7 s,
what is the speed of light in meters per second?

Example 1.3

Converting between Metric Units


The density of iron is 7.86 g/cm 3 under standard conditions. Convert this to kg/m3.

Strategy
We need to convert grams to kilograms and cubic centimeters to cubic meters. The conversion factors we need are
1 kg = 10 3 g and 1 cm = 10 −2 m. However, we are dealing with cubic centimeters (cm 3 = cm × cm × cm),

This OpenStax book is available for free at https://legacy.cnx.org/content/col12169/1.46


Chapter 1 | Introducing Physics 25

so we have to use the second conversion factor three times (that is, we need to cube it). The idea is still to multiply
by the conversion factors in such a way that they cancel the units we want to get rid of and introduce the units we
want to keep.
Solution

⎛ cm ⎞
3
g kg 7.86
10 3 g ⎝10 −2 m ⎠
7.86 3
× × = kg/m 3 = 7.86 × 10 3 kg/m 3
cm (10 3)(10 −6)

Significance
Remember, it’s always important to check the answer.
1. Be sure that each conversion factor is a fraction whose numerator and denominator are equal. In this case,
the first conversion factor has a numerator of 1 kg and a denominator of 103 g. The second conversion
factor has a numerator of 1 cm and a denominator of 10-2 m.
2. Be sure to cancel the units in the unit conversion correctly. We see that the gram (“g”) in the numerator
in 7.86 g/cm3 cancels the “g” in the denominator in the first conversion factor. Also, the three factors of
“cm” in the denominator in 7.86 g/cm3 cancel with the three factors of “cm” in the numerator that we get
by cubing the second conversion factor.
3. Check that the units of the final answer are the desired units. The problem asked for us to convert to
kilograms per cubic meter. After the cancellations just described, we see the only units we have left are
“kg” in the numerator and three factors of “m” in the denominator (that is, one factor of “m” cubed, or
“m3”). Therefore, the units on the final answer are correct.

1.3 Check Your Understanding We know from Figure 1.9 that the diameter of Earth is on the order of 107
m, so the order of magnitude of its surface area is 1014 m2. What is that in square kilometers (that is, km2)? (Try
doing this both by converting 107 m to km and then squaring it and then by converting 1014 m2 directly to
square kilometers. You should get the same answer both ways.)

Unit conversions may not seem very interesting, but not doing them can be costly. One famous example of this situation was
seen with the Mars Climate Orbiter. This probe was launched by NASA on December 11, 1998. On September 23, 1999,
while attempting to guide the probe into its planned orbit around Mars, NASA lost contact with it. Subsequent investigations
showed a piece of software called SM_FORCES (or “small forces”) was recording thruster performance data in the English
units of pound-seconds (lb-s). However, other pieces of software that used these values for course corrections expected them
to be recorded in the SI units of newton-seconds (N-s), as dictated in the software interface protocols. This error caused the
probe to follow a very different trajectory from what NASA thought it was following, which most likely caused the probe
either to burn up in the Martian atmosphere or to shoot out into space. This failure to pay attention to unit conversions cost
hundreds of millions of dollars, not to mention all the time invested by the scientists and engineers who worked on the
project.
1.4 Check Your Understanding Given that 1 lb (pound) is 4.45 N, were the numbers being output by
SM_FORCES too big or too small?

1.5 | Dimensional Analysis


Learning Objectives
By the end of this section, you will be able to:
• Find the dimensions of a mathematical expression involving physical quantities.
• Determine whether an equation involving physical quantities is dimensionally consistent.

The dimension of any physical quantity expresses its dependence on the base quantities as a product of symbols (or powers
26 Chapter 1 | Introducing Physics

of symbols) representing the base quantities. Table 1.3 lists the base quantities and the symbols used for their dimension.
For example, a measurement of length is said to have dimension L or L1, a measurement of mass has dimension M or
M1, and a measurement of time has dimension T or T1. Like units, dimensions obey the rules of algebra. Thus, area is
the product of two lengths and so has dimension L2, or length squared. Similarly, volume is the product of three lengths
and has dimension L3, or length cubed. Speed has dimension length over time, L/T or LT–1. Volumetric mass density has
dimension M/L3 or ML–3, or mass over length cubed. In general, the dimension of any physical quantity can be written as
L a M b T c I d Θ e N f J g for some powers a, b, c, d, e, f , and g. We can write the dimensions of a length in this form with
a = 1 and the remaining six powers all set equal to zero: L 1 = L 1 M 0 T 0 I 0 Θ 0 N 0 J 0. Any quantity with a dimension that
can be written so that all seven powers are zero (that is, its dimension is L 0 M 0 T 0 I 0 Θ 0 N 0 J 0 ) is called dimensionless
(or sometimes “of dimension 1,” because anything raised to the zero power is one). Physicists often call dimensionless
quantities pure numbers.

Base Quantity Symbol for Dimension


Length L
Mass M
Time T
Current I
Thermodynamic temperature Θ
Amount of substance N
Luminous intensity J

Table 1.3 Base Quantities and Their Dimensions

Physicists often use square brackets around the symbol for a physical quantity to represent the dimensions of that quantity.
For example, if r is the radius of a cylinder and h is its height, then we write [r] = L and [h] = L to indicate the
dimensions of the radius and height are both those of length, or L. Similarly, if we use the symbol A for the surface area of
a cylinder and V for its volume, then [A] = L2 and [V] = L3. If we use the symbol m for the mass of the cylinder and ρ
for the density of the material from which the cylinder is made, then [m] = M and [ρ] = ML −3.

The importance of the concept of dimension arises from the fact that any mathematical equation relating physical quantities
must be dimensionally consistent, which means the equation must obey the following rules:
• Every term in an expression must have the same dimensions; it does not make sense to add or subtract quantities of
differing dimension (think of the old saying: “You can’t add apples and oranges”). In particular, the expressions on
each side of the equality in an equation must have the same dimensions.
• The arguments of any of the standard mathematical functions such as trigonometric functions (such as sine and
cosine), logarithms, or exponential functions that appear in the equation must be dimensionless. These functions
require pure numbers as inputs and give pure numbers as outputs.
If either of these rules is violated, an equation is not dimensionally consistent and cannot possibly be a correct statement
of physical law. This simple fact can be used to check for typos or algebra mistakes, to help remember the various laws of
physics, and even to suggest the form that new laws of physics might take. This last use of dimensions is beyond the scope
of this text, but is something you will undoubtedly learn later in your academic career.

Example 1.4

Using Dimensions to Remember an Equation


Suppose we need the formula for the area of a circle for some computation. Like many people who learned
geometry too long ago to recall with any certainty, two expressions may pop into our mind when we think of
circles: πr 2 and 2πr. One expression is the circumference of a circle of radius r and the other is its area. But
which is which?

This OpenStax book is available for free at https://legacy.cnx.org/content/col12169/1.46


Chapter 1 | Introducing Physics 27

Strategy
One natural strategy is to look it up, but this could take time to find information from a reputable source. Besides,
even if we think the source is reputable, we shouldn’t trust everything we read. It is nice to have a way to double-
check just by thinking about it. Also, we might be in a situation in which we cannot look things up (such as during
a test). Thus, the strategy is to find the dimensions of both expressions by making use of the fact that dimensions
follow the rules of algebra. If either expression does not have the same dimensions as area, then it cannot possibly
be the correct equation for the area of a circle.
Solution
We know the dimension of area is L2. Now, the dimension of the expression πr 2 is

[πr 2] = [π] · [r] 2 = 1 · L 2 = L 2,

since the constant π is a pure number and the radius r is a length. Therefore, πr 2 has the dimension of area.
Similarly, the dimension of the expression 2πr is
[2πr] = [2] · [π] · [r] = 1 · 1 · L = L,
since the constants 2 and π are both dimensionless and the radius r is a length. We see that 2πr has the
dimension of length, which means it cannot possibly be an area.
We rule out 2πr because it is not dimensionally consistent with being an area. We see that πr 2 is dimensionally
consistent with being an area, so if we have to choose between these two expressions, πr 2 is the one to choose.
Significance
This may seem like kind of a silly example, but the ideas are very general. As long as we know the dimensions
of the individual physical quantities that appear in an equation, we can check to see whether the equation is
dimensionally consistent. On the other hand, knowing that true equations are dimensionally consistent, we can
match expressions from our imperfect memories to the quantities for which they might be expressions. Doing this
will not help us remember dimensionless factors that appear in the equations (for example, if you had accidentally
conflated the two expressions from the example into 2πr 2, then dimensional analysis is no help), but it does
help us remember the correct basic form of equations.

1.5 Check Your Understanding Suppose we want the formula for the volume of a sphere. The two
expressions commonly mentioned in elementary discussions of spheres are 4πr 2 and 4πr 3 /3. One is the
volume of a sphere of radius r and the other is its surface area. Which one is the volume?

Example 1.5

Checking Equations for Dimensional Consistency


Consider the physical quantities s, v, a, and t with dimensions [s] = L, [v] = LT −1, [a] = LT −2,
and [t] = T. Determine whether each of the following equations is dimensionally consistent: (a)
s = vt + 0.5at 2; (b) s = vt 2 + 0.5at; and (c) v = sin(at 2 /s).
Strategy
By the definition of dimensional consistency, we need to check that each term in a given equation has the same
dimensions as the other terms in that equation and that the arguments of any standard mathematical functions are
dimensionless.
Solution
a. There are no trigonometric, logarithmic, or exponential functions to worry about in this equation, so we
need only look at the dimensions of each term appearing in the equation. There are three terms, one in the
28 Chapter 1 | Introducing Physics

left expression and two in the expression on the right, so we look at each in turn:
[s] = L
[vt] = [v] · [t] = LT −1 · T = LT 0 = L
[0.5at 2] = [a] · [t] 2 = LT −2 · T 2 = LT 0 = L.

All three terms have the same dimension, so this equation is dimensionally consistent.
b. Again, there are no trigonometric, exponential, or logarithmic functions, so we only need to look at the
dimensions of each of the three terms appearing in the equation:
[s] = L
[vt 2] = [v] · [t] 2 = LT −1 · T 2 = LT
[at] = [a] · [t] = LT −2 · T = LT −1.

None of the three terms has the same dimension as any other, so this is about as far from being
dimensionally consistent as you can get. The technical term for an equation like this is nonsense.
c. This equation has a trigonometric function in it, so first we should check that the argument of the sine
function is dimensionless:
⎡at 2 ⎤ [a] · [t] 2 LT −2 · T 2 L
⎣ s ⎦ = [s] = L
= = 1.
L

The argument is dimensionless. So far, so good. Now we need to check the dimensions of each of the two
terms (that is, the left expression and the right expression) in the equation:
[v] = LT −1
⎡ ⎛at 2 ⎞⎤
⎣sin⎝ s ⎠⎦ = 1.
The two terms have different dimensions—meaning, the equation is not dimensionally consistent. This equation
is another example of “nonsense.”
Significance
If we are trusting people, these types of dimensional checks might seem unnecessary. But, rest assured, any
textbook on a quantitative subject such as physics (including this one) almost certainly contains some equations
with typos. Checking equations routinely by dimensional analysis save us the embarrassment of using an incorrect
equation. Also, checking the dimensions of an equation we obtain through algebraic manipulation is a great way
to make sure we did not make a mistake (or to spot a mistake, if we made one).

1.6 Check Your Understanding Is the equation v = at dimensionally consistent?

One further point that needs to be mentioned is the effect of the operations of calculus on dimensions. We have seen that
dimensions obey the rules of algebra, just like units, but what happens when we take the derivative of one physical quantity
with respect to another or integrate a physical quantity over another? The derivative of a function is just the slope of the line
tangent to its graph and slopes are ratios, so for physical quantities v and t, we have that the dimension of the derivative of
v with respect to t is just the ratio of the dimension of v over that of t:
⎡ dv ⎤ [v]
⎣ dt ⎦ = [t] .

Similarly, since integrals are just sums of products, the dimension of the integral of v with respect to t is simply the
dimension of v times the dimension of t:

⎣∫ vdt⎤⎦ = [v] · [t].

This OpenStax book is available for free at https://legacy.cnx.org/content/col12169/1.46


Chapter 1 | Introducing Physics 29

By the same reasoning, analogous rules hold for the units of physical quantities derived from other quantities by integration
or differentiation.

1.6 | Estimates and Fermi Calculations


Learning Objectives
By the end of this section, you will be able to:
• Estimate the values of physical quantities.

On many occasions, physicists, other scientists, and engineers need to make estimates for a particular quantity. Other
terms sometimes used are guesstimates, order-of-magnitude approximations, back-of-the-envelope calculations, or Fermi
calculations. (The physicist Enrico Fermi mentioned earlier was famous for his ability to estimate various kinds of data with
surprising precision.) Will that piece of equipment fit in the back of the car or do we need to rent a truck? How long will
this download take? About how large a current will there be in this circuit when it is turned on? How many houses could
a proposed power plant actually power if it is built? Note that estimating does not mean guessing a number or a formula
at random. Rather, estimation means using prior experience and sound physical reasoning to arrive at a rough idea of a
quantity’s value. Because the process of determining a reliable approximation usually involves the identification of correct
physical principles and a good guess about the relevant variables, estimating is very useful in developing physical intuition.
Estimates also allow us perform “sanity checks” on calculations or policy proposals by helping us rule out certain scenarios
or unrealistic numbers. They allow us to challenge others (as well as ourselves) in our efforts to learn truths about the world.
Many estimates are based on formulas in which the input quantities are known only to a limited precision. As you develop
physics problem-solving skills (which are applicable to a wide variety of fields), you also will develop skills at estimating.
You develop these skills by thinking more quantitatively and by being willing to take risks. As with any skill, experience
helps. Familiarity with dimensions (see Table 1.3) and units (see Table 1.1 and Table 1.2), and the scales of base
quantities (see Figure 1.9) also helps.
To make some progress in estimating, you need to have some definite ideas about how variables may be related. The
following strategies may help you in practicing the art of estimation:
• Get big lengths from smaller lengths. When estimating lengths, remember that anything can be a ruler. Thus,
imagine breaking a big thing into smaller things, estimate the length of one of the smaller things, and multiply to
get the length of the big thing. For example, to estimate the height of a building, first count how many floors it
has. Then, estimate how big a single floor is by imagining how many people would have to stand on each other’s
shoulders to reach the ceiling. Last, estimate the height of a person. The product of these three estimates is your
estimate of the height of the building. It helps to have memorized a few length scales relevant to the sorts of
problems you find yourself solving. For example, knowing some of the length scales in Figure 1.9 might come
in handy. Sometimes it also helps to do this in reverse—that is, to estimate the length of a small thing, imagine a
bunch of them making up a bigger thing. For example, to estimate the thickness of a sheet of paper, estimate the
thickness of a stack of paper and then divide by the number of pages in the stack. These same strategies of breaking
big things into smaller things or aggregating smaller things into a bigger thing can sometimes be used to estimate
other physical quantities, such as masses and times.
• Get areas and volumes from lengths. When dealing with an area or a volume of a complex object, introduce a simple
model of the object such as a sphere or a box. Then, estimate the linear dimensions (such as the radius of the sphere
or the length, width, and height of the box) first, and use your estimates to obtain the volume or area from standard
geometric formulas. If you happen to have an estimate of an object’s area or volume, you can also do the reverse;
that is, use standard geometric formulas to get an estimate of its linear dimensions.
• Get masses from volumes and densities. When estimating masses of objects, it can help first to estimate its volume
and then to estimate its mass from a rough estimate of its average density (recall, density has dimension mass over
length cubed, so mass is density times volume). For this, it helps to remember that the density of air is around 1 kg/
m3, the density of water is 103 kg/m3, and the densest everyday solids max out at around 104 kg/m3. Asking yourself
whether an object floats or sinks in either air or water gets you a ballpark estimate of its density. You can also do
this the other way around; if you have an estimate of an object’s mass and its density, you can use them to get an
estimate of its volume.
• If all else fails, bound it. For physical quantities for which you do not have a lot of intuition, sometimes the best
you can do is think something like: Well, it must be bigger than this and smaller than that. For example, suppose
30 Chapter 1 | Introducing Physics

you need to estimate the mass of a moose. Maybe you have a lot of experience with moose and know their average
mass offhand. If so, great. But for most people, the best they can do is to think something like: It must be bigger
than a person (of order 102 kg) and less than a car (of order 103 kg). If you need a single number for a subsequent
calculation, you can take the geometric mean of the upper and lower bound—that is, you multiply them together
and then take the square root. For the moose mass example, this would be
0.5
⎛ 2
⎝10 × 10 3⎞⎠ = 10 2.5 = 10 0.5 × 10 2 ≈ 3 × 10 2 kg.

The tighter the bounds, the better. Also, no rules are unbreakable when it comes to estimation. If you think the value
of the quantity is likely to be closer to the upper bound than the lower bound, then you may want to bump up your
estimate from the geometric mean by an order or two of magnitude.
• One “sig. fig.” is fine. There is no need to go beyond one significant figure when doing calculations to obtain an
estimate. In most cases, the order of magnitude is good enough. The goal is just to get in the ballpark figure, so keep
the arithmetic as simple as possible.
• Ask yourself: Does this make any sense? Last, check to see whether your answer is reasonable. How does it compare
with the values of other quantities with the same dimensions that you already know or can look up easily? If you
get some wacky answer (for example, if you estimate the mass of the Atlantic Ocean to be bigger than the mass of
Earth, or some time span to be longer than the age of the universe), first check to see whether your units are correct.
Then, check for arithmetic errors. Then, rethink the logic you used to arrive at your answer. If everything checks
out, you may have just proved that some slick new idea is actually bogus.

Example 1.6

Mass of Earth’s Oceans


Estimate the total mass of the oceans on Earth.
Strategy
We know the density of water is about 103 kg/m3, so we start with the advice to “get masses from densities
and volumes.” Thus, we need to estimate the volume of the planet’s oceans. Using the advice to “get areas and
volumes from lengths,” we can estimate the volume of the oceans as surface area times average depth, or V = AD.
We know the diameter of Earth from Figure 1.9 and we know that most of Earth’s surface is covered in water,
so we can estimate the surface area of the oceans as being roughly equal to the surface area of the planet. By
following the advice to “get areas and volumes from lengths” again, we can approximate Earth as a sphere and
use the formula for the surface area of a sphere of diameter d—that is, A = πd 2, to estimate the surface area of
the oceans. Now we just need to estimate the average depth of the oceans. For this, we use the advice: “If all else
fails, bound it.” We happen to know the deepest points in the ocean are around 10 km and that it is not uncommon
for the ocean to be deeper than 1 km, so we take the average depth to be around (10 3 × 10 4) 0.5 ≈ 3 × 10 3 m.
Now we just need to put it all together, heeding the advice that “one ‘sig. fig.’ is fine.”
Solution
We estimate the surface area of Earth (and hence the surface area of Earth’s oceans) to be roughly
A = πd 2 = π(10 7 m) 2 ≈ 3 × 10 14 m 2.

Next, using our average depth estimate of D = 3 × 10 3 m, which was obtained by bounding, we estimate the
volume of Earth’s oceans to be
V = AD = (3 × 10 14 m 2)(3 × 10 3 m) = 9 × 10 17 m 3.
Last, we estimate the mass of the world’s oceans to be
M = ρV = (10 3 kg/m 3)(9 × 10 17 m 3) = 9 × 10 20 kg.

Thus, we estimate that the order of magnitude of the mass of the planet’s oceans is 1021 kg.

This OpenStax book is available for free at https://legacy.cnx.org/content/col12169/1.46


Chapter 1 | Introducing Physics 31

Significance
To verify our answer to the best of our ability, we first need to answer the question: Does this make any sense?
From Figure 1.9, we see the mass of Earth’s atmosphere is on the order of 1019 kg and the mass of Earth is on
the order of 1025 kg. It is reassuring that our estimate of 1021 kg for the mass of Earth’s oceans falls somewhere
between these two. So, yes, it does seem to make sense. It just so happens that we did a search on the Web for
“mass of oceans” and the top search results all said 1.4 × 10 21 kg, which is the same order of magnitude as our
estimate. Now, rather than having to trust blindly whoever first put that number up on a website (most of the other
sites probably just copied it from them, after all), we can have a little more confidence in it.

1.7 Check Your Understanding Figure 1.9 says the mass of the atmosphere is 1019 kg. Assuming the
density of the atmosphere is 1 kg/m3, estimate the height of Earth’s atmosphere. Do you think your answer is an
underestimate or an overestimate? Explain why.

How many piano tuners are there in New York City? How many leaves are on that tree? If you are studying photosynthesis
or thinking of writing a smartphone app for piano tuners, then the answers to these questions might be of great interest to
you. Otherwise, you probably couldn’t care less what the answers are. However, these are exactly the sorts of estimation
problems that people in various tech industries have been asking potential employees to evaluate their quantitative reasoning
skills. If building physical intuition and evaluating quantitative claims do not seem like sufficient reasons for you to practice
estimation problems, how about the fact that being good at them just might land you a high-paying job?
For practice estimating relative lengths, areas, and volumes, check out this PhET
(https://openstaxcollege.org/l/21lengthgame) simulation, titled “Estimation.”

1.7 | Significant Figures


Learning Objectives
By the end of this section, you will be able to:
• Determine the correct number of significant figures for the result of a computation.
• Describe the relationship between the concepts of accuracy, precision, uncertainty, and
discrepancy.
• Calculate the percent uncertainty of a measurement, given its value and its uncertainty.
• Determine the uncertainty of the result of a computation involving quantities with given
uncertainties.

Figure 1.16 shows two instruments used to measure the mass of an object. The digital scale has mostly replaced the
double-pan balance in physics labs because it gives more accurate and precise measurements. But what exactly do we
mean by accurate and precise? Aren’t they the same thing? In this section we examine in detail the process of making and
reporting a measurement.
32 Chapter 1 | Introducing Physics

Figure 1.16 (a) A double-pan mechanical balance is used to compare different masses. Usually an object with unknown mass
is placed in one pan and objects of known mass are placed in the other pan. When the bar that connects the two pans is
horizontal, then the masses in both pans are equal. The “known masses” are typically metal cylinders of standard mass such as 1
g, 10 g, and 100 g. (b) Many mechanical balances, such as double-pan balances, have been replaced by digital scales, which can
typically measure the mass of an object more precisely. A mechanical balance may read only the mass of an object to the nearest
tenth of a gram, but many digital scales can measure the mass of an object up to the nearest thousandth of a gram. (credit a:
modification of work by Serge Melki; credit b: modification of work by Karel Jakubec)

Accuracy and Precision of a Measurement


Science is based on observation and experiment—that is, on measurements. Accuracy is how close a measurement is to the
accepted reference value for that measurement. For example, let’s say we want to measure the length of standard printer
paper. The packaging in which we purchased the paper states that it is 11.0 in. long. We then measure the length of the paper
three times and obtain the following measurements: 11.1 in., 11.2 in., and 10.9 in. These measurements are quite accurate
because they are very close to the reference value of 11.0 in. In contrast, if we had obtained a measurement of 12 in., our
measurement would not be very accurate. Notice that the concept of accuracy requires that an accepted reference value be
given.
The precision of measurements refers to how close the agreement is between repeated independent measurements (which
are repeated under the same conditions). Consider the example of the paper measurements. The precision of the
measurements refers to the spread of the measured values. One way to analyze the precision of the measurements is to
determine the range, or difference, between the lowest and the highest measured values. In this case, the lowest value was
10.9 in. and the highest value was 11.2 in. Thus, the measured values deviated from each other by, at most, 0.3 in. These
measurements were relatively precise because they did not vary too much in value. However, if the measured values had
been 10.9 in., 11.1 in., and 11.9 in., then the measurements would not be very precise because there would be significant
variation from one measurement to another. Notice that the concept of precision depends only on the actual measurements
acquired and does not depend on an accepted reference value.
The measurements in the paper example are both accurate and precise, but in some cases, measurements are accurate but
not precise, or they are precise but not accurate. Let’s consider an example of a GPS attempting to locate the position
of a restaurant in a city. Think of the restaurant location as existing at the center of a bull’s-eye target and think of each
GPS attempt to locate the restaurant as a black dot. In Figure 1.17(a), we see the GPS measurements are spread out far
apart from each other, but they are all relatively close to the actual location of the restaurant at the center of the target.
This indicates a low-precision, high-accuracy measuring system. However, in Figure 1.17(b), the GPS measurements are
concentrated quite closely to one another, but they are far away from the target location. This indicates a high-precision,
low-accuracy measuring system.

This OpenStax book is available for free at https://legacy.cnx.org/content/col12169/1.46


Chapter 1 | Introducing Physics 33

Figure 1.17 A GPS attempts to locate a restaurant at the center of the bull’s-eye. The
black dots represent each attempt to pinpoint the location of the restaurant. (a) The dots
are spread out quite far apart from one another, indicating low precision, but they are each
rather close to the actual location of the restaurant, indicating high accuracy. (b) The dots
are concentrated rather closely to one another, indicating high precision, but they are
rather far away from the actual location of the restaurant, indicating low accuracy. (credit
a and credit b: modification of works by Dark Evil)

Accuracy, Precision, Uncertainty, and Discrepancy


The precision of a measuring system is related to the uncertainty in the measurements whereas the accuracy is related
to the discrepancy from the accepted reference value. Uncertainty is a quantitative measure of how much your measured
values deviate from one another. There are many different methods of calculating uncertainty, each of which is appropriate
to different situations. Some examples include taking the range (that is, the biggest less the smallest) or finding the standard
deviation of the measurements. Discrepancy (or “measurement error”) is the difference between the measured value and a
given standard or expected value. If the measurements are not very precise, then the uncertainty of the values is high. If the
measurements are not very accurate, then the discrepancy of the values is high.
Recall our example of measuring paper length; we obtained measurements of 11.1 in., 11.2 in., and 10.9 in., and the accepted
value was 11.0 in. We might average the three measurements to say our best guess is 11.1 in.; in this case, our discrepancy
is 11.1 – 11.0 = 0.1 in., which provides a quantitative measure of accuracy. We might calculate the uncertainty in our best
guess by using the range of our measured values: 0.3 in. Then we would say the length of the paper is 11.1 in. plus or minus
0.3 in. The uncertainty in a measurement, A, is often denoted as δA (read “delta A”), so the measurement result would be
recorded as A ± δA. Returning to our paper example, the measured length of the paper could be expressed as 11.1 ± 0.3
in. Since the discrepancy of 0.1 in. is less than the uncertainty of 0.3 in., we might say the measured value agrees with the
accepted reference value to within experimental uncertainty.
Some factors that contribute to uncertainty in a measurement include the following:
• Limitations of the measuring device
• The skill of the person taking the measurement
• Irregularities in the object being measured
• Any other factors that affect the outcome (highly dependent on the situation)
In our example, such factors contributing to the uncertainty could be the smallest division on the ruler is 1/16 in., the person
using the ruler has bad eyesight, the ruler is worn down on one end, or one side of the paper is slightly longer than the other.
At any rate, the uncertainty in a measurement must be calculated to quantify its precision. If a reference value is known, it
makes sense to calculate the discrepancy as well to quantify its accuracy.
Percent uncertainty
Another method of expressing uncertainty is as a percent of the measured value. If a measurement A is expressed with
uncertainty δA, the percent uncertainty is defined as

Percent uncertainty = δA × 100%.


A
34 Chapter 1 | Introducing Physics

Example 1.7

Calculating Percent Uncertainty: A Bag of Apples


A grocery store sells 5-lb bags of apples. Let’s say we purchase four bags during the course of a month and weigh
the bags each time. We obtain the following measurements:
• Week 1 weight: 4.8 lb
• Week 2 weight: 5.3 lb
• Week 3 weight: 4.9 lb
• Week 4 weight: 5.4 lb
We then determine the average weight of the 5-lb bag of apples is 5.1 ± 0.2 lb. What is the percent uncertainty of
the bag’s weight?
Strategy
First, observe that the average value of the bag’s weight, A, is 5.1 lb. The uncertainty in this value, δA, is 0.2 lb.
We can use the following equation to determine the percent uncertainty of the weight:

Percent uncertainty = δA × 100%. (1.1)


A

Solution
Substitute the values into the equation:
Percent Uncertainty
δA × 100% = 0.2 lb × 100% = 3.9% ≈ 4%.
A 5.1 lb
Significance
We can conclude the average weight of a bag of apples from this store is 5.1 lb ± 4%. Notice the percent
uncertainty is dimensionless because the units of weight in δA = 0.2 lb canceled those inn A = 5.1 lb when we
took the ratio.

1.8 Check Your Understanding A high school track coach has just purchased a new stopwatch. The
stopwatch manual states the stopwatch has an uncertainty of ±0.05 s. Runners on the track coach’s team
regularly clock 100-m sprints of 11.49 s to 15.01 s. At the school’s last track meet, the first-place sprinter came
in at 12.04 s and the second-place sprinter came in at 12.07 s. Will the coach’s new stopwatch be helpful in
timing the sprint team? Why or why not?

Combining uncertainties in calculations


Addition or Subtraction of Quantities
What happens if two different measurements need to be combined by addition or subtraction to calculate a final quantity?
For example, suppose we want to determine the weight of a suitcase using a bathroom scale whose range can only measure
objects weighing more than 50 lb. One method is to first take a measurement of the total weight of a person standing on
the scale holding the suitcase. Let's call that measurement Wt. Perhaps we obtain a value of 193±2 lb. (The sources of the
uncertainty may come from multiple causes: how hard it was for the person to read the scale, the fact that the needle on the
scale may have been jiggling somewhat, etc.) Then, take another measurement of the person alone, Wp. Suppose that has a
value of 175±1 lb. (Since he was not holding the suitcase at the same time he read the scale, perhaps the uncertainty was
less this time.) The weight of the suitcase can obviously be calculated from Ws = Wt - Wp. And, 193 lb - 175 lb = 18 lb. But,
there must be some uncertainty in the weight of the suitcase, because the individual measurements that we subtracted each
had uncertainties.
How do we combine the uncertainties from the individual measurements, Wt ± δWt and Wp ± dWp to arrive at a final result,

This OpenStax book is available for free at https://legacy.cnx.org/content/col12169/1.46


Chapter 1 | Introducing Physics 35

Ws ± δWs? We've already seen that the value of Ws is found by simply subtracting Wt - Wp. But what do we do with the
uncertainties? The simplest way might be to just add the uncertainties, using the logic that they both contribute to the
uncertainty in the final quantity. However more sophisticated analysis reveals that, if the uncertainties in the individual
measurements were independent of one another, we are likely overestimating the uncertainty in our final result if we simply
add them. A more accurate final answer is obtained by taking the square root of the sum of the squares of the individual
uncertainties. That is:
Adding Uncertainties in Quadrature
δW s = δW t2 + δW p2 (1.2)

So, δW s = 2 2 + 1 2 = 5 = 2.236

In our example, then, the uncertainty in the weight of the suitcase, is 2.236 lb. However, we must follow the rules for
significant figures (discussed in detail below). Since the individual weight measurements are expressed to the nearest lb, we
will round this calculated number to the nearest lb, and express the weight of the suitcase as 18 ± 2 lb.
Multiplication or Division of Quantities
Uncertainty exists in anything calculated from measured quantities. For example, the area of a floor calculated from
measurements of its length and width has an uncertainty because the length and width each have uncertainties. How big is
the uncertainty in something you calculate by multiplication or division? In this case, the measurements may not even have
the same dimensions. In this case, assuming that the uncertainties in the individual measurements are independent of one
another, the percent uncertainty in a quantity calculated by multiplication or division is the quadrature sum of the percent
uncertainties in the items used to make the calculation. Equivalently, the relative uncertainty in a quantity calculated by
multiplication or division is the quadrature sum of the relative uncertainties in the items used to make the calculation. In our
case, if A = H × W, then
Adding Relative Uncertainties in Quadrature
δA = ⎛δL ⎞ + ⎛δW ⎞
2 2 (1.3)
A ⎝L⎠ ⎝W ⎠

For example, if a floor has a length of L = 4.00 m and a width of W = 3.00 m, with uncertainties of 1% and 2%, respectively,
then the area of the floor is 12.0 m2 and has an uncertainty of 1 2 + 2 2 = 5 = 2.236 %. (Expressed as an area, this is
0.268 m2 [ 12.0 m 2 × 0.02236 ], which we round to 0.3 m2 since the area of the floor is given to a tenth of a square meter.)
So, we express the area of the floor as A = 12.0 ± 0.3 m2.

Precision of Measuring Tools and Significant Figures


An important factor in the precision of measurements involves the precision of the measuring tool. In general, a precise
measuring tool is one that can measure values in very small increments. For example, a standard ruler can measure length to
the nearest millimeter whereas a caliper can measure length to the nearest 0.01 mm. The caliper is a more precise measuring
tool because it can measure extremely small differences in length. The more precise the measuring tool, the more precise
the measurements.
When we express measured values, we can only list as many digits as we measured initially with our measuring tool. For
example, if we use a standard ruler to measure the length of a stick, we may measure it to be 36.7 cm. We can’t express this
value as 36.71 cm because our measuring tool is not precise enough to measure a hundredth of a centimeter. It should be
noted that the last digit in a measured value has been estimated in some way by the person performing the measurement. For
example, the person measuring the length of a stick with a ruler notices the stick length seems to be somewhere in between
36.6 cm and 36.7 cm, and he or she must estimate the value of the last digit. Using the method of significant figures, the
rule is that the last digit written down in a measurement is the first digit with some uncertainty. To determine the number
of significant digits in a value, start with the first measured value at the left and count the number of digits through the last
digit written on the right. For example, the measured value 36.7 cm has three digits, or three significant figures. Significant
figures indicate the precision of the measuring tool used to measure a value.
Zeros
Special consideration is given to zeros when counting significant figures. The zeros in 0.053 are not significant because
they are placeholders that locate the decimal point. There are two significant figures in 0.053. The zeros in 10.053 are
not placeholders; they are significant. This number has five significant figures. The zeros in 1300 may or may not be
significant, depending on the style of writing numbers. They could mean the number is known to the last digit or they
could be placeholders. So 1300 could have two, three, or four significant figures. To avoid this ambiguity, we should write
36 Chapter 1 | Introducing Physics

1300 in scientific notation as 1.3 × 10 3, 1.30 × 10 3, or 1.300 × 10 3, depending on whether it has two, three, or four
significant figures. Zeros are significant except when they serve only as placeholders.
Significant figures in calculations
When combining measurements with different degrees of precision, the number of significant digits in the final answer can
be no greater than the number of significant digits in the least-precise measured value. There are two different rules, one
for multiplication and division and the other for addition and subtraction.
1. For multiplication and division, the result should have the same number of significant figures as the quantity
with the least number of significant figures entering into the calculation. For example, the area of a circle can be
calculated from its radius using A = πr2. Let’s see how many significant figures the area has if the radius has only
two—say, r = 1.2 m. Using a calculator with an eight-digit output, we would calculate
A = πr 2 = (3.1415927…) × (1.2 m) 2 = 4.5238934 m 2.

But because the radius has only two significant figures, it limits the calculated quantity to two significant figures, or
A = 4.5 m 2,

although π is good to at least eight digits.


2. For addition and subtraction, the answer can contain no more decimal places than the least-precise measurement.
Suppose we buy 7.56 kg of potatoes in a grocery store as measured with a scale with precision 0.01 kg, then we drop
off 6.052 kg of potatoes at your laboratory as measured by a scale with precision 0.001 kg. Then, we go home and
add 13.7 kg of potatoes as measured by a bathroom scale with precision 0.1 kg. How many kilograms of potatoes do
we now have and how many significant figures are appropriate in the answer? The mass is found by simple addition
and subtraction:

7.56 kg
−6.052 kg
+13.7 kg
= 15.2 kg.
15.208 kg

Next, we identify the least-precise measurement: 13.7 kg. This measurement is expressed to the 0.1 decimal place,
so our final answer must also be expressed to the 0.1 decimal place. Thus, the answer is rounded to the tenths place,
giving us 15.2 kg.
Significant figures in this text
In this text, most numbers are assumed to have three significant figures. Furthermore, consistent numbers of significant
figures are used in all worked examples. An answer given to three digits is based on input good to at least three digits,
for example. If the input has fewer significant figures, the answer will also have fewer significant figures. Care is also
taken that the number of significant figures is reasonable for the situation posed. In some topics, particularly in optics, more
accurate numbers are needed and we use more than three significant figures. Finally, if a number is exact, such as the two
in the formula for the circumference of a circle, C = 2πr, it does not affect the number of significant figures in a calculation.
Likewise, conversion factors such as 100 cm/1 m are considered exact and do not affect the number of significant figures in
a calculation.

1.8 | Solving Problems in Physics


Learning Objectives
By the end of this section, you will be able to:
• Describe the process for developing a problem-solving strategy.
• Explain how to find the numerical solution to a problem.
• Summarize the process for assessing the significance of the numerical solution to a problem.

This OpenStax book is available for free at https://legacy.cnx.org/content/col12169/1.46


Chapter 1 | Introducing Physics 37

Figure 1.18 Problem-solving skills are essential to your success in physics. (credit:
“scui3asteveo”/Flickr)

Problem-solving skills are clearly essential to success in a quantitative course in physics. More important, the ability
to apply broad physical principles—usually represented by equations—to specific situations is a very powerful form of
knowledge. It is much more powerful than memorizing a list of facts. Analytical skills and problem-solving abilities can be
applied to new situations whereas a list of facts cannot be made long enough to contain every possible circumstance. Such
analytical skills are useful both for solving problems in this text and for applying physics in everyday life.
As you are probably well aware, a certain amount of creativity and insight is required to solve problems. No rigid procedure
works every time. Creativity and insight grow with experience. With practice, the basics of problem solving become almost
automatic. One way to get practice is to work out the text’s examples for yourself as you read. Another is to work as
many end-of-section problems as possible, starting with the easiest to build confidence and then progressing to the more
difficult. After you become involved in physics, you will see it all around you, and you can begin to apply it to situations
you encounter outside the classroom, just as is done in many of the applications in this text.
Although there is no simple step-by-step method that works for every problem, the following three-stage process facilitates
problem solving and makes it more meaningful. The three stages are strategy, solution, and significance. This process is
used in examples throughout the book. Here, we look at each stage of the process in turn.

Strategy
Strategy is the beginning stage of solving a problem. The idea is to figure out exactly what the problem is and then develop
a strategy for solving it. Some general advice for this stage is as follows:
• Examine the situation to determine which physical principles are involved. It often helps to draw a simple sketch
at the outset. You often need to decide which direction is positive and note that on your sketch. When you have
identified the physical principles, it is much easier to find and apply the equations representing those principles.
Although finding the correct equation is essential, keep in mind that equations represent physical principles, laws of
nature, and relationships among physical quantities. Without a conceptual understanding of a problem, a numerical
solution is meaningless.
• Make a list of what is given or can be inferred from the problem as stated (identify the “knowns”). Many problems
are stated very succinctly and require some inspection to determine what is known. Drawing a sketch be very useful
at this point as well. Formally identifying the knowns is of particular importance in applying physics to real-world
situations. For example, the word stopped means the velocity is zero at that instant. Also, we can often take initial
time and position as zero by the appropriate choice of coordinate system.
• Identify exactly what needs to be determined in the problem (identify the unknowns). In complex problems,
especially, it is not always obvious what needs to be found or in what sequence. Making a list can help identify the
unknowns.
38 Chapter 1 | Introducing Physics

• Determine which physical principles can help you solve the problem. Since physical principles tend to be expressed
in the form of mathematical equations, a list of knowns and unknowns can help here. It is easiest if you can find
equations that contain only one unknown—that is, all the other variables are known—so you can solve for the
unknown easily. If the equation contains more than one unknown, then additional equations are needed to solve the
problem. In some problems, several unknowns must be determined to get at the one needed most. In such problems
it is especially important to keep physical principles in mind to avoid going astray in a sea of equations. You may
have to use two (or more) different equations to get the final answer.

Solution
The solution stage is when you do the math. Substitute the knowns (along with their units) into the appropriate equation
and obtain numerical solutions complete with units. That is, do the algebra, calculus, geometry, or arithmetic necessary to
find the unknown from the knowns, being sure to carry the units through the calculations. This step is clearly important
because it produces the numerical answer, along with its units. Notice, however, that this stage is only one-third of the
overall problem-solving process.

Significance
After having done the math in the solution stage of problem solving, it is tempting to think you are done. But, always
remember that physics is not math. Rather, in doing physics, we use mathematics as a tool to help us understand nature. So,
after you obtain a numerical answer, you should always assess its significance:
• Check your units. If the units of the answer are incorrect, then an error has been made and you should go back over
your previous steps to find it. One way to find the mistake is to check all the equations you derived for dimensional
consistency. However, be warned that correct units do not guarantee the numerical part of the answer is also correct.
• Check the answer to see whether it is reasonable. Does it make sense? This step is extremely important: –the goal
of physics is to describe nature accurately. To determine whether the answer is reasonable, check both its magnitude
and its sign, in addition to its units. The magnitude should be consistent with a rough estimate of what it should
be. It should also compare reasonably with magnitudes of other quantities of the same type. The sign usually tells
you about direction and should be consistent with your prior expectations. Your judgment will improve as you solve
more physics problems, and it will become possible for you to make finer judgments regarding whether nature is
described adequately by the answer to a problem. This step brings the problem back to its conceptual meaning. If
you can judge whether the answer is reasonable, you have a deeper understanding of physics than just being able to
solve a problem mechanically.
• Check to see whether the answer tells you something interesting. What does it mean? This is the flip side of the
question: Does it make sense? Ultimately, physics is about understanding nature, and we solve physics problems to
learn a little something about how nature operates. Therefore, assuming the answer does make sense, you should
always take a moment to see if it tells you something about the world that you find interesting. Even if the answer to
this particular problem is not very interesting to you, what about the method you used to solve it? Could the method
be adapted to answer a question that you do find interesting? In many ways, it is in answering questions such as
these science that progresses.

This OpenStax book is available for free at https://legacy.cnx.org/content/col12169/1.46


Chapter 1 | Introducing Physics 39

CHAPTER 1 REVIEW
KEY TERMS
accuracy the degree to which a measured value agrees with an accepted reference value for that measurement
base quantity physical quantity chosen by convention and practical considerations such that all other physical quantities
can be expressed as algebraic combinations of them
base unit standard for expressing the measurement of a base quantity within a particular system of units; defined by a
particular procedure used to measure the corresponding base quantity
conversion factor a ratio that expresses how many of one unit are equal to another unit
derived quantity physical quantity defined using algebraic combinations of base quantities
derived units units that can be calculated using algebraic combinations of the fundamental units
dimension expression of the dependence of a physical quantity on the base quantities as a product of powers of symbols
representing the base quantities; in general, the dimension of a quantity has the form L a M b T c I d Θ e N f J g for some
powers a, b, c, d, e, f, and g.
dimensionally consistent equation in which every term has the same dimensions and the arguments of any
mathematical functions appearing in the equation are dimensionless
dimensionless quantity with a dimension of L 0 M 0 T 0 I 0 Θ 0 N 0 J 0 = 1; also called quantity of dimension 1 or a pure
number
discrepancy the difference between the measured value and a given standard or expected value
English units system of measurement used in the United States; includes units of measure such as feet, gallons, and
pounds
estimation using prior experience and sound physical reasoning to arrive at a rough idea of a quantity’s value; sometimes
called an “order-of-magnitude approximation,” a “guesstimate,” a “back-of-the-envelope calculation”, or a “Fermi
calculation”
kilogram SI unit for mass, abbreviated kg
law description, using concise language or a mathematical formula, of a generalized pattern in nature supported by
scientific evidence and repeated experiments
meter SI unit for length, abbreviated m
method of adding percents the percent uncertainty in a quantity calculated by multiplication or division is the sum of
the percent uncertainties in the items used to make the calculation.
metric system system in which values can be calculated in factors of 10
model representation of something often too difficult (or impossible) to display directly
order of magnitude the size of a quantity as it relates to a power of 10
percent uncertainty the ratio of the uncertainty of a measurement to the measured value, expressed as a percentage
physical quantity characteristic or property of an object that can be measured or calculated from other measurements
physics science concerned with describing the interactions of energy, matter, space, and time; especially interested in
what fundamental mechanisms underlie every phenomenon
precision the degree to which repeated measurements agree with each other
second the SI unit for time, abbreviated s
SI units the international system of units that scientists in most countries have agreed to use; includes units such as
meters, liters, and grams
significant figures used to express the precision of a measuring tool used to measure a value
40 Chapter 1 | Introducing Physics

theory testable explanation for patterns in nature supported by scientific evidence and verified multiple times by various
groups of researchers
uncertainty a quantitative measure of how much measured values deviate from one another
units standards used for expressing and comparing measurements

KEY EQUATIONS
Percent uncertainty Percent uncertainty = δA × 100%
A

Adding uncertainties in quadrature If z = x ± y, then δz = δx 2 + δy 2

2 ⎛δy ⎞ 2
⎛δx ⎞
Adding relative uncertainties in quadrature If z = xy or z = x/y then δz
z = ⎝ x +⎝y ⎠

SUMMARY
1.2 The Scope and Scale of Physics
• Physics is about trying to find the simple laws that describe all natural phenomena.
• Physics operates on a vast range of scales of length, mass, and time. Scientists use the concept of the order of
magnitude of a number to track which phenomena occur on which scales. They also use orders of magnitude to
compare the various scales.
• Scientists attempt to describe the world by formulating models, theories, and laws.

1.3 Units and Standards


• Systems of units are built up from a small number of base units, which are defined by accurate and precise
measurements of conventionally chosen base quantities. Other units are then derived as algebraic combinations of
the base units.
• Two commonly used systems of units are English units and SI units. All scientists and most of the other people in
the world use SI, whereas nonscientists in the United States still tend to use English units.
• The SI base units of length, mass, and time are the meter (m), kilogram (kg), and second (s), respectively.
• SI units are a metric system of units, meaning values can be calculated by factors of 10. Metric prefixes may be
used with metric units to scale the base units to sizes appropriate for almost any application.

1.4 Unit Conversion


• To convert a quantity from one unit to another, multiply by conversions factors in such a way that you cancel the
units you want to get rid of and introduce the units you want to end up with.
• Be careful with areas and volumes. Units obey the rules of algebra so, for example, if a unit is squared we need two
factors to cancel it.

1.5 Dimensional Analysis


• The dimension of a physical quantity is just an expression of the base quantities from which it is derived.
• All equations expressing physical laws or principles must be dimensionally consistent. This fact can be used as an
aid in remembering physical laws, as a way to check whether claimed relationships between physical quantities are
possible, and even to derive new physical laws.

1.6 Estimates and Fermi Calculations


• An estimate is a rough educated guess at the value of a physical quantity based on prior experience and sound
physical reasoning. Some strategies that may help when making an estimate are as follows:

This OpenStax book is available for free at https://legacy.cnx.org/content/col12169/1.46


Chapter 1 | Introducing Physics 41

◦ Get big lengths from smaller lengths.


◦ Get areas and volumes from lengths.
◦ Get masses from volumes and densities.
◦ If all else fails, bound it.
◦ One “sig. fig.” is fine.
◦ Ask yourself: Does this make any sense?

1.7 Significant Figures


• Accuracy of a measured value refers to how close a measurement is to an accepted reference value. The discrepancy
in a measurement is the amount by which the measurement result differs from this value.
• Precision of measured values refers to how close the agreement is between repeated measurements. The uncertainty
of a measurement is a quantification of this.
• The precision of a measuring tool is related to the size of its measurement increments. The smaller the measurement
increment, the more precise the tool.
• Significant figures express the precision of a measuring tool.
• When multiplying or dividing measured values, the final answer can contain only as many significant figures as the
least-precise value.
• When adding or subtracting measured values, the final answer cannot contain more decimal places than the least-
precise value.

1.8 Solving Problems in Physics

The three stages of the process for solving physics problems used in this book are as follows:
• Strategy: Determine which physical principles are involved and develop a strategy for using them to solve the
problem.
• Solution: Do the math necessary to obtain a numerical solution complete with units.
• Significance: Check the solution to make sure it makes sense (correct units, reasonable magnitude and sign) and
assess its significance.

CONCEPTUAL QUESTIONS
6. Can the validity of a model be limited or must it be
1.2 The Scope and Scale of Physics universally valid? How does this compare with the required
1. What is physics? validity of a theory or a law?

2. Some have described physics as a “search for


simplicity.” Explain why this might be an appropriate 1.3 Units and Standards
description. 7. Identify some advantages of metric units.

3. If two different theories describe experimental 8. What are the SI base units of length, mass, and time?
observations equally well, can one be said to be more valid
than the other (assuming both use accepted rules of logic)? 9. What is the difference between a base unit and a derived
unit? (b) What is the difference between a base quantity and
4. What determines the validity of a theory? a derived quantity? (c) What is the difference between a
base quantity and a base unit?
5. Certain criteria must be satisfied if a measurement or
observation is to be believed. Will the criteria necessarily 10. For each of the following scenarios, refer to Figure
be as strict for an expected result as for an unexpected 1.9 and Table 1.2 to determine which metric prefix on
result? the meter is most appropriate for each of the following
scenarios. (a) You want to tabulate the mean distance from
42 Chapter 1 | Introducing Physics

the Sun for each planet in the solar system. (b) You want relationship between the accuracy and the discrepancy of a
to compare the sizes of some common viruses to design a measurement?
mechanical filter capable of blocking the pathogenic ones.
(c) You want to list the diameters of all the elements on the
periodic table. (d) You want to list the distances to all the 1.8 Solving Problems in Physics
stars that have now received any radio broadcasts sent from
12. What information do you need to choose which
Earth 10 years ago.
equation or equations to use to solve a problem?

1.7 Significant Figures 13. What should you do after obtaining a numerical
answer when solving a problem?
11. (a) What is the relationship between the precision
and the uncertainty of a measurement? (b) What is the

PROBLEMS
18. Roughly how many times longer than the mean life of
1.2 The Scope and Scale of Physics an extremely unstable atomic nucleus is the lifetime of a
14. Find the order of magnitude of the following physical human?
quantities. (a) The mass of Earth’s atmosphere:
5.1 × 10 18 kg; (b) The mass of the Moon’s atmosphere: 19. Calculate the approximate number of atoms in a
bacterium. Assume the average mass of an atom in the
25,000 kg; (c) The mass of Earth’s hydrosphere: bacterium is 10 times the mass of a proton.
1.4 × 10 21 kg; (d) The mass of Earth: 5.97 × 10 24 kg;
(e) The mass of the Moon: 7.34 × 10 22 kg; (f) The 20. (a) Calculate the number of cells in a hummingbird
assuming the mass of an average cell is 10 times the mass
Earth–Moon distance (semimajor axis): 3.84 × 10 8 m; (g) of a bacterium. (b) Making the same assumption, how
The mean Earth–Sun distance: 1.5 × 10 11 m; (h) The many cells are there in a human?

equatorial radius of Earth: 6.38 × 10 6 m; (i) The mass of 21. Assuming one nerve impulse must end before another
−31 can begin, what is the maximum firing rate of a nerve in
an electron: 9.11 × 10 kg; (j) The mass of a proton:
impulses per second?
−27
1.67 × 10 kg; (k) The mass of the Sun:
1.99 × 10 30 kg. 22. About how many floating-point operations can a
supercomputer perform each year?

15. Use the orders of magnitude you found in the previous


23. Roughly how many floating-point operations can a
problem to answer the following questions to within an
supercomputer perform in a human lifetime?
order of magnitude. (a) How many electrons would it take
to equal the mass of a proton? (b) How many Earths would
it take to equal the mass of the Sun? (c) How many
Earth–Moon distances would it take to cover the distance 1.3 Units and Standards
from Earth to the Sun? (d) How many Moon atmospheres 24. The following times are given using metric prefixes
would it take to equal the mass of Earth’s atmosphere? (e) on the base SI unit of time: the second. Rewrite them in
How many moons would it take to equal the mass of Earth? scientific notation without the prefix. For example, 47 Ts
(f) How many protons would it take to equal the mass of would be rewritten as 4.7 × 10 13 s. (a) 980 Ps; (b) 980 fs;
the Sun?
(c) 17 ns; (d) 577 μs.

For the remaining questions, you need to use Figure 1.9


to obtain the necessary orders of magnitude of lengths, 25. The following times are given in seconds. Use metric
masses, and times. prefixes to rewrite them so the numerical value is greater
16. Roughly how many heartbeats are there in a lifetime? than one but less than 1000. For example, 7.9 × 10 −2 s
could be written as either 7.9 cs or 79 ms. (a) 9.57 × 10 5 s;
17. A generation is about one-third of a lifetime. (b) 0.045 s; (c) 5.5 × 10 −7 s; (d) 3.16 × 10 7 s.
Approximately how many generations have passed since
the year 0 AD?
26. The following lengths are given using metric prefixes

This OpenStax book is available for free at https://legacy.cnx.org/content/col12169/1.46


Chapter 1 | Introducing Physics 43

on the base SI unit of length: the meter. Rewrite them in 35. Soccer fields vary in size. A large soccer field is 115
scientific notation without the prefix. For example, 4.2 Pm m long and 85.0 m wide. What is its area in square feet?
would be rewritten as 4.2 × 10 15 m. (a) 89 Tm; (b) 89 pm; (Assume that 1 m = 3.281 ft.)
(c) 711 mm; (d) 0.45 μm.
36. What is the height in meters of a person who is 6 ft 1.0
in. tall?
27. The following lengths are given in meters. Use metric
prefixes to rewrite them so the numerical value is bigger
37. Mount Everest, at 29,028 ft, is the tallest mountain on
than one but less than 1000. For example, 7.9 × 10 −2 m Earth. What is its height in kilometers? (Assume that 1 m =
could be written either as 7.9 cm or 79 mm. (a) 3.281 ft.)
7.59 × 10 7 m; (b) 0.0074 m; (c) 8.8 × 10 −11 m; (d)
38. The speed of sound is measured to be 342 m/s on
1.63 × 10 13 m.
a certain day. What is this measurement in kilometers per
hour?
28. The following masses are written using metric
prefixes on the gram. Rewrite them in scientific notation 39. Tectonic plates are large segments of Earth’s crust that
in terms of the SI base unit of mass: the kilogram. For move slowly. Suppose one such plate has an average speed
example, 40 Mg would be written as 4 × 10 4 kg. (a) 23 of 4.0 cm/yr. (a) What distance does it move in 1.0 s at
mg; (b) 320 Tg; (c) 42 ng; (d) 7 g; (e) 9 Pg. this speed? (b) What is its speed in kilometers per million
years?
29. The following masses are given in kilograms. Use
metric prefixes on the gram to rewrite them so the 40. The average distance between Earth and the Sun is
numerical value is bigger than one but less than 1000. 1.5 × 10 11 m. (a) Calculate the average speed of Earth in
For example, 7 × 10 −4 kg could be written as 70 cg or its orbit (assumed to be circular) in meters per second. (b)
What is this speed in miles per hour?
700 mg. (a) 3.8 × 10 −5 kg; (b) 2.3 × 10 17 kg; (c)
2.4 × 10 −11 kg; (d) 8 × 10 15 kg; (e) 4.2 × 10 −3 kg. 41. The density of nuclear matter is about 1018 kg/m3.
Given that 1 mL is equal in volume to cm3, what is the
density of nuclear matter in megagrams per microliter (that
is, Mg/μL )?
1.4 Unit Conversion
30. The volume of Earth is on the order of 1021 m3.
(a) What is this in cubic kilometers (km3)? (b) What is it 42. The density of aluminum is 2.7 g/cm3. What is the
in cubic miles (mi3)? (c) What is it in cubic centimeters density in kilograms per cubic meter?
(cm3)?
43. A commonly used unit of mass in the English system
31. The speed limit on some interstate highways is is the pound-mass, abbreviated lbm, where 1 lbm = 0.454
roughly 100 km/h. (a) What is this in meters per second? kg. What is the density of water in pound-mass per cubic
(b) How many miles per hour is this? foot?

32. A car is traveling at a speed of 33 m/s. (a) What is its 44. A furlong is 220 yd. A fortnight is 2 weeks. Convert
speed in kilometers per hour? (b) Is it exceeding the 90 km/ a speed of one furlong per fortnight to millimeters per
h speed limit? second.

33. In SI units, speeds are measured in meters per second 45. It takes 2π radians (rad) to get around a circle, which
(m/s). But, depending on where you live, you’re probably is the same as 360°. How many radians are in 1°?
more comfortable of thinking of speeds in terms of either
kilometers per hour (km/h) or miles per hour (mi/h). In 46. Light travels a distance of about 3 × 10 8 m/s. A
this problem, you will see that 1 m/s is roughly 4 km/h
light-minute is the distance light travels in 1 min. If the Sun
or 2 mi/h, which is handy to use when developing your
physical intuition. More precisely, show that (a) is 1.5 × 10 11 m from Earth, how far away is it in light-
1.0 m/s = 3.6 km/h and (b) 1.0 m/s = 2.2 mi/h. minutes?

34. American football is played on a 100-yd-long field, 47. A light-nanosecond is the distance light travels in 1 ns.
excluding the end zones. How long is the field in meters? Convert 1 ft to light-nanoseconds.
(Assume that 1 m = 3.281 ft.)
44 Chapter 1 | Introducing Physics

48. An electron has a mass of 9.11 × 10 −31 kg. A proton


1.6 Estimates and Fermi Calculations
has a mass of 1.67 × 10 −27 kg. What is the mass of a
56. Assuming the human body is made primarily of water,
proton in electron-masses? estimate the volume of a person.

49. A fluid ounce is about 30 mL. What is the volume of a 57. Assuming the human body is primarily made of water,
12 fl-oz can of soda pop in cubic meters? estimate the number of molecules in it. (Note that water has
a molecular mass of 18 g/mol and there are roughly 1024
atoms in a mole.)
1.5 Dimensional Analysis
50. A student is trying to remember some formulas from 58. Estimate the mass of air in a classroom.
geometry. In what follows, assume A is area, V is
volume, and all other variables are lengths. Determine 59. Estimate the number of molecules that make up Earth,
which formulas are dimensionally consistent. (a) assuming an average molecular mass of 30 g/mol. (Note
V = πr 2 h; (b) A = 2πr 2 + 2πrh; (c) V = 0.5bh; (d) there are on the order of 1024 objects per mole.)

V = πd 2; (e) V = πd 3 /6. 60. Estimate the surface area of a person.

51. Consider the physical quantities s, v, a, and t with 61. Roughly how many solar systems would it take to tile
dimensions [s] = L, [v] = LT −1, [a] = LT −2, and the disk of the Milky Way?
[t] = T. Determine whether each of the following
62. (a) Estimate the density of the Moon. (b) Estimate the
equations is dimensionally consistent. (a) v 2 = 2as; (b) diameter of the Moon. (c) Given that the Moon subtends
s = vt 2 + 0.5at 2; (c) v = s/t; (d) a = v/t. at an angle of about half a degree in the sky, estimate its
distance from Earth.

52. Consider the physical quantities m, s, v, a, 63. The average density of the Sun is on the order 103 kg/
and t with dimensions [m] = M, [s] = L, [v] = LT–1, m3. (a) Estimate the diameter of the Sun. (b) Given that the
[a] = LT–2, and [t] = T. Assuming each of the following Sun subtends at an angle of about half a degree in the sky,
equations is dimensionally consistent, find the dimension estimate its distance from Earth.
of the quantity on the left-hand side of the equation: (a) F =
ma; (b) K = 0.5mv2; (c) p = mv; (d) W = mas; (e) L = mvr. 64. Estimate the mass of a virus.

53. Suppose quantity s is a length and quantity t is a 65. A floating-point operation is a single arithmetic
time. Suppose the quantities v and a are defined by v operation such as addition, subtraction, multiplication, or
= ds/dt and a = dv/dt. (a) What is the dimension of v? division. (a) Estimate the maximum number of floating-
(b) What is the dimension of the quantity a? What are the point operations a human being could possibly perform in
a lifetime. (b) How long would it take a supercomputer to
dimensions of (c) ∫ vdt, (d) ∫ adt, and (e) da/dt?
perform that many floating-point operations?

54. Suppose [V] = L3, [ρ] = ML –3, and [t] = T. (a)


1.7 Significant Figures
What is the dimension of ∫ ρdV ? (b) What is the 66. Consider the equation 4000/400 = 10.0. Assuming the
number of significant figures in the answer is correct, what
dimension of dV/dt? (c) What is the dimension of
can you say about the number of significant figures in 4000
ρ(dV/dt) ?
and 400?

55. The arc length formula says the length s of arc 67. Suppose your bathroom scale reads your mass as 65
subtended by angle Ɵ in a circle of radius r is given by kg with a 3% uncertainty. What is the uncertainty in your
the equation s = r Ɵ . What are the dimensions of (a) s, mass (in kilograms)?
(b) r, and (c) Ɵ?
68. A good-quality measuring tape can be off by 0.50 cm
over a distance of 20 m. What is its percent uncertainty?

69. An infant’s pulse rate is measured to be 130 ± 5 beats/

This OpenStax book is available for free at https://legacy.cnx.org/content/col12169/1.46


Chapter 1 | Introducing Physics 45

min. What is the percent uncertainty in this measurement? Assuming the same percent uncertainty, what is the
uncertainty in a blood pressure measurement of 80 mm
70. (a) Suppose that a person has an average heart rate of Hg?
72.0 beats/min. How many beats does he or she have in 2.0
years? (b) In 2.00 years? (c) In 2.000 years? 76. A person measures his or her heart rate by counting the
number of beats in 30 s. If 40 ± 1 beats are counted in 30.0
71. A can contains 375 mL of soda. How much is left after ± 0.5 s, what is the heart rate and its uncertainty in beats per
308 mL is removed? minute?

72. State how many significant figures are proper in the 77. What is the area of a circle 3.102 cm in diameter?
results of the following calculations: (a)
(106.7)(98.2) / (46.210)(1.01); (b) (18.7) 2; (c) 78. Determine the number of significant figures in the

following measurements: (a) 0.0009, (b) 15,450.0, (c)
⎝1.60 × 10 −19⎞⎠(3712) 6×103, (d) 87.990, and (e) 30.42.

73. (a) How many significant figures are in the numbers 79. Perform the following calculations and express your
99 and 100.? (b) If the uncertainty in each number is 1, answer using the correct number of significant digits. (a)
what is the percent uncertainty in each? (c) Which is a A woman has two bags weighing 13.5 lb and one bag
more meaningful way to express the accuracy of these two with a weight of 10.2 lb. What is the total weight of the
numbers: significant figures or percent uncertainties? bags? (b) The force F on an object is equal to its mass m
multiplied by its acceleration a. If a wagon with mass 55
kg accelerates at a rate of 0.0255 m/s2, what is the force on
74. (a) If your speedometer has an uncertainty of 2.0 km/h the wagon? (The unit of force is called the newton and it is
at a speed of 90 km/h, what is the percent uncertainty? (b) expressed with the symbol N.)
If it has the same percent uncertainty when it reads 60 km/
h, what is the range of speeds you could be going?

75. (a) A person’s blood pressure is measured to be


120 ± 2 mm Hg. What is its percent uncertainty? (b)

ADDITIONAL PROBLEMS
80. Consider the equation y = mt +b, where the dimension is the uncertainty in the average speed?
of y is length and the dimension of t is time, and m and b
are constants. What are the dimensions and SI units of (a) 84. The sides of a small rectangular box are measured to
m and (b) b? be 1.80 ± 0.1 cm, 2.05 ± 0.02 cm, and 3.1 ± 0.1 cm long.
Calculate its volume and uncertainty in cubic centimeters.
81. Consider the equation
s = s 0 + v 0 t + a 0 t /2 + j 0 t /6 + S 0 t /24 + ct 5 /120,
2 3 4
85. When nonmetric units were used in the United
Kingdom, a unit of mass called the pound-mass (lbm) was
where s is a length and t is a time. What are the dimensions
used, where 1 lbm = 0.4539 kg. (a) If there is an uncertainty
and SI units of (a) s 0, (b) v 0, (c) a 0, (d) j 0, (e) S 0,
of 0.0001 kg in the pound-mass unit, what is its percent
and (f) c? uncertainty? (b) Based on that percent uncertainty, what
mass in pound-mass has an uncertainty of 1 kg when
82. (a) A car speedometer has a 5% uncertainty. What is converted to kilograms?
the range of possible speeds when it reads 90 km/h? (b)
Convert this range to miles per hour. Note 1 km = 0.6214 86. The length and width of a rectangular room are
mi. measured to be 3.955 ± 0.005 m and 3.050 ± 0.005 m.
Calculate the area of the room and its uncertainty in square
83. A marathon runner completes a 42.188-km course in 2 meters.
h, 30 min, and 12 s. There is an uncertainty of 25 m in the
distance traveled and an uncertainty of 1 s in the elapsed 87. A car engine moves a piston with a circular cross-
time. (a) Calculate the percent uncertainty in the distance. section of 7.500 ± 0.002 cm in diameter a distance of
(b) Calculate the percent uncertainty in the elapsed time. (c) 3.250 ± 0.001 cm to compress the gas in the cylinder. (a)
What is the average speed in meters per second? (d) What By what amount is the gas decreased in volume in cubic
46 Chapter 1 | Introducing Physics

centimeters? (b) Find the uncertainty in this volume.

CHALLENGE PROBLEMS
88. The first atomic bomb was detonated on July 16, 1945, of TNT” (abbreviated “t TNT”), where 1 t TNT is about
at the Trinity test site about 200 mi south of Los Alamos. 4.2 GJ. Convert your answer to (b) into kilotons of TNT
In 1947, the U.S. government declassified a film reel of the (that is, kt TNT). Compare your answer with the quick-
explosion. From this film reel, British physicist G. I. Taylor and-dirty estimate of 10 kt TNT made by physicist Enrico
was able to determine the rate at which the radius of the Fermi shortly after witnessing the explosion from what was
fireball from the blast grew. Using dimensional analysis, he thought to be a safe distance. (Reportedly, Fermi made his
was then able to deduce the amount of energy released in estimate by dropping some shredded bits of paper right
the explosion, which was a closely guarded secret at the before the remnants of the shock wave hit him and looked
time. Because of this, Taylor did not publish his results until to see how far they were carried by it.)
1950. This problem challenges you to recreate this famous
calculation. (a) Using keen physical insight developed from 89. The purpose of this problem is to show the entire
years of experience, Taylor decided the radius r of the concept of dimensional consistency can be summarized
fireball should depend only on time since the explosion, by the old saying “You can’t add apples and oranges.” If
t, the density of the air, ρ, and the energy of the initial you have studied power series expansions in a calculus
explosion, E. Thus, he made the educated guess that course, you know the standard mathematical functions such
r = kE a ρ b t c for some dimensionless constant k and some as trigonometric functions, logarithms, and exponential
functions can be expressed as infinite sums of the form
unknown exponents a, b, and c. Given that [E] = ML2T–2, ∞
determine the values of the exponents necessary to make ∑ an xn = a0 + a1 x + a2 x2 + a3 x3 + ⋯ , where
this equation dimensionally consistent. (Hint: Notice the n=0
equation implies that k = rE −a ρ −b t −c and that [k] = 1. the an are dimensionless constants for all
) (b) By analyzing data from high-energy conventional n = 0, 1, 2, ⋯ and x is the argument of the function.
explosives, Taylor found the formula he derived seemed to (If you have not studied power series in calculus yet, just
be valid as long as the constant k had the value 1.03. From trust us.) Use this fact to explain why the requirement
the film reel, he was able to determine many values of r that all terms in an equation have the same dimensions is
and the corresponding values of t. For example, he found sufficient as a definition of dimensional consistency. That
that after 25.0 ms, the fireball had a radius of 130.0 m. is, it actually implies the arguments of standard
Use these values, along with an average air density of 1.25 mathematical functions must be dimensionless, so it is not
kg/m3, to calculate the initial energy release of the Trinity really necessary to make this latter condition a separate
detonation in joules (J). (Hint: To get energy in joules, requirement of the definition of dimensional consistency as
you need to make sure all the numbers you substitute in we have done in this section.
are expressed in terms of SI base units.) (c) The energy
released in large explosions is often cited in units of “tons

This OpenStax book is available for free at https://legacy.cnx.org/content/col12169/1.46


Chapter 2 | The Universe at Its Limits 47

2 | THE UNIVERSE AT ITS


LIMITS
Chapter Outline
2.1 Consequences of Light Travel Time
2.2 A Tour of the Universe
2.3 The Universe on the Large Scale
2.4 The Universe of the Very Small
2.5 A Conclusion and a Beginning

Introduction
In astronomy we deal with distances on a scale you may never have thought about before, with numbers larger than any
you may have encountered. We adopt two approaches that make dealing with astronomical numbers a little bit easier. First,
we use a system for writing large and small numbers called scientific notation (or sometimes powers-of-ten notation). This
system is very appealing because it eliminates the many zeros that can seem overwhelming to the reader. In scientific
notation, if you want to write a number such as 500,000,000, you express it as 5 × 108. The small raised number after
the 10, called an exponent, keeps track of the number of places we had to move the decimal point to the left to convert
500,000,000 to 5. If you are encountering this system for the first time or would like a refresher, we suggest you look back
at Section 1.3 for more information. The second way we try to keep numbers simple is to use a consistent set of units—the
metric International System of Units, or SI (from the French Système International d’Unités). The SI system is summarized
in Appendix A.

Watch this brief PBS animation (https://openstax.org/l/30scinotation) that explains how scientific notation
works and why it’s useful.

A common unit astronomers use to describe distances in the universe is a light-year, which is the distance light travels
during one year. Because light always travels at the same speed, and because its speed turns out to be the fastest possible
speed in the universe, it makes a good standard for keeping track of distances. You might be confused because a “light-year”
seems to imply that we are measuring time, but this mix-up of time and distance is common in everyday life as well. For
example, when your friend asks where the movie theater is located, you might say “about 20 minutes from downtown.”
So, how many kilometers are there in a light-year? Light travels at the amazing pace of 3 × 105 kilometers per second (km/
s), which makes a light-year 9.46 × 1012 kilometers. You might think that such a large unit would reach the nearest star
easily, but the stars are far more remote than our imaginations might lead us to believe. Even the nearest star is 4.3 light-
years away—more than 40 trillion kilometers. Other stars visible to the unaided eye are hundreds to thousands of light-years
away (Figure 2.1).
48 Chapter 2 | The Universe at Its Limits

Figure 2.1 This beautiful cloud of cosmic raw material (gas


and dust from which new stars and planets are being made)
called the Orion Nebula is about 1400 light-years away. That’s a
distance of roughly 1.34 × 1016 kilometers—a pretty big
number. The gas and dust in this region are illuminated by the
intense light from a few extremely energetic adolescent stars.
(credit: NASA, ESA, M. Robberto (Space Telescope Science
Institute/ESA) and the Hubble Space Telescope Orion Treasury
Project Team)

Example 2.1.

Scientific Notation
In 2015, the richest human being on our planet had a net worth of $79.2 billion. Some might say this is an
astronomical sum of money. Express this amount in scientific notation.
Solution
$79.2 billion can be written $79,200,000,000. Expressed in scientific notation it becomes $7.92 × 1010.

Example 2.2.

Getting Familiar with a Light-Year


How many kilometers are there in a light-year?
Solution
The speed of light, c = 3 × 108 m/s. That is to say, light travels 3 × 108 m in 1 s. So, let’s calculate how far it goes
in a year. That's the meaning of "one light-year."
We'll use the same "Power of 1" idea from Chapter 1, i.e. that multiplying a quantity by any fraction whose value
is precisely 1.0 does not change the quantity.
3×10 8 m × ⎛ 60 s ⎞ × ⎛60 min ⎞ × ⎛ 24 hr ⎞ × ⎛365.24 day ⎞ = 9.46×10 15 m
1s ⎝1 min ⎠ ⎝ 1 hr ⎠ ⎝1 day ⎠ ⎝ 1 yr ⎠ 1 yr

And, to answer the original question precisely, we will convert that distance into units of km. The final unit
conversion gives us:

This OpenStax book is available for free at https://legacy.cnx.org/content/col12169/1.46


Chapter 2 | The Universe at Its Limits 49

⎛ ⎞
9.46×10 15 m × ⎝ 1 km ⎠ = 9.46×10 12 km
1000 m
That’s almost 10,000,000,000,000 km that light covers in a year. To help you imagine how long this distance is,
we’ll mention that a string 1 light-year long could fit around the circumference of Earth 236 million times.

2.1 | Consequences of Light Travel Time


Learning Objectives
After completing this section, you should be able to:
• Explain how looking farther out in space is also looking farther back in time

There is another reason the speed of light is such a natural unit of distance for astronomers. Information about the universe
comes to us almost exclusively through various forms of light, and all such light travels at the speed of light—that is, 1
light-year every year. This sets a limit on how quickly we can learn about events in the universe. If a star is 100 light-years
away, the light we see from it tonight left that star 100 years ago and is just now arriving in our neighborhood. The soonest
we can learn about any changes in that star is 100 years after the fact. For a star 500 light-years away, the light we detect
tonight left 500 years ago and is carrying 500-year-old news.
Because many of us are accustomed to instant news from the Internet, some might find this frustrating.
“You mean, when I see that star up there,” you ask, “I won’t know what’s actually happening there for another 500 years?”
But this isn’t the most helpful way to think about the situation. For astronomers, now is when the light reaches us here on
Earth. There is no way for us to know anything about that star (or other object) until its light reaches us.
But what at first may seem a great frustration is actually a tremendous benefit in disguise. If astronomers really want to
piece together what has happened in the universe since its beginning, they must find evidence about each epoch (or period
of time) of the past. Where can we find evidence today about cosmic events that occurred billions of years ago?
The delay in the arrival of light provides an answer to this question. The farther out in space we look, the longer the light has
taken to get here, and the longer ago it left its place of origin. By looking billions of light-years out into space, astronomers
are actually seeing billions of years into the past. In this way, we can reconstruct the history of the cosmos and get a sense
of how it has evolved over time.
This is one reason why astronomers strive to build telescopes that can collect more and more of the faint light in the
universe. The more light we collect, the fainter the objects we can observe. On average, fainter objects are farther away
and can, therefore, tell us about periods of time even deeper in the past. Instruments such as the Hubble Space Telescope
(Figure 2.2) and the Very Large Telescope in Chile (which you will learn about in the chapter on Astronomical
Instruments (https://legacy.cnx.org/content/m59802/latest/) ), are giving astronomers views of deep space and
deep time better than any we have had before.
50 Chapter 2 | The Universe at Its Limits

Figure 2.2 The Hubble Space Telescope, shown here in orbit


around Earth, is one of many astronomical instruments in space.
(credit: modification of work by European Space Agency)

2.2 | A Tour of the Universe


Learning Objectives
By the end of this section you will be able to:
• Know the definition of one astronomical unit (AU).
• Explain the distance scales involved as we move outward from our planet through the solar
system, to other stars in the Milky Way, and on to other galaxies.
• Describe the major bodies that make up our solar system.
• Describe the structure of the Milky Way galaxy.

We can now take a brief introductory tour of the universe as astronomers understand it today to get acquainted with the
types of objects and distances you will encounter throughout the text. We begin at home with Earth, a nearly spherical planet
about 13,000 kilometers in diameter (Figure 2.3). A space traveler entering our planetary system would easily distinguish
Earth from the other planets in our solar system by the large amount of liquid water that covers some two thirds of its crust.
If the traveler had equipment to receive radio or television signals, or came close enough to see the lights of our cities at
night, she would soon find signs that this watery planet has sentient life.

This OpenStax book is available for free at https://legacy.cnx.org/content/col12169/1.46


Chapter 2 | The Universe at Its Limits 51

Figure 2.3 This image shows the Western hemisphere as viewed from space 35,400 kilometers (about 22,000 miles) above
Earth. Data about the land surface from one satellite was combined with another satellite’s data about the clouds to create the
image. (credit: modification of work by R. Stockli, A. Nelson, F. Hasler, NASA/ GSFC/ NOAA/ USGS)

Our nearest astronomical neighbor is Earth’s satellite, commonly called the Moon. Figure 2.4 shows Earth and the Moon
drawn to scale on the same diagram. Notice how small we have to make these bodies to fit them on the page with the right
scale. The Moon’s distance from Earth is about 30 times Earth’s diameter, or approximately 384,000 kilometers, and it takes
about a month for the Moon to revolve around Earth. The Moon’s diameter is 3476 kilometers, about one fourth the size of
Earth.

Figure 2.4 This image shows Earth and the Moon shown to scale for both size and distance. (credit: modification of work by
NASA)

Light (or radio waves) takes 1.3 seconds to travel between Earth and the Moon. If you’ve seen videos of the Apollo flights
to the Moon, you may recall that there was a delay of about 3 seconds between the time Mission Control asked a question
and the time the astronauts responded. This was not because the astronomers were thinking slowly, but rather because it
took the radio waves almost 3 seconds to make the round trip.
Earth revolves around our star, the Sun, which is about 150 million kilometers away—approximately 400 times as far away
from us as the Moon. We call the average Earth–Sun distance an astronomical unit (AU) because, in the early days of
astronomy, it was the most important measuring standard. Light takes slightly more than 8 minutes to travel 1 astronomical
unit, which means the latest news we receive from the Sun is always 8 minutes old. The diameter of the Sun is about 1.5
million kilometers; Earth could fit comfortably inside one of the minor eruptions that occurs on the surface of our star. If
the Sun were reduced to the size of a basketball, Earth would be a small apple seed about 30 meters from the ball.
It takes Earth 1 year (3 × 107 seconds) to go around the Sun at our distance; to make it around, we must travel at
approximately 110,000 kilometers per hour. (If you, like many students, still prefer miles to kilometers, you might find the
following trick helpful. To convert kilometers to miles, just multiply kilometers by 0.6. Thus, 110,000 kilometers per hour
becomes 66,000 miles per hour.) Because gravity holds us firmly to Earth and there is no resistance to Earth’s motion in the
vacuum of space, we participate in this extremely fast-moving trip without being aware of it day to day.
Earth is only one of eight planets that revolve around the Sun. These planets, along with their moons and swarms of smaller
bodies such as dwarf planets, make up the solar system (Figure 2.5). A planet is defined as a body of significant size that
orbits a star and does not produce its own light. (If a large body consistently produces its own light, it is then called a star.)
Later in the book this definition will be modified a bit, but it is perfectly fine for now as you begin your voyage.
52 Chapter 2 | The Universe at Its Limits

Figure 2.5 The Sun, the planets, and some dwarf planets are shown with their sizes drawn to scale. The orbits of the planets are
much more widely separated than shown in this drawing. Notice the size of Earth compared to the giant planets. (credit:
modification of work by NASA)

We are able to see the nearby planets in our skies only because they reflect the light of our local star, the Sun. If the planets
were much farther away, the tiny amount of light they reflect would usually not be visible to us. The planets we have so
far discovered orbiting other stars were found from the pull their gravity exerts on their parent stars, or from the light they
block from their stars when they pass in front of them. We can’t see most of these planets directly, although a few are now
being imaged directly.
The Sun is our local star, and all the other stars are also enormous balls of glowing gas that generate vast amounts of energy
by nuclear reactions deep within. We will discuss the processes that cause stars to shine in more detail later in the book. The
other stars look faint only because they are so very far away. If we continue our basketball analogy, Proxima Centauri, the
nearest star beyond the Sun, which is 4.3 light-years away, would be almost 7000 kilometers from the basketball.
When you look up at a star-filled sky on a clear night, all the stars visible to the unaided eye are part of a single collection of
stars we call the Milky Way Galaxy, or simply the Galaxy. (When referring to the Milky Way, we capitalize Galaxy; when
talking about other galaxies of stars, we use lowercase galaxy.) The Sun is one of hundreds of billions of stars that make
up the Galaxy; its extent, as we will see, staggers the human imagination. Within a sphere 10 light-years in radius centered
on the Sun, we find roughly ten stars. Within a sphere 100 light-years in radius, there are roughly 10,000 (104) stars—far
too many to count or name—but we have still traversed only a tiny part of the Milky Way Galaxy. Within a 1000-light-year
sphere, we find some ten million (107) stars; within a sphere of 100,000 light-years, we finally encompass the entire Milky
Way Galaxy.
Our Galaxy looks like a giant disk with a small ball in the middle. If we could move outside our Galaxy and look down
on the disk of the Milky Way from above, it would probably resemble the galaxy in Figure 2.6, with its spiral structure
outlined by the blue light of hot adolescent stars.

This OpenStax book is available for free at https://legacy.cnx.org/content/col12169/1.46


Chapter 2 | The Universe at Its Limits 53

Figure 2.6 This galaxy of billions of stars, called by its


catalog number NGC 1073, is thought to be similar to our own
Milky Way Galaxy. Here we see the giant wheel-shaped system
with a bar of stars across its middle. (credit: NASA, ESA)

The Sun is somewhat less than 30,000 light-years from the center of the Galaxy, in a location with nothing much to
distinguish it. From our position inside the Milky Way Galaxy, we cannot see through to its far rim (at least not with
ordinary light) because the space between the stars is not completely empty. It contains a sparse distribution of gas (mostly
the simplest element, hydrogen) intermixed with tiny solid particles that we call interstellar dust. This gas and dust collect
into enormous clouds in many places in the Galaxy, becoming the raw material for future generations of stars. Figure 2.7
shows an image of the disk of the Galaxy as seen from our vantage point.
54 Chapter 2 | The Universe at Its Limits

Figure 2.7 Because we are inside the Milky Way Galaxy, we see its disk in cross-section flung
across the sky like a great milky white avenue of stars with dark “rifts” of dust. In this dramatic
image, part of it is seen above Trona Pinnacles in the California desert. (credit: Ian Norman)

Typically, the interstellar material is so extremely sparse that the space between stars is a much better vacuum than anything
we can produce in terrestrial laboratories. Yet, the dust in space, building up over thousands of light-years, can block the
light of more distant stars. Like the distant buildings that disappear from our view on a smoggy day in Los Angeles, the
more distant regions of the Milky Way cannot be seen behind the layers of interstellar smog. Luckily, astronomers have
found that stars and raw material shine with various forms of light, some of which do penetrate the smog, and so we have
been able to develop a pretty good map of the Galaxy.
Recent observations, however, have also revealed a rather surprising and disturbing fact. There appears to be more—much
more—to the Galaxy than meets the eye (or the telescope). From various investigations, we have evidence that much of our
Galaxy is made of material we cannot currently observe directly with our instruments. We therefore call this component of
the Galaxy dark matter. We know the dark matter is there by the pull its gravity exerts on the stars and raw material we can
observe, but what this dark matter is made of and how much of it exists remain a mystery. Furthermore, this dark matter is
not confined to our Galaxy; it appears to be an important part of other star groupings as well.
By the way, not all stars live by themselves, as the Sun does. Many are born in double or triple systems with two, three,
or more stars revolving about each other. Because the stars influence each other in such close systems, multiple stars
allow us to measure characteristics that we cannot discern from observing single stars. In a number of places, enough stars
have formed together that we recognized them as star clusters (Figure 2.8). Some of the largest of the star clusters that
astronomers have cataloged contain hundreds of thousands of stars and take up volumes of space hundreds of light-years
across.

This OpenStax book is available for free at https://legacy.cnx.org/content/col12169/1.46


Chapter 2 | The Universe at Its Limits 55

Figure 2.8 This large star cluster is known by its catalog


number, M9. It contains some 250,000 stars and is seen more
clearly from space using the Hubble Space Telescope. It is
located roughly 25,000 light-years away. (credit: NASA, ESA)

You may hear stars referred to as “eternal,” but in fact no star can last forever. Since the “business” of stars is making
energy, and energy production requires some sort of fuel to be used up, eventually all stars run out of fuel. This news should
not cause you to panic, though, because our Sun still has at least 5 or 6 billion years to go. Ultimately, the Sun and all stars
will die, and it is in their death throes that some of the most intriguing and important processes of the universe are revealed.
For example, we now know that many of the atoms in our bodies were once inside stars. These stars exploded at the ends of
their lives, recycling their material back into the reservoir of the Galaxy. In this sense, all of us are literally made of recycled
“star dust.”

2.3 | The Universe on the Large Scale


Learning Objectives
By the end of this section you will be able to:
• Name the galaxies nearest to our own, along with their approximate distances from Earth.
• Explain what is meant by a galactic cluster.
• Explain what is meant by a quasar.

In a very rough sense, you could think of the solar system as your house or apartment and the Galaxy as your town, made up
of many houses and buildings. In the twentieth century, astronomers were able to show that, just as our world is made up of
many, many towns, so the universe is made up of enormous numbers of galaxies. (We define the universe to be everything
that exists that is accessible to our observations.) Galaxies stretch as far into space as our telescopes can see, many billions
of them within the reach of modern instruments. When they were first discovered, some astronomers called galaxies island
universes, and the term is aptly descriptive; galaxies do look like islands of stars in the vast, dark seas of intergalactic space.
The nearest galaxy, discovered in 1993, is a small one that lies 75,000 light-years from the Sun in the direction of the
constellation Sagittarius, where the smog in our own Galaxy makes it especially difficult to discern. (A constellation, we
should note, is one of the 88 sections into which astronomers divide the sky, each named after a prominent star pattern
within it.) Beyond this Sagittarius dwarf galaxy lie two other small galaxies, about 160,000 light-years away. First recorded
by Magellan’s crew as he sailed around the world, these are called the Magellanic Clouds (Figure 2.9). All three of these
small galaxies are satellites of the Milky Way Galaxy, interacting with it through the force of gravity. Ultimately, all three
may even be swallowed by our much larger Galaxy, as other small galaxies have been over the course of cosmic time.
56 Chapter 2 | The Universe at Its Limits

Figure 2.9 This image shows both the Large Magellanic Cloud and the Small Magellanic Cloud above the telescopes of the
Atacama Large Millimeter/Submillimeter Array (ALMA) in the Atacama Desert of northern Chile. (credit: ESO, C. Malin)

The nearest large galaxy is a spiral quite similar to our own, located in the constellation of Andromeda, and is thus called
the Andromeda galaxy; it is also known by one of its catalog numbers, M31 (Figure 2.10). M31 is a little more than 2
million light-years away and, along with the Milky Way, is part of a small cluster of more than 50 galaxies referred to as the
Local Group.

This OpenStax book is available for free at https://legacy.cnx.org/content/col12169/1.46


Chapter 2 | The Universe at Its Limits 57

Figure 2.10 The Andromeda galaxy (M31) is a spiral-shaped collection of stars similar to our own Milky Way. (credit: Adam
Evans)

At distances of 10 to 15 million light-years, we find other small galaxy groups, and then at about 50 million light-years
there are more impressive systems with thousands of member galaxies. We have discovered that galaxies occur mostly in
clusters, both large and small (Figure 2.11).
58 Chapter 2 | The Universe at Its Limits

Figure 2.11 In this image, you can see part of a cluster of galaxies located about 60 million
light-years away in the constellation of Fornax. All the objects that are not pinpoints of light in
the picture are galaxies of billions of stars. (credit: ESO, J. Emerson, VISTA. Acknowledgment:
Cambridge Astronomical Survey Unit)

Some of the clusters themselves form into larger groups called superclusters. The Local Group is part of a supercluster of
galaxies, called the Virgo Supercluster, which stretches over a diameter of 110 million light-years. We are just beginning to
explore the structure of the universe at these enormous scales and are already encountering some unexpected findings.
At even greater distances, where many ordinary galaxies are too dim to see, we find quasars. These are brilliant centers of
galaxies, glowing with the light of an extraordinarily energetic process. The enormous energy of the quasars is produced by
gas that is heated to a temperature of millions of degrees as it falls toward a massive black hole and swirls around it. The
brilliance of quasars makes them the most distant beacons we can see in the dark oceans of space. They allow us to probe
the universe 10 billion light-years away or more, and thus 10 billion years or more in the past.
With quasars we can see way back close to the Big Bang explosion that marks the beginning of time. Beyond the quasars
and the most distant visible galaxies, we have detected the feeble glow of the explosion itself, filling the universe and thus
coming to us from all directions in space. The discovery of this “afterglow of creation” is considered to be one of the most
significant events in twentieth-century science, and we are still exploring the many things it has to tell us about the earliest
times of the universe.
Measurements of the properties of galaxies and quasars in remote locations require large telescopes, sophisticated light-
amplifying devices, and painstaking labor. Every clear night, at observatories around the world, astronomers and students
are at work on such mysteries as the birth of new stars and the large-scale structure of the universe, fitting their results into
the tapestry of our understanding.

This OpenStax book is available for free at https://legacy.cnx.org/content/col12169/1.46


Chapter 2 | The Universe at Its Limits 59

2.4 | The Universe of the Very Small


Learning Objectives
By the end of this section, you will be able to:
• Explain what is meant by an element.
• Name the most abundant elements in the universe.
• Explain the difference between atoms and molecules.

The foregoing discussion has likely impressed on you that the universe is extraordinarily large and extraordinarily empty.
On average, it is 10,000 times more empty than our Galaxy. Yet, as we have seen, even the Galaxy is mostly empty space.
The air we breathe has about 1019 atoms in each cubic centimeter—and we usually think of air as empty space. In the
interstellar gas of the Galaxy, there is about one atom in every cubic centimeter. Intergalactic space is filled so sparsely that
to find one atom, on average, we must search through a cubic meter of space. Most of the universe is fantastically empty;
places that are dense, such as the human body, are tremendously rare.
Even our most familiar solids are mostly space. If we could take apart such a solid, piece by piece, we would eventually
reach the tiny molecules from which it is formed. Molecules are the smallest particles into which any matter can be divided
while still retaining its chemical properties. A molecule of water (H2O), for example, consists of two hydrogen atoms and
one oxygen atom bonded together.
Molecules, in turn, are built of atoms, which are the smallest particles of an element that can still be identified as that
element. For example, an atom of gold is the smallest possible piece of gold. Nearly 100 different kinds of atoms (elements)
exist in nature. Most of them are rare, and only a handful account for more than 99% of everything with which we come in
contact. The most abundant elements in the cosmos today are listed in Table 2.1; think of this table as the “greatest hits”
of the universe when it comes to elements.

The Cosmically Abundant Elements


Element[1] Symbol Number of Atoms per
Million Hydrogen Atoms
Hydrogen H 1,000,000
Helium He 80,000
Carbon C 450
Nitrogen N 92
Oxygen O 740
Neon Ne 130
Magnesium Mg 40
Silicon Si 37
Sulfur S 19
Iron Fe 32

Table 2.1

All atoms consist of a central, positively charged nucleus surrounded by negatively charged electrons. The bulk of the matter
in each atom is found in the nucleus, which consists of positive protons and electrically neutral neutrons all bound tightly
together in a very small space. Each element is defined by the number of protons in its atoms. Thus, any atom with 6 protons
in its nucleus is called carbon, any with 50 protons is called tin, and any with 70 protons is called ytterbium. (For a list of
the elements, see Appendix F.)

1. This list of elements is arranged in order of the atomic number, which is the number of protons in each
nucleus.
60 Chapter 2 | The Universe at Its Limits

The distance from an atomic nucleus to its electrons is typically 100,000 times the size of the nucleus itself. This is why
we say that even solid matter is mostly space. The typical atom is far emptier than the solar system out to Neptune. (The
distance from Earth to the Sun, for example, is only 100 times the size of the Sun.) This is one reason atoms are not like
miniature solar systems.
Remarkably, physicists have discovered that everything that happens in the universe, from the smallest atomic nucleus to
the largest superclusters of galaxies, can be explained through the action of only four forces: gravity, electromagnetism
(which combines the actions of electricity and magnetism), and two forces that act at the nuclear level. The fact that there
are four forces (and not a million, or just one) has puzzled physicists and astronomers for many years and has led to a quest
for a unified picture of nature.

To construct an atom, particle by particle, check out this guided animation (https://openstax.org/l/
30buildanatom) for building an atom.

2.5 | A Conclusion and a Beginning


Learning Objectives
By the end of this section you will be able to:
• Using the analogy of compressing the age of the Universe into one calendar year, state the
major events in its history and roughly when each one occurred.

If you are new to astronomy, you have probably reached the end of our brief tour in this chapter with mixed emotions. On
the one hand, you may be fascinated by some of the new ideas you’ve read about and you may be eager to learn more. On
the other hand, you may be feeling a bit overwhelmed by the number of topics we have covered, and the number of new
words and ideas we have introduced. Learning astronomy is a little like learning a new language: at first it seems there are
so many new expressions that you’ll never master them all, but with practice, you soon develop facility with them.
At this point you may also feel a bit small and insignificant, dwarfed by the cosmic scales of distance and time. But, there
is another way to look at what you have learned from our first glimpses of the cosmos. Let us consider the history of the
universe from the Big Bang to today and compress it, for easy reference, into a single year. (We have borrowed this idea
from Carl Sagan’s 1997 Pulitzer Prize-winning book, The Dragons of Eden.)
On this scale, the Big Bang happened at the first moment of January 1, and this moment, when you are reading this chapter
would be the end of the very last second of December 31. When did other events in the development of the universe happen
in this “cosmic year?” Our solar system formed around September 10, and the oldest rocks we can date on Earth go back to
the third week in September (Figure 2.12).

This OpenStax book is available for free at https://legacy.cnx.org/content/col12169/1.46


Chapter 2 | The Universe at Its Limits 61

Figure 2.12 On a cosmic calendar, where the time since the Big Bang is compressed into 1 year, creatures we would call
human do not emerge on the scene until the evening of December 31. (credit: February: modification of work by NASA, JPL-
Caltech, W. Reach (SSC/Caltech); March: modification of work by ESA, Hubble and NASA, Acknowledgement: Giles
Chapdelaine; April: modification of work by NASA, ESA, CFHT, CXO, M.J. Jee (University of California, Davis), A. Mahdavi
(San Francisco State University); May: modification of work by NASA, JPL-Caltech; June: modification of work by NASA/
ESA; July: modification of work by NASA, JPL-Caltech, Harvard-Smithsonian; August: modification of work by NASA, JPL-
Caltech, R. Hurt (SSC-Caltech); September: modification of work by NASA; October: modification of work by NASA;
November: modification of work by Dénes Emőke)

Where does the origin of human beings fall during the course of this cosmic year? The answer turns out to be the evening
of December 31. The invention of the alphabet doesn’t occur until the fiftieth second of 11:59 p.m. on December 31. And
the beginnings of modern astronomy are a mere fraction of a second before the New Year. Seen in a cosmic context, the
amount of time we have had to study the stars is minute, and our success in piecing together as much of the story as we
have is remarkable.
Certainly our attempts to understand the universe are not complete. As new technologies and new ideas allow us to gather
more and better data about the cosmos, our present picture of astronomy will very likely undergo many changes. Still, as
you read our current progress report on the exploration of the universe, take a few minutes every once in a while just to
savor how much you have already learned.

For Further Exploration


Books
Miller, Ron, and William Hartmann. The Grand Tour: A Traveler’s Guide to the Solar System. 3rd ed. Workman, 2005. This
volume for beginners is a colorfully illustrated voyage among the planets.
Sagan, Carl. Cosmos. Ballantine, 2013 [1980]. This tome presents a classic overview of astronomy by an astronomer who
had a true gift for explaining things clearly. (You can also check out Sagan’s television series Cosmos: A Personal Voyage
and Neil DeGrasse Tyson’s current series Cosmos: A Spacetime Odyssey.)
Tyson, Neil DeGrasse, and Don Goldsmith. Origins: Fourteen Billion Years of Cosmic Evolution. Norton, 2004. This book
provides a guided tour through the beginnings of the universe, galaxies, stars, planets, and life.
Websites
If you enjoyed the beautiful images in this chapter (and there are many more fabulous photos to come in other chapters), you
may want to know where you can obtain and download such pictures for your own enjoyment. (Many astronomy images
62 Chapter 2 | The Universe at Its Limits

are from government-supported instruments or projects, paid for by tax dollars, and therefore are free of copyright laws.)
Here are three resources we especially like:
• Astronomy Picture of the Day: apod.nasa.gov/apod/astropix.html. Two space scientists scour the Internet and select
one beautiful astronomy image to feature each day. Their archives range widely, from images of planets and nebulae
to rockets and space instruments; they also have many photos of the night sky. The search function (see the menu
on the bottom of the page) works quite well for finding something specific among the many years’ worth of daily
images.
• Hubble Space Telescope Images: www.hubblesite.org/newscenter/archive/browse/images. Starting at this page, you
can select from among hundreds of Hubble pictures by subject or by date. Note that many of the images have
supporting pictures with them, such as diagrams, animations, or comparisons. Excellent captions and background
information are provided. Other ways to approach these images are through the more public-oriented Hubble
Gallery (www.hubblesite.org/gallery) and the European homepage (www.spacetelescope.org/images).
• National Aeronautics and Space Administration’s (NASA’s) Planetary Photojournal: photojournal.jpl.nasa.gov. This
site features thousands of images from planetary exploration, with captions of varied length. You can select images
by world, feature name, date, or catalog number, and download images in a number of popular formats. However,
only NASA mission images are included. Note the Photojournal Search option on the menu at the top of the
homepage to access ways to search their archives.
Videos
Cosmic Voyage: www.youtube.com/watch?v=qxXf7AJZ73A. This video presents a portion of Cosmic Voyage, narrated by
Morgan Freeman (8:34).
Powers of Ten: www.youtube.com/watch?v=0fKBhvDjuy0. This classic short video is a much earlier version of Powers of
Ten, narrated by Philip Morrison (9:00).
The Known Universe: www.youtube.com/watch?v=17jymDn0W6U. This video tour from the American Museum of
Natural History has realistic animation, music, and captions (6:30).
Wanderers: apod.nasa.gov/apod/ap141208.html. This video provides a tour of the solar system, with narrative by Carl
Sagan, imagining other worlds with dramatically realistic paintings (3:50).

This OpenStax book is available for free at https://legacy.cnx.org/content/col12169/1.46


Chapter 3 | Straight-Line Motion 63

3 | STRAIGHT-LINE MOTION

Figure 3.1 A JR Central L0 series five-car maglev (magnetic levitation) train undergoing a test run on the Yamanashi Test
Track. The maglev train’s motion can be described using kinematics, the subject of this chapter. (credit: modification of work by
“Maryland GovPics”/Flickr)

Chapter Outline
3.1 Position, Displacement and Average Velocity
3.2 Instantaneous Velocity and Speed
3.3 Average and Instantaneous Acceleration
3.4 Motion with Constant Acceleration
3.5 Free Fall

Introduction
Our universe is full of objects in motion. From the stars, planets, and galaxies; to the motion of people and animals; down
to the microscopic scale of atoms and molecules—everything in our universe is in motion. We can describe motion using
the two disciplines of kinematics and dynamics. We will eventually study dynamics, which is concerned with the causes of
motion, in Newton’s Laws of Motion (https://legacy.cnx.org/content/m58294/latest/) ; but, there is much to be
learned about motion without referring to what causes it, and this is the study of kinematics. Kinematics involves describing
motion through properties such as position, time, velocity, and acceleration.
A full treatment of kinematics considers motion in two and three dimensions. For now, we discuss motion in one dimension,
which provides us with the tools necessary to study multidimensional motion. A good example of an object undergoing one-
dimensional motion is the maglev (magnetic levitation) train depicted at the beginning of this chapter. As it travels, say, from
Tokyo to Kyoto, it is at different positions along the track at various times in its journey, and therefore has displacements,
or changes in position. It also has a variety of velocities along its path and it undergoes accelerations (changes in velocity).
With the skills learned in this chapter we can calculate these quantities and average velocity. All these quantities can be
described using kinematics, without knowing the train’s mass or the forces involved.
64 Chapter 3 | Straight-Line Motion

3.1 | Position, Displacement and Average Velocity


Learning Objectives
By the end of this section, you will be able to:
• Define position, displacement, and distance traveled.
• Calculate the total displacement given the position as a function of time.
• Determine the total distance traveled.
• Calculate the average velocity given the displacement and elapsed time.

When you’re in motion, the basic questions to ask are: Where are you? Where are you going? How fast are you getting
there? The answers to these questions require that you specify your position, your displacement, and your average
velocity—the terms we define in this section.

Position
To describe the motion of an object, you must first be able to describe its position (x): where it is at any particular time.
More precisely, we need to specify its position relative to a convenient frame of reference. A frame of reference is an
arbitrary set of axes from which the position and motion of an object are described. Earth is often used as a frame of
reference, and we often describe the position of an object as it relates to stationary objects on Earth. For example, a rocket
launch could be described in terms of the position of the rocket with respect to Earth as a whole, whereas a cyclist’s position
could be described in terms of where she is in relation to the buildings she passes Figure 3.2. In other cases, we use
reference frames that are not stationary but are in motion relative to Earth. To describe the position of a person in an airplane,
for example, we use the airplane, not Earth, as the reference frame. To describe the position of an object undergoing one-
dimensional motion, we often use the variable x. Later in the chapter, during the discussion of free fall, we use the variable
y.

Figure 3.2 These cyclists in Vietnam can be described by


their position relative to buildings or a canal. Their motion can
be described by their change in position, or displacement, in a
frame of reference. (credit: Suzan Black)

Displacement
If an object moves relative to a frame of reference—for example, if a professor moves to the right relative to a whiteboard
Figure 3.3—then the object’s position changes. This change in position is called displacement. The word displacement
implies that an object has moved, or has been displaced. Although position is the numerical value of x along a straight line
where an object might be located, displacement gives the change in position along this line. Since displacement indicates
direction, it is a vector and can be either positive or negative, depending on the choice of positive direction. Also, an analysis
of motion can have many displacements embedded in it. If right is positive and an object moves 2 m to the right, then 4 m
to the left, the individual displacements are 2 m and −4 m, respectively.

This OpenStax book is available for free at https://legacy.cnx.org/content/col12169/1.46


Chapter 3 | Straight-Line Motion 65

Figure 3.3 A professor paces left and right while lecturing. Her position relative to Earth
is given by x. The +2.0-m displacement of the professor relative to Earth is represented by
an arrow pointing to the right.

Displacement
Displacement Δx is the change in position of an object:
Δx = x f − x 0, (3.1)

where Δx is displacement, x f is the final position, and x 0 is the initial position.

We use the uppercase Greek letter delta (Δ) to mean “change in” whatever quantity follows it; thus, Δx means change in
position (final position less initial position). We always solve for displacement by subtracting initial position x 0 from final
position x f . Note that the SI unit for displacement is the meter, but sometimes we use kilometers or other units of length.
Keep in mind that when units other than meters are used in a problem, you may need to convert them to meters to complete
the calculation (see Appendix B).
Objects in motion can also have a series of displacements. In the previous example of the pacing professor, the individual
displacements are 2 m and −4 m, giving a total displacement of −2 m. We define total displacement Δx Total , as the sum
of the individual displacements, and express this mathematically with the equation

Δx Total = ∑ Δx i, (3.2)

where Δx i are the individual displacements. In the earlier example,

Δx 1 = x 1 − x 0 = 2 − 0 = 2 m.

Similarly,
Δx 2 = x 2 − x 1 = −2 − (2) = −4 m.
66 Chapter 3 | Straight-Line Motion

Thus,
Δx Total = Δx 1 + Δx 2 = 2 − 4 = −2 m.

The total displacement is 2 − 4 = −2 m to the left, or in the negative direction. It is also useful to calculate the magnitude
of the displacement, or its size. The magnitude of the displacement is always positive. This is the absolute value of
the displacement, because displacement is a vector and cannot have a negative value of magnitude. In our example, the
magnitude of the total displacement is 2 m, whereas the magnitudes of the individual displacements are 2 m and 4 m.
The magnitude of the total displacement should not be confused with the distance traveled. Distance traveled x Total , is the
total length of the path traveled between two positions. In the previous problem, the distance traveled is the sum of the
magnitudes of the individual displacements:
x Total = |Δx 1| + |Δx 2| = 2 + 4 = 6 m.

Average Velocity
To calculate the other physical quantities in kinematics we must introduce the time variable. The time variable allows us not
only to state where the object is (its position) during its motion, but also how fast it is moving. How fast an object is moving
is given by the rate at which the position changes with time.
For each position x i , we assign a particular time t i . If the details of the motion at each instant are not important, the rate

is usually expressed as the average velocity v . This vector quantity is simply the total displacement between two points
divided by the time taken to travel between them. The time taken to travel between two points is called the elapsed time
Δt .

Average Velocity
If x 1 and x 2 are the positions of an object at times t 1 and t 2 , respectively, then

– Displacement between two points (3.3)


Average velocity = v =
Elapsed time between two points
x −x
v = Δx = t 2 − t 1 .

Δt 2 1

It is important to note that the average velocity is a vector and can be negative, depending on positions x 1 and x 2 .

Example 3.1

Delivering Flyers
Jill sets out from her home to deliver flyers for her yard sale, traveling due east along her street lined with houses.
At 0.5 km and 9 minutes later she runs out of flyers and has to retrace her steps back to her house to get more.
This takes an additional 9 minutes. After picking up more flyers, she sets out again on the same path, continuing
where she left off, and ends up 1.0 km from her house. This third leg of her trip takes 15 minutes. At this point
she turns back toward her house, heading west. After 1.75 km and 25 minutes she stops to rest.
a. What is Jill’s total displacement to the point where she stops to rest?
b. What is the magnitude of the final displacement?
c. What is the average velocity during her entire trip?
d. What is the total distance traveled?
e. Make a graph of position versus time.
A sketch of Jill’s movements is shown in Figure 3.4.

This OpenStax book is available for free at https://legacy.cnx.org/content/col12169/1.46


Chapter 3 | Straight-Line Motion 67

Figure 3.4 Timeline of Jill’s movements.

Strategy
The problem contains data on the various legs of Jill’s trip, so it would be useful to make a table of the physical
quantities. We are given position and time in the wording of the problem so we can calculate the displacements
and the elapsed time. We take east to be the positive direction. From this information we can find the total
displacement and average velocity. Jill’s home is the starting point x 0 . The following table gives Jill’s time and
position in the first two columns, and the displacements are calculated in the third column.

Time ti (min) Position xi (km) Displacement Δxi (km)

t0 = 0 x0 = 0 Δx 0 = 0

t1 = 9 x 1 = 0.5 Δx 1 = x 1 − x 0 = 0.5

t 2 = 18 x2 = 0 Δx 2 = x 2 − x 1 = −0.5

t 3 = 33 x 3 = 1.0 Δx 3 = x 3 − x 2 = 1.0

t 4 = 58 x 4 = −0.75 Δx 4 = x 4 − x 3 = −1.75

Solution
a. From the above table, the total displacement is

∑ Δx i = 0.5 − 0.5 + 1.0 − 1.75 km = −0.75 km.


b. The magnitude of the total displacement is |−0.75| km = 0.75 km .

Total displacement – −0.75 km


c. Average velocity = =v= = −0.013 km/min
Elapsed time 58 min

d. The total distance traveled (sum of magnitudes of individual displacements) is


x Total = ∑ |Δx i| = 0.5 + 0.5 + 1.0 + 1.75 km = 3.75 km .

e. We can graph Jill’s position versus time as a useful aid to see the motion; the graph is shown in Figure
3.5.
68 Chapter 3 | Straight-Line Motion

Figure 3.5 This graph depicts Jill’s position versus time. The
average velocity is the slope of a line connecting the initial and
final points.

Significance
Jill’s total displacement is −0.75 km, which means at the end of her trip she ends up 0.75 km due west of her
home. The average velocity means if someone was to walk due west at 0.013 km/min starting at the same time
Jill left her home, they both would arrive at the final stopping point at the same time. Note that if Jill were to
end her trip at her house, her total displacement would be zero, as well as her average velocity. The total distance
traveled during the 58 minutes of elapsed time for her trip is 3.75 km.

3.1 Check Your Understanding A cyclist rides 3 km west and then turns around and rides 2 km east. (a)
What is his displacement? (b) What is the distance traveled? (c) What is the magnitude of his displacement?

3.2 | Instantaneous Velocity and Speed


Learning Objectives
By the end of this section, you will be able to:
• Explain the difference between average velocity and instantaneous velocity.
• Describe the difference between velocity and speed.
• Calculate the instantaneous velocity given the mathematical equation for the velocity.
• Calculate the speed given the instantaneous velocity.

We have now seen how to calculate the average velocity between two positions. However, since objects in the real world
move continuously through space and time, we would like to find the velocity of an object at any single point. We can find
the velocity of the object anywhere along its path by using some fundamental principles of calculus. This section gives us

This OpenStax book is available for free at https://legacy.cnx.org/content/col12169/1.46


Chapter 3 | Straight-Line Motion 69

better insight into the physics of motion and will be useful in later chapters.

Instantaneous Velocity
The quantity that tells us how fast an object is moving anywhere along its path is the instantaneous velocity, usually
called simply velocity. It is the average velocity between two points on the path in the limit that the time (and therefore the
displacement) between the two points approaches zero. To illustrate this idea mathematically, we need to express position x
as a continuous function of t denoted by x(t). The expression for the average velocity between two points using this notation
– x(t 2) − x(t 1)
is v = t 2 − t 1 . To find the instantaneous velocity at any position, we let t 1 = t and t 2 = t + Δt . After inserting
these expressions into the equation for the average velocity and taking the limit as Δt → 0 , we find the expression for the
instantaneous velocity:
x(t + Δt) − x(t) dx(t)
v(t) = lim = .
Δt → 0 Δt dt

The Calculus of Instantaneous Velocity


The instantaneous velocity of an object is the limit of the average velocity as the elapsed time approaches zero, or the
derivative of x with respect to t:

v(t) = d x(t). (3.4)


dt
Because many students reading this book may not yet have learned how to take derivatives of functions in their
calculus course, we will postpone such use of calculus unit later chapters.

Like average velocity, instantaneous velocity is a vector with dimension of length per time. The instantaneous velocity at
a specific time point t 0 is the rate of change of the position function, which is the slope of the position function x(t) at

t 0 . Figure 3.6 shows how the average velocity v = Δx between two times approaches the instantaneous velocity at t 0.

Δt
The instantaneous velocity is shown at time t 0 , which happens to be at the maximum of the position function. The slope
of the position graph is zero at this point, and thus the instantaneous velocity is zero. At other times, t 1, t 2 , and so on,
the instantaneous velocity is not zero because the slope of the position graph would be positive or negative. If the position
function had a minimum, the slope of the position graph would also be zero, giving an instantaneous velocity of zero there
as well. Thus, the zeros of the velocity function give the minimum and maximum of the position function.
70 Chapter 3 | Straight-Line Motion

Figure 3.6 In a graph of position versus time, the


instantaneous velocity is the slope of the tangent line at a given
x − xi
point. The average velocities v = Δx = f

between
Δt tf − ti
times Δt = t 6 − t 1, Δt = t 5 − t 2, and Δt = t 4 − t 3 are
shown. When Δt → 0 , the average velocity approaches the
instantaneous velocity at t = t 0 .

Example 3.2

Finding Velocity from a Position-Versus-Time Graph


Given the position-versus-time graph of Figure 3.7, find the velocity-versus-time graph.

Figure 3.7 The object starts out in the positive direction, stops
for a short time, and then reverses direction, heading back
toward the origin. Notice that the object comes to rest
instantaneously, which would require an infinite force. Thus, the
graph is an approximation of motion in the real world. (The
concept of force is discussed in Newton’s Laws of Motion
(https://legacy.cnx.org/content/m58294/latest/) .)

This OpenStax book is available for free at https://legacy.cnx.org/content/col12169/1.46


Chapter 3 | Straight-Line Motion 71

Strategy
The graph contains three straight lines during three time intervals. We find the velocity during each time interval
by taking the slope of the line using the grid.
Solution

Time interval 0 s to 0.5 s: v = Δx = 0.5 m − 0.0 m = 1.0 m/s



Δt 0.5 s − 0.0 s

Time interval 0.5 s to 1.0 s: v = Δx = 0.0 m − 0.0 m = 0.0 m/s



Δt 1.0 s − 0.5 s

Time interval 1.0 s to 2.0 s: v = Δx = 0.0 m − 0.5 m = −0.5 m/s



Δt 2.0 s − 1.0 s
The graph of these values of velocity versus time is shown in Figure 3.8.

Figure 3.8 The velocity is positive for the first part of the trip,
zero when the object is stopped, and negative when the object
reverses direction.

Significance
During the time interval between 0 s and 0.5 s, the object’s position is moving away from the origin and the
position-versus-time curve has a positive slope. At any point along the curve during this time interval, we can
find the instantaneous velocity by taking its slope, which is +1 m/s, as shown in Figure 3.8. In the subsequent
time interval, between 0.5 s and 1.0 s, the position doesn’t change and we see the slope is zero. From 1.0 s to 2.0
s, the object is moving back toward the origin and the slope is −0.5 m/s. The object has reversed direction and has
a negative velocity.

Speed
In everyday language, most people use the terms speed and velocity interchangeably. In physics, however, they do not have
the same meaning and are distinct concepts. One major difference is that speed has no direction; that is, speed is a scalar.
We can calculate the average speed by finding the total distance traveled divided by the elapsed time:

Average speed = –s = Total distance . (3.5)


Elapsed time

Average speed is not necessarily the same as the magnitude of the average velocity, which is found by dividing the
magnitude of the total displacement by the elapsed time. For example, if a trip starts and ends at the same location, the total
displacement is zero, and therefore the average velocity is zero. The average speed, however, is not zero, because the total
72 Chapter 3 | Straight-Line Motion

distance traveled is greater than zero. If we take a road trip of 300 km and need to be at our destination at a certain time,
then we would be interested in our average speed.
However, we can calculate the instantaneous speed from the magnitude of the instantaneous velocity:

Instantaneous speed = |v(t)|. (3.6)

If a particle is moving along the x-axis at +7.0 m/s and another particle is moving along the same axis at −7.0 m/s, they have
different velocities, but both have the same speed of 7.0 m/s. Some typical speeds are shown in the following table.

Speed m/s mi/h


Continental drift 10 −7 2 × 10 −7
Brisk walk 1.7 3.9
Cyclist 4.4 10
Sprint runner 12.2 27
Rural speed limit 24.6 56
Official land speed record 341.1 763
Speed of sound at sea level 343 768
Space shuttle on reentry 7800 17,500
Escape velocity of Earth* 11,200 25,000
Orbital speed of Earth around the Sun 29,783 66,623
Speed of light in a vacuum 299,792,458 670,616,629

Table 3.1 Speeds of Various Objects *Escape velocity is the velocity at


which an object must be launched so that it overcomes Earth’s gravity and is
not pulled back toward Earth.

3.3 | Average and Instantaneous Acceleration


Learning Objectives
By the end of this section, you will be able to:
• Calculate the average acceleration between two points in time.
• Calculate the instantaneous acceleration given the functional form of velocity.
• Explain the vector nature of instantaneous acceleration and velocity.
• Explain the difference between average acceleration and instantaneous acceleration.
• Find instantaneous acceleration at a specified time on a graph of velocity versus time.

The importance of understanding acceleration spans our day-to-day experience, as well as the vast reaches of outer space
and the tiny world of subatomic physics. In everyday conversation, to accelerate means to speed up; applying the brake
pedal causes a vehicle to slow down. We are familiar with the acceleration of our car, for example. The greater the
acceleration, the greater the change in velocity over a given time. Acceleration is widely seen in experimental physics. In
linear particle accelerator experiments, for example, subatomic particles are accelerated to very high velocities in collision
experiments, which tell us information about the structure of the subatomic world as well as the origin of the universe.
In space, cosmic rays are subatomic particles that have been accelerated to very high energies in supernovas (exploding
massive stars) and active galactic nuclei. It is important to understand the processes that accelerate cosmic rays because
these rays contain highly penetrating radiation that can damage electronics flown on spacecraft, for example.

This OpenStax book is available for free at https://legacy.cnx.org/content/col12169/1.46


Chapter 3 | Straight-Line Motion 73

Average Acceleration
The formal definition of acceleration is consistent with these notions just described, but is more inclusive.

Average Acceleration
Average acceleration is the rate at which velocity changes:
–a = Δv = v f − v 0 , (3.7)
Δt tf − t0

where −
a is average acceleration, v is velocity, and t is time. (The bar over the a means average acceleration.)

Because acceleration is velocity in meters divided by time in seconds, the SI units for acceleration are often abbreviated m/
s2—that is, meters per second squared or meters per second per second. This literally means by how many meters per second
the velocity changes every second. Recall that velocity is a vector—it has both magnitude and direction—which means that
a change in velocity can be a change in magnitude (or speed), but it can also be a change in direction. For example, if a
runner traveling at 10 km/h due east slows to a stop, reverses direction, continues her run at 10 km/h due west, her velocity
has changed as a result of the change in direction, although the magnitude of the velocity is the same in both directions.
Thus, acceleration occurs when velocity changes in magnitude (an increase or decrease in speed) or in direction, or both.

Acceleration as a Vector
Acceleration is a vector in the same direction as the change in velocity, Δv . Since velocity is a vector, it can change
in magnitude or in direction, or both. Acceleration is, therefore, a change in speed or direction, or both.

Keep in mind that although acceleration is in the direction of the change in velocity, it is not always in the direction of
motion. When an object slows down, its acceleration is opposite to the direction of its motion. Although this is commonly
referred to as deceleration Figure 3.9, we say the train is accelerating in a direction opposite to its direction of motion.

Figure 3.9 A subway train in Sao Paulo, Brazil, decelerates as


it comes into a station. It is accelerating in a direction opposite
to its direction of motion. (credit: Yusuke Kawasaki)

The term deceleration can cause confusion in our analysis because it is not a vector and it does not point to a specific
direction with respect to a coordinate system, so we do not use it. Acceleration is a vector, so we must choose the appropriate
sign for it in our chosen coordinate system. In the case of the train in Figure 3.9, acceleration is in the negative direction
in the chosen coordinate system, so we say the train is undergoing negative acceleration.
If an object in motion has a velocity in the positive direction with respect to a chosen origin and it acquires a constant
negative acceleration, the object eventually comes to a rest and reverses direction. If we wait long enough, the object passes
through the origin going in the opposite direction. This is illustrated in Figure 3.10.
74 Chapter 3 | Straight-Line Motion

Figure 3.10 An object in motion with a velocity vector toward the east under
negative acceleration comes to a rest and reverses direction. It passes the origin going
in the opposite direction after a long enough time.

Example 3.3

Calculating Average Acceleration: A Racehorse Leaves the Gate


A racehorse coming out of the gate accelerates from rest to a velocity of 15.0 m/s due west in 1.80 s. What is its
average acceleration?

Figure 3.11 Racehorses accelerating out of the gate. (credit:


Jon Sullivan)

Strategy
First we draw a sketch and assign a coordinate system to the problem Figure 3.12. This is a simple problem, but
it always helps to visualize it. Notice that we assign east as positive and west as negative. Thus, in this case, we
have negative velocity.

Figure 3.12 Identify the coordinate system, the given information, and what you want to
determine.

We can solve this problem by identifying Δv and Δt from the given information, and then calculating the
v −v
average acceleration directly from the equation –
a = Δv = t f − t 0 .
Δt f 0

This OpenStax book is available for free at https://legacy.cnx.org/content/col12169/1.46


Chapter 3 | Straight-Line Motion 75

Solution
First, identify the knowns: v 0 = 0, v f = −15.0 m/s (the negative sign indicates direction toward the west), Δt =
1.80 s.
Second, find the change in velocity. Since the horse is going from zero to –15.0 m/s, its change in velocity equals
its final velocity:
Δv = v f − v 0 = v f = −15.0 m/s.

Last, substitute the known values ( Δv and Δt ) and solve for the unknown –
a:
–a = Δv = −15.0 m/s = −8.33m/s 2.
Δt 1.80 s
Significance
The negative sign for acceleration indicates that acceleration is toward the west. An acceleration of 8.33 m/s2 due
west means the horse increases its velocity by 8.33 m/s due west each second; that is, 8.33 meters per second per
second, which we write as 8.33 m/s2. This is truly an average acceleration, because the ride is not smooth. We
see later that an acceleration of this magnitude would require the rider to hang on with a force nearly equal to his
weight.

3.2 Check Your Understanding Protons in a linear accelerator are accelerated from rest to 2.0 × 10 7 m/s
in 10–4 s. What is the average acceleration of the protons?

Instantaneous Acceleration
Instantaneous acceleration a, or acceleration at a specific instant in time, is obtained using the same process discussed
for instantaneous velocity. That is, we calculate the average velocity between two points in time separated by Δt and let
Δt approach zero. The result is the derivative of the velocity function v(t), which is instantaneous acceleration and is
expressed mathematically as

a(t) = d v(t). (3.8)


dt

Thus, similar to velocity being the derivative of the position function, instantaneous acceleration is the derivative of the
velocity function. We can show this graphically in the same way as instantaneous velocity. In Figure 3.13, instantaneous
acceleration at time t0 is the slope of the tangent line to the velocity-versus-time graph at time t0. We see that average
acceleration –
a = Δv approaches instantaneous acceleration as Δt approaches zero. Also in part (a) of the figure, we see
Δt
that velocity has a maximum when its slope is zero. This time corresponds to the zero of the acceleration function. In
part (b), instantaneous acceleration at the minimum velocity is shown, which is also zero, since the slope of the curve is
zero there, too. Thus, for a given velocity function, the zeros of the acceleration function give either the minimum or the
maximum velocity.
76 Chapter 3 | Straight-Line Motion

Figure 3.13 In a graph of velocity versus time, instantaneous acceleration is the slope of the tangent line. (a)
v −v
Shown is average acceleration –
a = Δv = t f − t i between times Δt = t 6 − t 1, Δt = t 5 − t 2 , and
Δt f i
Δt = t 4 − t 3 . When Δt → 0 , the average acceleration approaches instantaneous acceleration at time t0. In view
(a), instantaneous acceleration is shown for the point on the velocity curve at maximum velocity. At this point,
instantaneous acceleration is the slope of the tangent line, which is zero. At any other time, the slope of the tangent
line—and thus instantaneous acceleration—would not be zero. (b) Same as (a) but shown for instantaneous
acceleration at minimum velocity.

To illustrate this concept, let’s look at two examples. First, a simple example is shown of a velocity-versus-time graph, to
find acceleration graphically. This graph is depicted in Figure 3.14(a), which is a straight line. The corresponding graph of
acceleration versus time is found from the slope of velocity and is shown in Figure 3.14(b). In this example, the velocity
function is a straight line with a constant slope, thus acceleration is a constant.

Figure 3.14 (a, b) The velocity-versus-time graph is linear and has a negative constant slope (a) that is equal to
acceleration, shown in (b).

3.3 Check Your Understanding An airplane lands on a runway traveling east. Describe its acceleration.

Getting a Feel for Acceleration


You are probably used to experiencing acceleration when you step into an elevator, or step on the gas pedal in your car.

This OpenStax book is available for free at https://legacy.cnx.org/content/col12169/1.46


Chapter 3 | Straight-Line Motion 77

However, acceleration is happening to many other objects in our universe with which we don’t have direct contact. Table
3.2 presents the acceleration of various objects. We can see the magnitudes of the accelerations extend over many orders of
magnitude.

Acceleration Value (m/s2)


High-speed train 0.25
Elevator 2
Cheetah 5
Object in a free fall without air resistance near the surface of Earth 9.8
Space shuttle maximum during launch 29
Parachutist peak during normal opening of parachute 59
F16 aircraft pulling out of a dive 79
Explosive seat ejection from aircraft 147
Sprint missile 982
Fastest rocket sled peak acceleration 1540
Jumping flea 3200
Baseball struck by a bat 30,000
Closing jaws of a trap-jaw ant 1,000,000
Proton in the large Hadron collider 1.9 × 10 9

Table 3.2 Typical Values of Acceleration (credit: Wikipedia: Orders of Magnitude


(acceleration))

In this table, we see that typical accelerations vary widely with different objects and have nothing to do with object size
or how massive it is. Acceleration can also vary widely with time during the motion of an object. A drag racer has a large
acceleration just after its start, but then it tapers off as the vehicle reaches a constant velocity. Its average acceleration can be
quite different from its instantaneous acceleration at a particular time during its motion. Figure 3.15 compares graphically
average acceleration with instantaneous acceleration for two very different motions.
78 Chapter 3 | Straight-Line Motion

Figure 3.15 Graphs of instantaneous acceleration versus time for two different one-dimensional motions. (a) Acceleration
varies only slightly and is always in the same direction, since it is positive. The average over the interval is nearly the same as
the acceleration at any given time. (b) Acceleration varies greatly, perhaps representing a package on a post office conveyor
belt that is accelerated forward and backward as it bumps along. It is necessary to consider small time intervals (such as from
0–1.0 s) with constant or nearly constant acceleration in such a situation.

Learn about position, velocity, and acceleration graphs. Move the little man back and forth with a mouse and plot
his motion. Set the position, velocity, or acceleration and let the simulation move the man for you. Visit this link
(https://openstaxcollege.org/l/21movmansimul) to use the moving man simulation.

3.4 | Motion with Constant Acceleration


Learning Objectives
By the end of this section, you will be able to:
• Identify which equations of motion are to be used to solve for unknowns.
• Use appropriate equations of motion to solve a two-body pursuit problem.

You might guess that the greater the acceleration of, say, a car moving away from a stop sign, the greater the car’s
displacement in a given time. But, we have not developed a specific equation that relates acceleration and displacement. In
this section, we look at some convenient equations for kinematic relationships, starting from the definitions of displacement,
velocity, and acceleration. We first investigate a single object in motion, called single-body motion. Then we investigate the
motion of two objects, called two-body pursuit problems.

Notation
First, let us make some simplifications in notation. Taking the initial time to be zero, as if time is measured with a stopwatch,
is a great simplification. Since elapsed time is Δt = t f − t 0 , taking t 0 = 0 means that Δt = t f , the final time on the
stopwatch. When initial time is taken to be zero, we use the subscript 0 to denote initial values of position and velocity.
That is, x 0 is the initial position and v 0 is the initial velocity. We put no subscripts on the final values. That is, t is the
final time, x is the final position, and v is the final velocity. This gives a simpler expression for elapsed time, Δt = t . It also
simplifies the expression for x displacement, which is now Δx = x − x 0 . Also, it simplifies the expression for change in
velocity, which is now Δv = v − v 0 . To summarize, using the simplified notation, with the initial time taken to be zero,

This OpenStax book is available for free at https://legacy.cnx.org/content/col12169/1.46


Chapter 3 | Straight-Line Motion 79

Δt = t
Δx = x − x 0
Δv = v − v 0,

where the subscript 0 denotes an initial value and the absence of a subscript denotes a final value in whatever motion is
under consideration.
We now make the important assumption that acceleration is constant. This assumption allows us to avoid using calculus to
find instantaneous acceleration. Since acceleration is constant, the average and instantaneous accelerations are equal—that
is,
–a = a = constant.

Thus, we can use the symbol a for acceleration at all times. Assuming acceleration to be constant does not seriously limit
the situations we can study nor does it degrade the accuracy of our treatment. For one thing, acceleration is constant in
a great number of situations. Furthermore, in many other situations we can describe motion accurately by assuming a
constant acceleration equal to the average acceleration for that motion. Lastly, for motion during which acceleration changes
drastically, such as a car accelerating to top speed and then braking to a stop, motion can be considered in separate parts,
each of which has its own constant acceleration.

Displacement and Position from Velocity


To get our first two equations, we start with the definition of average velocity:

v = Δx .

Δt
Substituting the simplified notation for Δx and Δt yields
– x − x0
v= t .
Solving for x gives us


x = x 0 + vt, (3.9)

where the average velocity is

– v0 + v (3.10)
v= .
2

– v +v
The equation v = 0 reflects the fact that when acceleration is constant, v is just the simple average of the initial and
2
final velocities. Figure 3.16 illustrates this concept graphically. In part (a) of the figure, acceleration is constant, with
velocity increasing at a constant rate. The average velocity during the 1-h interval from 40 km/h to 80 km/h is 60 km/h:
– v 0 + v 40 km/h + 80 km/h
v= = = 60 km/h.
2 2
In part (b), acceleration is not constant. During the 1-h interval, velocity is closer to 80 km/h than 40 km/h. Thus, the average
velocity is greater than in part (a).
80 Chapter 3 | Straight-Line Motion

Figure 3.16 (a) Velocity-versus-time graph with constant acceleration showing the initial and final velocities v 0 and v .

The average velocity is 1 (v 0 + v) = 60km/h . (b) Velocity-versus-time graph with an acceleration that changes with time.
2
The average velocity is not given by 1 (v 0 + v) , but is greater than 60 km/h.
2

Solving for Final Velocity from Acceleration and Time


We can derive another useful equation by manipulating the definition of acceleration:

a = Δv .
Δt
Substituting the simplified notation for Δv and Δt gives us
v − v0
a= t (constant a).

Solving for v yields

v = v 0 + at (constant a). (3.11)

Example 3.4

Calculating Final Velocity


An airplane lands with an initial velocity of 70.0 m/s and then decelerates at 1.50 m/s2 for 40.0 s. What is its final
velocity?
Strategy
First, we identify the knowns: v 0 = 70 m/s, a = −1.50 m/s 2, t = 40 s .

Second, we identify the unknown; in this case, it is final velocity v f .

Last, we determine which equation to use. To do this we figure out which kinematic equation gives the unknown
in terms of the knowns. We calculate the final velocity using Equation 3.11, v = v 0 + at .

Solution
Substitute the known values and solve:

This OpenStax book is available for free at https://legacy.cnx.org/content/col12169/1.46


Chapter 3 | Straight-Line Motion 81

v = v 0 + at = 70.0 m/s + ⎛⎝−1.50 m/s 2⎞⎠(40.0 s) = 10.0 m/s.

Figure 3.17 is a sketch that shows the acceleration and velocity vectors.

Figure 3.17 The airplane lands with an initial velocity of 70.0 m/s and slows to a final velocity of 10.0 m/s before
heading for the terminal. Note the acceleration is negative because its direction is opposite to its velocity, which is
positive.

Significance
The final velocity is much less than the initial velocity, as desired when slowing down, but is still positive (see
figure). With jet engines, reverse thrust can be maintained long enough to stop the plane and start moving it
backward, which is indicated by a negative final velocity, but is not the case here.

In addition to being useful in problem solving, the equation v = v 0 + at gives us insight into the relationships among
velocity, acceleration, and time. We can see, for example, that
• Final velocity depends on how large the acceleration is and how long it lasts
• If the acceleration is zero, then the final velocity equals the initial velocity (v = v0), as expected (in other words,
velocity is constant)
• If a is negative, then the final velocity is less than the initial velocity
All these observations fit our intuition. Note that it is always useful to examine basic equations in light of our intuition and
experience to check that they do indeed describe nature accurately.

Solving for Final Position with Constant Acceleration


We can combine the previous equations to find a third equation that allows us to calculate the final position of an object
experiencing constant acceleration. We start with
v = v 0 + at.

Adding v 0 to each side of this equation and dividing by 2 gives

v0 + v
= v 0 + 1 at.
2 2
v0 + v –
Since = v for constant acceleration, we have
2

v = v 0 + 1 at.

2
– –
Now we substitute this expression for v into the equation for displacement, x = x 0 + vt , yielding

x = x 0 + v 0 t + 1 at 2 (constant a). (3.12)


2
82 Chapter 3 | Straight-Line Motion

Example 3.5

Calculating Displacement of an Accelerating Object


Dragsters can achieve an average acceleration of 26.0 m/s2. Suppose a dragster accelerates from rest at this rate
for 5.56 s Figure 3.18. How far does it travel in this time?

Figure 3.18 U.S. Army Top Fuel pilot Tony “The Sarge”
Schumacher begins a race with a controlled burnout. (credit: Lt.
Col. William Thurmond. Photo Courtesy of U.S. Army.)

Strategy
First, let’s draw a sketch Figure 3.19. We are asked to find displacement, which is x if we take x 0 to be zero.
(Think about x 0 as the starting line of a race. It can be anywhere, but we call it zero and measure all other

positions relative to it.) We can use the equation x = x 0 + v 0 t + 1 at 2 when we identify v 0 , a , and t from the
2
statement of the problem.

Figure 3.19 Sketch of an accelerating dragster.

Solution
First, we need to identify the knowns. Starting from rest means that v 0 = 0 , a is given as 26.0 m/s2 and t is given
as 5.56 s.
Second, we substitute the known values into the equation to solve for the unknown:

x = x 0 + v 0 t + 1 at 2.
2

This OpenStax book is available for free at https://legacy.cnx.org/content/col12169/1.46


Chapter 3 | Straight-Line Motion 83

Since the initial position and velocity are both zero, this equation simplifies to

x = 1 at 2.
2
Substituting the identified values of a and t gives

x = 1 (26.0 m/s 2)(5.56 s) 2 = 402 m.


2
Significance
If we convert 402 m to miles, we find that the distance covered is very close to one-quarter of a mile, the standard
distance for drag racing. So, our answer is reasonable. This is an impressive displacement to cover in only 5.56
s, but top-notch dragsters can do a quarter mile in even less time than this. If the dragster were given an initial
velocity, this would add another term to the distance equation. If the same acceleration and time are used in the
equation, the distance covered would be much greater.

What else can we learn by examining the equation x = x 0 + v 0 t + 1 at 2 ? We can see the following relationships:
2
• Displacement depends on the square of the elapsed time when acceleration is not zero. In Example 3.5, the
dragster covers only one-fourth of the total distance in the first half of the elapsed time.

• If acceleration is zero, then initial velocity equals average velocity (v 0 = v) , and

x = x 0 + v 0 t + 1 at 2 becomes x = x 0 + v 0 t.
2

Solving for Final Velocity from Distance and Acceleration


A fourth useful equation can be obtained from another algebraic manipulation of previous equations. If we solve
v = v 0 + at for t, we get
v − v0
t= a .
– v +v –
Substituting this and v = 0 into x = x 0 + vt , we get
2

v 2 = v 20 + 2a(x − x 0) (constant a). (3.13)

Example 3.6

Calculating Final Velocity


Calculate the final velocity of the dragster in Example 3.5 without using information about time.
Strategy
The equation v 2 = v 20 + 2a(x − x 0) is ideally suited to this task because it relates velocities, acceleration, and
displacement, and no time information is required.
Solution
First, we identify the known values. We know that v0 = 0, since the dragster starts from rest. We also know that x
− x0 = 402 m (this was the answer in Example 3.5). The average acceleration was given by a = 26.0 m/s2.
Second, we substitute the knowns into the equation v 2 = v 20 + 2a(x − x 0) and solve for v:

v 2 = 0 + 2⎛⎝26.0 m/s 2⎞⎠(402 m).


84 Chapter 3 | Straight-Line Motion

Thus,
v 2 = 2.09 × 10 4 m
/s 2

v = 2.09 × 10 4 m 2 /s 2 = 145 m/s.


Significance
A velocity of 145 m/s is about 522 km/h, or about 324 mi/h, but even this breakneck speed is short of the record
for the quarter mile. Also, note that a square root has two values; we took the positive value to indicate a velocity
in the same direction as the acceleration.

An examination of the equation v 2 = v 20 + 2a(x − x 0) can produce additional insights into the general relationships among
physical quantities:
• The final velocity depends on how large the acceleration is and the distance over which it acts.
• For a fixed acceleration, a car that is going twice as fast doesn’t simply stop in twice the distance. It takes much
farther to stop. (This is why we have reduced speed zones near schools.)

Putting Equations Together


In the following examples, we continue to explore one-dimensional motion, but in situations requiring slightly more
algebraic manipulation. The examples also give insight into problem-solving techniques. The note that follows is provided
for easy reference to the equations needed. Be aware that these equations are not independent. In many situations we have
two unknowns and need two equations from the set to solve for the unknowns. We need as many equations as there are
unknowns to solve a given situation.

Summary of Kinematic Equations (constant a)



x = x 0 + vt
– v +v
v= 0
2
v = v 0 + at
x = x 0 + v 0 t + 1 at 2
2
v 2 = v 20 + 2a(x − x 0)

Before we get into the examples, let’s look at some of the equations more closely to see the behavior of acceleration at
extreme values. Rearranging Equation 3.11, we have
v − v0
a= t .
From this we see that, for a finite time, if the difference between the initial and final velocities is small, the acceleration is
small, approaching zero in the limit that the initial and final velocities are equal. On the contrary, in the limit t → 0 for a
finite difference between the initial and final velocities, acceleration becomes infinite.
Similarly, rearranging Equation 3.13, we can express acceleration in terms of velocities and displacement:
v 2 − v 20
a= .
2(x − x 0)

Thus, for a finite difference between the initial and final velocities acceleration becomes infinite in the limit the
displacement approaches zero. Acceleration approaches zero in the limit the difference in initial and final velocities
approaches zero for a finite displacement.

This OpenStax book is available for free at https://legacy.cnx.org/content/col12169/1.46


Chapter 3 | Straight-Line Motion 85

Example 3.7

How Far Does a Car Go?


On dry concrete, a car can decelerate at a rate of 7.00 m/s2, whereas on wet concrete it can decelerate at only 5.00
m/s2. Find the distances necessary to stop a car moving at 30.0 m/s (about 110 km/h) on (a) dry concrete and (b)
wet concrete. (c) Repeat both calculations and find the displacement from the point where the driver sees a traffic
light turn red, taking into account his reaction time of 0.500 s to get his foot on the brake.
Strategy
First, we need to draw a sketch Figure 3.20. To determine which equations are best to use, we need to list all the
known values and identify exactly what we need to solve for.

Figure 3.20 Sample sketch to visualize deceleration and stopping distance of a car.

Solution
a. First, we need to identify the knowns and what we want to solve for. We know that v0 = 30.0 m/s, v = 0,
and a = −7.00 m/s2 (a is negative because it is in a direction opposite to velocity). We take x0 to be zero.
We are looking for displacement Δx , or x − x0.
Second, we identify the equation that will help us solve the problem. The best equation to use is
v 2 = v 20 + 2a(x − x 0).

This equation is best because it includes only one unknown, x. We know the values of all the other
variables in this equation. (Other equations would allow us to solve for x, but they require us to know the
stopping time, t, which we do not know. We could use them, but it would entail additional calculations.)
Third, we rearrange the equation to solve for x:
v 2 − v 20
x − x0 =
2a

and substitute the known values:


0 2 − (30.0 m/s) 2
x−0= .
2(−7.00m/s 2)

Thus,
x = 64.3 m on dry concrete.
b. This part can be solved in exactly the same manner as (a). The only difference is that the acceleration is
−5.00 m/s2. The result is
x wet = 90.0 m on wet concrete.
c. When the driver reacts, the stopping distance is the same as it is in (a) and (b) for dry and wet concrete.
So, to answer this question, we need to calculate how far the car travels during the reaction time, and then
86 Chapter 3 | Straight-Line Motion

add that to the stopping time. It is reasonable to assume the velocity remains constant during the driver’s
reaction time.

To do this, we, again, identify the knowns and what we want to solve for. We know that v = 30.0 m/s ,
t reaction = 0.500 s , and a reaction = 0 . We take x 0-reaction to be zero. We are looking for x reaction .

Second, as before, we identify the best equation to use. In this case, x = x 0 + vt works well because the
only unknown value is x, which is what we want to solve for.
Third, we substitute the knowns to solve the equation:
x = 0 + (30.0 m/s)(0.500 s) = 15.0 m.

This means the car travels 15.0 m while the driver reacts, making the total displacements in the two cases
of dry and wet concrete 15.0 m greater than if he reacted instantly.
Last, we then add the displacement during the reaction time to the displacement when braking (Figure
3.21),
x braking + x reaction = x total,

and find (a) to be 64.3 m + 15.0 m = 79.3 m when dry and (b) to be 90.0 m + 15.0 m = 105 m when wet.

Figure 3.21 The distance necessary to stop a car varies greatly, depending on road conditions and driver reaction
time. Shown here are the braking distances for dry and wet pavement, as calculated in this example, for a car
traveling initially at 30.0 m/s. Also shown are the total distances traveled from the point when the driver first sees a
light turn red, assuming a 0.500-s reaction time.

Significance
The displacements found in this example seem reasonable for stopping a fast-moving car. It should take longer
to stop a car on wet pavement than dry. It is interesting that reaction time adds significantly to the displacements,
but more important is the general approach to solving problems. We identify the knowns and the quantities to be
determined, then find an appropriate equation. If there is more than one unknown, we need as many independent
equations as there are unknowns to solve. There is often more than one way to solve a problem. The various parts
of this example can, in fact, be solved by other methods, but the solutions presented here are the shortest.

Example 3.8

Calculating Time
Suppose a car merges into freeway traffic on a 200-m-long ramp. If its initial velocity is 10.0 m/s and it accelerates
at 2.00 m/s2, how long does it take the car to travel the 200 m up the ramp? (Such information might be useful to
a traffic engineer.)

This OpenStax book is available for free at https://legacy.cnx.org/content/col12169/1.46


Chapter 3 | Straight-Line Motion 87

Strategy
First, we draw a sketch Figure 3.22. We are asked to solve for time t. As before, we identify the known quantities
to choose a convenient physical relationship (that is, an equation with one unknown, t.)

Figure 3.22 Sketch of a car accelerating on a freeway ramp.

Solution
Again, we identify the knowns and what we want to solve for. We know that x 0 = 0,
v 0 = 10 m/s, a = 2.00 m/s 2 , and x = 200 m.

We need to solve for t. The equation x = x 0 + v 0 t + 1 at 2 works best because the only unknown in the equation
2
is the variable t, for which we need to solve. From this insight we see that when we input the knowns into the
equation, we end up with a quadratic equation.
We need to rearrange the equation to solve for t, then substituting the knowns into the equation:

200 m = 0 m + (10.0 m/s)t + 1 ⎛⎝2.00 m/s 2⎞⎠t 2.


2
We then simplify the equation. The units of meters cancel because they are in each term. We can get the units of
seconds to cancel by taking t = t s, where t is the magnitude of time and s is the unit. Doing so leaves
200 = 10t + t 2.
We then use the quadratic formula to solve for t,
t 2 + 10t − 200 = 0
2
t = −b ± b − 4ac ,
2a

which yields two solutions: t = 10.0 and t = −20.0. A negative value for time is unreasonable, since it would mean
the event happened 20 s before the motion began. We can discard that solution. Thus,
t = 10.0 s.
Significance
Whenever an equation contains an unknown squared, there are two solutions. In some problems both solutions are
meaningful; in others, only one solution is reasonable. The 10.0-s answer seems reasonable for a typical freeway
on-ramp.

3.4 Check Your Understanding A manned rocket accelerates at a rate of 20 m/s2 during launch. How long
does it take the rocket to reach a velocity of 400 m/s?
88 Chapter 3 | Straight-Line Motion

Example 3.9

Acceleration of a Spaceship
A spaceship has left Earth’s orbit and is on its way to the Moon. It accelerates at 20 m/s2 for 2 min and covers a
distance of 1000 km. What are the initial and final velocities of the spaceship?
Strategy
We are asked to find the initial and final velocities of the spaceship. Looking at the kinematic equations, we see
that one equation will not give the answer. We must use one kinematic equation to solve for one of the velocities
and substitute it into another kinematic equation to get the second velocity. Thus, we solve two of the kinematic
equations simultaneously.
Solution

First we solve for v 0 using x = x 0 + v 0 t + 1 at 2 = 1 at 2 :


2 2

x − x 0 = v 0 t + 1 at
2
6 1
1.0 × 10 m = v 0(120.0 s) + (20.0m/s 2)(120.0 s) 2
2
v 0 = 7133.3 m/s.

Then we substitute v 0 into v = v 0 + at to solve for the final velocity:

v = v 0 + at = 7133.3 m/s + (20.0 m/s 2)(120.0 s) = 9533.3 m/s.

Significance
There are six variables in displacement, time, velocity, and acceleration that describe motion in one dimension.
The initial conditions of a given problem can be many combinations of these variables. Because of this diversity,
solutions may not be easy as simple substitutions into one of the equations. This example illustrates that solutions
to kinematics may require solving two simultaneous kinematic equations.

With the basics of kinematics established, we can go on to many other interesting examples and applications. In the process
of developing kinematics, we have also glimpsed a general approach to problem solving that produces both correct answers
and insights into physical relationships. The next level of complexity in our kinematics problems involves the motion of
two interrelated bodies, called two-body pursuit problems.

Two-Body Pursuit Problems


Up until this point we have looked at examples of motion involving a single body. Even for the problem with two cars
and the stopping distances on wet and dry roads, we divided this problem into two separate problems to find the answers.
In a two-body pursuit problem, the motions of the objects are coupled—meaning, the unknown we seek depends on the
motion of both objects. To solve these problems we write the equations of motion for each object and then solve them
simultaneously to find the unknown. This is illustrated in Figure 3.23.

Figure 3.23 A two-body pursuit scenario where car 2 has a constant velocity and car 1 is
behind with a constant acceleration. Car 1 catches up with car 2 at a later time.

The time and distance required for car 1 to catch car 2 depends on the initial distance car 1 is from car 2 as well as the
velocities of both cars and the acceleration of car 1. The kinematic equations describing the motion of both cars must be
solved to find these unknowns.

This OpenStax book is available for free at https://legacy.cnx.org/content/col12169/1.46


Chapter 3 | Straight-Line Motion 89

Consider the following example.

Example 3.10

Cheetah Catching a Gazelle


A cheetah waits in hiding behind a bush. The cheetah spots a gazelle running past at 10 m/s. At the instant the
gazelle passes the cheetah, the cheetah accelerates from rest at 4 m/s2 to catch the gazelle. (a) How long does it
take the cheetah to catch the gazelle? (b) What is the displacement of the gazelle and cheetah?
Strategy
We use the set of equations for constant acceleration to solve this problem. Since there are two objects in
motion, we have separate equations of motion describing each animal. But what links the equations is a common
parameter that has the same value for each animal. If we look at the problem closely, it is clear the common
parameter to each animal is their position x at a later time t. Since they both start at x 0 = 0 , their displacements
are the same at a later time t, when the cheetah catches up with the gazelle. If we pick the equation of motion that
solves for the displacement for each animal, we can then set the equations equal to each other and solve for the
unknown, which is time.
Solution
a. Equation for the gazelle: The gazelle has a constant velocity, which is its average velocity, since it is not
accelerating. Therefore, we use Equation 3.9 with x 0 = 0 :
– –
x = x 0 + vt = vt.

Equation for the cheetah: The cheetah is accelerating from rest, so we use Equation 3.12 with x 0 = 0
and v 0 = 0 :

x = x 0 + v 0 t + 1 at 2 = 1 at 2.
2 2

Now we have an equation of motion for each animal with a common parameter, which can be eliminated
to find the solution. In this case, we solve for t:

x = vt = 1 at 2

2

2v
t= a.

The gazelle has a constant velocity of 10 m/s, which is its average velocity. The acceleration of the
cheetah is 4 m/s2. Evaluating t, the time for the cheetah to reach the gazelle, we have
– 2(10)
t = 2v
a = 4 = 5 s.
b. To get the displacement, we use either the equation of motion for the cheetah or the gazelle, since they
should both give the same answer.
Displacement of the cheetah:

x = 1 at 2 = 1 (4)(5) 2 = 50 m.
2 2

Displacement of the gazelle:



x = vt = 10(5) = 50 m.

We see that both displacements are equal, as expected.


Significance
It is important to analyze the motion of each object and to use the appropriate kinematic equations to describe the
90 Chapter 3 | Straight-Line Motion

individual motion. It is also important to have a good visual perspective of the two-body pursuit problem to see
the common parameter that links the motion of both objects.

3.5 Check Your Understanding A bicycle has a constant velocity of 10 m/s. A person starts from rest and
runs to catch up to the bicycle in 30 s. What is the acceleration of the person?

3.5 | Free Fall


Learning Objectives
By the end of this section, you will be able to:
• Use the kinematic equations with the variables y and g to analyze free-fall motion.
• Describe how the values of the position, velocity, and acceleration change during a free fall.
• Solve for the position, velocity, and acceleration as functions of time when an object is in a free
fall.

An interesting application of Equation 3.4 through Equation 3.13 is called free fall, which describes the motion of an
object falling in a gravitational field, such as near the surface of Earth or other celestial objects of planetary size. Let’s
assume the body is falling in a straight line perpendicular to the surface, so its motion is one-dimensional. For example, we
can estimate the depth of a vertical mine shaft by dropping a rock into it and listening for the rock to hit the bottom. But
“falling,” in the context of free fall, does not necessarily imply the body is moving from a greater height to a lesser height.
If a ball is thrown upward, the equations of free fall apply equally to its ascent as well as its descent.

Gravity
The most remarkable and unexpected fact about falling objects is that if air resistance and friction are negligible, then in
a given location all objects fall toward the center of Earth with the same constant acceleration, independent of their mass.
This experimentally determined fact is unexpected because we are so accustomed to the effects of air resistance and friction
that we expect light objects to fall slower than heavy ones. Until Galileo Galilei (1564–1642) proved otherwise, people
believed that a heavier object has a greater acceleration in a free fall. We now know this is not the case. In the absence of
air resistance, heavy objects arrive at the ground at the same time as lighter objects when dropped from the same height
Figure 3.24.

This OpenStax book is available for free at https://legacy.cnx.org/content/col12169/1.46


Chapter 3 | Straight-Line Motion 91

Figure 3.24 A hammer and a feather fall with the same constant acceleration if air resistance is negligible. This is a general
characteristic of gravity not unique to Earth, as astronaut David R. Scott demonstrated in 1971 on the Moon, where the
acceleration from gravity is only 1.67 m/s2 and there is no atmosphere.

In the real world, air resistance can cause a lighter object to fall slower than a heavier object of the same size. A tennis ball
reaches the ground after a baseball dropped at the same time. (It might be difficult to observe the difference if the height is
not large.) Air resistance opposes the motion of an object through the air, and friction between objects—such as between
clothes and a laundry chute or between a stone and a pool into which it is dropped—also opposes motion between them.
For the ideal situations of these first few chapters, an object falling without air resistance or friction is defined to be in
free fall. The force of gravity causes objects to fall toward the center of Earth. The acceleration of free-falling objects
is therefore called acceleration due to gravity. Acceleration due to gravity is constant, which means we can apply the
kinematic equations to any falling object where air resistance and friction are negligible. This opens to us a broad class of
interesting situations.
Acceleration due to gravity is so important that its magnitude is given its own symbol, g. It is constant at any given location
on Earth and has the average value
g = 9.81 m/s 2 (or 32.2 ft/s 2).

Although g varies from 9.78 m/s2 to 9.83 m/s2, depending on latitude, altitude, underlying geological formations, and
local topography, let’s use an average value of 9.8 m/s2 rounded to two significant figures in this text unless specified
otherwise. Neglecting these effects on the value of g as a result of position on Earth’s surface, as well as effects resulting
from Earth’s rotation, we take the direction of acceleration due to gravity to be downward (toward the center of Earth).
In fact, its direction defines what we call vertical. Note that whether acceleration a in the kinematic equations has the
value +g or −g depends on how we define our coordinate system. If we define the upward direction as positive, then
a = −g = −9.8 m/s 2, and if we define the downward direction as positive, then a = g = 9.8 m/s 2 .

One-Dimensional Motion Involving Gravity


The best way to see the basic features of motion involving gravity is to start with the simplest situations and then progress
toward more complex ones. So, we start by considering straight up-and-down motion with no air resistance or friction.
These assumptions mean the velocity (if there is any) is vertical. If an object is dropped, we know the initial velocity is zero
when in free fall. When the object has left contact with whatever held or threw it, the object is in free fall. When the object
is thrown, it has the same initial speed in free fall as it did before it was released. When the object comes in contact with the
ground or any other object, it is no longer in free fall and its acceleration of g is no longer valid. Under these circumstances,
the motion is one-dimensional and has constant acceleration of magnitude g. We represent vertical displacement with the
symbol y.
92 Chapter 3 | Straight-Line Motion

Kinematic Equations for Objects in Free Fall


We assume here that acceleration equals −g (with the positive direction upward).
v = v 0 − gt (3.14)

y = y 0 + v 0 t − 1 gt 2 (3.15)
2
v 2 = v 20 − 2g(y − y 0) (3.16)

Problem-Solving Strategy: Free Fall


1. Decide on the sign of the acceleration of gravity. In Equation 3.14 through Equation 3.16, acceleration
g is negative, which says the positive direction is upward and the negative direction is downward. In some
problems, it may be useful to have acceleration g as positive, indicating the positive direction is downward.
2. Draw a sketch of the problem. This helps visualize the physics involved.
3. Record the knowns and unknowns from the problem description. This helps devise a strategy for selecting the
appropriate equations to solve the problem.
4. Decide which of Equation 3.14 through Equation 3.16 are to be used to solve for the unknowns.

Example 3.11

Free Fall of a Ball


Figure 3.25 shows the positions of a ball, at 1-s intervals, with an initial velocity of 4.9 m/s downward, that is
thrown from the top of a 98-m-high building. (a) How much time elapses before the ball reaches the ground? (b)
What is the velocity when it arrives at the ground?

Figure 3.25 The positions and velocities at 1-s intervals of a


ball thrown downward from a tall building at 4.9 m/s.

Strategy
Choose the origin at the top of the building with the positive direction upward and the negative direction
downward. To find the time when the position is −98 m, we use Equation 3.15, with
y 0 = 0, v 0 = −4.9 m/s, and g = 9.8 m/s 2 .

Solution
a. Substitute the given values into the equation:

This OpenStax book is available for free at https://legacy.cnx.org/content/col12169/1.46


Chapter 3 | Straight-Line Motion 93

y = y 0 + v 0 t − 1 gt 2
2
−98.0 m = 0 − (4.9 m/s)t − 1 (9.8 m/s 2)t 2.
2

This simplifies to
t 2 + t − 20 = 0.

This is a quadratic equation with roots t = −5.0s and t = 4.0s . The positive root is the one we are
interested in, since time t = 0 is the time when the ball is released at the top of the building. (The time
t = −5.0s represents the fact that a ball thrown upward from the ground would have been in the air for
5.0 s when it passed by the top of the building moving downward at 4.9 m/s.)
b. Using Equation 3.14, we have
v = v 0 − gt = −4.9 m/s − (9.8m/s 2)(4.0 s) = −44.1 m/s.

Significance
For situations when two roots are obtained from a quadratic equation in the time variable, we must look at the
physical significance of both roots to determine which is correct. Since t = 0 corresponds to the time when
the ball was released, the negative root would correspond to a time before the ball was released, which is not
physically meaningful. When the ball hits the ground, its velocity is not immediately zero, but as soon as the ball
interacts with the ground, its acceleration is not g and it accelerates with a different value over a short time to zero
velocity. This problem shows how important it is to establish the correct coordinate system and to keep the signs
of g in the kinematic equations consistent.

Example 3.12

Vertical Motion of a Baseball


A batter hits a baseball straight upward at home plate and the ball is caught 5.0 s after it is struck Figure 3.26.
(a) What is the initial velocity of the ball? (b) What is the maximum height the ball reaches? (c) How long does it
take to reach the maximum height? (d) What is the acceleration at the top of its path? (e) What is the velocity of
the ball when it is caught? Assume the ball is hit and caught at the same location.

Figure 3.26 A baseball hit straight up is caught by the catcher 5.0 s later.
94 Chapter 3 | Straight-Line Motion

Strategy
Choose a coordinate system with a positive y-axis that is straight up and with an origin that is at the spot where
the ball is hit and caught.
Solution
a. Equation 3.15 gives

y = y 0 + v 0 t − 1 gt 2
2

0 = 0 + v 0(5.0 s) − 1 ⎛⎝9.8 m/s 2⎞⎠(5.0 s) 2,


2

which gives v 0 = 24.5 m/sec .

b. At the maximum height, v = 0 . With v 0 = 24.5 m/s , Equation 3.16 gives

v 2 = v 20 − 2g(y − y 0)

0 = (24.5 m/s) 2 − 2(9.8m/s 2)(y − 0)

or
y = 30.6 m.
c. To find the time when v = 0 , we use Equation 3.14:
v = v 0 − gt

0 = 24.5 m/s − (9.8m/s 2)t.

This gives t = 2.5 s . Since the ball rises for 2.5 s, the time to fall is 2.5 s.
d. The acceleration is 9.8 m/s2 everywhere, even when the velocity is zero at the top of the path. Although
the velocity is zero at the top, it is changing at the rate of 9.8 m/s2 downward.
e. The velocity at t = 5.0s can be determined with Equation 3.14:
v = v 0 − gt
= 24.5 m/s − 9.8m/s 2(5.0 s)
= −24.5 m/s.
Significance
The ball returns with the speed it had when it left. This is a general property of free fall for any initial velocity. We
used a single equation to go from throw to catch, and did not have to break the motion into two segments, upward
and downward. We are used to thinking of the effect of gravity is to create free fall downward toward Earth. It is
important to understand, as illustrated in this example, that objects moving upward away from Earth are also in a
state of free fall.

3.6 Check Your Understanding A chunk of ice breaks off a glacier and falls 30.0 m before it hits the water.
Assuming it falls freely (there is no air resistance), how long does it take to hit the water? Which quantity
increases faster, the speed of the ice chunk or its distance traveled?

This OpenStax book is available for free at https://legacy.cnx.org/content/col12169/1.46


Chapter 3 | Straight-Line Motion 95

Example 3.13

Rocket Booster
A small rocket with a booster blasts off and heads straight upward. When at a height of 5.0 km and velocity of
200.0 m/s, it releases its booster. (a) What is the maximum height the booster attains? (b) What is the velocity of
the booster at a height of 6.0 km? Neglect air resistance.

Figure 3.27 A rocket releases its booster at a given height and


velocity. How high and how fast does the booster go?

Strategy
We need to select the coordinate system for the acceleration of gravity, which we take as negative downward. We
are given the initial velocity of the booster and its height. We consider the point of release as the origin. We know
the velocity is zero at the maximum position within the acceleration interval; thus, the velocity of the booster is
zero at its maximum height, so we can use this information as well. From these observations, we use Equation
3.16, which gives us the maximum height of the booster. We also use Equation 3.16 to give the velocity at 6.0
km. The initial velocity of the booster is 200.0 m/s.
Solution
a. From Equation 3.16, v 2 = v 20 − 2g(y − y 0) . With v = 0 and y 0 = 0 , we can solve for y:

v 20 (2.0 × 10 2 m/s) 2
y= = = 2040.8 m.
−2g −2(9.8 m/s 2)

This solution gives the maximum height of the booster in our coordinate system, which has its origin at
the point of release, so the maximum height of the booster is roughly 7.0 km.
b. An altitude of 6.0 km corresponds to y = 1.0 × 10 3 m in the coordinate system we are using. The other
96 Chapter 3 | Straight-Line Motion

initial conditions are y 0 = 0, and v 0 = 200.0 m/s .


We have, from Equation 3.16,
v 2 = (200.0 m/s) 2 − 2(9.8 m/s 2)(1.0 × 10 3 m) ⇒ v = ± 142.8 m/s.
Significance
We have both a positive and negative solution in (b). Since our coordinate system has the positive direction
upward, the +142.8 m/s corresponds to a positive upward velocity at 6000 m during the upward leg of the
trajectory of the booster. The value v = −142.8 m/s corresponds to the velocity at 6000 m on the downward leg.
This example is also important in that an object is given an initial velocity at the origin of our coordinate system,
but the origin is at an altitude above the surface of Earth, which must be taken into account when forming the
solution.

Visit this site (https://openstaxcollege.org/l/21equatgraph) to learn about graphing polynomials. The


shape of the curve changes as the constants are adjusted. View the curves for the individual terms (for example, y
= bx) to see how they add to generate the polynomial curve.

This OpenStax book is available for free at https://legacy.cnx.org/content/col12169/1.46


Chapter 3 | Straight-Line Motion 97

CHAPTER 3 REVIEW
KEY TERMS
acceleration due to gravity acceleration of an object as a result of gravity
average acceleration the rate of change in velocity; the change in velocity over time
average speed the total distance traveled divided by elapsed time
average velocity the displacement divided by the time over which displacement occurs
displacement the change in position of an object
distance traveled the total length of the path traveled between two positions
elapsed time the difference between the ending time and the beginning time
free fall the state of movement that results from gravitational force only
instantaneous acceleration acceleration at a specific point in time
instantaneous speed the absolute value of the instantaneous velocity
instantaneous velocity the velocity at a specific instant or time point
kinematics the description of motion through properties such as position, time, velocity, and acceleration
position the location of an object at a particular time
total displacement the sum of individual displacements over a given time period
two-body pursuit problem a kinematics problem in which the unknowns are calculated by solving the kinematic
equations simultaneously for two moving objects

KEY EQUATIONS
Displacement Δx = x f − x i

Total displacement Δx Total = ∑ Δx i

–v = Δx = x 2 − x 1
Average velocity Δt t2 − t1

dx(t)
Instantaneous velocity v(t) =
dt

Average speed Average speed = –s = Total distance


Elapsed time

Instantaneous speed Instantaneous speed = |v(t)|

–a = Δv = v f − v 0
Average acceleration Δt t f − t0

dv(t)
Instantaneous acceleration a(t) =
dt

Position from average velocity x = x 0 + –vt

–v = v 0 + v
Average velocity 2

Velocity from acceleration v = v 0 + at (constant a)


98 Chapter 3 | Straight-Line Motion

Position from velocity and acceleration x = x 0 + v 0 t + 1 at 2 (constant a)


2

Velocity from distance v 2 = v 20 + 2a(x − x 0) (constant a)

Velocity of free fall v = v 0 − gt (positive upward)

Height of free fall y = y 0 + v 0 t − 1 gt 2


2

Velocity of free fall from height v 2 = v 20 − 2g(y − y 0)

SUMMARY
3.1 Position, Displacement and Average Velocity
• Kinematics is the description of motion without considering its causes. In this chapter, it is limited to motion along
a straight line, called one-dimensional motion.
• Displacement is the change in position of an object. The SI unit for displacement is the meter. Displacement has
direction as well as magnitude.
• Distance traveled is the total length of the path traveled between two positions.
• Time is measured in terms of change. The time between two position points x 1 and x 2 is Δt = t 2 − t 1 . Elapsed
time for an event is Δt = t f − t 0 , where t f is the final time and t 0 is the initial time. The initial time is often taken
to be zero.

• Average velocity v is defined as displacement divided by elapsed time. If x 1, t 1 and x 2, t 2 are two position time
points, the average velocity between these points is
x −x
v = Δx = t 2 − t 1 .

Δt 2 1

3.2 Instantaneous Velocity and Speed


• Instantaneous velocity gives the velocity at any point in time during a particle’s motion.
• Instantaneous velocity is a vector and can be negative.
• Instantaneous speed is found by taking the absolute value of instantaneous velocity, and it is always positive.
• Average speed is total distance traveled divided by elapsed time.
• The slope of a position-versus-time graph at a specific time gives instantaneous velocity at that time.

3.3 Average and Instantaneous Acceleration


• Acceleration is the rate at which velocity changes. Acceleration is a vector; it has both a magnitude and direction.
The SI unit for acceleration is meters per second squared.
• Acceleration can be caused by a change in the magnitude or the direction of the velocity, or both.
• Instantaneous acceleration a(t) is a continuous function of time and gives the acceleration at any specific time during
the motion. It is calculated from the derivative of the velocity function. Instantaneous acceleration is the slope of
the velocity-versus-time graph.
• Negative acceleration (sometimes called deceleration) is acceleration in the negative direction in the chosen
coordinate system.

3.4 Motion with Constant Acceleration


• When analyzing one-dimensional motion with constant acceleration, identify the known quantities and choose the
appropriate equations to solve for the unknowns. Either one or two of the kinematic equations are needed to solve
for the unknowns, depending on the known and unknown quantities.

This OpenStax book is available for free at https://legacy.cnx.org/content/col12169/1.46


Chapter 3 | Straight-Line Motion 99

• Two-body pursuit problems always require two equations to be solved simultaneously for the unknowns.

3.5 Free Fall


• An object in free fall experiences constant acceleration if air resistance is negligible.
• On Earth, all free-falling objects have an acceleration g due to gravity, which averages g = 9.81 m/s 2 .

• For objects in free fall, the upward direction is normally taken as positive for displacement, velocity, and
acceleration.

CONCEPTUAL QUESTIONS
quantities the same?
3.1 Position, Displacement and Average
Velocity 10. How are instantaneous velocity and instantaneous
speed related to one another? How do they differ?
1. Give an example in which there are clear distinctions
among distance traveled, displacement, and magnitude of
displacement. Identify each quantity in your example
specifically. 3.3 Average and Instantaneous Acceleration
11. Is it possible for speed to be constant while
2. Under what circumstances does distance traveled equal acceleration is not zero?
magnitude of displacement? What is the only case in which
magnitude of displacement and displacement are exactly 12. Is it possible for velocity to be constant while
the same? acceleration is not zero? Explain.

3. Bacteria move back and forth using their flagella 13. Give an example in which velocity is zero yet
(structures that look like little tails). Speeds of up to 50 μm/ acceleration is not.
s (50 × 10−6 m/s) have been observed. The total distance
traveled by a bacterium is large for its size, whereas its 14. If a subway train is moving to the left (has a negative
displacement is small. Why is this? velocity) and then comes to a stop, what is the direction of
its acceleration? Is the acceleration positive or negative?
4. Give an example of a device used to measure time and
identify what change in that device indicates a change in 15. Plus and minus signs are used in one-dimensional
time. motion to indicate direction. What is the sign of an
acceleration that reduces the magnitude of a negative
5. Does a car’s odometer measure distance traveled or velocity? Of a positive velocity?
displacement?

6. During a given time interval the average velocity of 3.4 Motion with Constant Acceleration
an object is zero. What can you say conclude about its 16. When analyzing the motion of a single object, what
displacement over the time interval? is the required number of known physical variables that
are needed to solve for the unknown quantities using the
kinematic equations?
3.2 Instantaneous Velocity and Speed
7. There is a distinction between average speed and the 17. State two scenarios of the kinematics of single object
magnitude of average velocity. Give an example that where three known quantities require two kinematic
illustrates the difference between these two quantities. equations to solve for the unknowns.

8. Does the speedometer of a car measure speed or


velocity? 3.5 Free Fall
18. What is the acceleration of a rock thrown straight
9. If you divide the total distance traveled on a car trip upward on the way up? At the top of its flight? On the way
(as determined by the odometer) by the elapsed time of down? Assume there is no air resistance.
the trip, are you calculating average speed or magnitude of
average velocity? Under what circumstances are these two 19. An object that is thrown straight up falls back to Earth.
100 Chapter 3 | Straight-Line Motion

This is one-dimensional motion. (a) When is its velocity 21. The severity of a fall depends on your speed when
zero? (b) Does its velocity change direction? (c) Does the you strike the ground. All factors but the acceleration from
acceleration have the same sign on the way up as on the gravity being the same, how many times higher could a safe
way down? fall on the Moon than on Earth (gravitational acceleration
on the Moon is about one-sixth that of the Earth)?
20. Suppose you throw a rock nearly straight up at a
coconut in a palm tree and the rock just misses the coconut 22. How many times higher could an astronaut jump on
on the way up but hits the coconut on the way down. the Moon than on Earth if her takeoff speed is the same
Neglecting air resistance and the slight horizontal variation in both locations (gravitational acceleration on the Moon is
in motion to account for the hit and miss of the coconut, about on-sixth of that on Earth)?
how does the speed of the rock when it hits the coconut
on the way down compare with what it would have been if 23. When given the acceleration function, what additional
it had hit the coconut on the way up? Is it more likely to information is needed to find the velocity function and
dislodge the coconut on the way up or down? Explain. position function?

PROBLEMS
average velocity of the blast wave? b) Compare this with
3.1 Position, Displacement and Average the speed of sound, which is 343 m/s at sea level.
Velocity
24. Consider a coordinate system in which the positive x
3.2 Instantaneous Velocity and Speed
axis is directed upward vertically. What are the positions of
a particle (a) 5.0 m directly above the origin and (b) 2.0 m 30. A woodchuck runs 20 m to the right in 5 s, then turns
below the origin? and runs 10 m to the left in 3 s. (a) What is the average
velocity of the woodchuck? (b) What is its average speed?
25. A car is 2.0 km west of a traffic light at t = 0 and 5.0
km east of the light at t = 6.0 min. Assume the origin of the 31. Sketch the velocity-versus-time graph from the
coordinate system is the light and the positive x direction following position-versus-time graph.
is eastward. (a) What are the car’s position vectors at these
two times? (b) What is the car’s displacement between 0
min and 6.0 min?

26. The Shanghai maglev train connects Longyang Road


to Pudong International Airport, a distance of 30 km. The
journey takes 8 minutes on average. What is the maglev
train’s average velocity?

27. The position of a particle moving along the x-axis is


given by x(t) = 4.0 − 2.0t m. (a) At what time does the
particle cross the origin? (b) What is the displacement of
the particle between t = 3.0 s and t = 6.0 s ?

28. A cyclist rides 8.0 km east for 20 minutes, then he


turns and heads west for 8 minutes and 3.2 km. Finally,
he rides east for 16 km, which takes 40 minutes. (a) What
is the final displacement of the cyclist? (b) What is his
average velocity? 32. Sketch the velocity-versus-time graph from the
following position-versus-time graph.
29. On February 15, 2013, a superbolide meteor (brighter
than the Sun) entered Earth’s atmosphere over
Chelyabinsk, Russia, and exploded at an altitude of 23.5
km. Eyewitnesses could feel the intense heat from the
fireball, and the blast wave from the explosion blew out
windows in buildings. The blast wave took approximately 2
minutes 30 seconds to reach ground level. (a) What was the

This OpenStax book is available for free at https://legacy.cnx.org/content/col12169/1.46


Chapter 3 | Straight-Line Motion 101

33. Given the following velocity-versus-time graph,


sketch the position-versus-time graph.

37. A commuter backs her car out of her garage with an


acceleration of 1.40 m/s2. (a) How long does it take her to
reach a speed of 2.00 m/s? (b) If she then brakes to a stop
in 0.800 s, what is her acceleration?

38. Assume an intercontinental ballistic missile goes from


rest to a suborbital speed of 6.50 km/s in 60.0 s (the actual
speed and time are classified). What is its average
acceleration in meters per second and in multiples of g
(9.80 m/s2)?

39. An airplane, starting from rest, moves down the


3.3 Average and Instantaneous Acceleration runway at constant acceleration for 18 s and then takes off
34. A cheetah can accelerate from rest to a speed of 30.0 at a speed of 60 m/s. What is the average acceleration of the
m/s in 7.00 s. What is its acceleration? plane?

35. Dr. John Paul Stapp was a U.S. Air Force officer who
studied the effects of extreme acceleration on the human 3.4 Motion with Constant Acceleration
body. On December 10, 1954, Stapp rode a rocket sled, 40. A particle moves in a straight line at a constant
accelerating from rest to a top speed of 282 m/s (1015 velocity of 30 m/s. What is its displacement between t = 0
km/h) in 5.00 s and was brought jarringly back to rest in and t = 5.0 s?
only 1.40 s. Calculate his (a) acceleration in his direction
of motion and (b) acceleration opposite to his direction 41. A particle moves in a straight line with an initial
of motion. Express each in multiples of g (9.80 m/s2) by velocity of 30 m/s and a constant acceleration of 30 m/s2. If
taking its ratio to the acceleration of gravity. at t = 0, x = 0 and v = 0 , what is the particle’s position
at t = 5 s?
36. Sketch the acceleration-versus-time graph from the
following velocity-versus-time graph.
42. A particle moves in a straight line with an initial
velocity of 30 m/s and constant acceleration 30 m/s2. (a)
What is its displacement at t = 5 s? (b) What is its velocity
at this same time?

43. (a) Sketch a graph of velocity versus time


corresponding to the graph of displacement versus time
given in the following figure. (b) Identify the time or times
(ta, tb, tc, etc.) at which the instantaneous velocity has the
greatest positive value. (c) At which times is it zero? (d) At
102 Chapter 3 | Straight-Line Motion

which times is it negative? 49. (a) A light-rail commuter train accelerates at a rate of
1.35 m/s2. How long does it take to reach its top speed of
80.0 km/h, starting from rest? (b) The same train ordinarily
decelerates at a rate of 1.65 m/s2. How long does it take
to come to a stop from its top speed? (c) In emergencies,
the train can decelerate more rapidly, coming to rest from
80.0 km/h in 8.30 s. What is its emergency acceleration in
meters per second squared?

50. While entering a freeway, a car accelerates from rest


at a rate of 2.04 m/s2 for 12.0 s. (a) Draw a sketch of the
situation. (b) List the knowns in this problem. (c) How
far does the car travel in those 12.0 s? To solve this part,
first identify the unknown, then indicate how you chose
the appropriate equation to solve for it. After choosing
the equation, show your steps in solving for the unknown,
44. (a) Sketch a graph of acceleration versus time check your units, and discuss whether the answer is
corresponding to the graph of velocity versus time given in reasonable. (d) What is the car’s final velocity? Solve for
the following figure. (b) Identify the time or times (ta, tb, this unknown in the same manner as in (c), showing all
tc, etc.) at which the acceleration has the greatest positive steps explicitly.
value. (c) At which times is it zero? (d) At which times is it
negative? 51. Unreasonable results At the end of a race, a runner
decelerates from a velocity of 9.00 m/s at a rate of 2.00 m/
s2. (a) How far does she travel in the next 5.00 s? (b) What
is her final velocity? (c) Evaluate the result. Does it make
sense?

52. Blood is accelerated from rest to 30.0 cm/s in a


distance of 1.80 cm by the left ventricle of the heart. (a)
Make a sketch of the situation. (b) List the knowns in
this problem. (c) How long does the acceleration take? To
solve this part, first identify the unknown, then discuss how
you chose the appropriate equation to solve for it. After
choosing the equation, show your steps in solving for the
unknown, checking your units. (d) Is the answer reasonable
45. A particle has a constant acceleration of 6.0 m/s2. when compared with the time for a heartbeat?
(a) If its initial velocity is 2.0 m/s, at what time is its
displacement 5.0 m? (b) What is its velocity at that time? 53. During a slap shot, a hockey player accelerates the
puck from a velocity of 8.00 m/s to 40.0 m/s in the same
46. At t = 10 s, a particle is moving from left to right direction. If this shot takes 3.33 × 10 −2 s , what is the
with a speed of 5.0 m/s. At t = 20 s, the particle is moving
distance over which the puck accelerates?
right to left with a speed of 8.0 m/s. Assuming the particle’s
acceleration is constant, determine (a) its acceleration, (b)
its initial velocity, and (c) the instant when its velocity is 54. A powerful motorcycle can accelerate from rest to
zero. 26.8 m/s (100 km/h) in only 3.90 s. (a) What is its average
acceleration? (b) How far does it travel in that time?
47. A well-thrown ball is caught in a well-padded mitt. If
55. Freight trains can produce only relatively small
the acceleration of the ball is 2.10 × 10 4 m/s 2 , and 1.85
accelerations. (a) What is the final velocity of a freight train
ms (1 ms = 10 −3 s) elapses from the time the ball first that accelerates at a rate of 0.0500 m/s 2 for 8.00 min,
touches the mitt until it stops, what is the initial velocity of starting with an initial velocity of 4.00 m/s? (b) If the train
the ball?
can slow down at a rate of 0.550 m/s 2 , how long will it
take to come to a stop from this velocity? (c) How far will
48. A bullet in a gun is accelerated from the firing it travel in each case?
chamber to the end of the barrel at an average rate of
6.20 × 10 5 m/s 2 for 8.10 × 10 −4 s . What is its muzzle 56. A fireworks shell is accelerated from rest to a velocity
velocity (that is, its final velocity)? of 65.0 m/s over a distance of 0.250 m. (a) Calculate the

This OpenStax book is available for free at https://legacy.cnx.org/content/col12169/1.46


Chapter 3 | Straight-Line Motion 103

acceleration. (b) How long did the acceleration last?


3.5 Free Fall
57. A swan on a lake gets airborne by flapping its wings 63. Calculate the displacement and velocity at times of
and running on top of the water. (a) If the swan must reach (a) 0.500 s, (b) 1.00 s, (c) 1.50 s, and (d) 2.00 s for a ball
a velocity of 6.00 m/s to take off and it accelerates from thrown straight up with an initial velocity of 15.0 m/s. Take
rest at an average rate of 0.35 m/s 2 , how far will it travel the point of release to be y 0 = 0 .
before becoming airborne? (b) How long does this take?
64. Calculate the displacement and velocity at times of (a)
58. A woodpecker’s brain is specially protected from large 0.500 s, (b) 1.00 s, (c) 1.50 s, (d) 2.00 s, and (e) 2.50 s for
accelerations by tendon-like attachments inside the skull. a rock thrown straight down with an initial velocity of 14.0
While pecking on a tree, the woodpecker’s head comes to m/s from the Verrazano Narrows Bridge in New York City.
a stop from an initial velocity of 0.600 m/s in a distance The roadway of this bridge is 70.0 m above the water.
of only 2.00 mm. (a) Find the acceleration in meters per
second squared and in multiples of g, where g = 9.80 m/ 65. A basketball referee tosses the ball straight up for the
s2. (b) Calculate the stopping time. (c) The tendons cradling starting tip-off. At what velocity must a basketball player
the brain stretch, making its stopping distance 4.50 mm leave the ground to rise 1.25 m above the floor in an
(greater than the head and, hence, less acceleration of the attempt to get the ball?
brain). What is the brain’s acceleration, expressed in
multiples of g?
66. A rescue helicopter is hovering over a person whose
boat has sunk. One of the rescuers throws a life preserver
59. An unwary football player collides with a padded straight down to the victim with an initial velocity of 1.40
goalpost while running at a velocity of 7.50 m/s and comes m/s and observes that it takes 1.8 s to reach the water. (a)
to a full stop after compressing the padding and his body List the knowns in this problem. (b) How high above the
0.350 m. (a) What is his acceleration? (b) How long does water was the preserver released? Note that the downdraft
the collision last? of the helicopter reduces the effects of air resistance on the
falling life preserver, so that an acceleration equal to that of
60. A care package is dropped out of a cargo plane and gravity is reasonable.
lands in the forest. If we assume the care package speed on
impact is 54 m/s (123 mph), then what is its acceleration? 67. Unreasonable results A dolphin in an aquatic show
Assume the trees and snow stops it over a distance of 3.0 jumps straight up out of the water at a velocity of 15.0
m. m/s. (a) List the knowns in this problem. (b) How high
does his body rise above the water? To solve this part, first
61. An express train passes through a station. It enters note that the final velocity is now a known, and identify
with an initial velocity of 22.0 m/s and decelerates at a rate its value. Then, identify the unknown and discuss how
of 0.150 m/s 2 as it goes through. The station is 210.0 m you chose the appropriate equation to solve for it. After
long. (a) How fast is it going when the nose leaves the choosing the equation, show your steps in solving for the
station? (b) How long is the nose of the train in the station? unknown, checking units, and discuss whether the answer
(c) If the train is 130 m long, what is the velocity of the end is reasonable. (c) How long a time is the dolphin in the air?
of the train as it leaves? (d) When does the end of the train Neglect any effects resulting from his size or orientation.
leave the station?
68. A diver bounces straight up from a diving board,
62. Unreasonable results Dragsters can actually reach a avoiding the diving board on the way down, and falls feet
top speed of 145.0 m/s in only 4.45 s. (a) Calculate the first into a pool. She starts with a velocity of 4.00 m/s and
average acceleration for such a dragster. (b) Find the final her takeoff point is 1.80 m above the pool. (a) What is her
velocity of this dragster starting from rest and accelerating highest point above the board? (b) How long a time are her
at the rate found in (a) for 402.0 m (a quarter mile) without feet in the air? (c) What is her velocity when her feet hit the
using any information on time. (c) Why is the final velocity water?
greater than that used to find the average acceleration?
(Hint: Consider whether the assumption of constant 69. (a) Calculate the height of a cliff if it takes 2.35 s for
acceleration is valid for a dragster. If not, discuss whether a rock to hit the ground when it is thrown straight up from
the acceleration would be greater at the beginning or end the cliff with an initial velocity of 8.00 m/s. (b) How long a
of the run and what effect that would have on the final time would it take to reach the ground if it is thrown straight
velocity.) down with the same speed?

70. A very strong, but inept, shot putter puts the shot
straight up vertically with an initial velocity of 11.0 m/s.
104 Chapter 3 | Straight-Line Motion

How long a time does he have to get out of the way if the in Victoria, Australia, a hiker hears a rock break loose from
shot was released at a height of 2.20 m and he is 1.80 m a height of 105.0 m. He can’t see the rock right away, but
tall? then does, 1.50 s later. (a) How far above the hiker is the
rock when he can see it? (b) How much time does he have
71. You throw a ball straight up with an initial velocity of to move before the rock hits his head?
15.0 m/s. It passes a tree branch on the way up at a height
of 7.0 m. How much additional time elapses before the ball 74. There is a 250-m-high cliff at Half Dome in Yosemite
passes the tree branch on the way back down? National Park in California. Suppose a boulder breaks loose
from the top of this cliff. (a) How fast will it be going when
72. A kangaroo can jump over an object 2.50 m high. (a) it strikes the ground? (b) Assuming a reaction time of 0.300
Considering just its vertical motion, calculate its vertical s, how long a time will a tourist at the bottom have to get
speed when it leaves the ground. (b) How long a time is it out of the way after hearing the sound of the rock breaking
in the air? loose (neglecting the height of the tourist, which would
become negligible anyway if hit)? The speed of sound is
335.0 m/s on this day.
73. Standing at the base of one of the cliffs of Mt. Arapiles

ADDITIONAL PROBLEMS
75. Professional baseball player Nolan Ryan could pitch of 4.0 × 10 5 m/s. It enters a region 5.0 cm long where it
a baseball at approximately 160.0 km/h. At that average
velocity, how long did it take a ball thrown by Ryan to undergoes an acceleration of 6.0 × 10 12 m/s 2 along the
reach home plate, which is 18.4 m from the pitcher’s same straight line. (a) What is the electron’s velocity when
mound? Compare this with the average reaction time of a it emerges from this region? b) How long does the electron
human to a visual stimulus, which is 0.25 s. take to cross the region?

76. An airplane leaves Chicago and makes the 3000-km 82. An ambulance driver is rushing a patient to the
trip to Los Angeles in 5.0 h. A second plane leaves Chicago hospital. While traveling at 72 km/h, she notices the traffic
one-half hour later and arrives in Los Angeles at the same light at the upcoming intersections has turned amber. To
time. Compare the average velocities of the two planes. reach the intersection before the light turns red, she must
Ignore the curvature of Earth and the difference in altitude travel 50 m in 2.0 s. (a) What minimum acceleration must
between the two cities. the ambulance have to reach the intersection before the
light turns red? (b) What is the speed of the ambulance
77. Unreasonable Results A cyclist rides 16.0 km east, when it reaches the intersection?
then 8.0 km west, then 8.0 km east, then 32.0 km west,
and finally 11.2 km east. If his average velocity is 24 km/ 83. A motorcycle that is slowing down uniformly covers
h, how long did it take him to complete the trip? Is this a 2.0 successive km in 80 s and 120 s, respectively. Calculate
reasonable time? (a) the acceleration of the motorcycle and (b) its velocity at
the beginning and end of the 2-km trip.
78. An object has an acceleration of +1.2 cm/s 2 . At
84. A cyclist travels from point A to point B in 10 min.
t = 4.0 s , its velocity is −3.4 cm/s . Determine the During the first 2.0 min of her trip, she maintains a uniform
object’s velocities at t = 1.0 s and t = 6.0 s .
acceleration of 0.090 m/s 2 . She then travels at constant
velocity for the next 5.0 min. Next, she decelerates at a
79. A particle moving at constant acceleration has constant rate so that she comes to a rest at point B 3.0 min
velocities of 2.0 m/s at t = 2.0 s and −7.6 m/s at later. (a) Sketch the velocity-versus-time graph for the trip.
t = 5.2 s. What is the acceleration of the particle? (b) What is the acceleration during the last 3 min? (c) How
far does the cyclist travel?
80. A train is moving up a steep grade at constant velocity
(see following figure) when its caboose breaks loose and 85. Two trains are moving at 30 m/s in opposite directions
starts rolling freely along the track. After 5.0 s, the caboose on the same track. The engineers see simultaneously that
is 30 m behind the train. What is the acceleration of the they are on a collision course and apply the brakes when
caboose? they are 1000 m apart. Assuming both trains have the same
Figure shows a train moving up a hill. acceleration, what must this acceleration be if the trains are
to stop just short of colliding?
81. An electron is moving in a straight line with a velocity

This OpenStax book is available for free at https://legacy.cnx.org/content/col12169/1.46


Chapter 3 | Straight-Line Motion 105

86. A 10.0-m-long truck moving with a constant velocity 94. A coin is dropped from a hot-air balloon that is 300
of 97.0 km/h passes a 3.0-m-long car moving with a m above the ground and rising at 10.0 m/s upward. For the
constant velocity of 80.0 km/h. How much time elapses coin, find (a) the maximum height reached, (b) its position
between the moment the front of the truck is even with the and velocity 4.00 s after being released, and (c) the time
back of the car and the moment the back of the truck is even before it hits the ground.
with the front of the car?
Top drawing shows passenger car with a speed of 80 95. A soft tennis ball is dropped onto a hard floor from
kilometers per hour in front of the truck with the speed of a height of 1.50 m and rebounds to a height of 1.10 m.
97 kilometers per hour. Middle drawing shows passenger (a) Calculate its velocity just before it strikes the floor. (b)
car with a speed of 80 kilometers per hour parallel to the Calculate its velocity just after it leaves the floor on its way
truck with the speed of 97 kilometers per hour. Bottom back up. (c) Calculate its acceleration during contact with
drawing shows passenger car with a speed of 80 kilometers
the floor if that contact lasts 3.50 ms (3.50 × 10 −3 s) (d)
per hour behind the truck with a speed of 97 kilometers per
hour. How much did the ball compress during its collision with
the floor, assuming the floor is absolutely rigid?
87. A police car waits in hiding slightly off the highway.
A speeding car is spotted by the police car doing 40 m/s. At 96. Unreasonable results. A raindrop falls from a cloud
the instant the speeding car passes the police car, the police 100 m above the ground. Neglect air resistance. What is
car accelerates from rest at 4 m/s2 to catch the speeding car. the speed of the raindrop when it hits the ground? Is this a
How long does it take the police car to catch the speeding reasonable number?
car?
97. Compare the time in the air of a basketball player who
88. Pablo is running in a half marathon at a velocity of jumps 1.0 m vertically off the floor with that of a player
3 m/s. Another runner, Jacob, is 50 meters behind Pablo who jumps 0.3 m vertically.
with the same velocity. Jacob begins to accelerate at 0.05
m/s2. (a) How long does it take Jacob to catch Pablo? (b) 98. Suppose that a person takes 0.5 s to react and move his
What is the distance covered by Jacob? (c) What is the final hand to catch an object he has dropped. (a) How far does
velocity of the Jacob? the object fall on Earth, where g = 9.8 m/s 2 ? (b) How far
does the object fall on the Moon, where the acceleration
89. Unreasonable results A runner approaches the finish due to gravity is 1/6 of that on Earth?
line and is 75 m away; her average speed at this position is
8 m/s. She decelerates at this point at 0.5 m/s2. How long
does it take her to cross the finish line from 75 m away? Is 99. A hot-air balloon rises from ground level at a constant
this reasonable? velocity of 3.0 m/s. One minute after liftoff, a sandbag is
dropped accidentally from the balloon. Calculate (a) the
time it takes for the sandbag to reach the ground and (b) the
90. An airplane accelerates at 5.0 m/s2 for 30.0 s. During velocity of the sandbag when it hits the ground.
this time, it covers a distance of 10.0 km. What are the
initial and final velocities of the airplane?
100. (a) A world record was set for the men’s 100-m dash
in the 2008 Olympic Games in Beijing by Usain Bolt of
91. Compare the distance traveled of an object that Jamaica. Bolt “coasted” across the finish line with a time
undergoes a change in velocity that is twice its initial of 9.69 s. If we assume that Bolt accelerated for 3.00 s to
velocity with an object that changes its velocity by four reach his maximum speed, and maintained that speed for
times its initial velocity over the same time period. The the rest of the race, calculate his maximum speed and his
accelerations of both objects are constant. acceleration. (b) During the same Olympics, Bolt also set
the world record in the 200-m dash with a time of 19.30 s.
92. An object is moving east with a constant velocity and Using the same assumptions as for the 100-m dash, what
is at position x 0 at time t 0 = 0 . (a) With what acceleration was his maximum speed for this race?
must the object have for its total displacement to be zero at
a later time t ? (b) What is the physical interpretation of the 101. An object is dropped from a height of 75.0 m above
solution in the case for t → ∞ ? ground level. (a) Determine the distance traveled during the
first second. (b) Determine the final velocity at which the
object hits the ground. (c) Determine the distance traveled
93. A ball is thrown straight up. It passes a 2.00-m-high
during the last second of motion before hitting the ground.
window 7.50 m off the ground on its path up and takes
1.30 s to go past the window. What was the ball’s initial
velocity? 102. A steel ball is dropped onto a hard floor from a height
of 1.50 m and rebounds to a height of 1.45 m. (a) Calculate
its velocity just before it strikes the floor. (b) Calculate its
106 Chapter 3 | Straight-Line Motion

velocity just after it leaves the floor on its way back up. 103. An object is dropped from a roof of a building of
(c) Calculate its acceleration during contact with the floor height h. During the last second of its descent, it drops a
if that contact lasts 0.0800 ms (8.00 × 10 −5 s) (d) How distance h/3. Calculate the height of the building.
much did the ball compress during its collision with the
floor, assuming the floor is absolutely rigid?

CHALLENGE PROBLEMS
104. In a 100-m race, the winner is timed at 11.2 s. The traveled at 11.8 m/s until the finish line. What was the
second-place finisher’s time is 11.6 s. How far is the difference in finish time in seconds between the winner
second-place finisher behind the winner when she crosses and runner-up? How far back was the runner-up when the
the finish line? Assume the velocity of each runner is winner crossed the finish line?
constant throughout the race.
106. In 1967, New Zealander Burt Munro set the world
105. A cyclist sprints at the end of a race to clinch a record for an Indian motorcycle, on the Bonneville Salt
victory. She has an initial velocity of 11.5 m/s and Flats in Utah, of 295.38 km/h. The one-way course was
accelerates at a rate of 0.500 m/s2 for 7.00 s. (a) What is her 8.00 km long. Acceleration rates are often described by the
final velocity? (b) The cyclist continues at this velocity to time it takes to reach 96.0 km/h from rest. If this time was
the finish line. If she is 300 m from the finish line when she 4.00 s and Burt accelerated at this rate until he reached his
starts to accelerate, how much time did she save? (c) The maximum speed, how long did it take Burt to complete the
second-place winner was 5.00 m ahead when the winner course?
started to accelerate, but he was unable to accelerate, and

This OpenStax book is available for free at https://legacy.cnx.org/content/col12169/1.46


Chapter 4 | Circular Motion as One-Dimensional Motion 107

4 | CIRCULAR MOTION AS
ONE-DIMENSIONAL
MOTION

Figure 4.1 Brazos wind farm in west Texas. As of 2012, wind farms in the US had a power output of 60 gigawatts, enough
capacity to power 15 million homes for a year. (credit: modification of work by “ENERGY.GOV”/Flickr)

Chapter Outline
4.1 Rotational Variables
4.2 Rotation with Constant Angular Acceleration
4.3 Relating Angular and Translational Quantities

Introduction
In Straight-Line Motion, we described motion (kinematics) in one dimension, and introduced important concepts like
displacement, velocity and acceleration. The type of motion we considered there is called translational motion, because the
moving object undergoes a translation from place to place in space. However, we know from everyday life that rotational
motion is also very important and that many objects that move have both translation and rotation. The wind turbines in our
chapter opening image are a prime example of how rotational motion impacts our daily lives, as the market for clean energy
sources continues to grow.
As we shall discover in Section 8.4, the objects that we study in astronomy often move (translate) in paths that follow
the shapes of conic sections (ellipses, parabolas, hyperbolas or circles). Those paths are the mathematical solutions to the
equations of motion which arise in the study of objects subjected to the force of gravity. A circular path is the simplest,
special case of this kind of motion.
Before considering rotating objects, we will begin by examining the idealization of a single point object moving in a path
that is shaped as a perfect circle. Of course, such an object is actually translating in a two dimensional manner. But, as we
will see, a mathematical analogy will allow us to describe that object using the kinematic equations of one dimensional
motion.
108 Chapter 4 | Circular Motion as One-Dimensional Motion

Figure 4.2 shows the orbital paths of some objects in our solar system. The motions of the planets around the Sun, although
they actually move along elliptical paths, can (at least initially) be approximated as being circular motion.

Figure 4.2 We see the orbits of typical comets and asteroids compared with those of the
planets Mercury, Venus, Earth, Mars, and Jupiter (shown in black). Shown in red are three
comets: Halley, Kopff, and Encke. In blue are the four largest asteroids: Ceres, Pallas, Vesta,
and Hygeia. While the orbits of the comets and asteroids are obviously elliptical, those of the
planets can be well approximated as circular.

Finally, to conclude the chapter, we will address the fixed-axis rotation of an extended object. Fixed-axis rotation describes
the rotation around a fixed axis of a rigid body; that is, an object that does not deform as it moves. We will show how to
apply all of the ideas we’ve developed up to this point about rotational and translational motion to such an object rotating
around a fixed axis.

4.1 | Rotational Variables


Learning Objectives
By the end of this section, you will be able to:
• Describe the physical meaning of rotational variables as applied to fixed-axis rotation
• Explain how angular velocity is related to tangential speed
• Calculate the instantaneous angular velocity given the angular position function
• Find the angular velocity and angular acceleration in a rotating system
• Calculate the average angular acceleration when the angular velocity is changing
• Calculate the instantaneous angular acceleration given the angular velocity function

So far in this text, we have only studied translational motion, including the variables that describe it: displacement, velocity,

This OpenStax book is available for free at https://legacy.cnx.org/content/col12169/1.46


Chapter 4 | Circular Motion as One-Dimensional Motion 109

and acceleration. Now we expand our description of motion to rotation—specifically, rotational motion about a fixed axis.
We will find that rotational motion is described by a set of related variables similar to those we used in translational motion.

Angular Velocity
Circular motion is, simply, motion in a perfectly circular path. Although this is the simplest case of rotational motion, it is
very useful for many situations, and we use it here to introduce rotational variables.
In Figure 4.3, we show (in red) a particle moving in a circle. The coordinate system is fixed and serves as a frame of
reference to define the particle’s position. Its position from the origin of the circle to the particle sweeps out the angle θ
, which increases in the counterclockwise direction as the particle moves along its circular path. The angle θ is called the
angular position of the particle. As the particle moves in its circular path, it also traces an arc length s.

Figure 4.3 A particle follows a circular path. As it moves


counterclockwise, it sweeps out a positive angle θ with respect
to the x-axis and traces out an arc length s.

The angle is related to the radius of the circle and the arc length by

θ = rs . (4.1)

The angle θ , the angular position of the particle along its path, has units of radians (rad). There are 2π radians in 360°.
Note that the radian measure is a ratio of length measurements, and therefore is a dimensionless quantity. As the particle
moves along its circular path, its angular position changes and it undergoes angular displacements Δθ.
Note again that the angle θ is measured counterclockwise starting from the positive x axis.
Now, let's note an interesting bit of mathematical simplification. Although this particle is moving in two dimensions, x and
y, we know that those coordinates can be written as:
x = rcosθ (4.2)
y = rsinθ (4.3)

However, the fact that the particle is constrained to move along a circular path, where r remains constant, means that its
only degree of motional freedom is the angle θ. So, in a mathematical sense, it is only moving in one dimension, and that
dimension is represented by the coordinate θ.
We can thus describe its location completely by simply stating the value of the coordinate θ, its angular position. Its motion
will be completely described by determining its angular displacements, Δθ , as time goes on. This means that we have
a system that is mathematically analogous to the one studied in Chapter 3, the only difference being that, there, the
dimension was called x (or y). Here, the dimension is called θ.
110 Chapter 4 | Circular Motion as One-Dimensional Motion

The magnitude of the angular velocity, denoted by ω , is the time rate of change of the angle θ as the particle moves in its
circular path. The instantaneous angular velocity is defined as the limit in which Δt → 0 in the average angular velocity
– = Δθ :
ω
Δt

ω = lim Δθ = dθ , (4.4)
Δt → 0 Δt dt

where θ is the angle of rotation (Figure 4.3). The units of angular velocity are radians per second (rad/s). Angular velocity
can also be referred to as the rotation rate in radians per second. In many situations, we are given the rotation rate in
revolutions/s or cycles/s. To find the angular velocity, we must multiply revolutions/s by 2π , since there are 2π radians in
one complete revolution. Since the direction of a positive angle in a circle is counterclockwise, we take counterclockwise
rotations as being positive and clockwise rotations as negative. (This choice is analogous to the one we made for horizontal,
translational motion. In that situation, moving from left to right was defined as moving in the positive motion, but moving
from right to left was in the negative direction.)

Connections to a Rotating Rigid Object


All of the arguments we have just made (and the mathematics we have developed) apply perfectly to the description of the
motion of any single point on an object that happens to be rotating about some fixed axis. Take a look at the rotating disk in
Figure 4.4. It shows two particles, 1 and 2, that are located at different distances ( r 1 and r 2 ) from the axis of rotation.

Figure 4.4 Two particles on a rotating disk

Which properties of motion are the same for these two points, and which are different? If you've ever ridden on a merry-go-
round, you know that the speed, v 2 , of a person located at point r 2 will be greater than the speed, v 1 , of a person located
at r 1 . Your experience tells you that, the farther you are from the axis of the rotating object, the faster you are moving. We
call v 1 and v 2 the tangential speeds, because they are moving instantaneously in a direction that is tangent to the circle
of their motion.
But, how do the angular speeds of the two particles (or people) compare? Since it will take each one exactly the same
amount of time to complete one rotation, their angular speed, ω , must be the same.
We can see how angular velocity is related to the tangential speed of the particle by going back to the definition of the
tangential displacement or arc length
s = rθ.
Noting that the radius r is a constant, we have

This OpenStax book is available for free at https://legacy.cnx.org/content/col12169/1.46


Chapter 4 | Circular Motion as One-Dimensional Motion 111

v t = Δs = rΔθ = r Δθ = rω
Δt Δt Δt
So, simply put, for any point located at a distance r from the axis of rotation, the tangential speed

v t = rω. (4.5)

That is, the tangential speed of the particle is its angular velocity times the radius of the circle traced out by its motion.
From Equation 4.5, we see that the tangential speed of the particle increases with its distance from the axis of rotation
for a constant angular velocity. This effect is shown in Figure 4.4. So it is for our two particles placed at different radii on
a rotating disk with a constant angular velocity. As the disk rotates, the tangential speed increases linearly with the radius
from the axis of rotation. In Figure 4.4, we see that v 1 = r 1 ω 1 and v 2 = r 2 ω 2 . But the disk has a constant angular
v v ⎛r ⎞
velocity, so ω 1 = ω 2 . This means r 1 = r 2 or v 2 = ⎝r 2 ⎠v 1 . Thus, since r 2 > r 1 , v 2 > v 1 .
1 2 1

Example 4.1

Rotation of a Flywheel
A flywheel rotates such that it sweeps out an angle at the rate of θ = ωt = (45.0 rad/s)t radians. The wheel
rotates counterclockwise when viewed in the plane of the page. (a) What is the angular velocity of the flywheel?
(b) How many radians does the flywheel rotate through in 30 s? (c) What is the tangential speed of a point on the
flywheel 10 cm from the axis of rotation?
Strategy
The functional form of the angular position of the flywheel is given in the problem as θ(t) = ωt , so we can find
the angular speed by inspection. It is just 45 rad/s. To find the angular displacement of the flywheel during 30
s, we seek the angular displacement Δθ , where the change in angular position is between 0 and 30 s. To find
the tangential speed of a point at a distance from the axis of rotation, we multiply its distance times the angular
velocity of the flywheel.
Solution
a. ω = 45 rad/s . We see that the angular velocity is a constant.
b. Δθ = θ(30 s) − θ(0 s) = 45.0(30.0 s) − 45.0(0 s) = 1350.0 rad .
c. v t = rω = (0.1 m)(45.0 rad/s) = 4.5 m/s .

Significance
In 30 s, the flywheel has rotated through quite a number of revolutions, about 215 if we divide the angular
displacement by 2π . A massive flywheel can be used to store energy in this way, if the losses due to friction are
minimal. Recent research has considered superconducting bearings on which the flywheel rests, with zero energy
loss due to friction.

Angular Acceleration
We have just discussed angular velocity for circular motion, but not all circular motion is at a uniform velocity. Envision
an ice skater spinning with his arms outstretched—when he pulls his arms inward, his angular velocity increases. Or think
about a computer’s hard disk slowing to a halt as the angular velocity decreases. We will explore these situations later, but
we can already see a need to define an angular acceleration for describing situations where ω changes. The faster the
change in ω , the greater the angular acceleration. We define the instantaneous angular acceleration α as the derivative
of angular velocity with respect to time:
2 (4.6)
α = lim Δω = dω = d 2θ ,
Δt → 0 Δt dt dt
112 Chapter 4 | Circular Motion as One-Dimensional Motion

– = Δω as Δt → 0 .
where we have taken the limit of the average angular acceleration, α
Δt

The units of angular acceleration are (rad/s)/s, or rad/s 2 .


We can relate the tangential acceleration of a point on a rotating body at a distance from the axis of rotation in the same way
that we related the tangential speed to the angular velocity. Again noting that the radius r is constant, we obtain
Δa t rΔω
= r Δω = rα
(4.7)
at = =
Δt Δt Δt
Therefore, for any point located at a distance r from the axis of rotation, the tangential acceleration is the distance from the
rotational axis times the angular acceleration.

a t = rα. (4.8)

Let’s apply these ideas to the analysis of a few simple fixed-axis rotation scenarios. Before doing so, we present a problem-
solving strategy that can be applied to rotational kinematics: the description of rotational motion.

Problem-Solving Strategy: Rotational Kinematics


1. Examine the situation to determine that rotational kinematics (rotational motion) is involved.
2. Identify exactly what needs to be determined in the problem (identify the unknowns). A sketch of the situation
is useful.
3. Make a complete list of what is given or can be inferred from the problem as stated (identify the knowns).
4. Solve the appropriate equation or equations for the quantity to be determined (the unknown). It can be useful
to think in terms of a translational analog, because by now you are familiar with the equations of translational
motion.
5. Substitute the known values along with their units into the appropriate equation and obtain numerical solutions
complete with units. Be sure to use units of radians for angles.
6. Check your answer to see if it is reasonable: Does your answer make sense?

Now let’s apply this problem-solving strategy to a few specific examples.

Example 4.2

A Spinning Bicycle Wheel


A bicycle mechanic mounts a bicycle on the repair stand and starts the rear wheel spinning from rest to a final
angular velocity of 250 rpm in 5.00 s. (a) Calculate the average angular acceleration in rad/s 2 . (b) If she now
hits the brakes, causing an angular acceleration of −87.3 rad/s 2 , how long does it take the wheel to stop?
Strategy
– = Δω because the final angular
The average angular acceleration can be found directly from its definition α
Δt
velocity and time are given. We see that Δω = ω final − ω initial = 250 rev/min and Δt is 5.00 s. For part (b),
we know the angular acceleration and the initial angular velocity. We can find the stopping time by using the
definition of average angular acceleration and solving for Δt , yielding

Δt = Δω
α .

This OpenStax book is available for free at https://legacy.cnx.org/content/col12169/1.46


Chapter 4 | Circular Motion as One-Dimensional Motion 113

Solution
a. Entering known information into the definition of angular acceleration, we get
– = Δω = 250 rpm .
α
Δt 5.00 s

Because Δω is in revolutions per minute (rpm) and we want the standard units of rad/s 2 for angular
acceleration, we need to convert from rpm to rad/s:

Δω = 250 rev · 2πrev


rad · 1 min = 26.2 rad .
s
min 60 s

Entering this quantity into the expression for α , we get

α = Δω = 26.2 rad/s = 5.24 rad/s 2.


Δt 5.00 s
b. Here the angular velocity decreases from 26.2 rad/s (250 rpm) to zero, so that Δω is −26.2 rad/s, and α
is given to be –87.3 rad/s 2 . Thus,

Δt = −26.2 rad/s2 = 0.300 s.


−87.3 rad/s
Significance
Note that the angular acceleration as the mechanic spins the wheel is small and positive; it takes 5 s to produce an
appreciable angular velocity. When she hits the brake, the angular acceleration is large and negative. The angular
velocity quickly goes to zero.

4.1 Check Your Understanding The fan blades on a turbofan jet engine (shown below) accelerate from rest
up to a rotation rate of 40.0 rev/s in 20 s. The increase in angular velocity of the fan is constant in time. (The
GE90-110B1 turbofan engine mounted on a Boeing 777, as shown, is currently the largest turbofan engine in
the world, capable of thrusts of 330–510 kN.)
(a) What is the average angular acceleration?
(b) What is the instantaneous angular acceleration at any time during the first 20 s?

Figure 4.5 (credit: “Bubinator”/ Wikimedia Commons)

We now have a basic vocabulary for discussing fixed-axis rotational kinematics and relationships between rotational
variables. We discuss more definitions and connections in the next section.
114 Chapter 4 | Circular Motion as One-Dimensional Motion

4.2 | Rotation with Constant Angular Acceleration


Learning Objectives
By the end of this section, you will be able to:
• Select from the kinematic equations for rotational motion with constant angular acceleration the
appropriate equations to solve for unknowns in the analysis of systems undergoing fixed-axis
rotation

In the preceding section, we defined the rotational variables of angular displacement, angular velocity, and angular
acceleration. In this section, we work with these definitions to derive relationships among these variables and use these
relationships to analyze rotational motion for a rigid body about a fixed axis under a constant angular acceleration. (By
the analogy previously established, these relationships also hold for any point-like object undergoing circular motion.)
This analysis forms the basis for rotational kinematics. If the angular acceleration is constant, the equations of rotational
kinematics simplify, similar to the equations of linear kinematics discussed in Section 3.4. We can then use this simplified
set of equations to describe many applications in physics and engineering where the angular acceleration of the system is
constant.

Kinematics of Rotational Motion


Using our intuition, we can begin to see how the rotational quantities θ, ω, α , and t are related to one another. For
example, we saw in the preceding section that if a flywheel has an angular acceleration in the same direction as its angular
velocity vector, its angular velocity increases with time and its angular displacement also increases. On the contrary, if the
angular acceleration is opposite to the angular velocity vector, its angular velocity decreases with time. We can describe
these physical situations and many others with a consistent set of rotational kinematic equations under a constant angular
acceleration. The method to investigate rotational motion in this way is called kinematics of rotational motion.
To begin, we note that if the system is rotating under a constant acceleration, then the average angular velocity follows a
simple relation because the angular velocity is increasing linearly with time. The average angular velocity is just half the
sum of the initial and final values:

– = ω0 + ωf .
ω (4.9)
2

From the definition of the average angular velocity, we can find an equation that relates the angular position, average
angular velocity, and time:
– = Δθ .
ω
Δt
Solving for θ , we have

– t,
θf = θ0 + ω (4.10)

where we have set t 0 = 0 . This equation can be very useful if we know the average angular velocity of the system. Then we
could find the angular displacement over a given time period. Next, we find an equation relating ω , α , and t. To determine
this equation, we start with the definition of angular acceleration:

α = Δω .
Δt
We rearrange this to get αΔt = Δω . In uniform rotational motion, the angular acceleration is constant so this becomes:
α(t − t 0) = ω f − ω 0 (4.11)

This OpenStax book is available for free at https://legacy.cnx.org/content/col12169/1.46


Chapter 4 | Circular Motion as One-Dimensional Motion 115

Setting t 0 = 0 , we have

αt = ω f − ω 0.

We rearrange this to obtain

ω f = ω 0 + αt, (4.12)

where ω 0 is the initial angular velocity. Equation 4.12 is the rotational counterpart to the linear kinematics equation
v f = v 0 + at . With Equation 4.12, we can find the angular velocity of an object at any specified time t given the initial
angular velocity and the angular acceleration.
Again by analogy to the discussion in Section 3.4, where we have set the initial time t 0 = 0 , we can obtain another
equation:

θ f = θ 0 + ω 0 t + 1 αt 2. (4.13)
2

Equation 4.13 is the rotational counterpart to the linear kinematics equation found in Section 3.4 for position as a
function of time. This equation gives us the angular position of a rotating rigid body at any time t given the initial conditions
(initial angular position and initial angular velocity) and the angular acceleration.
We can find an equation that is independent of time by solving for t in Equation 4.12 and substituting into Equation
4.13. Equation 4.13 becomes

⎛ω f − ω 0 ⎞ 1 ⎛ω f − ω 0 ⎞2
θf = θ0 + ω0 ⎝ α ⎠ + 2 α⎝ α ⎠
ω 0 ω f ω 20 1 ω 2f ω 0 ω f 1 ω 20
= θ0 + α − α +2 α − α +2 α
ω2 ω2
= θ 0 + 1 αf − 1 α0 ,
2 2
ω 2f − ω 20
θf − θ0 =

or

ω 2f = ω 20 + 2α(Δθ). (4.14)

Equation 4.10 through Equation 4.14 describe fixed-axis rotation for constant acceleration and are summarized in
Table 4.1.

Angular displacement from average angular velocity –t


θf = θ0 + ω

Angular velocity from angular acceleration ω f = ω 0 + αt

Angular displacement from angular velocity and angular acceleration θ f = θ 0 + ω 0 t + 1 αt 2


2
Angular velocity from angular displacement and angular acceleration ω 2f = ω 0 2 + 2α(Δθ)

Table 4.1 Kinematic Equations


116 Chapter 4 | Circular Motion as One-Dimensional Motion

Applying the Equations for Rotational Motion


Now we can apply the key kinematic relations for rotational motion to some simple examples to get a feel for how the
equations can be applied to everyday situations.

Example 4.3

Calculating the Acceleration of a Fishing Reel


A deep-sea fisherman hooks a big fish that swims away from the boat, pulling the fishing line from his fishing
reel. The whole system is initially at rest, and the fishing line unwinds from the reel at a radius of 4.50 cm from
its axis of rotation. The reel is given an angular acceleration of 110 rad/s 2 for 2.00 s (Figure 4.6).
(a) What is the final angular velocity of the reel after 2 s?
(b) How many revolutions does the reel make?

Figure 4.6 Fishing line coming off a rotating reel moves


linearly.

Strategy
Identify the knowns and compare with the kinematic equations for constant acceleration. Look for the appropriate
equation that can be solved for the unknown, using the knowns given in the problem description.
Solution
a. We are given α and t and want to determine ω . The most straightforward equation to use is
ω f = ω 0 + αt , since all terms are known besides the unknown variable we are looking for. We are given
that ω 0 = 0 (it starts from rest), so

ω f = 0 + (110 rad/s 2)(2.00 s) = 220 rad/s.


b. We are asked to find the number of revolutions. Because 1 rev = 2π rad , we can find the number of
revolutions by finding θ in radians. We are given α and t, and we know ω 0 is zero, so we can obtain
θ by using

θ f = θ i + ω i t + 1 αt 2
2
= 0 + 0 + (0.500)⎛⎝110 rad/s 2⎞⎠(2.00 s) 2 = 220 rad.

Converting radians to revolutions gives

Number of rev = (220 rad) 1 rev = 35.0 rev.


2π rad

This OpenStax book is available for free at https://legacy.cnx.org/content/col12169/1.46


Chapter 4 | Circular Motion as One-Dimensional Motion 117

Significance
This example illustrates that relationships among rotational quantities are highly analogous to those among linear
quantities. The answers to the questions are realistic. After unwinding for two seconds, the reel is found to spin at
220 rad/s, which is 2100 rpm. (No wonder reels sometimes make high-pitched sounds.)

In the preceding example, we considered a fishing reel with a positive angular acceleration. Now let us consider what
happens with a negative angular acceleration.

Example 4.4

Calculating the Duration When the Fishing Reel Slows Down and Stops
Now the fisherman applies a brake to the spinning reel, achieving an angular acceleration of −300 rad/s 2 . How
long does it take the reel to come to a stop?
Strategy
We are asked to find the time t for the reel to come to a stop. The initial and final conditions are different from
those in the previous problem, which involved the same fishing reel. Now we see that the initial angular velocity is
ω 0 = 220 rad/s and the final angular velocity ω is zero. The angular acceleration is given as α = −300 rad/s 2.
Examining the available equations, we see all quantities but t are known in ω f = ω 0 + αt , making it easiest to
use this equation.
Solution
The equation states
ω f = ω 0 + αt.

We solve the equation algebraically for t and then substitute the known values as usual, yielding
ω f − ω 0 0 − 220.0 rad/s
t= α = = 0.733 s.
−300.0 rad/s 2
Significance
Note that care must be taken with the signs that indicate the directions of various quantities. Also, note that the
time to stop the reel is fairly small because the acceleration is rather large. Fishing lines sometimes snap because
of the accelerations involved, and fishermen often let the fish swim for a while before applying brakes on the reel.
A tired fish is slower, requiring a smaller acceleration.

4.2 Check Your Understanding A centrifuge used in DNA extraction spins at a maximum rate of 7000 rpm,
producing a “g-force” on the sample that is 6000 times the force of gravity. If the centrifuge takes 10 seconds to
come to rest from the maximum spin rate: (a) What is the angular acceleration of the centrifuge? (b) What is the
angular displacement of the centrifuge during this time?

Example 4.5

Angular Acceleration of a Propeller


Figure 4.7 shows a graph of the angular velocity of a propeller on an aircraft as a function of time. Its angular
velocity starts at 30 rad/s and drops linearly to 0 rad/s over the course of 5 seconds. (a) Find the angular
acceleration of the object. (b) Find the angle through which the propeller rotates during these 5 seconds.
118 Chapter 4 | Circular Motion as One-Dimensional Motion

Figure 4.7 A graph of the angular velocity of a propeller


versus time.

Strategy
a. Since the angular velocity varies linearly with time, we know that the angular acceleration is constant and
does not depend on the time variable. The angular acceleration is the slope of the angular velocity vs. time
graph, α = dω . To calculate the slope, we read directly from Figure 4.7, and see that ω 0 = 30 rad/s
dt
at t = 0 s and ω f = 0 rad/s at t = 5 s .

b. We use the equation ω 2f = ω 20 + 2α(Δθ). to calculate the angular displacement, Δθ .

Solution
a. Calculating the slope, we get
ω−ω (0 − 30.0) rad/s
α= t−t 0 = = −6.0 rad/s 2.
0 (5.0 − 0) s

b. θ f = θ 0 + ω 0 t + 1 αt 2.
2

Setting θ 0 = 0 , we have

θ 0 = (30.0 rad/s)(5.0 s) + 1 (−6.0 rad/s 2)(5.0 rad/s) 2 = 150.0 − 75.0 = 75.0 rad.
2

4.3 | Relating Angular and Translational Quantities


Learning Objectives
By the end of this section, you will be able to:
• Given the linear kinematic equation, write the corresponding rotational kinematic equation
• Calculate the linear distances, velocities, and accelerations of points on a rotating system given
the angular velocities and accelerations

In this section, as we consider the motion of some point on a rotating object, we will relate each of the rotational variables

This OpenStax book is available for free at https://legacy.cnx.org/content/col12169/1.46


Chapter 4 | Circular Motion as One-Dimensional Motion 119

to the translational variables defined in Straight-Line Motion. This will complete our ability to describe rigid-body
rotations.

Angular vs. Linear Variables


In Section 4.1, we introduced angular kinematic variables. If we compare the angular definitions with the definitions
of linear kinematic variables from Straight-Line Motion, we find that there is a mapping of the linear variables to the
rotational ones. Linear position, velocity, and acceleration have their rotational counterparts, as we can see when we write
them side by side:

Linear Rotational
Position x θ
Velocity v = dx ω = dθ
dt dt
Acceleration a = dv α = dω
dt dt

Let’s compare the linear and rotational variables individually. The linear variable of position has physical units of meters,
whereas the angular position variable has dimensionless units of radians, as can be seen from the definition of θ = rs , which
is the ratio of two lengths. The linear velocity has units of m/s, and its counterpart, the angular velocity, has units of rad/
s. In Section 4.1, we saw in the case of circular motion that the linear tangential speed of a particle at a radius r from the
axis of rotation is related to the angular velocity by the relation v t = rω . This could also apply to points on a rigid body
rotating about a fixed axis. Here, we consider only circular motion. (In circular motion, both uniform and nonuniform, there
exists another acceleration, the centripetal acceleration, which will be discussed in Section 7.4.)

Relationships between Rotational and Translational Motion


We can look at two relationships between rotational and translational motion.
1. Generally speaking, the linear kinematic equations have their rotational counterparts. Table 4.2 lists the four linear
kinematic equations and the corresponding rotational counterpart. The two sets of equations look similar to each
other, but describe two different physical situations, that is, rotation and translation.

Rotational Translational
θ =θ + ω
f
–t
0 x = x + –v t
0

ω f = ω 0 + αt v f = v 0 + at

θ f = θ 0 + ω 0 t + 1 αt 2 x f = x 0 + v 0 t + 1 at 2
2 2

ω 2f = ω 20 + 2α(Δθ) v 2f = v 20 + 2a(Δx)

Table 4.2 Rotational and Translational Kinematic


Equations

2. The second correspondence has to do with relating linear and rotational variables in the special case of circular
motion. Importantly, any object moving in a circular path possesses both an angular velocity and a tangential linear
velocity. It possesses both an angular acceleration and a tangential linear acceleration. This is shown in Table 4.3,
where in the third column, we have listed the connecting equation that relates the linear variable to the rotational
variable for a rotating point located a distance r from its axis of rotation. The rotational variables of angular
velocity and acceleration have subscripts that indicate their definition in circular motion.
120 Chapter 4 | Circular Motion as One-Dimensional Motion

Rotational Translational Relationship ( r = radius)

θ s θ = rs

ω vt v
ω = rt

α at a
α = rt

Table 4.3 Rotational and Translational Quantities for an Object


in Circular Motion

Example 4.6

Linear Acceleration of a Centrifuge


A centrifuge has a radius of 20 cm and accelerates from a maximum rotation rate of 10,000 rpm to rest in 30
seconds under a constant angular acceleration. It is rotating counterclockwise. What is the magnitude of the
tangential acceleration of a point at the tip of the centrifuge?
Strategy
With the information given, we can calculate the angular acceleration, which then will allow us to find the
tangential acceleration.
Solution
The angular acceleration is
ω − ω 0 0 − (1.0 × 10 4)2π/60.0 s(rad/s)
α= t = = −34.9 rad/s 2.
30.0 s
Therefore, the tangential acceleration is
a t = rα = 0.2 m(−34.9 rad/s 2) = −7.0 m/s 2.

4.3 Check Your Understanding A boy jumps on a merry-go-round with a radius of 5 m that is at rest. It
starts accelerating at a constant rate up to an angular velocity of 5 rad/s in 20 seconds. What is the distance
travelled by the boy?

Check out this PhET simulation (https://openstaxcollege.org/l/21rotatingdisk) to change the parameters


of a rotating disk (the initial angle, angular velocity, and angular acceleration), and place bugs at different radial
distances from the axis. The simulation then lets you explore how circular motion relates to the bugs’ xy-position,
velocity, and acceleration using vectors or graphs.

This OpenStax book is available for free at https://legacy.cnx.org/content/col12169/1.46


Chapter 4 | Circular Motion as One-Dimensional Motion 121

CHAPTER 4 REVIEW
KEY TERMS
angular acceleration time rate of change of angular velocity
angular position angle a body has rotated through in a fixed coordinate system
angular velocity time rate of change of angular position
instantaneous angular acceleration derivative of angular velocity with respect to time
instantaneous angular velocity derivative of angular position with respect to time
kinematics of rotational motion describes the relationships among rotation angle, angular velocity, angular
acceleration, and time

SUMMARY
4.1 Rotational Variables
• The angular position θ of a rotating body is the angle the body has rotated through in a fixed coordinate system,
which serves as a frame of reference.
• The angular velocity of a rotating body about a fixed axis is defined as ω(rad/s) , the rotational rate of the body in
radians per second. The instantaneous angular velocity of a rotating body ω = lim Δω = dθ is the derivative
Δt → 0 Δt dt
with respect to time of the angular position θ , found by taking the limit Δt → 0 in the average angular velocity
– = Δθ . The angular velocity relates v to the tangential speed of a point on the rotating body through the
ω
Δt t

relation v t = rω , where r is the radius to the point and v t is the tangential speed at the given point.

• If the system’s angular velocity is not constant, then the system has an angular acceleration. The average angular
– = Δω . The
acceleration over a given time interval is the change in angular velocity over this time interval, α
Δt
instantaneous angular acceleration is the time derivative of angular velocity, α = lim Δω = dω . The sign of the
Δt → 0 Δt dt
angular acceleration α is found by examining the angular velocity. If a rotation rate of a rotating body is decreasing,
the angular acceleration is in the opposite direction to ω . If the rotation rate is increasing, the angular acceleration
is in the same direction as ω .
• The tangential acceleration of a point at a radius from the axis of rotation is the angular acceleration times the radius
to the point.

4.2 Rotation with Constant Angular Acceleration


• The kinematics of rotational motion describes the relationships among rotation angle (angular position), angular
velocity, angular acceleration, and time.
• For a constant angular acceleration, the angular velocity varies linearly. Therefore, the average angular velocity is
1/2 the initial plus final angular velocity over a given time period:
– = ω0 + ωf .
ω
2
• We derived a set of kinematics equations for rotational motion (with constant angular acceleration) that are
mathematically identical to the set of equations from Section 3.4 for straight-line motion (with constant
acceleration). The only differences are the substitution of the rotational-motion variables θ , ω and α for the
translational-motion variables x , v and a . The equations are summarized in the following table:
122 Chapter 4 | Circular Motion as One-Dimensional Motion

Displacement from average Velocity –t


θf = θ0 + ω x f = x 0 + –v t

Velocity from Acceleration ω f = ω 0 + αt v f = v 0 + at

Displacement from Velocity and Acceleration θ f = θ 0 + ω 0 t + 1 αt 2 x f = x 0 + v 0 t + 1 at 2


2 2
Velocity from Displacement and Acceleration ω 2f = ω 0 2 + 2α(Δθ) v 2f = v 0 2 + 2a(Δx)

Table 4.4 Comparison of the Kinematic Equations for Rotational and Translational Motion

4.3 Relating Angular and Translational Quantities


• The linear kinematic equations have their rotational counterparts such that there is a mapping
x → θ, v → ω, a → α .
• An object undergoing circular motion undergoes both angular and translational displacements, and possesses both
angular and tangential velocities and accelerations. Their interrelationships are shown in Table 4.3

CONCEPTUAL QUESTIONS
time variable?
4.1 Rotational Variables
1. A clock is mounted on the wall. As you look at it, what 5. If a rigid body has a constant angular acceleration, what
is the direction of the angular velocity vector of the second is the functional form of the angular position?
hand?
6. If the angular acceleration of a rigid body is zero, what
2. What is the value of the angular acceleration of the is the functional form of the angular velocity?
second hand of the clock on the wall?
7. A massless tether with a masses tied to both ends
3. A baseball bat is swung. Do all points on the bat have rotates about a fixed axis through the center. Can the total
the same angular velocity? The same tangential speed? acceleration of the tether/mass combination be zero if the
angular velocity is constant?

4.2 Rotation with Constant Angular


Acceleration
4. If a rigid body has a constant angular acceleration, what
is the functional form of the angular velocity in terms of the

PROBLEMS
11. A particle moves 3.0 m along a circle of radius 1.5
4.1 Rotational Variables m. (a) Through what angle does it rotate? (b) If the particle
8. Calculate the angular velocity of Earth. makes this trip in 1.0 s at a constant speed, what is its
angular velocity?
9. A track star runs a 400-m race on a 400-m circular track
in 45 s. What is his angular velocity assuming a constant 12. A compact disc rotates at 500 rev/min. If the diameter
speed? of the disc is 120 mm, (a) what is the tangential speed of a
point at the edge of the disc? (b) At a point halfway to the
center of the disc?
10. A wheel rotates at a constant rate of
2.0 × 10 3 rev/min . (a) What is its angular velocity in 13. Unreasonable results. The propeller of an aircraft is
radians per second? (b) Through what angle does it turn in spinning at 10 rev/s when the pilot shuts off the engine.
10 s? Express the solution in radians and degrees. The propeller reduces its angular velocity at a constant

This OpenStax book is available for free at https://legacy.cnx.org/content/col12169/1.46


Chapter 4 | Circular Motion as One-Dimensional Motion 123

2.0 rad/s 2 for a time period of 40 s. What is the rotation acceleration of 1.0 rad/s 2 ; at t = 0 its angular velocity
rate of the propeller in 40 s? Is this a reasonable situation? is 2.0 rad/s. (a) Determine the disk’s angular velocity at
t = 5.0 s . (b) What is the angle it has rotated through
14. A gyroscope slows from an initial rate of 32.0 rad/s at during this time? (c) What is the tangential acceleration of
a rate of 0.700 rad/s 2 . How long does it take to come to a point on the disk at t = 5.0 s ?
rest?
23. The angular velocity vs. time for a fan on a hovercraft
15. On takeoff, the propellers on a UAV (unmanned aerial is shown below. (a) What is the angle through which the fan
vehicle) increase their angular velocity from rest at a rate of blades rotate in the first 8 seconds? (b) Verify your result
ω = (25.0t) rad/s for 3.0 s. (a) What is the instantaneous using the kinematic equations.
angular velocity of the propellers at t = 2.0 s ? (b) What is
the angular acceleration?

4.2 Rotation with Constant Angular


Acceleration
16. A wheel has a constant angular acceleration of
5.0 rad/s 2 . Starting from rest, it turns through 300 rad.
(a) What is its final angular velocity? (b) How much time
elapses while it turns through the 300 radians? 24. A rod of length 20 cm has two beads attached to its
ends. The rod with beads starts rotating from rest. If the
beads are to have a tangential speed of 20 m/s in 7 s, what
17. During a 6.0-s time interval, a flywheel with a constant
is the angular acceleration of the rod to achieve this?
angular acceleration turns through 500 radians that acquire
an angular velocity of 100 rad/s. (a) What is the angular
velocity at the beginning of the 6.0 s? (b) What is the
angular acceleration of the flywheel? 4.3 Relating Angular and Translational
Quantities
18. The angular velocity of a rotating rigid body increases 25. At its peak, a tornado is 60.0 m in diameter and carries
from 500 to 1500 rev/min in 120 s. (a) What is the angular 500 km/h winds. What is its angular velocity in revolutions
acceleration of the body? (b) Through what angle does it per second?
turn in this 120 s?
26. An ultracentrifuge accelerates from rest to 100,000
19. A flywheel slows from 600 to 400 rev/min while rpm in 2.00 min. (a) What is the average angular
rotating through 40 revolutions. (a) What is the angular
acceleration of the flywheel? (b) How much time elapses acceleration in rad/s 2 ? (b) What is the tangential
during the 40 revolutions? acceleration of a point 9.50 cm from the axis of rotation?
(c) What is the total distance traveled by a point 9.5 cm
from the axis of rotation of the ultracentrifuge?
20. A wheel 1.0 m in diameter rotates with an angular
acceleration of 4.0 rad/s 2 . (a) If the wheel’s initial angular
27. A wind turbine is rotating counterclockwise at 0.5 rev/
velocity is 2.0 rad/s, what is its angular velocity after 10 s? s and slows to a stop in 10 s. Its blades are 20 m in length.
(b) Through what angle does it rotate in the 10-s interval? (a) What is the initial tangential speed of a point at the tip
(c) What are the tangential speed and acceleration of a point of one of its blades? (b) What is the angular acceleration of
on the rim of the wheel at the end of the 10-s interval? the turbine?

21. A vertical wheel with a diameter of 50 cm starts 28. What is (a) the angular speed and (b) the linear speed
from rest and rotates with a constant angular acceleration of a point on Earth’s surface at latitude 30° N. Take the
of 5.0 rad/s 2 around a fixed axis through its center radius of the Earth to be 6309 km. (c) At what latitude
counterclockwise. (a) Where is the point that is initially at would your linear speed be 10 m/s?
the bottom of the wheel at t = 10 s? (b) What is the point’s
linear acceleration at this instant? 29. A bicycle wheel with radius 0.3m rotates from rest
to 3 rev/s in 5 s. What is the magnitude of the tangential
22. A circular disk of radius 10 cm has a constant angular acceleration of a point on the outside edge of the wheel?
124 Chapter 4 | Circular Motion as One-Dimensional Motion

This OpenStax book is available for free at https://legacy.cnx.org/content/col12169/1.46


Chapter 5 | Relativistic Kinematics in One Dimension 125

5 | RELATIVISTIC
KINEMATICS IN ONE
DIMENSION
Chapter Outline
5.1 Invariance of Physical Laws
5.2 Relativity of Simultaneity
5.3 Time Dilation
5.4 Length Contraction
5.5 The Lorentz Transformation
5.6 Relativistic Velocity Transformation

Introduction

Figure 5.1 Special relativity explains how time passes slightly differently on Earth and within the rapidly moving global
positioning satellite (GPS). GPS units in vehicles could not find their correct location on Earth without taking this correction into
account. (credit: USAF)

Our description of motion (kinematics), thus far, has hopefully obeyed your common sense ideas about such things. For
example, when we use the equation

t = Δx
v
it hopefully seems sensible to you that, if you are moving at a constant speed, v , then the time t it takes you to travel some
distance Δx is directly proportional to that distance, but inversely proportional to your speed.
Here's another bit of common-sense kinematics: Suppose that on a winter day you are pulling a child on a sled, which is
moving to the right at a speed of v = 1 m/s. The child takes a snowball, and throws it forward at a speed u' = 1.5 m/s
126 Chapter 5 | Relativistic Kinematics in One Dimension

relative to herself, you and the sled. She's trying to hit her friend, who is standing at rest just in front of you (see Figure 1).
At what speed, u , does her friend observe the snowball to be moving?

Figure 5.2 Classically, velocities add like ordinary numbers in one-dimensional motion.
Here the girl throws a snowball forward and then backward from a sled. The velocity of the
sled relative to the Earth is v=1.0 m/s . The velocity of the snowball relative to the truck is
u′ , while its velocity relative to the Earth is u . Classically, u=v+u′ .
(Credit: OpenStax College Physics. "Relativistic Addition of Velocities." OpenStax-CNX.
September 12, 2013. https://legacy.cnx.org/content/m42540/1.6/)

The answer, of course, is child's play! (Sorry...) But your common sense tells you that it's moving at 2.5 m/s, and if you
think about an equation that tells that story, it's just
u = u' + v
We can see that this equation is still valid if the child throws the snowball in the opposite direction. In that case, the value
of u' = -1.5 m/s, with the minus sign indicating that the direction of her throw is in the negative x direction. The equation
yields u = -0.5 m/s. Thus, to the stationary friend ahead, the snowball appears to be moving away to the left at a speed of
0.5 m/s.
This classical addition of velocities may be common sense, but early in the 20th century it was shown to be wrong for
objects that travel at or near the speed of light. And, although our everyday experience does not seem to include objects that
move that fast, as Figure 5.1 shows, there are examples where the so-called relativistic corrections do matter. In particular,
as we study astrophysics (where the distance scales can be enormous) and the motion of elementary particles (which do
move near the speed of light), we must incorporate the theory of relativity.
The special theory of relativity was proposed in 1905 by Albert Einstein (1879–1955). It describes how time, space, and
physical phenomena appear in different frames of reference that are moving at constant velocity with respect to each other.
This differs from Einstein’s later work on general relativity, which deals with any frame of reference, including accelerated
frames.
The theory of relativity led to a profound change in the way we perceive space and time. The “common sense” rules that
we use to relate space and time measurements in the Newtonian worldview differ seriously from the correct rules at speeds
near the speed of light. For example, the special theory of relativity tells us that measurements of length and time intervals
are not the same in reference frames moving relative to one another. A particle might be observed to have a lifetime of

This OpenStax book is available for free at https://legacy.cnx.org/content/col12169/1.46


Chapter 5 | Relativistic Kinematics in One Dimension 127

1.0 × 10 −8 s in one reference frame, but a lifetime of 2.0 × 10 −8 s in another; and an object might be measured to be
2.0 m long in one frame and 3.0 m long in another frame. These effects are usually significant only at speeds comparable to
the speed of light, but even at the much lower speeds of the global positioning satellite, which requires extremely accurate
time measurements to function, the different lengths of the same distance in different frames of reference are significant
enough that they need to be taken into account.
The modifications of kinematics in special relativity do not invalidate classical theories or require their replacement. Instead,
the equations of relativistic mechanics differ meaningfully from those of classical mechanics only for objects moving at
relativistic speeds (i.e., speeds less than, but comparable to, the speed of light). In the macroscopic world that you encounter
in your daily life, the relativistic equations reduce to classical equations, and the predictions of classical mechanics agree
closely enough with experimental results to disregard relativistic corrections.

5.1 | Invariance of Physical Laws


Learning Objectives
By the end of this section, you will be able to:
• Describe the theoretical and experimental issues that Einstein’s theory of special relativity
addressed.
• State the two postulates of the special theory of relativity.

Suppose you calculate the hypotenuse of a right triangle given the base angles and adjacent sides. Whether you calculate
the hypotenuse from one of the sides and the cosine of the base angle, or from the Pythagorean theorem, the results
should agree. Predictions based on different principles of physics must also agree, whether we consider them principles of
mechanics or principles of electromagnetism.
Albert Einstein pondered a disagreement between predictions based on electromagnetism and on assumptions made in
classical mechanics. Specifically, suppose an observer measures the velocity of a light pulse in the observer’s own rest
frame; that is, in the frame of reference in which the observer is at rest. According to the assumptions long considered
obvious in classical mechanics, if an observer measures a velocity → v in one frame of reference, and that frame of
reference is moving with velocity →
u past a second reference frame, an observer in the second frame measures the original

velocity as v′ = → v + → u . This sum of velocities is often referred to as Galilean relativity. If this principle is correct,
the pulse of light that the observer measures as traveling with speed c travels at speed c + u measured in the frame of the
second observer. If we reasonably assume that the laws of electrodynamics are the same in both frames of reference, then the
predicted speed of light (in vacuum) in both frames should be c = 1/ ε 0 μ 0. Each observer should measure the same speed
of the light pulse with respect to that observer’s own rest frame. To reconcile difficulties of this kind, Einstein constructed
his special theory of relativity, which introduced radical new ideas about time and space that have since been confirmed
experimentally.

Inertial Frames
All velocities are measured relative to some frame of reference. For example, a car’s motion is measured relative to its
starting position on the road it travels on; a projectile’s motion is measured relative to the surface from which it is launched;
and a planet’s orbital motion is measured relative to the star it orbits. The frames of reference in which mechanics takes the
simplest form are those that are not accelerating. Newton’s first law, the law of inertia, holds exactly in such a frame.

Inertial Reference Frame


An inertial frame of reference is a reference frame in which a body at rest remains at rest and a body in motion moves
at a constant speed in a straight line unless acted upon by an outside force.

For example, to a passenger inside a plane flying at constant speed and constant altitude, physics seems to work exactly
the same as when the passenger is standing on the surface of Earth. When the plane is taking off, however, matters are
somewhat more complicated. In this case, the passenger at rest inside the plane concludes that a net force F on an object
is not equal to the product of mass and acceleration, ma. Instead, F is equal to ma plus a fictitious force. This situation is
128 Chapter 5 | Relativistic Kinematics in One Dimension

not as simple as in an inertial frame. The term “special” in “special relativity” refers to dealing only with inertial frames of
reference. Einstein’s later theory of general relativity deals with all kinds of reference frames, including accelerating, and
therefore non-inertial, reference frames.

Einstein’s First Postulate


Not only are the principles of classical mechanics simplest in inertial frames, but they are the same in all inertial frames.
Einstein based the first postulate of his theory on the idea that this is true for all the laws of physics, not merely those in
mechanics.

First Postulate of Special Relativity


The laws of physics are the same in all inertial frames of reference.

This postulate denies the existence of a special or preferred inertial frame. The laws of nature do not give us a way to
endow any one inertial frame with special properties. For example, we cannot identify any inertial frame as being in a state
of “absolute rest.” We can only determine the relative motion of one frame with respect to another.
There is, however, more to this postulate than meets the eye. The laws of physics include only those that satisfy this
postulate. We will see that the definitions of energy and momentum must be altered to fit this postulate. Another outcome
of this postulate is the famous equation E = mc 2, which relates energy to mass.

Einstein’s Second Postulate


The second postulate upon which Einstein based his theory of special relativity deals with the speed of light. Late in the
nineteenth century, the major tenets of classical physics were well established. Two of the most important were the laws
of electromagnetism and Newton’s laws. Investigations such as Young’s double-slit experiment in the early 1800s had
convincingly demonstrated that light is a wave. Maxwell’s equations of electromagnetism implied that electromagnetic
waves travel at c = 3.00×10 8 m/s in a vacuum, but they do not specify the frame of reference in which light has this
speed. Many types of waves were known, and all travelled in some medium. Scientists therefore assumed that some medium
carried the light, even in a vacuum, and that light travels at a speed c relative to that medium (often called “the aether”).
Starting in the mid-1880s, the American physicist A.A. Michelson, later aided by E.W. Morley, made a series of direct
measurements of the speed of light. They intended to deduce from their data the speed v at which Earth was moving through
the mysterious medium for light waves. The speed of light measured on Earth should have been c + v when Earth’s motion
was opposite to the medium’s flow at speed u past the Earth, and c – v when Earth was moving in the same direction as the
medium. The results of their measurements were startling.

Michelson-Morley Experiment
The Michelson-Morley experiment demonstrated that the speed of light in a vacuum is independent of the motion of
Earth about the Sun.

The eventual conclusion derived from this result is that light, unlike mechanical waves such as sound, does not need a
medium to carry it. Furthermore, the Michelson-Morley results implied that the speed of light c is independent of the motion
of the source relative to the observer. That is, everyone observes light to move at speed c regardless of how they move
relative to the light source or to one another. For several years, many scientists tried unsuccessfully to explain these results
within the framework of Newton’s laws.
In addition, there was a contradiction between the principles of electromagnetism and the assumption made in Newton’s
laws about relative velocity. Classically, the velocity of an object in one frame of reference and the velocity of that object in
a second frame of reference relative to the first should combine like simple vectors to give the velocity seen in the second
frame. If that were correct, then two observers moving at different speeds would see light traveling at different speeds.
Imagine what a light wave would look like to a person traveling along with it (in vacuum) at a speed c. If such a motion
were possible, then the wave would be stationary relative to the observer. It would have electric and magnetic fields whose
strengths varied with position but were constant in time. This is not allowed by Maxwell’s equations. So either Maxwell’s
equations are different in different inertial frames, or an object with mass cannot travel at speed c. Einstein concluded that
the latter is true: An object with mass cannot travel at speed c. Maxwell’s equations are correct, but Newton’s addition of
velocities is not correct for light.

This OpenStax book is available for free at https://legacy.cnx.org/content/col12169/1.46


Chapter 5 | Relativistic Kinematics in One Dimension 129

Not until 1905, when Einstein published his first paper on special relativity, was the currently accepted conclusion reached.
Based mostly on his analysis that the laws of electricity and magnetism would not allow another speed for light, and only
slightly aware of the Michelson-Morley experiment, Einstein detailed his second postulate of special relativity.

Second Postulate of Special Relativity


Light travels in a vacuum with the same speed c in any direction in all inertial frames.

In other words, the speed of light has the same definite speed for any observer, regardless of the relative motion of the
source. This deceptively simple and counterintuitive postulate, along with the first postulate, leave all else open for change.
Among the changes are the loss of agreement on the time between events, the variation of distance with speed, and the
realization that matter and energy can be converted into one another. We describe these concepts in the following sections.
5.1 Check Your Understanding Explain how special relativity differs from general relativity.

5.2 | Relativity of Simultaneity


Learning Objectives
By the end of this section, you will be able to:
• Show from Einstein's postulates that two events measured as simultaneous in one inertial
frame are not necessarily simultaneous in all inertial frames.
• Describe how simultaneity is a relative concept for observers in different inertial frames in
relative motion.

Do time intervals depend on who observes them? Intuitively, it seems that the time for a process, such as the elapsed time
for a foot race (Figure 5.3), should be the same for all observers. In everyday experiences, disagreements over elapsed time
have to do with the accuracy of measuring time. No one would be likely to argue that the actual time interval was different
for the moving runner and for the stationary clock displayed. Carefully considering just how time is measured, however,
shows that elapsed time does depends on the relative motion of an observer with respect to the process being measured.
130 Chapter 5 | Relativistic Kinematics in One Dimension

Figure 5.3 Elapsed time for a foot race is the same for all observers, but at relativistic speeds,
elapsed time depends on the motion of the observer relative to the location where the process
being timed occurs. (credit: "Jason Edward Scott Bain"/Flickr)

Consider how we measure elapsed time. If we use a stopwatch, for example, how do we know when to start and stop the
watch? One method is to use the arrival of light from the event. For example, if you’re in a moving car and observe the light
arriving from a traffic signal change from green to red, you know it’s time to step on the brake pedal. The timing is more
accurate if some sort of electronic detection is used, avoiding human reaction times and other complications.
Now suppose two observers use this method to measure the time interval between two flashes of light from flash lamps
that are a distance apart (Figure 5.4). An observer A is seated midway on a rail car with two flash lamps at opposite sides
equidistant from her. A pulse of light is emitted from each flash lamp and moves toward observer A, shown in frame (a)
of the figure. The rail car is moving rapidly in the direction indicated by the velocity vector in the diagram. An observer B
standing on the platform is facing the rail car as it passes and observes both flashes of light reach him simultaneously, as
shown in frame (c). He measures the distances from where he saw the pulses originate, finds them equal, and concludes that
the pulses were emitted simultaneously.
However, because of Observer A’s motion, the pulse from the right of the railcar, from the direction the car is moving,
reaches her before the pulse from the left, as shown in frame (b). She also measures the distances from within her frame of
reference, finds them equal, and concludes that the pulses were not emitted simultaneously.
The two observers reach conflicting conclusions about whether the two events at well-separated locations were
simultaneous. Both frames of reference are valid, and both conclusions are valid. Whether two events at separate locations
are simultaneous depends on the motion of the observer relative to the locations of the events.

This OpenStax book is available for free at https://legacy.cnx.org/content/col12169/1.46


Chapter 5 | Relativistic Kinematics in One Dimension 131

Figure 5.4 (a) Two pulses of light are emitted simultaneously relative to observer B. (c) The pulses reach
observer B’s position simultaneously. (b) Because of A’s motion, she sees the pulse from the right first and
concludes the bulbs did not flash simultaneously. Both conclusions are correct.

Here, the relative velocity between observers affects whether two events a distance apart are observed to be simultaneous.
Simultaneity is not absolute. We might have guessed (incorrectly) that if light is emitted simultaneously, then two observers
halfway between the sources would see the flashes simultaneously. But careful analysis shows this cannot be the case if the
speed of light is the same in all inertial frames.
This type of thought experiment (in German, “Gedankenexperiment”) shows that seemingly obvious conclusions must be
changed to agree with the postulates of relativity. The validity of thought experiments can only be determined by actual
observation, and careful experiments have repeatedly confirmed Einstein’s theory of relativity.
132 Chapter 5 | Relativistic Kinematics in One Dimension

5.3 | Time Dilation


Learning Objectives
By the end of this section, you will be able to:
• Explain how time intervals can be measured differently in different reference frames.
• Describe how to distinguish a proper time interval from a dilated time interval.
• Describe the significance of the muon experiment.
• Explain why the twin paradox is not a contradiction.
• Calculate time dilation given the speed of an object in a given frame.

The analysis of simultaneity shows that Einstein’s postulates imply an important effect: Time intervals have different values
when measured in different inertial frames. Suppose, for example, an astronaut measures the time it takes for a pulse of
light to travel a distance perpendicular to the direction of his ship’s motion (relative to an earthbound observer), bounce off
a mirror, and return (Figure 5.5). How does the elapsed time that the astronaut measures in the spacecraft compare with
the elapsed time that an earthbound observer measures by observing what is happening in the spacecraft?
Examining this question leads to a profound result. The elapsed time for a process depends on which observer is measuring
it. In this case, the time measured by the astronaut (within the spaceship where the astronaut is at rest) is smaller than
the time measured by the earthbound observer (to whom the astronaut is moving). The time elapsed for the same process
is different for the observers, because the distance the light pulse travels in the astronaut’s frame is smaller than in the
earthbound frame, as seen in Figure 5.5. Light travels at the same speed in each frame, so it takes more time to travel the
greater distance in the earthbound frame.

This OpenStax book is available for free at https://legacy.cnx.org/content/col12169/1.46


Chapter 5 | Relativistic Kinematics in One Dimension 133

Figure 5.5 (a) An astronaut measures the time Δτ for light to travel distance 2D in the astronaut’s frame. (b) A NASA
scientist on Earth sees the light follow the longer path 2s and take a longer time Δt. (c) These triangles are used to find the
relationship between the two distances D and s.

Time Dilation
Time dilation is the lengthening of the time interval between two events for an observer in an inertial frame that is
moving with respect to the rest frame of the events (in which the events occur at the same location).

To quantitatively compare the time measurements in the two inertial frames, we can relate the distances in Figure 5.5 to
each other, then express each distance in terms of the time of travel (respectively either Δt or Δτ ) of the pulse in the
corresponding reference frame. The resulting equation can then be solved for Δt in terms of Δτ.
The lengths D and L in Figure 5.5 are the sides of a right triangle with hypotenuse s. From the Pythagorean theorem,
s 2 = D 2 + L 2.
The lengths 2s and 2L are, respectively, the distances that the pulse of light and the spacecraft travel in time Δt in the
earthbound observer’s frame. The length D is the distance that the light pulse travels in time Δτ in the astronaut’s frame.
This gives us three equations:
2s = cΔt; 2L = vΔt; 2D = cΔτ.
134 Chapter 5 | Relativistic Kinematics in One Dimension

Note that we used Einstein’s second postulate by taking the speed of light to be c in both inertial frames. We substitute these
results into the previous expression from the Pythagorean theorem:
s2 = D2 + L2
⎛ Δt ⎞ ⎛ ⎞ ⎛ ⎞
2 2 2
⎝c 2 ⎠ = ⎝c Δτ ⎠ + ⎝v Δt ⎠ .
2 2
Then we rearrange to obtain
(cΔt) 2 − (vΔt) 2 = (cΔτ) 2.
Finally, solving for Δt in terms of Δτ gives us

Δt = Δτ . (5.1)
1 − (v/c) 2

This is equivalent to
Δt = γΔτ,

where γ is the relativistic factor (often called the Lorentz factor) given by

γ= 1 (5.2)
1− v2
c2

and v and c are the speeds of the moving observer and light, respectively.
Note the asymmetry between the two measurements. Only one of them is a measurement of the time interval between two
events—the emission and arrival of the light pulse—at the same position. It is a measurement of the time interval in the
rest frame of a single clock. The measurement in the earthbound frame involves comparing the time interval between two
events that occur at different locations. The time interval between events that occur at a single location has a separate name
to distinguish it from the time measured by the earthbound observer, and we use the separate symbol Δτ to refer to it
throughout this chapter.

Proper Time
The proper time interval Δτ between two events is the time interval measured by an observer for whom both events
occur at the same location.

The equation relating Δt and Δτ is truly remarkable. First, as stated earlier, elapsed time is not the same for different
observers moving relative to one another, even though both are in inertial frames. A proper time interval Δτ for an
observer who, like the astronaut, is moving with the apparatus, is smaller than the time interval for other observers. It is the
smallest possible measured time between two events. The earthbound observer sees time intervals within the moving system
as dilated (i.e., lengthened) relative to how the observer moving relative to Earth sees them within the moving system.
Alternatively, according to the earthbound observer, less time passes between events within the moving frame. Note that the
shortest elapsed time between events is in the inertial frame in which the observer sees the events (e.g., the emission and
arrival of the light signal) occur at the same point.
This time effect is real and is not caused by inaccurate clocks or improper measurements. Time-interval measurements of
the same event differ for observers in relative motion. The dilation of time is an intrinsic property of time itself. All clocks
moving relative to an observer, including biological clocks, such as a person’s heartbeat, or aging, are observed to run more
slowly compared with a clock that is stationary relative to the observer.

This OpenStax book is available for free at https://legacy.cnx.org/content/col12169/1.46


Chapter 5 | Relativistic Kinematics in One Dimension 135

Note that if the relative velocity is much less than the speed of light (v<<c), then v 2 /c 2 is extremely small, and the
elapsed times Δt and Δτ are nearly equal. At low velocities, physics based on modern relativity approaches classical
physics—everyday experiences involve very small relativistic effects. However, for speeds near the speed of light, v 2 /c 2
is close to one, so 1 − v 2/c 2 is very small and Δt becomes significantly larger than Δτ.

Half-Life of a Muon
There is considerable experimental evidence that the equation Δt = γΔτ is correct. One example is found in cosmic ray
particles that continuously rain down on Earth from deep space. Some collisions of these particles with nuclei in the upper
atmosphere result in short-lived particles called muons. The half-life (amount of time for half of a material to decay) of a
muon is 1.52 μs when it is at rest relative to the observer who measures the half-life. This is the proper time interval Δτ.
This short time allows very few muons to reach Earth’s surface and be detected if Newtonian assumptions about time and
space were correct. However, muons produced by cosmic ray particles have a range of velocities, with some moving near
the speed of light. It has been found that the muon’s half-life as measured by an earthbound observer ( Δt ) varies with
velocity exactly as predicted by the equation Δt = γΔτ. The faster the muon moves, the longer it lives. We on Earth see
the muon last much longer than its half-life predicts within its own rest frame. As viewed from our frame, the muon decays
more slowly than it does when at rest relative to us. A far larger fraction of muons reach the ground as a result.
Before we present the first example of solving a problem in relativity, we state a strategy you can use as a guideline for
these calculations.

Problem-Solving Strategy: Relativity


1. Make a list of what is given or can be inferred from the problem as stated (identify the knowns). Look in
particular for information on relative velocity v.
2. Identify exactly what needs to be determined in the problem (identify the unknowns).
3. Make certain you understand the conceptual aspects of the problem before making any calculations (express
the answer as an equation). Decide, for example, which observer sees time dilated or length contracted before
working with the equations or using them to carry out the calculation. If you have thought about who sees
what, who is moving with the event being observed, who sees proper time, and so on, you will find it much
easier to determine if your calculation is reasonable.
4. Determine the primary type of calculation to be done to find the unknowns identified above (do the
calculation). You will find the section summary helpful in determining whether a length contraction, relativistic
kinetic energy, or some other concept is involved.

Note that you should not round off during the calculation. As noted in the text, you must often perform your calculations to
many digits to see the desired effect. You may round off at the very end of the problem solution, but do not use a rounded
number in a subsequent calculation. Also, check the answer to see if it is reasonable: Does it make sense? This may be more
difficult for relativity, which has few everyday examples to provide experience with what is reasonable. But you can look
for velocities greater than c or relativistic effects that are in the wrong direction (such as a time contraction where a dilation
was expected).
136 Chapter 5 | Relativistic Kinematics in One Dimension

Example 5.1

Time Dilation in a High-Speed Vehicle


The Hypersonic Technology Vehicle 2 (HTV-2) is an experimental rocket vehicle capable of traveling at 21,000
km/h (5830 m/s). If an electronic clock in the HTV-2 measures a time interval of exactly 1-s duration, what would
observers on Earth measure the time interval to be?
Strategy
Apply the time dilation formula to relate the proper time interval of the signal in HTV-2 to the time interval
measured on the ground.
Solution
a. Identify the knowns: Δτ = 1 s; v = 5830 m/s.

b. Identify the unknown: Δt.


c. Express the answer as an equation:

Δt = γΔτ = Δτ .
2
1 − v2
c
d. Do the calculation. Use the expression for γ to determine Δt from Δτ :

Δt = 1s
⎛ 5830 m/s ⎞
2
1−⎝
3.00 × 10 8 m/s ⎠
= 1.000000000189 s
= 1 s + 1.89 × 10 −10 s.
Significance
The very high speed of the HTV-2 is still only 10-5 times the speed of light. Relativistic effects for the HTV-2 are
negligible for almost all purposes, but are not zero.

Example 5.2

What Speeds are Relativistic?


How fast must a vehicle travel for 1 second of time measured on a passenger’s watch in the vehicle to differ by
1% for an observer measuring it from the ground outside?
Strategy
Use the time dilation formula to find v/c for the given ratio of times.
Solution
a. Identify the known:
Δτ = 1 .
Δt 1.01
b. Identify the unknown: v/c.
c. Express the answer as an equation:

This OpenStax book is available for free at https://legacy.cnx.org/content/col12169/1.46


Chapter 5 | Relativistic Kinematics in One Dimension 137

Δt = γΔτ = 1 Δτ
1 − v 2/c 2
Δτ = 1 − v 2/c 2
Δt
⎛Δτ ⎞
2 2
⎝ Δt ⎠ = 1 − v2
c
v = 1 − (Δτ/Δt) 2.
c
d. Do the calculation:
v = 1 − (1/1.01) 2
c
= 0.14.
Significance
The result shows that an object must travel at very roughly 10% of the speed of light for its motion to produce
significant relativistic time dilation effects.

Example 5.3

Calculating Δt for a Relativistic Event


Suppose a cosmic ray colliding with a nucleus in Earth’s upper atmosphere produces a muon that has a velocity
v = 0.950c. The muon then travels at constant velocity and lives 2.20 μs as measured in the muon’s frame of
reference. (You can imagine this as the muon’s internal clock.) How long does the muon live as measured by an
earthbound observer (Figure 5.6)?

Figure 5.6 A muon in Earth’s atmosphere lives longer as measured by an


earthbound observer than as measured by the muon’s internal clock.

As we will discuss later, in the muon’s reference frame, it travels a shorter distance than measured in Earth’s
reference frame.
Strategy
A clock moving with the muon measures the proper time of its decay process, so the time we are given is
Δτ = 2.20μs. The earthbound observer measures Δt as given by the equation Δt = γΔτ. Because the velocity
is given, we can calculate the time in Earth’s frame of reference.
138 Chapter 5 | Relativistic Kinematics in One Dimension

Solution
a. Identify the knowns: v = 0.950c, Δτ = 2.20μs.

b. Identify the unknown: Δt.


c. Express the answer as an equation. Use:
Δt = γΔτ

with

γ= 1 .
2
1 − v2
c
d. Do the calculation. Use the expression for γ to determine Δt from Δτ :

Δt = γΔτ
= 1 Δτ
2
1 − v2
c
2.20μs
=
1 − (0.950) 2
= 7.05 μs.

Remember to keep extra significant figures until the final answer.


Significance
One implication of this example is that because γ = 3.20 at 95.0% of the speed of light (v = 0.950c), the
relativistic effects are significant. The two time intervals differ by a factor of 3.20, when classically they would
be the same. Something moving at 0.950c is said to be highly relativistic.

Example 5.4

Relativistic Television
A non-flat screen, older-style television display (Figure 5.7) works by accelerating electrons over a short
distance to relativistic speed, and then using electromagnetic fields to control where the electron beam strikes a
fluorescent layer at the front of the tube. Suppose the electrons travel at 6.00 × 10 7 m/s through a distance of
0.200 m from the start of the beam to the screen. (a) What is the time of travel of an electron in the rest frame of
the television set? (b) What is the electron’s time of travel in its own rest frame?

This OpenStax book is available for free at https://legacy.cnx.org/content/col12169/1.46


Chapter 5 | Relativistic Kinematics in One Dimension 139

Figure 5.7 The electron beam in a cathode ray tube television display.

Strategy for (a)


(a) Calculate the time from vt = d. Even though the speed is relativistic, the calculation is entirely in one frame
of reference, and relativity is therefore not involved.
Solution
a. Identify the knowns:
v = 6.00 × 10 7 m/s; d = 0.200 m.
b. Identify the unknown: the time of travel Δt.
c. Express the answer as an equation:

Δt = dv .
d. Do the calculation:

t = 0.200 m
6.00 × 10 7 m/s
= 3.33 × 10 −9 s.
Significance
The time of travel is extremely short, as expected. Because the calculation is entirely within a single frame of
reference, relativity is not involved, even though the electron speed is close to c.
Strategy for (b)
(b) In the frame of reference of the electron, the vacuum tube is moving and the electron is stationary. The
electron-emitting cathode leaves the electron and the front of the vacuum tube strikes the electron with the
electron at the same location. Therefore we use the time dilation formula to relate the proper time in the electron
rest frame to the time in the television frame.
140 Chapter 5 | Relativistic Kinematics in One Dimension

Solution
a. Identify the knowns (from part a):
Δt = 3.33 × 10 −9 s; v = 6.00 × 10 7 m/s; d = 0.200 m.
b. Identify the unknown: τ.
c. Express the answer as an equation:

Δt = γΔτ = Δτ
1 − v 2/c 2
Δτ = Δt 1 − v 2/c 2.
d. Do the calculation:

⎛ ⎞
2
7
Δτ = ⎛⎝3.33 × 10 −9 s⎞⎠ 1 − 6.00 × 108 m/s
⎝3.00 × 10 m/s ⎠
= 3.26 × 10 −9 s.
Significance
The time of travel is shorter in the electron frame of reference. Because the problem requires finding the time
interval measured in different reference frames for the same process, relativity is involved. If we had tried to
calculate the time in the electron rest frame by simply dividing the 0.200 m by the speed, the result would be
slightly incorrect because of the relativistic speed of the electron.

5.2 Check Your Understanding What is γ if v = 0.650c ?

The Twin Paradox


An intriguing consequence of time dilation is that a space traveler moving at a high velocity relative to Earth would age
less than the astronaut’s earthbound twin. This is often known as the twin paradox. Imagine the astronaut moving at such a
velocity that γ = 30.0, as in Figure 5.8. A trip that takes 2.00 years in her frame would take 60.0 years in the earthbound
twin’s frame. Suppose the astronaut travels 1.00 year to another star system, briefly explores the area, and then travels 1.00
year back. An astronaut who was 40 years old at the start of the trip would be would be 42 when the spaceship returns.
Everything on Earth, however, would have aged 60.0 years. The earthbound twin, if still alive, would be 100 years old.
The situation would seem different to the astronaut in Figure 5.8. Because motion is relative, the spaceship would seem to
be stationary and Earth would appear to move. (This is the sensation you have when flying in a jet.) Looking out the window
of the spaceship, the astronaut would see time slow down on Earth by a factor of γ = 30.0. Seen from the spaceship, the
earthbound sibling will have aged only 2/30, or 0.07, of a year, whereas the astronaut would have aged 2.00 years.

This OpenStax book is available for free at https://legacy.cnx.org/content/col12169/1.46


Chapter 5 | Relativistic Kinematics in One Dimension 141

Figure 5.8 The twin paradox consists of the conflicting


conclusions about which twin ages more as a result of a long
space journey at relativistic speed.

The paradox here is that the two twins cannot both be correct. As with all paradoxes, conflicting conclusions come from
a false premise. In fact, the astronaut’s motion is significantly different from that of the earthbound twin. The astronaut
accelerates to a high velocity and then decelerates to view the star system. To return to Earth, she again accelerates and
decelerates. The spacecraft is not in a single inertial frame to which the time dilation formula can be directly applied. That
is, the astronaut twin changes inertial references. The earthbound twin does not experience these accelerations and remains
in the same inertial frame. Thus, the situation is not symmetric, and it is incorrect to claim that the astronaut observes the
same effects as her twin. The lack of symmetry between the twins will be still more evident when we analyze the journey
later in this chapter in terms of the path the astronaut follows through four-dimensional space-time.
In 1971, American physicists Joseph Hafele and Richard Keating verified time dilation at low relative velocities by flying
extremely accurate atomic clocks around the world on commercial aircraft. They measured elapsed time to an accuracy of
a few nanoseconds and compared it with the time measured by clocks left behind. Hafele and Keating’s results were within
experimental uncertainties of the predictions of relativity. Both special and general relativity had to be taken into account,
because gravity and accelerations were involved as well as relative motion.
5.3 Check Your Understanding a. A particle travels at 1.90 × 10 8 m/s and lives 2.10 × 10 −8 s when at
rest relative to an observer. How long does the particle live as viewed in the laboratory?
b. Spacecraft A and B pass in opposite directions at a relative speed of 4.00 × 10 7 m/s. An internal clock in
spacecraft A causes it to emit a radio signal for 1.00 s. The computer in spacecraft B corrects for the beginning
and end of the signal having traveled different distances, to calculate the time interval during which ship A was
emitting the signal. What is the time interval that the computer in spacecraft B calculates?
142 Chapter 5 | Relativistic Kinematics in One Dimension

5.4 | Length Contraction


Learning Objectives
By the end of this section, you will be able to:
• Explain how simultaneity and length contraction are related.
• Describe the relation between length contraction and time dilation and use it to derive the
length-contraction equation.

The length of the train car in Figure 5.9 is the same for all the passengers. All of them would agree on the simultaneous
location of the two ends of the car and obtain the same result for the distance between them. But simultaneous events in one
inertial frame need not be simultaneous in another. If the train could travel at relativistic speeds, an observer on the ground
would see the simultaneous locations of the two endpoints of the car at a different distance apart than observers inside the
car. Measured distances need not be the same for different observers when relativistic speeds are involved.

Figure 5.9 People might describe distances differently, but at


relativistic speeds, the distances really are different. (credit:
“russavia”/Flickr)

Proper Length
Two observers passing each other always see the same value of their relative speed. Even though time dilation implies that
the train passenger and the observer standing alongside the tracks measure different times for the train to pass, they still
agree that relative speed, which is distance divided by elapsed time, is the same. If an observer on the ground and one on the
train measure a different time for the length of the train to pass the ground observer, agreeing on their relative speed means
they must also see different distances traveled.
The muon discussed in m58563 (https://legacy.cnx.org/content/m58563/latest/#fs-id1167793912924) illustrates
this concept (Figure 5.10). To an observer on Earth, the muon travels at 0.950c for 7.05 μs from the time it is produced
until it decays. Therefore, it travels a distance relative to Earth of:
L 0 = vΔt = (0.950)(3.00 × 10 8 m/s)(7.05 × 10 −6 s⎞⎠ = 2.01 km.

In the muon frame, the lifetime of the muon is 2.20 μs. In this frame of reference, the Earth, air, and ground have only
enough time to travel:
L = vΔτ = (0.950)(3.00 × 10 8 m/s)(2.20 × 10 −6 s) km = 0.627 km.
The distance between the same two events (production and decay of a muon) depends on who measures it and how they are
moving relative to it.

This OpenStax book is available for free at https://legacy.cnx.org/content/col12169/1.46


Chapter 5 | Relativistic Kinematics in One Dimension 143

Proper Length
Proper length L 0 is the distance between two points measured by an observer who is at rest relative to both of the
points.

The earthbound observer measures the proper length L 0 because the points at which the muon is produced and decays are
stationary relative to Earth. To the muon, Earth, air, and clouds are moving, so the distance L it sees is not the proper length.

Figure 5.10 (a) The earthbound observer sees the muon travel 2.01 km. (b) The same path has length 0.627 km seen from the
muon’s frame of reference. The Earth, air, and clouds are moving relative to the muon in its frame, and have smaller lengths
along the direction of travel.

Length Contraction
To relate distances measured by different observers, note that the velocity relative to the earthbound observer in our muon
example is given by
L0
v= .
Δt
The time relative to the earthbound observer is Δt, because the object being timed is moving relative to this observer. The
velocity relative to the moving observer is given by

v= L.
Δτ
The moving observer travels with the muon and therefore observes the proper time Δτ. The two velocities are identical;
thus,
L0
= L.
Δt Δτ
We know that Δt = γΔτ. Substituting this equation into the relationship above gives

L (5.3)
L = γ0 .

Substituting for γ gives an equation relating the distances measured by different observers.

Length Contraction
Length contraction is the decrease in the measured length of an object from its proper length when measured in a
reference frame that is moving with respect to the object:
2 (5.4)
L = L0 1 − v2
c
where L 0 is the length of the object in its rest frame, and L is the length in the frame moving with velocity v.

If we measure the length of anything moving relative to our frame, we find its length L to be smaller than the proper length
144 Chapter 5 | Relativistic Kinematics in One Dimension

L 0 that would be measured if the object were stationary. For example, in the muon’s rest frame, the distance Earth moves
between where the muon was produced and where it decayed is shorter than the distance traveled as seen from the Earth’s
frame. Those points are fixed relative to Earth but are moving relative to the muon. Clouds and other objects are also
contracted along the direction of motion as seen from muon’s rest frame.
Thus, two observers measure different distances along their direction of relative motion, depending on which one is
measuring distances between objects at rest.
But what about distances measured in a direction perpendicular to the relative motion? Imagine two observers moving along
their x-axes and passing each other while holding meter sticks vertically in the y-direction. Figure 5.11 shows two meter
sticks M and M′ that are at rest in the reference frames of two boys S and S′, respectively. A small paintbrush is attached
to the top (the 100-cm mark) of stick M′. Suppose that S′ is moving to the right at a very high speed v relative to S, and the
sticks are oriented so that they are perpendicular, or transverse, to their relative velocity vector. The sticks are held so that
as they pass each other, their lower ends (the 0-cm marks) coincide. Assume that when S looks at his stick M afterwards,
he finds a line painted on it, just below the top of the stick. Because the brush is attached to the top of the other boy’s stick
M′, S can only conclude that stick M′ is less than 1.0 m long.

Figure 5.11 Meter sticks M and M′ are stationary in the reference


frames of observers S and S′, respectively. As the sticks pass, a
small brush attached to the 100-cm mark of M′ paints a line on M.

Now when the boys approach each other, S′, like S, sees a meter stick moving toward him with speed v. Because their
situations are symmetric, each boy must make the same measurement of the stick in the other frame. So, if S measures stick
M′ to be less than 1.0 m long, S′ must measure stick M to be also less than 1.0 m long, and S′ must see his paintbrush
pass over the top of stick M and not paint a line on it. In other words, after the same event, one boy sees a painted line on a
stick, while the other does not see such a line on that same stick!
Einstein’s first postulate requires that the laws of physics (as, for example, applied to painting) predict that S and S′, who
are both in inertial frames, make the same observations; that is, S and S′ must either both see a line painted on stick M, or
both not see that line. We are therefore forced to conclude our original assumption that S saw a line painted below the top
of his stick was wrong! Instead, S finds the line painted right at the 100-cm mark on M. Then both boys will agree that a
line is painted on M, and they will also agree that both sticks are exactly 1 m long. We conclude then that measurements of
a transverse length must be the same in different inertial frames.

Example 5.5

Calculating Length Contraction


Suppose an astronaut, such as the twin in the twin paradox discussion, travels so fast that γ = 30.00. (a) The
astronaut travels from Earth to the nearest star system, Alpha Centauri, 4.300 light years (ly) away as measured
by an earthbound observer. How far apart are Earth and Alpha Centauri as measured by the astronaut? (b) In terms
of c, what is the astronaut’s velocity relative to Earth? You may neglect the motion of Earth relative to the sun
(Figure 5.12).

This OpenStax book is available for free at https://legacy.cnx.org/content/col12169/1.46


Chapter 5 | Relativistic Kinematics in One Dimension 145

Figure 5.12 (a) The earthbound observer measures the proper distance between Earth and Alpha Centauri. (b) The
astronaut observes a length contraction because Earth and Alpha Centauri move relative to her ship. She can travel this
shorter distance in a smaller time (her proper time) without exceeding the speed of light.

Strategy
First, note that a light year (ly) is a convenient unit of distance on an astronomical scale—it is the distance light
travels in a year. For part (a), the 4.300-ly distance between Alpha Centauri and Earth is the proper distance
L 0, because it is measured by an earthbound observer to whom both stars are (approximately) stationary. To
the astronaut, Earth and Alpha Centauri are moving past at the same velocity, so the distance between them is the
contracted length L. In part (b), we are given γ, so we can find v by rearranging the definition of γ to express v
in terms of c.
Solution for (a)
For part (a):
a. Identify the knowns: L 0 = 4.300 ly; γ = 30.00.

b. Identify the unknown: L.


L
c. Express the answer as an equation: L = γ0 .

d. Do the calculation:
L
L = γ0
4.300 ly
=
30.00
= 0.1433 ly.
146 Chapter 5 | Relativistic Kinematics in One Dimension

Solution for (b)


For part (b):
a. Identify the known: γ = 30.00.

b. Identify the unknown: v in terms of c.


c. Express the answer as an equation. Start with:

γ= 1 .
2
1 − v2
c

Then solve for the unknown v/c by first squaring both sides and then rearranging:

γ2 = 1
2
1 − v2
c
v2 = 1 − 1
c2 γ2
v = 1 − 12 .
c
γ
d. Do the calculation:
v = 1− 1
c
γ2
= 1− 1
(30.00) 2
= 0.99944

or
v = 0.9994 c.
Significance
Remember not to round off calculations until the final answer, or you could get erroneous results. This is
especially true for special relativity calculations, where the differences might only be revealed after several
decimal places. The relativistic effect is large here ⎛⎝γ = 30.00⎞⎠, and we see that v is approaching (not equaling)
the speed of light. Because the distance as measured by the astronaut is so much smaller, the astronaut can travel
it in much less time in her frame.

People traveling at extremely high velocities could cover very large distances (thousands or even millions of light years) and
age only a few years on the way. However, like emigrants in past centuries who left their home, these people would leave the
Earth they know forever. Even if they returned, thousands to millions of years would have passed on Earth, obliterating most
of what now exists. There is also a more serious practical obstacle to traveling at such velocities; immensely greater energies
would be needed to achieve such high velocities than classical physics predicts can be attained. This will be discussed later
in the chapter.
Why don’t we notice length contraction in everyday life? The distance to the grocery store does not seem to depend on
2
whether we are moving or not. Examining the equation L = L 0 1 − v 2 , we see that at low velocities (v<<c), the
c
lengths are nearly equal, which is the classical expectation. But length contraction is real, if not commonly experienced. For
example, a charged particle such as an electron traveling at relativistic velocity has electric field lines that are compressed
along the direction of motion as seen by a stationary observer (Figure 5.13). As the electron passes a detector, such as a
coil of wire, its field interacts much more briefly, an effect observed at particle accelerators such as the 3-km-long Stanford
Linear Accelerator (SLAC). In fact, to an electron traveling down the beam pipe at SLAC, the accelerator and Earth are all
moving by and are length contracted. The relativistic effect is so great that the accelerator is only 0.5 m long to the electron.
It is actually easier to get the electron beam down the pipe, because the beam does not have to be as precisely aimed to get

This OpenStax book is available for free at https://legacy.cnx.org/content/col12169/1.46


Chapter 5 | Relativistic Kinematics in One Dimension 147

down a short pipe as it would to get down a pipe 3 km long. This, again, is an experimental verification of the special theory
of relativity.

Figure 5.13 The electric field lines of a high-velocity charged


particle are compressed along the direction of motion by length
contraction, producing an observably different signal as the
particle goes through a coil.

5.4 Check Your Understanding A particle is traveling through Earth’s atmosphere at a speed of 0.750c. To
an earthbound observer, the distance it travels is 2.50 km. How far does the particle travel as viewed from the
particle’s reference frame?

5.5 | The Lorentz Transformation


Learning Objectives
• Describe the Galilean transformation of classical mechanics, relating the position, time,
velocities, and accelerations measured in different inertial frames
• Derive the corresponding Lorentz transformation equations, which, in contrast to the Galilean
transformation, are consistent with special relativity
• Explain the Lorentz transformation and many of the features of relativity in terms of four-
dimensional space-time

We have used the postulates of relativity to examine, in particular examples, how observers in different frames of reference
measure different values for lengths and the time intervals. We can gain further insight into how the postulates of relativity
change the Newtonian view of time and space by examining the transformation equations that give the space and time
coordinates of events in one inertial reference frame in terms of those in another. We first examine how position and
time coordinates transform between inertial frames according to the view in Newtonian physics. Then we examine how
this has to be changed to agree with the postulates of relativity. Finally, we examine the resulting Lorentz transformation
equations and some of their consequences in terms of four-dimensional space-time diagrams, to support the view that the
consequences of special relativity result from the properties of time and space itself, rather than electromagnetism.

The Galilean Transformation Equations


An event is specified by its location and time (x, y, z, t) relative to one particular inertial frame of reference S. As an
example, (x, y, z, t) could denote the position of a particle at time t, and we could be looking at these positions for many
different times to follow the motion of the particle. Suppose a second frame of reference S′ moves with velocity v with
respect to the first. For simplicity, assume this relative velocity is along the x-axis. The relation between the time and
coordinates in the two frames of reference is then
x = x′ + vt, y = y′, z = z′.

Implicit in these equations is the assumption that time measurements made by observers in both S and S′ are the same.
That is,
t = t′.
148 Chapter 5 | Relativistic Kinematics in One Dimension

These four equations are known collectively as the Galilean transformation.


We can obtain the Galilean velocity and acceleration transformation equations by differentiating these equations with
respect to time. We use u for the velocity of a particle throughout this chapter to distinguish it from v, the relative velocity
of two reference frames. Note that, for the Galilean transformation, the increment of time used in differentiating to calculate
the particle velocity is the same in both frames, dt = dt′. Differentiation yields
u x = u′x + v, u y = u′y, u z = u′z

We denote the velocity of the particle by u rather than v to avoid confusion with the velocity v of one frame of reference
with respect to the other. Velocities in each frame differ by the velocity that one frame has as seen from the other frame.

The Lorentz Transformation Equations


The Galilean transformation nevertheless violates Einstein’s postulates, because the velocity equations state that a pulse of
light moving with speed c along the x-axis would travel at speed c − v in the other inertial frame. Specifically, the spherical
pulse has radius r = ct at time t in the unprimed frame, and also has radius r′ = ct′ at time t′ in the primed frame.
Expressing these relations in Cartesian coordinates gives
x2 + y2 + z2 − c2 t2 = 0
x′ 2 + y′ 2 + z′ 2 − c 2 t′ 2 = 0.

The left-hand sides of the two expressions can be set equal because both are zero. Because y = y′ and z = z′, we obtain

x 2 − c 2 t 2 = x′ 2 − c 2 t′ 2. (5.5)

This cannot be satisfied for nonzero relative velocity v of the two frames if we assume the Galilean transformation results
in t = t′ with x = x′ + vt′.
To find the correct set of transformation equations, assume the two coordinate systems S and S′ in Figure 5.14. First
suppose that an event occurs at (x′, 0, 0, t′) in S′ and at (x, 0, 0, t) in S, as depicted in the figure.

Figure 5.14 An event occurs at (x, 0, 0, t) in S and at


(x′, 0, 0, t′) in S′. The Lorentz transformation equations relate
events in the two systems.

Suppose that at the instant that the origins of the coordinate systems in S and S′ coincide, a flash bulb emits a spherically
spreading pulse of light starting from the origin. At time t, an observer in S finds the origin of S′ to be at x = vt. With
the help of a friend in S, the S′ observer also measures the distance from the event to the origin of S′ and finds it to be
x′ 1 − v 2 /c 2. This follows because we have already shown the postulates of relativity to imply length contraction. Thus
the position of the event in S is

x = vt + x′ 1 − v 2 /c 2
and

This OpenStax book is available for free at https://legacy.cnx.org/content/col12169/1.46


Chapter 5 | Relativistic Kinematics in One Dimension 149

x′ = x − vt .
1 − v 2 /c 2
The postulates of relativity imply that the equation relating distance and time of the spherical wave front:
x2 + y2 + z2 − c2 t2 = 0

must apply both in terms of primed and unprimed coordinates, which was shown above to lead to Equation 5.5:
x 2 − c 2 t 2 = x′ 2 − c 2 t′ 2.
We combine this with the equation relating x and x′ to obtain the relation between t and t′ :
2
t′ = t − vx/c .
2 2
1 − v /c
The equations relating the time and position of the events as seen in S are then
2
t = t′ + vx′/c
2 2
1 − v /c
x = x′ + vt′
1 − v 2 /c 2
y = y′
z = z′.
This set of equations, relating the position and time in the two inertial frames, is known as the Lorentz transformation.
They are named in honor of H.A. Lorentz (1853–1928), who first proposed them. Interestingly, he justified the
transformation on what was eventually discovered to be a fallacious hypothesis. The correct theoretical basis is Einstein’s
special theory of relativity.
The reverse transformation expresses the variables in S in terms of those in S′. Simply interchanging the primed and
unprimed variables and substituting gives:

t′ = t − vx/c 2
1 − v 2 /c 2
x′ = x − vt
1 − v 2 /c 2
y′ = y
z′ = z.

Example 5.6

Using the Lorentz Transformation for Time


Spacecraft S′ is on its way to Alpha Centauri when Spacecraft S passes it at relative speed c/2. The captain of
S′ sends a radio signal that lasts 1.2 s according to that ship’s clock. Use the Lorentz transformation to find the
time interval of the signal measured by the communications officer of spaceship S.
Solution
a. Identify the known: Δt′ = t 2′ − t 1′ = 1.2 s; Δx′ = x′ 2 − x′ 1 = 0.

b. Identify the unknown: Δt = t 2 − t 1.

c. Express the answer as an equation. The time signal starts as ⎛⎝x′, t 1′⎞⎠ and stops at ⎛⎝x′, t 2′⎞⎠. Note
that the x′ coordinate of both events is the same because the clock is at rest in S′. Write the first
Lorentz transformation equation in terms of Δt = t 2 − t 1, Δx = x 2 − x 1, and similarly for the primed
coordinates, as:
150 Chapter 5 | Relativistic Kinematics in One Dimension

2
Δt = Δt′ + vΔx′/c .
2
1 − v2
c

Because the position of the clock in S′ is fixed, Δx′ = 0, and the time interval Δt becomes:

Δt = Δt′ .
2
1 − v2
c
d. Do the calculation.
With Δt′ = 1.2 s this gives:

Δt = 1.2 s = 1.6 s.
⎛1 ⎞
2
1− ⎝2 ⎠

Note that the Lorentz transformation reproduces the time dilation equation.

Example 5.7

Using the Lorentz Transformation for Length


A surveyor measures a street to be L = 100 m long in Earth frame S. Use the Lorentz transformation to obtain an
expression for its length measured from a spaceship S′, moving by at speed 0.20c, assuming the x coordinates
of the two frames coincide at time t = 0.
Solution
a. Identify the known: L = 100 m; v = 0.20c; Δτ = 0.

b. Identify the unknown: L′.


c. Express the answer as an equation. The surveyor in frame S has measured the two ends of the stick
simultaneously, and found them at rest at x 2 and x 1 a distance L = x 2 − x 1 = 100 m apart. The
spaceship crew measures the simultaneous location of the ends of the sticks in their frame. To relate
the lengths recorded by observers in S′ and S, respectively, write the second of the four Lorentz
transformation equations as:
x 2 − vt x 1 − vt
x′ 2 − x′ 1 = −
2 2
1 − v 2 /c 2
1 − v /c
x2 − x1
L
= = .
1 − v 2 /c 2 1 − v 2 /c 2
d. Do the calculation. Because x′ 2 − x′ 1 = 100 m, the length of the moving stick is equal to:

L′ = (100 m) 1 − v 2/c 2
= (100 m) 1 − (0.20) 2
= 98.0 m.

Note that the Lorentz transformation gave the length contraction equation for the street.

Example 5.8

Lorentz Transformation and Simultaneity


The observer shown in Figure 5.15 standing by the railroad tracks sees the two bulbs flash simultaneously at

This OpenStax book is available for free at https://legacy.cnx.org/content/col12169/1.46


Chapter 5 | Relativistic Kinematics in One Dimension 151

both ends of the 26 m long passenger car when the middle of the car passes him at a speed of c/2. Find the
separation in time between when the bulbs flashed as seen by the train passenger seated in the middle of the car.

Figure 5.15 An person watching a train go by observes two bulbs flash simultaneously at opposite ends
of a passenger car. There is another passenger inside of the car observing the same flashes but from a
different perspective.

Solution
a. Identify the known: Δt = 0.
Note that the spatial separation of the two events is between the two lamps, not the distance of the lamp
to the passenger.
b. Identify the unknown: Δt′ = t′2 − t′1.
Again, note that the time interval is between the flashes of the lamps, not between arrival times for
reaching the passenger.
c. Express the answer as an equation:
2
Δt = Δt′ + vΔx′/c .
1 − v 2 /c 2
d. Do the calculation:
Δt′ + 2c (26 m)/c 2
0 =
1 − v 2 /c 2
Δt′ = − 26 m/s = − ⎛ 26 m/s8
2c 2⎝3.00×10 m/s⎞⎠
Δt′ = −4.33×10 −8 s.
Significance
The sign indicates that the event with the larger x 2′, namely, the flash from the right, is seen to occur first in the
S′ frame, as found earlier for this example, so that t 2 < t 1.

Space-time
Relativistic phenomena can be analyzed in terms of events in a four-dimensional space-time. When phenomena such as the
twin paradox, time dilation, length contraction, and the dependence of simultaneity on relative motion are viewed in this
way, they are seen to be characteristic of the nature of space and time, rather than specific aspects of electromagnetism.
In three-dimensional space, positions are specified by three coordinates on a set of Cartesian axes, and the displacement of
one point from another is given by:

Δx, Δy, Δz⎞⎠ = (x 2 − x 1, y 2 − y 1, z 2 − z 1).

The distance Δr between the points is

Δr 2 = (Δx) 2 + ⎛⎝Δy⎞⎠ 2 + (Δz) 2.


152 Chapter 5 | Relativistic Kinematics in One Dimension

The distance Δr is invariant under a rotation of axes. If a new set of Cartesian axes rotated around the origin relative to the
original axes are used, each point in space will have new coordinates in terms of the new axes, but the distance Δr′ given
by
Δr′ 2 = (Δx′) 2 + ⎛⎝Δy′⎞⎠ 2 + (Δz′) 2.

That has the same value that Δr 2 had. Something similar happens with the Lorentz transformation in space-time.
Define the separation between two events, each given by a set of x, y, z¸ and ct along a four-dimensional Cartesian system
of axes in space-time, as
Δx, Δy, Δz, cΔt⎞⎠ = ⎛⎝x 2 − x 1, y 2 − y 1, z 2 − z 1, c(t 2 − t 1)⎞⎠.

Also define the space-time interval Δs between the two events as

Δs 2 = (Δx) 2 + ⎛⎝Δy⎞⎠ 2 + (Δz) 2 − (cΔt) 2.

The space-time interval is invariant under the Lorentz transformation. This follows from the postulates of relativity, and can
be seen also by substitution of the previous Lorentz transformation equations into the expression for the space-time interval:
Δs 2 = (Δx) 2 + ⎛⎝Δy⎞⎠ 2 + (Δz) 2 − (cΔt) 2
⎛ Δt′ + vΔx′ ⎞2
⎛ ⎞
2
⎜ c2 ⎟
= ⎜ Δx′ + vΔt′ ⎟ + ⎛⎝Δy′⎞⎠ 2 + (Δz′) 2 − ⎜c
⎝ 1 − v /c ⎠
2 2 2 2⎟
⎝ 1 − v /c ⎠
= (Δx′) 2 + ⎛⎝Δy′⎞⎠ 2 + (Δz′) 2 − (cΔt′) 2
= Δs′ 2.
5.5 Check Your Understanding Show that if a time increment dt elapses for an observer who sees the
particle moving with velocity v, it corresponds to a proper time particle increment for the particle of dτ = γdt.

The light cone


We can deal with the difficulty of visualizing and sketching graphs in four dimensions by imagining the three spatial
coordinates to be represented collectively by a horizontal axis, and the vertical axis to be the ct-axis. Starting with a
particular event in space-time as the origin of the space-time graph shown, the world line of a particle that remains at rest at
the initial location of the event at the origin then is the time axis. Any plane through the time axis parallel to the spatial axes
contains all the events that are simultaneous with each other and with the intersection of the plane and the time axis, as seen
in the rest frame of the event at the origin.
It is useful to picture a light cone on the graph, formed by the world lines of all light beams passing through the origin event
A, as shown in Figure 5.16. The light cone, according to the postulates of relativity, has sides at an angle of 45° if the time
axis is measured in units of ct, and, according to the postulates of relativity, the light cone remains the same in all inertial
frames. Because the event A is arbitrary, every point in the space-time diagram has a light cone associated with it.

This OpenStax book is available for free at https://legacy.cnx.org/content/col12169/1.46


Chapter 5 | Relativistic Kinematics in One Dimension 153

Figure 5.16 The light cone consists of all the world lines
followed by light from the event A at the vertex of the cone.

Consider now the world line of a particle through space-time. Any world line outside of the cone, such as one passing from
A through C, would involve speeds greater than c, and would therefore not be possible.
The twin paradox seen in space-time
The twin paradox discussed earlier involves an astronaut twin traveling at near light speed to a distant star system, and
returning to Earth. Because of time dilation, the space twin is predicted to age much less than the earthbound twin. This
seems paradoxical because we might have expected at first glance for the relative motion to be symmetrical and naively
thought it possible to also argue that the earthbound twin should age less.
To analyze this in terms of a space-time diagram, assume that the origin of the axes used is fixed in Earth. The world line of
the earthbound twin is then along the time axis.
The world line of the astronaut twin, who travels to the distant star and then returns, must deviate from a straight line path
in order to allow a return trip. As seen in Figure 5.17, the circumstances of the two twins are not at all symmetrical. Their
paths in space-time are of manifestly different length. Specifically, the world line of the earthbound twin has length 2cΔt,
which then gives the proper time that elapses for the earthbound twin as 2Δt. The distance to the distant star system is
Δx = vΔt. The proper time that elapses for the space twin is 2Δτ where

c 2 Δτ 2 = (cΔt) 2 − (Δx) 2.
This is considerably shorter than the proper time for the earthbound twin by the ratio
2 2 2 2
cΔτ = (cΔt) − (Δx) = (cΔt) − (vΔt)
cΔt (cΔt) 2 (cΔt) 2
2
= 1 − v 2 = 1γ .
c
consistent with the time dilation formula. The twin paradox is therefore seen to be no paradox at all. The situation of the
two twins is not symmetrical in the space-time diagram. The only surprise is perhaps that the seemingly longer path on the
space-time diagram corresponds to the smaller proper time interval.
154 Chapter 5 | Relativistic Kinematics in One Dimension

Figure 5.17 The space twin and the earthbound twin, in the
twin paradox example, follow world lines of different length
through space-time.

Lorentz transformations in space-time


We have already noted how the Lorentz transformation leaves
Δs 2 = (Δx) 2 + ⎛⎝Δy⎞⎠ 2 + (Δz) 2 − (cΔt) 2

unchanged and corresponds to a rotation of axes in the four-dimensional space-time. If the S and S′ frames are in relative
motion along their shared x-direction the space and time axes of S′ are rotated by an angle α as seen from S, in the way
shown in shown in Figure 5.18, where:
tanα = vc = β.

The line labeled “v = c” at 45° to the x-axis corresponds to the edge of the light cone, and is unaffected by the Lorentz
transformation, in accordance with the second postulate of relativity. The “v = c” line, and the light cone it represents, are
the same for both the S and S′ frame of reference.

Figure 5.18 The Lorentz transformation results in new space


and time axes rotated in a scissors-like way with respect to the
original axes.

Simultaneity
Simultaneity of events at separated locations depends on the frame of reference used to describe them, as given by the
scissors-like “rotation” to new time and space coordinates as described. If two events have the same t values in the unprimed
frame of reference, they need not have the same values measured along the ct′-axis, and would then not be simultaneous
in the primed frame.
As a specific example, consider the near-light-speed train in which flash lamps at the two ends of the car have flashed
simultaneously in the frame of reference of an observer on the ground. The space-time graph is shown Figure 5.19. The

This OpenStax book is available for free at https://legacy.cnx.org/content/col12169/1.46


Chapter 5 | Relativistic Kinematics in One Dimension 155

flashes of the two lamps are represented by the dots labeled “Left flash lamp” and “Right flash lamp” that lie on the light
cone in the past. The world line of both pulses travel along the edge of the light cone to arrive at the observer on the ground
simultaneously. Their arrival is the event at the origin. They therefore had to be emitted simultaneously in the unprimed
frame, as represented by the point labeled as t(both). But time is measured along the ct′-axis in the frame of reference of
the observer seated in the middle of the train car. So in her frame of reference, the emission event of the bulbs labeled as t′
(left) and t′ (right) were not simultaneous.

Figure 5.19 The train example revisited. The flashes occur at the same time
t(both) along the time axis of the ground observer, but at different times, along
the t′ time axis of the passenger.

In terms of the space-time diagram, the two observers are merely using different time axes for the same events because
they are in different inertial frames, and the conclusions of both observers are equally valid. As the analysis in terms of the
space-time diagrams further suggests, the property of how simultaneity of events depends on the frame of reference results
from the properties of space and time itself, rather than from anything specifically about electromagnetism.

5.6 | Relativistic Velocity Transformation


Learning Objectives
By the end of this section, you will be able to:
• Derive the equations consistent with special relativity for transforming velocities in one inertial
frame of reference into another.
• Apply the velocity transformation equations to objects moving at relativistic speeds.
• Examine how the combined velocities predicted by the relativistic transformation equations
compare with those expected classically.

Do you remember the snowball-throwing child we discussed in the Introduction to this chapter? In that example, we
added velocities using the common-sense method known as the Galilean velocity transformation. However, the relativistic
156 Chapter 5 | Relativistic Kinematics in One Dimension

addition of velocities is quite different. Before explaining how velocities add in special relativity, it is important to briefly
consider a classical example where the motion is not confined to one dimension. (While a complete discussion of the
classical addition of velocities in more than one dimension will be presented in Section 7.5, a simple example here will
help us understand what it is about relativity that is so fundamentally different.)
Classical Relativity
Let us consider an example of what two different observers see in a situation analyzed long ago by Galileo. Suppose a
sailor at the top of a mast on a moving ship drops his binoculars. Where will it hit the deck? Will it hit at the base of the
mast, or will it hit behind the mast because the ship is moving forward? The answer is that if air resistance is negligible,
the binoculars will hit at the base of the mast at a point directly below its point of release. Now let us consider what two
different observers see when the binoculars drop. One observer is on the ship and the other on shore. The binoculars have no
horizontal velocity relative to the observer on the ship, and so he sees them fall straight down the mast. (See Figure 5.20.)
To the observer on shore, the binoculars and the ship have the same horizontal velocity, so both move the same distance
forward while the binoculars are falling. This observer sees the curved path shown in Figure 5.20. Although the paths look
different to the different observers, each sees the same result—the binoculars hit at the base of the mast and not behind it.
To get the correct description, it is crucial to correctly specify the velocities relative to the observer.

Figure 5.20 Classical relativity. The same motion as viewed


by two different observers. An observer on the moving ship sees
the binoculars dropped from the top of its mast fall straight
down. An observer on shore sees the binoculars take the curved
path, moving forward with the ship. Both observers see the
binoculars strike the deck at the base of the mast. The initial
horizontal velocity is different relative to the two observers.
(The ship is shown moving rather fast to emphasize the effect.)
(Source: "Addition of Velocities", OpenStax College,
https://legacy.cnx.org/content/m42045/1.13/)

If we analyze this situation classically, it is obvious that the two observers will measure different horizontal velocities for
the binoculars. After all, the boat is moving horizontally. But, classically, how would the two observers compare the vertical
velocity of the binoculars. And how long would each observer think that it took the binoculars to travel from the sailor's
hand to the deck of the boat?
The classical answer is that the two observers would measure exactly the same vertical motion (free fall) and would measure
the same time of fall from the sailor's hand to the deck of the boat. This is common sense! The purely horizontal motion of
the boat does not affect the vertical motion of the binoculars.
But, remember that the Lorentz transformation shows that even the measurement of time will be different for two observers
in the relativistic case. So, as we will see, even velocities in directions other than the direction of the observers' relative
motion will be measured differently by two different observers. This is definitely not common sense. But it is the truth of
special relativity.

This OpenStax book is available for free at https://legacy.cnx.org/content/col12169/1.46


Chapter 5 | Relativistic Kinematics in One Dimension 157

Velocity Transformations
Imagine a car traveling at night along a straight road, as in Figure 5.21. The driver sees the light leaving the headlights
at speed c within the car’s frame of reference. If the Galilean transformation applied to light, then the light from the car’s
headlights would approach the pedestrian at a speed u = v + c, contrary to Einstein’s postulates.

Figure 5.21 According to experimental results and the second postulate of relativity, light from the car’s headlights moves
away from the car at speed c and toward the observer on the sidewalk at speed c.

Both the distance traveled and the time of travel are different in the two frames of reference, and they must differ in a way
that makes the speed of light the same in all inertial frames. The correct rules for transforming velocities from one frame to
another can be obtained from the Lorentz transformation equations.

Relativistic Transformation of Velocity


Suppose an object P is moving at constant velocity u = ⎛⎝u′x, u′y, u′z⎞⎠ as measured in the S′ frame. The S′ frame is moving
along its x′-axis at velocity v. In an increment of time dt′ , the particle is displaced by dx′ along the x′-axis. Applying
the Lorentz transformation equations gives the corresponding increments of time and displacement in the unprimed axes:
dt = γ⎛⎝dt′ + vdx′ /c 2⎞⎠
dx = γ(dx′ + vdt′)
dy = dy′
dz = dz′.
The velocity components of the particle seen in the unprimed coordinate system are then
dx′
dx = γ(dx′ + vdt′) = dt′ + v
dt γ⎛⎝dt′ + vdx′/c 2⎞⎠ 1 + v2 dx′
dt′
c
dy′
dy dy′ dt′
= ⎛ =
dt γ⎝dt′ + vdx′/c 2⎞⎠ γ⎛1 + v dx′ ⎞
⎝ 2 c dt′ ⎠
dz′
dz = dz′ = dt′ .
dt γ⎛⎝dt′ + vdx′/c 2⎞⎠ ⎛ ⎞
γ⎝1 + v2 dx′
c dt′ ⎠
158 Chapter 5 | Relativistic Kinematics in One Dimension

We thus obtain the equations for the velocity components of the object as seen in frame S:
⎛ u′x + v ⎞⎟ ⎛ u′ / γ ⎞ ⎛ u′ / γ ⎞
ux = ⎜ ⎜ ⎟, u z = ⎜ ⎟.
y z
, u y =
⎝1 + vu′x / c 2 ⎠ ⎝1 + vu′x / c 2 ⎠ ⎝1 + vu′x / c 2 ⎠

Compare this with how the Galilean transformation of classical mechanics says the velocities transform, by adding simply
as vectors:
u x = u′x + u, u y = u′y, u z = u′.
z

When the relative velocity of the frames is much smaller than the speed of light, that is, when v ≪ c, the special relativity
velocity addition law reduces to the Galilean velocity law. When the speed v of S′ relative to S is comparable to the speed
of light, the relativistic velocity addition law gives a much smaller result than the classical (Galilean) velocity addition
does.

Example 5.9

Velocity Transformation Equations for Light


Suppose a spaceship heading directly toward Earth at half the speed of light sends a signal to us on a laser-
produced beam of light (Figure 5.22). Given that the light leaves the ship at speed c as observed from the ship,
calculate the speed at which it approaches Earth.

Figure 5.22 How fast does a light signal approach Earth if sent from a
spaceship traveling at 0.500c?

Strategy
Because the light and the spaceship are moving at relativistic speeds, we cannot use simple velocity addition.
Instead, we determine the speed at which the light approaches Earth using relativistic velocity addition.
Solution
a. Identify the knowns: v = 0.500c; u′ = c.

b. Identify the unknown: u.

c. Express the answer as an equation: u = v + vu′


u′ .
1+
c2

d. Do the calculation:

u = v + vu′
u′
1+ 2
c
= 0.500c +c
1 + (0.500c)(c)
2 c
(0.500 + 1)c
=
⎛c 2 + 0.500c 2 ⎞
⎝ c2 ⎠
= c.
Significance
Relativistic velocity addition gives the correct result. Light leaves the ship at speed c and approaches Earth at

This OpenStax book is available for free at https://legacy.cnx.org/content/col12169/1.46


Chapter 5 | Relativistic Kinematics in One Dimension 159

speed c. The speed of light is independent of the relative motion of source and observer, whether the observer is
on the ship or earthbound.

Velocities cannot add to greater than the speed of light, provided that v is less than c and u′ does not exceed c. The
following example illustrates that relativistic velocity addition is not as symmetric as classical velocity addition.

Example 5.10

Relativistic Package Delivery


Suppose the spaceship in the previous example approaches Earth at half the speed of light and shoots a canister
at a speed of 0.750c (Figure 5.23). (a) At what velocity does an earthbound observer see the canister if it is shot
directly toward Earth? (b) If it is shot directly away from Earth?

Figure 5.23 A canister is fired at 0.7500c toward Earth or away from Earth.

Strategy
Because the canister and the spaceship are moving at relativistic speeds, we must determine the speed of the
canister by an earthbound observer using relativistic velocity addition instead of simple velocity addition.
Solution for (a)
a. Identify the knowns: v = 0.500c; u′ = 0.750c.

b. Identify the unknown: u.

c. Express the answer as an equation: u = v + vu′


u′ .
1+
c2

d. Do the calculation:

u = v + vu′
u′
1+ 2
c
= 0.500c + 0.750c
1 + (0.500c)(0.750c)
2
c
= 0.909c.
Solution for (b)
a. Identify the knowns: v = 0.500c; u′ = −0.750c.

b. Identify the unknown: u.

c. Express the answer as an equation: u = v + vu′


u′ .
1+
c2

d. Do the calculation:
160 Chapter 5 | Relativistic Kinematics in One Dimension

u = v + vu′
u′
1+ 2
c
0.500c + (−0.750c)
=
1 + (0.500c)(−0.750c)
2 c
= −0.400c.
Significance
The minus sign indicates a velocity away from Earth (in the opposite direction from v), which means the canister
is heading toward Earth in part (a) and away in part (b), as expected. But relativistic velocities do not add as
simply as they do classically. In part (a), the canister does approach Earth faster, but at less than the vector
sum of the velocities, which would give 1.250c. In part (b), the canister moves away from Earth at a velocity
of −0.400c, which is faster than the −0.250c expected classically. The differences in velocities are not even
symmetric: In part (a), an observer on Earth sees the canister and the ship moving apart at a speed of 0.409c, and
at a speed of 0.900c in part (b).

5.6 Check Your Understanding Distances along a direction perpendicular to the relative motion of the two
frames are the same in both frames. Why then are velocities perpendicular to the x-direction different in the two
frames?

This OpenStax book is available for free at https://legacy.cnx.org/content/col12169/1.46


Chapter 5 | Relativistic Kinematics in One Dimension 161

CHAPTER 5 REVIEW
KEY TERMS
classical (Galilean) velocity addition method of adding velocities when v<<c; velocities add like regular numbers
in one-dimensional motion: u = v + u′, where v is the velocity between two observers, u is the velocity of an object
relative to one observer, and u′ is the velocity relative to the other observer

event occurrence in space and time specified by its position and time coordinates (x, y, z, t) measured relative to a frame
of reference
first postulate of special relativity laws of physics are the same in all inertial frames of reference
Galilean relativity if an observer measures a velocity in one frame of reference, and that frame of reference is moving
with a velocity past a second reference frame, an observer in the second frame measures the original velocity as the
vector sum of these velocities
Galilean transformation relation between position and time coordinates of the same events as seen in different
reference frames, according to classical mechanics
inertial frame of reference reference frame in which a body at rest remains at rest and a body in motion moves at a
constant speed in a straight line unless acted on by an outside force
length contraction decrease in observed length of an object from its proper length L 0 to length L when its length is
observed in a reference frame where it is traveling at speed v
Lorentz transformation relation between position and time coordinates of the same events as seen in different
reference frames, according to the special theory of relativity
Michelson-Morley experiment investigation performed in 1887 that showed that the speed of light in a vacuum is the
same in all frames of reference from which it is viewed
proper length L 0; the distance between two points measured by an observer who is at rest relative to both of the points;
for example, earthbound observers measure proper length when measuring the distance between two points that are
stationary relative to Earth
proper time Δτ is the time interval measured by an observer who sees the beginning and end of the process that the
time interval measures occur at the same location
relativistic velocity addition method of adding velocities of an object moving at a relativistic speeds
rest frame frame of reference in which the observer is at rest
second postulate of special relativity light travels in a vacuum with the same speed c in any direction in all inertial
frames
special theory of relativity theory that Albert Einstein proposed in 1905 that assumes all the laws of physics have the
same form in every inertial frame of reference, and that the speed of light is the same within all inertial frames
time dilation lengthening of the time interval between two events when seen in a moving inertial frame rather than the
rest frame of the events (in which the events occur at the same location)
world line path through space-time

KEY EQUATIONS
Δt = Δτ = γτ
2
Time dilation 1 − v2
c

γ= 1
2
Lorentz factor 1 − v2
c
162 Chapter 5 | Relativistic Kinematics in One Dimension

2 L
Length contraction L = L 0 1 − v 2 = γ0
c

Galilean transformation x = x′ + vt, y = y′, z = z′, t = t′


2
Lorentz transformation t = t′ + vx′/c
2 2
1 − v /c

x= x′ + vt′
1 − v 2 /c 2
y = y′

z = z′
2
Inverse Lorentz transformation t′ = t − vx/c
2 2
1 − v /c

x′ = x − vt
1 − v 2 /c 2
y′ = y

z′ = z

Space-time interval (Δs) 2 = (Δx) 2 + ⎛⎝Δy⎞⎠ 2 + (Δz) 2 − c 2 (Δt) 2

⎛ u′x + v ⎞⎟ ⎛ u′ /γ ⎞ ⎛ u′ /γ ⎞
ux = ⎜ ⎜ ⎟, u z = ⎜ ⎟
y z
Relativistic velocity addition , u y =
⎝1 + vu′x /c 2 ⎠ ⎝1 + vu′x /c 2 ⎠ ⎝1 + vu′x /c 2 ⎠

SUMMARY
5.1 Invariance of Physical Laws
• Relativity is the study of how observers in different reference frames measure the same event.
• Modern relativity is divided into two parts. Special relativity deals with observers in uniform (unaccelerated)
motion, whereas general relativity includes accelerated relative motion and gravity. Modern relativity is consistent
with all empirical evidence thus far and, in the limit of low velocity and weak gravitation, gives close agreement
with the predictions of classical (Galilean) relativity.
• An inertial frame of reference is a reference frame in which a body at rest remains at rest and a body in motion
moves at a constant speed in a straight line unless acted upon by an outside force.
• Modern relativity is based on Einstein’s two postulates. The first postulate of special relativity is that the laws of
physics are the same in all inertial frames of reference. The second postulate of special relativity is that the speed of
light c is the same in all inertial frames of reference, independent of the relative motion of the observer and the light
source.
• The Michelson-Morley experiment demonstrated that the speed of light in a vacuum is independent of the motion
of Earth about the sun.

5.2 Relativity of Simultaneity


• Two events are defined to be simultaneous if an observer measures them as occurring at the same time (such as by
receiving light from the events).
• Two events at locations a distance apart that are simultaneous for an observer at rest in one frame of reference are
not necessarily simultaneous for an observer at rest in a different frame of reference.

This OpenStax book is available for free at https://legacy.cnx.org/content/col12169/1.46


Chapter 5 | Relativistic Kinematics in One Dimension 163

5.3 Time Dilation


• Two events are defined to be simultaneous if an observer measures them as occurring at the same time. They are not
necessarily simultaneous to all observers—simultaneity is not absolute.
• Time dilation is the lengthening of the time interval between two events when seen in a moving inertial frame rather
than the rest frame of the events (in which the events occur at the same location).
• Observers moving at a relative velocity v do not measure the same elapsed time between two events. Proper time
Δτ is the time measured in the reference frame where the start and end of the time interval occur at the same
location. The time interval Δt measured by an observer who sees the frame of events moving at speed v is related
to the proper time interval Δτ of the events by the equation:

Δt = Δτ = γΔτ,
2
1 − v2
c

where

γ= 1 .
2
1 − v2
c
• The premise of the twin paradox is faulty because the traveling twin is accelerating. The journey is not symmetrical
for the two twins.
• Time dilation is usually negligible at low relative velocities, but it does occur, and it has been verified by
experiment.
• The proper time is the shortest measure of any time interval. Any observer who is moving relative to the system
being observed measures a time interval longer than the proper time.

5.4 Length Contraction


• All observers agree upon relative speed.
• Distance depends on an observer’s motion. Proper length L 0 is the distance between two points measured by an
observer who is at rest relative to both of the points.
• Length contraction is the decrease in observed length of an object from its proper length L 0 to length L when its
length is observed in a reference frame where it is traveling at speed v.
• The proper length is the longest measurement of any length interval. Any observer who is moving relative to the
system being observed measures a length shorter than the proper length.

5.5 The Lorentz Transformation


• The Galilean transformation equations describe how, in classical nonrelativistic mechanics, the position, velocity,
and accelerations measured in one frame appear in another. Lengths remain unchanged and a single universal time
scale is assumed to apply to all inertial frames.
• Newton’s laws of mechanics obey the principle of having the same form in all inertial frames under a Galilean
transformation, given by
x = x′ + vt, y = y′, z = z′, t = t′.

The concept that times and distances are the same in all inertial frames in the Galilean transformation, however, is
inconsistent with the postulates of special relativity.
• The relativistically correct Lorentz transformation equations are
164 Chapter 5 | Relativistic Kinematics in One Dimension

Lorentz transformation Inverse Lorentz transformation


2 2
t = t′ + vx′/c t′ = t − vx/c
2 2 2 2
1 − v /c 1 − v /c

x= x′ + vt′ x′ = x − vt
1 − v 2 /c 2 1 − v 2 /c 2
y = y′ y′ = y
z = z′ z′ = z

We can obtain these equations by requiring an expanding spherical light signal to have the same shape and speed of
growth, c, in both reference frames.
• Relativistic phenomena can be explained in terms of the geometrical properties of four-dimensional space-time, in
which Lorentz transformations correspond to rotations of axes.
• The Lorentz transformation corresponds to a space-time axis rotation, similar in some ways to a rotation of space
axes, but in which the invariant spatial separation is given by Δs rather than distances Δr, and that the Lorentz
transformation involving the time axis does not preserve perpendicularity of axes or the scales along the axes.
• The analysis of relativistic phenomena in terms of space-time diagrams supports the conclusion that these
phenomena result from properties of space and time itself, rather than from the laws of electromagnetism.

5.6 Relativistic Velocity Transformation


• With classical velocity addition, velocities add like regular numbers in one-dimensional motion: u = v + u′,
where v is the velocity between two observers, u is the velocity of an object relative to one observer, and u′ is the
velocity relative to the other observer.
• Velocities cannot add to be greater than the speed of light.
• Relativistic velocity addition describes the velocities of an object moving at a relativistic velocity.

CONCEPTUAL QUESTIONS
6. (a) How could you travel far into the future of Earth
5.1 Invariance of Physical Laws without aging significantly? (b) Could this method also
1. Which of Einstein’s postulates of special relativity allow you to travel into the past?
includes a concept that does not fit with the ideas of
classical physics? Explain.
5.4 Length Contraction
2. Is Earth an inertial frame of reference? Is the sun? 7. To whom does an object seem greater in length, an
Justify your response. observer moving with the object or an observer moving
relative to the object? Which observer measures the
3. When you are flying in a commercial jet, it may appear object’s proper length?
to you that the airplane is stationary and Earth is moving
beneath you. Is this point of view valid? Discuss briefly. 8. Relativistic effects such as time dilation and length
contraction are present for cars and airplanes. Why do these
effects seem strange to us?
5.3 Time Dilation
4. (a) Does motion affect the rate of a clock as measured 9. Suppose an astronaut is moving relative to Earth at
by an observer moving with it? (b) Does motion affect how a significant fraction of the speed of light. (a) Does he
an observer moving relative to a clock measures its rate? observe the rate of his clocks to have slowed? (b) What
change in the rate of earthbound clocks does he see? (c)
Does his ship seem to him to shorten? (d) What about the
5. To whom does the elapsed time for a process seem
distance between two stars that lie in the direction of his
to be longer, an observer moving relative to the process
motion? (e) Do he and an earthbound observer agree on his
or an observer moving with the process? Which observer
velocity relative to Earth?
measures the interval of proper time?

This OpenStax book is available for free at https://legacy.cnx.org/content/col12169/1.46


Chapter 5 | Relativistic Kinematics in One Dimension 165

PROBLEMS
observer moving with it? Base your calculation on its
5.3 Time Dilation velocity relative to the Earth and the time it lives (proper
10. (a) What is γ if v = 0.250c ? (b) If v = 0.500c ? time). (c) Verify that these two distances are related through
length contraction γ = 3.20.

11. (a) What is γ if v = 0.100c ? (b) If v = 0.900c ?


21. (a) How long would the muon in m58563
(https://legacy.cnx.org/content/m58563/latest/#fs-
12. Particles called π -mesons are produced by accelerator id1167794027273) have lived as observed on Earth if its
beams. If these particles travel at 2.70 × 10 8 m/s and live velocity was 0.0500c ? (b) How far would it have traveled
as observed on Earth? (c) What distance is this in the
2.60 × 10 −8 s when at rest relative to an observer, how muon’s frame?
long do they live as viewed in the laboratory?
22. Unreasonable Results A spaceship is heading
13. Suppose a particle called a kaon is created by cosmic directly toward Earth at a velocity of 0.800c. The astronaut
radiation striking the atmosphere. It moves by you at on board claims that he can send a canister toward the Earth
0.980c, and it lives 1.24 × 10 −8 s when at rest relative at 1.20c relative to Earth. (a) Calculate the velocity the
to an observer. How long does it live as you observe it? canister must have relative to the spaceship. (b) What is
unreasonable about this result? (c) Which assumptions are
unreasonable or inconsistent?
14. A neutral π -meson is a particle that can be created
by accelerator beams. If one such particle lives
1.40 × 10 −16 s as measured in the laboratory, and 5.5 The Lorentz Transformation
0.840 × 10 −16 s when at rest relative to an observer, what 23. Describe the following physical occurrences as events,
is its velocity relative to the laboratory? that is, in the form (x, y, z, t): (a) A postman rings a doorbell
of a house precisely at noon. (b) At the same time as the
15. A neutron lives 900 s when at rest relative to an doorbell is rung, a slice of bread pops out of a toaster that
observer. How fast is the neutron moving relative to an is located 10 m from the door in the east direction from the
observer who measures its life span to be 2065 s? door. (c) Ten seconds later, an airplane arrives at the airport,
which is 10 km from the door in the east direction and 2 km
to the south.
16. If relativistic effects are to be less than 1%, then
γ must be less than 1.01. At what relative velocity is
24. Describe what happens to the angle α = tan(v/c),
γ = 1.01 ?
and therefore to the transformed axes in Figure 5.18, as
the relative velocity v of the S and S′ frames of reference
17. If relativistic effects are to be less than 3%, then approaches c.
γ must be less than 1.03. At what relative velocity is
γ = 1.03 ? 25. Describe the shape of the world line on a space-
time diagram of (a) an object that remains at rest at a
specific position along the x-axis; (b) an object that moves
at constant velocity u in the x-direction; (c) an object that
5.4 Length Contraction
begins at rest and accelerates at a constant rate of in the
18. A spaceship, 200 m long as seen on board, moves by positive x-direction.
the Earth at 0.970c. What is its length as measured by an
earthbound observer?
26. A man standing still at a train station watches two boys
throwing a baseball in a moving train. Suppose the train is
19. How fast would a 6.0 m-long sports car have to be moving east with a constant speed of 20 m/s and one of the
going past you in order for it to appear only 5.5 m long? boys throws the ball with a speed of 5 m/s with respect to
himself toward the other boy, who is 5 m west from him.
20. (a) How far does the muon in m58563 What is the velocity of the ball as observed by the man on
(https://legacy.cnx.org/content/m58563/latest/#fs- the station?
id1167794027273) travel according to the earthbound
observer? (b) How far does it travel as viewed by an 27. When observed from the sun at a particular instant,
166 Chapter 5 | Relativistic Kinematics in One Dimension

Earth and Mars appear to move in opposite directions with S′ is moving in the positive direction along the positive
speeds 108,000 km/h and 86,871 km/h, respectively. What x-axis with a constant speed v and observes the same two
is the speed of Mars at this instant when observed from events in his frame. The origins of the two frames coincide
Earth? at t = t′ = 0. (a) Find the positions and timings of these
two events in the frame S′ (a) according to the Galilean
28. A man is running on a straight road perpendicular to transformation, and (b) according to the Lorentz
a train track and away from the track at a speed of 12 m/ transformation.
s. The train is moving with a speed of 30 m/s with respect
to the track. What is the speed of the man with respect to a
passenger sitting at rest in the train?
5.6 Relativistic Velocity Transformation
29. A man is running on a straight road that makes 30° 33. If two spaceships are heading directly toward each
other at 0.800c, at what speed must a canister be shot from
with the train track. The man is running in the direction on
the first ship to approach the other at 0.999c as seen by the
the road that is away from the track at a speed of 12 m/s.
second ship?
The train is moving with a speed of 30 m/s with respect to
the track. What is the speed of the man with respect to a
passenger sitting at rest in the train? 34. Two planets are on a collision course, heading directly
toward each other at 0.250c. A spaceship sent from one
planet approaches the second at 0.750c as seen by the
30. In a frame at rest with respect to the billiard table, a
second planet. What is the velocity of the ship relative to
billiard ball of mass m moving with speed v strikes another
the first planet?
billiard ball of mass m at rest. The first ball comes to rest
after the collision while the second ball takes off with speed
v in the original direction of the motion of the first ball. 35. When a missile is shot from one spaceship toward
This shows that momentum is conserved in this frame. (a) another, it leaves the first at 0.950c and approaches the
Now, describe the same collision from the perspective of a other at 0.750c. What is the relative velocity of the two
frame that is moving with speed v in the direction of the ships?
motion of the first ball. (b) Is the momentum conserved in
this frame? 36. What is the relative velocity of two spaceships if one
fires a missile at the other at 0.750c and the other observes
31. In a frame at rest with respect to the billiard table, it to approach at 0.950c?
two billiard balls of same mass m are moving toward each
other with the same speed v. After the collision, the two 37. Prove that for any relative velocity v between two
balls come to rest. (a) Show that momentum is conserved observers, a beam of light sent from one to the other will
in this frame. (b) Now, describe the same collision from approach at speed c (provided that v is less than c, of
the perspective of a frame that is moving with speed v course).
in the direction of the motion of the first ball. (c) Is the
momentum conserved in this frame? 38. Show that for any relative velocity v between two
observers, a beam of light projected by one directly away
32. In a frame S, two events are observed: event 1: a from the other will move away at the speed of light
pion is created at rest at the origin and event 2: the pion (provided that v is less than c, of course).
disintegrates after time τ . Another observer in a frame

ADDITIONAL PROBLEMS
39. (a) At what relative velocity is γ = 1.50 ? (b) At what a high-velocity space probe indicate that 24.0 h have passed
relative velocity is γ = 100 ? on board. (b) What is unreasonable about this result? (c)
Which assumptions are unreasonable or inconsistent?

40. (a) At what relative velocity is γ = 2.00 ? (b) At what 42. (a) How long does it take the astronaut in Example
relative velocity is γ = 10.0 ? 5.5 to travel 4.30 ly at 0.99944c (as measured by the
earthbound observer)? (b) How long does it take according
to the astronaut? (c) Verify that these two times are related
41. Unreasonable Results (a) Find the value of γ
through time dilation with γ = 30.00 as given.
required for the following situation. An earthbound
observer measures 23.9 h to have passed while signals from
43. (a) How fast would an athlete need to be running

This OpenStax book is available for free at https://legacy.cnx.org/content/col12169/1.46


Chapter 5 | Relativistic Kinematics in One Dimension 167

for a 100- m race to look 100 yd long? (b) Is the answer 51. An observer sees two events 1.5 × 10 −8 s apart at
consistent with the fact that relativistic effects are difficult a separation of 800 m. How fast must a second observer
to observe in ordinary circumstances? Explain. be moving relative to the first to see the two events occur
simultaneously?
44. (a) Find the value of γ for the following situation. An
astronaut measures the length of his spaceship to be 100 m, 52. An observer standing by the railroad tracks sees two
while an earthbound observer measures it to be 25.0 m. (b) bolts of lightning strike the ends of a 500-m-long train
What is the speed of the spaceship relative to Earth? simultaneously at the instant the middle of the train passes
him at 50 m/s. Use the Lorentz transformation to find the
45. A clock in a spaceship runs one-tenth the rate at which time between the lightning strikes as measured by a
an identical clock on Earth runs. What is the speed of the passenger seated in the middle of the train.
spaceship?
53. Two astronomical events are observed from Earth to
46. An astronaut has a heartbeat rate of 66 beats per occur at a time of 1 s apart and a distance separation of
minute as measured during his physical exam on Earth. The 1.5 × 10 9 m from each other. (a) Determine whether
heartbeat rate of the astronaut is measured when he is in separation of the two events is space like or time like. (b)
a spaceship traveling at 0.5c with respect to Earth by an State what this implies about whether it is consistent with
observer (A) in the ship and by an observer (B) on Earth. special relativity for one event to have caused the other?
(a) Describe an experimental method by which observer B
on Earth will be able to determine the heartbeat rate of the 54. Two astronomical events are observed from Earth to
astronaut when the astronaut is in the spaceship. (b) What occur at a time of 0.30 s apart and a distance separation of
will be the heartbeat rate(s) of the astronaut reported by
observers A and B?
2.0 × 10 9 m from each other. How fast must a spacecraft
travel from the site of one event toward the other to make
the events occur at the same time when measured in the
47. A spaceship (A) is moving at speed c/2 with respect
frame of reference of the spacecraft?
to another spaceship (B). Observers in A and B set their
clocks so that the event at (x, y, z, t) of turning on a laser in
spaceship B has coordinates (0, 0, 0, 0) in A and also (0, 0, 55. A spacecraft starts from being at rest at the origin and
0, 0) in B. An observer at the origin of B turns on the laser accelerates at a constant rate g, as seen from Earth, taken to
at t = 0 and turns it off at t = τ in his time. What is the be an inertial frame, until it reaches a speed of c/2. (a) Show
time duration between on and off as seen by an observer in that the increment of proper time is related to the elapsed
A? time in Earth’s frame by: dτ = 1 − v 2/c 2dt.
(b) Find an expression for the elapsed time to reach speed
48. Same two observers as in the preceding exercise, but c/2 as seen in Earth’s frame. (c) Use the relationship in (a)
now we look at two events occurring in spaceship A. A to obtain a similar expression for the elapsed proper time to
photon arrives at the origin of A at its time t = 0 and reach c/2 as seen in the spacecraft, and determine the ratio
another photon arrives at (x = 1.00 m, 0, 0) at t = 0 in of the time seen from Earth with that on the spacecraft to
the frame of ship A. (a) Find the coordinates and times of reach the final speed.
the two events as seen by an observer in frame B. (b) In
which frame are the two events simultaneous and in which 56. (a) All but the closest galaxies are receding from our
frame are they are not simultaneous? own Milky Way Galaxy. If a galaxy 12.0 × 10 9 ly away
is receding from us at 0.900c, at what velocity relative to
49. Same two observers as in the preceding exercises. A us must we send an exploratory probe to approach the other
rod of length 1 m is laid out on the x-axis in the frame of galaxy at 0.990c as measured from that galaxy? (b) How
B from origin to (x = 1.00 m, 0, 0). What is the length of long will it take the probe to reach the other galaxy as
the rod observed by an observer in the frame of spaceship measured from Earth? You may assume that the velocity of
A? the other galaxy remains constant. (c) How long will it then
take for a radio signal to be beamed back? (All of this is
50. An observer at origin of inertial frame S sees a possible in principle, but not practical.)
flashbulb go off at x = 150 km, y = 15.0 km, and
57. Suppose a spaceship heading straight toward the Earth
z = 1.00 km at time t = 4.5 × 10 −4 s. At what time and
at 0.750c can shoot a canister at 0.500c relative to the ship.
position in the S ′ system did the flash occur, if S ′ is (a) What is the velocity of the canister relative to Earth, if it
moving along shared x-direction with S at a velocity is shot directly at Earth? (b) If it is shot directly away from
v = 0.6c ? Earth?
168 Chapter 5 | Relativistic Kinematics in One Dimension

58. Repeat the preceding problem with the ship heading 60. (a) Suppose the speed of light were only 3000 m/s.
directly away from Earth. A jet fighter moving toward a target on the ground at 800
m/s shoots bullets, each having a muzzle velocity of 1000
59. If a spaceship is approaching the Earth at 0.100c and m/s. What are the bullets’ velocity relative to the target?
a message capsule is sent toward it at 0.100c relative to (b) If the speed of light was this small, would you observe
Earth, what is the speed of the capsule relative to the ship? relativistic effects in everyday life? Discuss.

FOR FURTHER EXPLORATION


https://www.youtube.com/watch?v=wteiuxyqtoM
5.3 Time Dilation YouTube animation expalining the relativity of
Videos simultanaeity (2:03)

Time Dilation; an experiment with mu-mesons (1960 Time Dilation - Albert Einstein and the Theory of
Educational Physics) https://www.youtube.com/ Relativity: https://www.youtube.com/
watch?v=tbsdrHlLfVQ&t=4s A classic 1960 educational watch?v=KHjpBjgIMVk&t=13s YouTube animation
video showing the verification of time dilation in the muon- explaining time dilation (1:22)
lifetime experiment (35:53)
Simultaneity - Albert Einstein and the Theory of Relativity

This OpenStax book is available for free at https://legacy.cnx.org/content/col12169/1.46


Chapter 6 | Introduction to Vectors 169

6 | INTRODUCTION TO
VECTORS

Figure 6.1 A signpost gives information about distances and directions to towns or to other locations relative to the location of
the signpost. Distance is a scalar quantity. Knowing the distance alone is not enough to get to the town; we must also know the
direction from the signpost to the town. The direction, together with the distance, is a vector quantity commonly called the
displacement vector. A signpost, therefore, gives information about displacement vectors from the signpost to towns. (credit:
modification of work by “studio tdes”/Flickr)

Chapter Outline
6.1 Scalars and Vectors
6.2 Coordinate Systems and Components of a Vector
6.3 Algebra of Vectors

Introduction
Vectors are essential to physics and engineering. Many fundamental physical quantities are vectors, including displacement,
velocity, force, and electric and magnetic vector fields. Scalar products of vectors define other fundamental scalar physical
quantities, such as energy. Vector products of vectors define still other fundamental vector physical quantities, such as torque
and angular momentum. In other words, vectors are a component part of physics in much the same way as sentences are a
component part of literature.
In introductory physics, vectors are Euclidean quantities that have geometric representations as arrows in one dimension (in
a line), in two dimensions (in a plane), or in three dimensions (in space). They can be added, subtracted, or multiplied. In
this chapter, we explore elements of vector algebra for applications in mechanics and in electricity and magnetism. Vector
operations also have numerous generalizations in other branches of physics.
170 Chapter 6 | Introduction to Vectors

6.1 | Scalars and Vectors


Learning Objectives
By the end of this section, you will be able to:
• Describe the difference between vector and scalar quantities.
• Identify the magnitude and direction of a vector.
• Explain the effect of multiplying a vector quantity by a scalar.
• Describe how one-dimensional vector quantities are added or subtracted.
• Explain the geometric construction for the addition or subtraction of vectors in a plane.
• Distinguish between a vector equation and a scalar equation.

Many familiar physical quantities can be specified completely by giving a single number and the appropriate unit. For
example, “a class period lasts 50 min” or “the gas tank in my car holds 65 L” or “the distance between two posts is 100
m.” A physical quantity that can be specified completely in this manner is called a scalar quantity. Scalar is a synonym of
“number.” Time, mass, distance, length, volume, temperature, and energy are examples of scalar quantities.
Scalar quantities that have the same physical units can be added or subtracted according to the usual rules of algebra for
numbers. For example, a class ending 10 min earlier than 50 min lasts 50 min − 10 min = 40 min . Similarly, a 60-cal
serving of corn followed by a 200-cal serving of donuts gives 60 cal + 200 cal = 260 cal of energy. When we multiply
a scalar quantity by a number, we obtain the same scalar quantity but with a larger (or smaller) value. For example, if
yesterday’s breakfast had 200 cal of energy and today’s breakfast has four times as much energy as it had yesterday, then
today’s breakfast has 4(200 cal) = 800 cal of energy. Two scalar quantities can also be multiplied or divided by each other
to form a derived scalar quantity. For example, if a train covers a distance of 100 km in 1.0 h, its speed is 100.0 km/1.0 h =
27.8 m/s, where the speed is a derived scalar quantity obtained by dividing distance by time.
Many physical quantities, however, cannot be described completely by just a single number of physical units. For example,
when the U.S. Coast Guard dispatches a ship or a helicopter for a rescue mission, the rescue team must know not only the
distance to the distress signal, but also the direction from which the signal is coming so they can get to its origin as quickly
as possible. Physical quantities specified completely by giving a number of units (magnitude) and a direction are called
vector quantities. Examples of vector quantities include displacement, velocity, position, force, and torque. In the language
of mathematics, physical vector quantities are represented by mathematical objects called vectors (Figure 6.2). We can
add or subtract two vectors, and we can multiply a vector by a scalar or by another vector, but we cannot divide by a vector.
The operation of division by a vector is not defined.

Figure 6.2 We draw a vector from the initial point or origin


(called the “tail” of a vector) to the end or terminal point (called
the “head” of a vector), marked by an arrowhead. Magnitude is
the length of a vector and is always a positive scalar quantity.
(credit: modification of work by Cate Sevilla)

Let’s examine vector algebra using a graphical method to be aware of basic terms and to develop a qualitative
understanding. In practice, however, when it comes to solving physics problems, we use analytical methods, which we’ll
see in the next section. Analytical methods are more simple computationally and more accurate than graphical methods.
From now on, to distinguish between a vector and a scalar quantity, we adopt the common convention that a letter in bold

This OpenStax book is available for free at https://legacy.cnx.org/content/col12169/1.46


Chapter 6 | Introduction to Vectors 171

type with an arrow above it denotes a vector, and a letter without an arrow denotes a scalar. For example, a distance of 2.0
km, which is a scalar quantity, is denoted by d = 2.0 km, whereas a displacement of 2.0 km in some direction, which is a

vector quantity, is denoted by d .
Suppose you tell a friend on a camping trip that you have discovered a terrific fishing hole 6 km from your tent. It is unlikely
your friend would be able to find the hole easily unless you also communicate the direction in which it can be found with
respect to your campsite. You may say, for example, “Walk about 6 km northeast from my tent.” The key concept here is
that you have to give not one but two pieces of information—namely, the distance or magnitude (6 km) and the direction
(northeast).
Displacement is a general term used to describe a change in position, such as during a trip from the tent to the fishing hole.
Displacement is an example of a vector quantity. If you walk from the tent (location A) to the hole (location B), as shown

in Figure 6.3, the vector D , representing your displacement, is drawn as the arrow that originates at point A and ends

at point B. The arrowhead marks the end of the vector. The direction of the displacement vector D is the direction of the

arrow. The length of the arrow represents the magnitude D of vector D . Here, D = 6 km. Since the magnitude of a vector
is its length, which is a positive number, the magnitude is also indicated by placing the absolute value notation around the

| |

symbol that denotes the vector; so, we can write equivalently that D ≡ D . To solve a vector problem graphically, we

need to draw the vector D to scale. For example, if we assume 1 unit of distance (1 km) is represented in the drawing
by a line segment of length u = 2 cm, then the total displacement in this example is represented by a vector of length
d = 6u = 6(2 cm) = 12 cm , as shown in Figure 6.4. Notice that here, to avoid confusion, we used D = 6 km to denote
the magnitude of the actual displacement and d = 12 cm to denote the length of its representation in the drawing.

Figure 6.3 The displacement vector from point A (the initial


position at the campsite) to point B (the final position at the
fishing hole) is indicated by an arrow with origin at point A and
end at point B. The displacement is the same for any of the
actual paths (dashed curves) that may be taken between points A
and B.
172 Chapter 6 | Introduction to Vectors


Figure 6.4 A displacement D of magnitude 6 km is drawn
to scale as a vector of length 12 cm when the length of 2 cm
represents 1 unit of displacement (which in this case is 1 km).

Suppose your friend walks from the campsite at A to the fishing pond at B and then walks back: from the fishing pond at

B to the campsite at A. The magnitude of the displacement vector D AB from A to B is the same as the magnitude of the

displacement vector D BA from B to A (it equals 6 km in both cases), so we can write D AB = D BA . However, vector
→ → → →
D AB is not equal to vector D BA because these two vectors have different directions: D AB ≠ D BA . In Figure

6.3, vector D BA would be represented by a vector with an origin at point B and an end at point A, indicating vector
→ →
D BA points to the southwest, which is exactly 180° opposite to the direction of vector D AB . We say that vector
→ → → →
D BA is antiparallel to vector D AB and write D AB = − D BA , where the minus sign indicates the antiparallel
direction.
Two vectors that have identical directions are said to be parallel vectors—meaning, they are parallel to each other. Two
→ → → →
parallel vectors A and B are equal, denoted by A = B , if and only if they have equal magnitudes
| | | |
→ →
A = B .
Two vectors with directions perpendicular to each other are said to be orthogonal vectors. These relations between vectors
are illustrated in Figure 6.5.

This OpenStax book is available for free at https://legacy.cnx.org/content/col12169/1.46


Chapter 6 | Introduction to Vectors 173

→ →
Figure 6.5 Various relations between two vectors A and B . (a)
→ → → →
A ≠ B because A ≠ B . (b) A ≠ B because they are not
→ →
parallel and A ≠ B . (c) A ≠ − A because they have different

| | |
→ →
| →
directions (even though A = − A = A) . (d) A = B

because they are parallel and have identical magnitudes A = B. (e)


→ →
A ≠ B because they have different directions (are not parallel);
here, their directions differ by 90° —meaning, they are orthogonal.

6.1 Check Your Understanding Two motorboats named Alice and Bob are moving on a lake. Given the
information about their velocity vectors in each of the following situations, indicate whether their velocity
vectors are equal or otherwise. (a) Alice moves north at 6 knots and Bob moves west at 6 knots. (b) Alice moves
west at 6 knots and Bob moves west at 3 knots. (c) Alice moves northeast at 6 knots and Bob moves south at 3
knots. (d) Alice moves northeast at 6 knots and Bob moves southwest at 6 knots. (e) Alice moves northeast at 2
knots and Bob moves closer to the shore northeast at 2 knots.

Algebra of Vectors in One Dimension


Vectors can be multiplied by scalars, added to other vectors, or subtracted from other vectors. We can illustrate these vector
concepts using an example of the fishing trip seen in Figure 6.6.
174 Chapter 6 | Introduction to Vectors

Figure 6.6 Displacement vectors for a fishing trip. (a) Stopping to rest at point C while walking from camp (point A) to the
pond (point B). (b) Going back for the dropped tackle box (point D). (c) Finishing up at the fishing pond.

Suppose your friend departs from point A (the campsite) and walks in the direction to point B (the fishing pond), but,
along the way, stops to rest at some point C located three-quarters of the distance between A and B, beginning from

point A (Figure 6.6(a)). What is his displacement vector D AC when he reaches point C? We know that if he walks

all the way to B, his displacement vector relative to A is D AB , which has magnitude D AB = 6 km and a direction
of northeast. If he walks only a 0.75 fraction of the total distance, maintaining the northeasterly direction, at point C he
must be 0.75D AB = 4.5 km away from the campsite at A. So, his displacement vector at the rest point C has magnitude

D AC = 4.5 km = 0.75D AB and is parallel to the displacement vector D AB . All of this can be stated succinctly in the
form of the following vector equation:
→ →
D AC = 0.75 D AB.

In a vector equation, both sides of the equation are vectors. The previous equation is an example of a vector multiplied by a

positive scalar (number) α = 0.75 . The result, D AC , of such a multiplication is a new vector with a direction parallel to

the direction of the original vector D AB .
→ → →
In general, when a vector A is multiplied by a positive scalar α , the result is a new vector B that is parallel to A :

→ → (6.1)
B =α A .

The magnitude
| |
→ →
| |
B of this new vector is obtained by multiplying the magnitude A of the original vector, as expressed
by the scalar equation:

B = |α| A. (6.2)

In a scalar equation, both sides of the equation are numbers. Equation 6.2 is a scalar equation because the magnitudes
of vectors are scalar quantities (and positive numbers). If the scalar α is negative in the vector equation Equation 6.1,
then the magnitude
| |
→ →
B of the new vector is still given by Equation 6.2, but the direction of the new vector B is

This OpenStax book is available for free at https://legacy.cnx.org/content/col12169/1.46


Chapter 6 | Introduction to Vectors 175


antiparallel to the direction of A . These principles are illustrated in Figure 6.7(a) by two examples where the length
→ → →
of vector A is 1.5 units. When α = 2 , the new vector B = 2 A has length B = 2A = 3.0 units (twice as long
→ →
as the original vector) and is parallel to the original vector. When α = −2 , the new vector C = −2 A has length
C = | − 2| A = 3.0 units (twice as long as the original vector) and is antiparallel to the original vector.

Figure 6.7 Algebra of vectors in one dimension. (a) Multiplication by a



scalar. (b) Addition of two vectors ( R is called the resultant of vectors
→ → →
A and B ) . (c) Subtraction of two vectors ( D is the difference of
→ →
vectors A and B ) .

Now suppose your fishing buddy departs from point A (the campsite), walking in the direction to point B (the fishing
hole), but he realizes he lost his tackle box when he stopped to rest at point C (located three-quarters of the distance
between A and B, beginning from point A). So, he turns back and retraces his steps in the direction toward the campsite
and finds the box lying on the path at some point D only 1.2 km away from point C (see Figure 6.6(b)). What is his
→ →
displacement vector D AD when he finds the box at point D? What is his displacement vector D DB from point D to the
→ →
hole? We have already established that at rest point C his displacement vector is D AC = 0.75 D AB . Starting at point

C, he walks southwest (toward the campsite), which means his new displacement vector D CD from point C to point

→ →

D is antiparallel to D AB . Its magnitude
| →
D CD is

| D CD = 1.2 km = 0.2D AB , so his second displacement vector is

D CD = −0.2 D AB . His total displacement D AD relative to the campsite is the vector sum of the two displacement
→ →
vectors: vector D AC (from the campsite to the rest point) and vector D CD (from the rest point to the point where he
finds his box):

→ → → (6.3)
D AD = D AC + D CD.

The vector sum of two (or more) vectors is called the resultant vector or, for short, the resultant. When the vectors on the

right-hand-side of Equation 6.3 are known, we can find the resultant D AD as follows:
→ → → → → → → (6.4)
D AD = D AC + D CD = 0.75 D AB − 0.2 D AB = (0.75 − 0.2) D AB = 0.55 D AB.

When your friend finally reaches the pond at B, his displacement vector D AB from point A is the vector sum of his
176 Chapter 6 | Introduction to Vectors

→ →
displacement vector D AD from point A to point D and his displacement vector D DB from point D to the fishing hole:
→ → → →
D AB = D AD + D DB (see Figure 6.6(c)). This means his displacement vector D DB is the difference of two
vectors:
→ → → → → (6.5)
D DB = D AB − D AD = D AB + (− D AD).

Notice that a difference of two vectors is nothing more than a vector sum of two vectors because the second term in
→ →
Equation 6.5 is vector − D AD (which is antiparallel to D AD) . When we substitute Equation 6.4 into Equation
6.5, we obtain the second displacement vector:
→ → → → → → → (6.6)
D DB = D AB − D AD = D AB − 0.55 D AB = (1.0 − 0.55) D AB = 0.45 D AB.

This result means your friend walked D DB = 0.45D AB = 0.45(6.0 km) = 2.7 km from the point where he finds his tackle
box to the fishing hole.
→ →
When vectors A and B lie along a line (that is, in one dimension), such as in the camping example, their resultant
→ → → → → →
R = A + B and their difference D = A − B both lie along the same direction. We can illustrate the addition
or subtraction of vectors by drawing the corresponding vectors to scale in one dimension, as shown in Figure 6.7.
→ →
To illustrate the resultant when A and B are two parallel vectors, we draw them along one line by placing the origin
of one vector at the end of the other vector in head-to-tail fashion (see Figure 6.7(b)). The magnitude of this resultant

is the sum of their magnitudes: R = A + B. The direction of the resultant is parallel to both vectors. When vector A

is antiparallel to vector B , we draw them along one line in either head-to-head fashion (Figure 6.7(c)) or tail-to-
tail fashion. The magnitude of the vector difference, then, is the absolute value D = | A − B| of the difference of their

magnitudes. The direction of the difference vector D is parallel to the direction of the longer vector.
In general, in one dimension—as well as in higher dimensions, such as in a plane or in space—we can add any number of
vectors and we can do so in any order because the addition of vectors is commutative,

→ → → → (6.7)
A + B = B + A ,

and associative,

→ → → → → → (6.8)
( A + B ) + C = A + ( B + C ).

Moreover, multiplication by a scalar is distributive:

→ → → (6.9)
α 1 A + α 2 A = (α 1 + α 2) A .

We used the distributive property in Equation 6.4 and Equation 6.6.


When adding many vectors in one dimension, it is convenient to use the concept of a unit vector. A unit vector, which
is denoted by a letter symbol with a hat, such as ^
u , has a magnitude of one and does not have any physical unit so that

||
^ →
u ≡ u = 1 . The only role of a unit vector is to specify direction. For example, instead of saying vector D AB has a

This OpenStax book is available for free at https://legacy.cnx.org/content/col12169/1.46


Chapter 6 | Introduction to Vectors 177

magnitude of 6.0 km and a direction of northeast, we can introduce a unit vector ^


u that points to the northeast and say

succinctly that D AB = (6.0 km) ^ u . Then the southwesterly direction is simply given by the unit vector − ^
u . In this way,
the displacement of 6.0 km in the southwesterly direction is expressed by the vector

D BA = (−6.0 km) ^
u.

Example 6.1

A Ladybug Walker
A long measuring stick rests against a wall in a physics laboratory with its 200-cm end at the floor. A ladybug
lands on the 100-cm mark and crawls randomly along the stick. It first walks 15 cm toward the floor, then it walks
56 cm toward the wall, then it walks 3 cm toward the floor again. Then, after a brief stop, it continues for 25 cm
toward the floor and then, again, it crawls up 19 cm toward the wall before coming to a complete rest (Figure
6.8). Find the vector of its total displacement and its final resting position on the stick.
Strategy

If we choose the direction along the stick toward the floor as the direction of unit vector ^
u , then the direction
toward the floor is + ^
u and the direction toward the wall is − ^
u . The ladybug makes a total of five
displacements:

D 1 = (15 cm)( + ^
u ),

D 2 = (56 cm)(− ^
u ),

D 3 = (3 cm)( + ^
u ),
→ ^
D 4 = (25 cm)( + u ), and

D 5 = (19 cm)(− ^
u ).

The total displacement D is the resultant of all its displacement vectors.

Figure 6.8 Five displacements of the ladybug. Note that in this schematic drawing,
magnitudes of displacements are not drawn to scale. (credit: modification of work by
“Persian Poet Gal”/Wikimedia Commons)
178 Chapter 6 | Introduction to Vectors

Solution
The resultant of all the displacement vectors is
→ → → → → →
D = D 1+ D 2+ D 3+ D 4+ D 5
= (15 cm)( + ^
u ) + (56 cm)(− ^
u ) + (3 cm)( + ^
u ) + (25 cm)( + ^
u ) + (19 cm)(− ^
u)
= (15 − 56 + 3 + 25 − 19)cm ^u
= −32 cm ^
u.
In this calculation, we use the distributive law given by Equation 6.9. The result reads that the total
displacement vector points away from the 100-cm mark (initial landing site) toward the end of the meter stick
that touches the wall. The end that touches the wall is marked 0 cm, so the final position of the ladybug is at the
(100 – 32)cm = 68-cm mark.

6.2 Check Your Understanding A cave diver enters a long underwater tunnel. When her displacement with
respect to the entry point is 20 m, she accidentally drops her camera, but she doesn’t notice it missing until she
is some 6 m farther into the tunnel. She swims back 10 m but cannot find the camera, so she decides to end the
dive. How far from the entry point is she? Taking the positive direction out of the tunnel, what is her
displacement vector relative to the entry point?

Algebra of Vectors in Two Dimensions


When vectors lie in a plane—that is, when they are in two dimensions—they can be multiplied by scalars, added to other
vectors, or subtracted from other vectors in accordance with the general laws expressed by Equation 6.1, Equation 6.2,
Equation 6.7, and Equation 6.8. However, the addition rule for two vectors in a plane becomes more complicated than
the rule for vector addition in one dimension. We have to use the laws of geometry to construct resultant vectors, followed
by trigonometry to find vector magnitudes and directions. This geometric approach is commonly used in navigation
(Figure 6.9). In this section, we need to have at hand two rulers, a triangle, a protractor, a pencil, and an eraser for drawing
vectors to scale by geometric constructions.

Figure 6.9 In navigation, the laws of geometry are used to draw resultant displacements on
nautical maps.

This OpenStax book is available for free at https://legacy.cnx.org/content/col12169/1.46


Chapter 6 | Introduction to Vectors 179

For a geometric construction of the sum of two vectors in a plane, we follow the parallelogram rule. Suppose two vectors
→ →
A and B are at the arbitrary positions shown in Figure 6.10. Translate either one of them in parallel to the beginning
of the other vector, so that after the translation, both vectors have their origins at the same point. Now, at the end of vector
→ → → →
A we draw a line parallel to vector B and at the end of vector B we draw a line parallel to vector A (the dashed
lines in Figure 6.10). In this way, we obtain a parallelogram. From the origin of the two vectors we draw a diagonal that is
→ → → →
the resultant R of the two vectors: R = A + B (Figure 6.10(a)). The other diagonal of this parallelogram is the
→ → →
vector difference of the two vectors D = A − B , as shown in Figure 6.10(b). Notice that the end of the difference

vector is placed at the end of vector A .

Figure 6.10 The parallelogram rule for the addition of two vectors. Make the parallel translation of each vector to a point
where their origins (marked by the dot) coincide and construct a parallelogram with two sides on the vectors and the other

two sides (indicated by dashed lines) parallel to the vectors. (a) Draw the resultant vector R along the diagonal of the
parallelogram from the common point to the opposite corner. Length R of the resultant vector is not equal to the sum of the
→ → →
magnitudes of the two vectors. (b) Draw the difference vector D = A − B along the diagonal connecting the ends of
→ → →
the vectors. Place the origin of vector D at the end of vector B and the end (arrowhead) of vector D at the end of

vector A . Length D of the difference vector is not equal to the difference of magnitudes of the two vectors.

It follows from the parallelogram rule that neither the magnitude of the resultant vector nor the magnitude of the difference
vector can be expressed as a simple sum or difference of magnitudes A and B, because the length of a diagonal cannot be
expressed as a simple sum of side lengths. When using a geometric construction to find magnitudes
| | | |
→ →
R and D , we
have to use trigonometry laws for triangles, which may lead to complicated algebra. There are two ways to circumvent this
algebraic complexity. One way is to use the method of components, which we examine in the next section. The other way is
to draw the vectors to scale, as is done in navigation, and read approximate vector lengths and angles (directions) from the
graphs. In this section we examine the second approach.
If we need to add three or more vectors, we repeat the parallelogram rule for the pairs of vectors until we find the resultant
of all of the resultants. For three vectors, for example, we first find the resultant of vector 1 and vector 2, and then we
find the resultant of this resultant and vector 3. The order in which we select the pairs of vectors does not matter because
the operation of vector addition is commutative and associative (see Equation 6.7 and Equation 6.8). Before we state a
general rule that follows from repetitive applications of the parallelogram rule, let’s look at the following example.
Suppose you plan a vacation trip in Florida. Departing from Tallahassee, the state capital, you plan to visit your uncle
Joe in Jacksonville, see your cousin Vinny in Daytona Beach, stop for a little fun in Orlando, see a circus performance
in Tampa, and visit the University of Florida in Gainesville. Your route may be represented by five displacement vectors
→ → → → →
A , B , C , D , and E , which are indicated by the red vectors in Figure 6.11. What is your total displacement
when you reach Gainesville? The total displacement is the vector sum of all five displacement vectors, which may be
found by using the parallelogram rule four times. Alternatively, recall that the displacement vector has its beginning at
the initial position (Tallahassee) and its end at the final position (Gainesville), so the total displacement vector can be
drawn directly as an arrow connecting Tallahassee with Gainesville (see the green vector in Figure 6.11). When we use

the parallelogram rule four times, the resultant R we obtain is exactly this green vector connecting Tallahassee with
→ → → → → →
Gainesville: R = A + B + C + D + E .
180 Chapter 6 | Introduction to Vectors

Figure 6.11 When we use the parallelogram rule four times, we obtain the resultant vector
→ → → → → →
R = A + B + C + D + E , which is the green vector connecting Tallahassee with Gainesville.

Drawing the resultant vector of many vectors can be generalized by using the following tail-to-head geometric
→ → → → →
construction. Suppose we want to draw the resultant vector R of four vectors A , B , C , and D (Figure
6.12(a)). We select any one of the vectors as the first vector and make a parallel translation of a second vector to a position
where the origin (“tail”) of the second vector coincides with the end (“head”) of the first vector. Then, we select a third
vector and make a parallel translation of the third vector to a position where the origin of the third vector coincides with
the end of the second vector. We repeat this procedure until all the vectors are in a head-to-tail arrangement like the one

shown in Figure 6.12. We draw the resultant vector R by connecting the origin (“tail”) of the first vector with the end
(“head”) of the last vector. The end of the resultant vector is at the end of the last vector. Because the addition of vectors is
associative and commutative, we obtain the same resultant vector regardless of which vector we choose to be first, second,
third, or fourth in this construction.

This OpenStax book is available for free at https://legacy.cnx.org/content/col12169/1.46


Chapter 6 | Introduction to Vectors 181

Figure 6.12 Tail-to-head method for drawing the resultant vector


→ → → → →
R = A + B + C + D . (a) Four vectors of different magnitudes and
directions. (b) Vectors in (a) are translated to new positions where the origin (“tail”) of
one vector is at the end (“head”) of another vector. The resultant vector is drawn from
the origin (“tail”) of the first vector to the end (“head”) of the last vector in this
arrangement.

Example 6.2

Geometric Construction of the Resultant


→ → →
The three displacement vectors A , B , and C in Figure 6.13 are specified by their magnitudes A = 10.0,
B = 7.0, and C = 8.0, respectively, and by their respective direction angles with the horizontal direction α = 35° ,
β = −110° , and γ = 30° . The physical units of the magnitudes are centimeters. Choose a convenient scale and
→ → → → → →
use a ruler and a protractor to find the following vector sums: (a) R = A + B , (b) D = A − B , and
→ → → →
(c) S = A − 3 B + C .

Figure 6.13 Vectors used in Example 6.2 and in the Check Your Understanding feature that follows.

Strategy
In geometric construction, to find a vector means to find its magnitude and its direction angle with the horizontal
direction. The strategy is to draw to scale the vectors that appear on the right-hand side of the equation and
construct the resultant vector. Then, use a ruler and a protractor to read the magnitude of the resultant and the
direction angle. For parts (a) and (b) we use the parallelogram rule. For (c) we use the tail-to-head method.
Solution
→ →
For parts (a) and (b), we attach the origin of vector B to the origin of vector A , as shown in Figure 6.14,
→ →
and construct a parallelogram. The shorter diagonal of this parallelogram is the sum A + B . The longer of
182 Chapter 6 | Introduction to Vectors

→ →
the diagonals is the difference A − B . We use a ruler to measure the lengths of the diagonals, and a protractor

to measure the angles with the horizontal. For the resultant R , we obtain R = 5.8 cm and θ R ≈ 0° . For the

difference D , we obtain D = 16.2 cm and θ D = 49.3° , which are shown in Figure 6.14.

Figure 6.14 Using the parallelogram rule to solve (a) (finding the resultant, red) and (b) (finding
the difference, blue).


For (c), we can start with vector −3 B and draw the remaining vectors tail-to-head as shown in Figure 6.15.
In vector addition, the order in which we draw the vectors is unimportant, but drawing the vectors to scale is very

important. Next, we draw vector S from the origin of the first vector to the end of the last vector and place the
→ →
arrowhead at the end of S . We use a ruler to measure the length of S , and find that its magnitude is
S = 36.9 cm. We use a protractor and find that its direction angle is θ S = 52.9° . This solution is shown in Figure
6.15.

Figure 6.15 Using the tail-to-head method to solve (c)



(finding vector S , green).

This OpenStax book is available for free at https://legacy.cnx.org/content/col12169/1.46


Chapter 6 | Introduction to Vectors 183

6.3 → → →
Check Your Understanding Using the three displacement vectors A , B , and F in Figure

6.13, choose a convenient scale, and use a ruler and a protractor to find vector G given by the vector
→ → → →
equation G = A + 2 B − F .

Observe the addition of vectors in a plane by visiting this vector calculator (https://openstaxcollege.org/l/
21compveccalc) and this Phet simulation (https://openstaxcollege.org/l/21phetvecaddsim) .

6.2 | Coordinate Systems and Components of a Vector


Learning Objectives
By the end of this section, you will be able to:
• Describe vectors in two and three dimensions in terms of their components, using unit vectors
along the axes.
• Distinguish between the vector components of a vector and the scalar components of a vector.
• Explain how the magnitude of a vector is defined in terms of the components of a vector.
• Identify the direction angle of a vector in a plane.
• Explain the connection between polar coordinates and Cartesian coordinates in a plane.

Vectors are usually described in terms of their components in a coordinate system. Even in everyday life we naturally invoke
the concept of orthogonal projections in a rectangular coordinate system. For example, if you ask someone for directions to
a particular location, you will more likely be told to go 40 km east and 30 km north than 50 km in the direction 37° north
of east.
In a rectangular (Cartesian) xy-coordinate system in a plane, a point in a plane is described by a pair of coordinates (x, y).

In a similar fashion, a vector A in a plane is described by a pair of its vector coordinates. The x-coordinate of vector
→ →
A is called its x-component and the y-coordinate of vector A is called its y-component. The vector x-component is
→ →
a vector denoted by A x . The vector y-component is a vector denoted by A y . In the Cartesian system, the x and y
vector components of a vector are the orthogonal projections of this vector onto the x- and y-axes, respectively. In this way,
following the parallelogram rule for vector addition, each vector on a Cartesian plane can be expressed as the vector sum of
its vector components:
→ → → (6.10)
A = A x + A y.
→ →
As illustrated in Figure 6.16, vector A is the diagonal of the rectangle where the x-component A x is the side parallel
→ →
to the x-axis and the y-component A y is the side parallel to the y-axis. Vector component A x is orthogonal to vector

component A y .
184 Chapter 6 | Introduction to Vectors


Figure 6.16 Vector A in a plane in the Cartesian coordinate
system is the vector sum of its vector x- and y-components. The

x-vector component A x is the orthogonal projection of vector
→ →
A onto the x-axis. The y-vector component A y is the

orthogonal projection of vector A onto the y-axis. The numbers
A x and A y that multiply the unit vectors are the scalar components
of the vector.

^
It is customary to denote the positive direction on the x-axis by the unit vector i and the positive direction on the y-axis
^ ^ ^
by the unit vector j . Unit vectors of the axes, i and j , define two orthogonal directions in the plane. As shown in
Figure 6.16, the x- and y- components of a vector can now be written in terms of the unit vectors of the axes:
⎧→ ^ (6.11)
⎨→
A x = Ax i

⎩ A y = Ay j .
^

→ → →
The vectors A x and A y defined by Equation 6.11 are the vector components of vector A . The numbers A x and

A y that define the vector components in Equation 6.11 are the scalar components of vector A . Combining Equation
6.10 with Equation 6.11, we obtain the component form of a vector:

→ ^ ^ (6.12)
A = Ax i + Ay j .

If we know the coordinates b(x b, y b) of the origin point of a vector (where b stands for “beginning”) and the coordinates
e(x e, y e) of the end point of a vector (where e stands for “end”), we can obtain the scalar components of a vector simply
by subtracting the origin point coordinates from the end point coordinates:

⎧ A x = xe − xb (6.13)

⎩A y = y e − y b.

This OpenStax book is available for free at https://legacy.cnx.org/content/col12169/1.46


Chapter 6 | Introduction to Vectors 185

Example 6.3

Displacement of a Mouse Pointer


A mouse pointer on the display monitor of a computer at its initial position is at point (6.0 cm, 1.6 cm) with
respect to the lower left-side corner. If you move the pointer to an icon located at point (2.0 cm, 4.5 cm), what is
the displacement vector of the pointer?
Strategy
The origin of the xy-coordinate system is the lower left-side corner of the computer monitor. Therefore, the unit
^ ^
vector i on the x-axis points horizontally to the right and the unit vector j on the y-axis points vertically
upward. The origin of the displacement vector is located at point b(6.0, 1.6) and the end of the displacement vector
is located at point e(2.0, 4.5). Substitute the coordinates of these points into Equation 6.13 to find the scalar

components D x and D y of the displacement vector D . Finally, substitute the coordinates into Equation
6.12 to write the displacement vector in the vector component form.
Solution
We identify x b = 6.0 , x e = 2.0 , y b = 1.6 , and y e = 4.5 , where the physical unit is 1 cm. The scalar x- and
y-components of the displacement vector are
D x = x e − x b = (2.0 − 6.0)cm = −4.0 cm,
D y = y e − y b = (4.5 − 1.6)cm = + 2.9 cm.

The vector component form of the displacement vector is


→ ^ ^ ^ ^ ^ ^ (6.14)
D = D x i + D y j = (−4.0 cm) i + (2.9 cm) j = (−4.0 i + 2.9 j )cm.

This solution is shown in Figure 6.17.

Figure 6.17 The graph of the displacement vector. The vector points from
the origin point at b to the end point at e.

Significance
Notice that the physical unit—here, 1 cm—can be placed either with each component immediately before the unit
vector or globally for both components, as in Equation 6.14. Often, the latter way is more convenient because
it is simpler.
186 Chapter 6 | Introduction to Vectors

→ ^ ^
The vector x-component D x = −4.0 i = 4.0(− i ) of the displacement vector has the magnitude

| || ||

D x
^ ^
||
= − 4.0 i = 4.0 because the magnitude of the unit vector is i = 1 . Notice, too, that the direction

^
of the x-component is − i , which is antiparallel to the direction of the +x-axis; hence, the x-component vector
→ →
D x points to the left, as shown in Figure 6.17. The scalar x-component of vector D is D x = −4.0 .
→ ^
Similarly, the vector y-component D y = + 2.9 j of the displacement vector has magnitude

| ||||

D

^
y
^ ^
||
= 2.9 j = 2.9 because the magnitude of the unit vector is j = 1 . The direction of the y-component


is + j , which is parallel to the direction of the +y-axis. Therefore, the y-component vector D y points up, as
→ →
seen in Figure 6.17. The scalar y-component of vector D is D y = + 2.9 . The displacement vector D is
the resultant of its two vector components.
The vector component form of the displacement vector Equation 6.14 tells us that the mouse pointer has been
moved on the monitor 4.0 cm to the left and 2.9 cm upward from its initial position.

6.4 Check Your Understanding A blue fly lands on a sheet of graph paper at a point located 10.0 cm to the
right of its left edge and 8.0 cm above its bottom edge and walks slowly to a point located 5.0 cm from the left
edge and 5.0 cm from the bottom edge. Choose the rectangular coordinate system with the origin at the lower
left-side corner of the paper and find the displacement vector of the fly. Illustrate your solution by graphing.


When we know the scalar components A x and A y of a vector A , we can find its magnitude A and its direction angle
θ A . The direction angle—or direction, for short—is the angle the vector forms with the positive direction on the x-axis.
The angle θ A is measured in the counterclockwise direction from the +x-axis to the vector (Figure 6.18). Because the
lengths A, A x , and A y form a right triangle, they are related by the Pythagorean theorem:

A 2 = A 2x + A 2y ⇔ A = A 2x + A 2y. (6.15)

This equation works even if the scalar components of a vector are negative. The direction angle θ A of a vector is defined
via the tangent function of angle θ A in the triangle shown in Figure 6.18:

Ay ⎛A y⎞ (6.16)
tan θ A =
Ax
⇒ θ A = tan −1
⎝ A x ⎠.

This OpenStax book is available for free at https://legacy.cnx.org/content/col12169/1.46


Chapter 6 | Introduction to Vectors 187


Figure 6.18 For vector A , its magnitude A and its direction
angle θ A are related to the magnitudes of its scalar components
because A, A x , and A y form a right triangle.

When the vector lies either in the first quadrant or in the fourth quadrant, where component A x is positive (Figure 6.19),
the angle θ in Equation 6.16) is identical to the direction angle θ A . For vectors in the fourth quadrant, angle θ is
negative, which means that for these vectors, direction angle θ A is measured clockwise from the positive x-axis. Similarly,
for vectors in the second quadrant, angle θ is negative. When the vector lies in either the second or third quadrant, where
component A x is negative, the direction angle is θ A = θ + 180° (Figure 6.19).

Figure 6.19 Scalar components of a vector may be positive or negative.


Vectors in the first quadrant (I)have both scalar components positive and
vectors in the third quadrant have both scalar components negative. For
vectors in quadrants II and III, the direction angle of a vector is
θ A = θ + 180° .
188 Chapter 6 | Introduction to Vectors

Example 6.4

Magnitude and Direction of the Displacement Vector


You move a mouse pointer on the display monitor from its initial position at point (6.0 cm, 1.6 cm) to an icon
located at point (2.0 cm, 4.5 cm). What is the magnitude and direction of the displacement vector of the pointer?
Strategy

In Example 6.3, we found the displacement vector D of the mouse pointer (see Equation 6.14). We identify
its scalar components D x = −4.0 cm and D y = + 2.9 cm and substitute into Equation 6.15 and Equation
6.16 to find the magnitude D and direction θ D , respectively.

Solution

The magnitude of vector D is

D = D 2x + D 2y = (−4.0 cm) 2 + (2.9 cm) 2 = (4.0) 2 + (2.9) 2 cm = 4.9 cm.

The direction angle is


D y +2.9 cm
tan θ = = = −0.725 ⇒ θ = tan −1 (−0.725) = −35.9°.
D x −4.0 cm

Vector D lies in the second quadrant, so its direction angle is
θ D = θ + 180° = −35.9° + 180° = 144.1°.

6.5 Check Your Understanding If the displacement vector of a blue fly walking on a sheet of graph paper is
→ ^ ^
D = (−5.00 i − 3.00 j )cm , find its magnitude and direction.

In many applications, the magnitudes and directions of vector quantities are known and we need to find the resultant of
many vectors. For example, imagine 400 cars moving on the Golden Gate Bridge in San Francisco in a strong wind. Each
car gives the bridge a different push in various directions and we would like to know how big the resultant push can possibly
be. We have already gained some experience with the geometric construction of vector sums, so we know the task of finding
the resultant by drawing the vectors and measuring their lengths and angles may become intractable pretty quickly, leading
to huge errors. Worries like this do not appear when we use analytical methods. The very first step in an analytical approach
is to find vector components when the direction and magnitude of a vector are known.
Let us return to the right triangle in Figure 6.18. The quotient of the adjacent side A x to the hypotenuse A is the cosine
function of direction angle θ A , A x/A = cos θ A , and the quotient of the opposite side A y to the hypotenuse A is the sine
function of θ A , A y/A = sin θ A . When magnitude A and direction θ A are known, we can solve these relations for the
scalar components:

⎧A x = A cos θ A (6.17)
⎨ .
⎩ A y = A sin θ A

When calculating vector components with Equation 6.17, care must be taken with the angle. The direction angle θ A
of a vector is the angle measured counterclockwise from the positive direction on the x-axis to the vector. The clockwise
measurement gives a negative angle.

This OpenStax book is available for free at https://legacy.cnx.org/content/col12169/1.46


Chapter 6 | Introduction to Vectors 189

Example 6.5

Components of Displacement Vectors


A rescue party for a missing child follows a search dog named Trooper. Trooper wanders a lot and makes many
trial sniffs along many different paths. Trooper eventually finds the child and the story has a happy ending, but his
displacements on various legs seem to be truly convoluted. On one of the legs he walks 200.0 m southeast, then
he runs north some 300.0 m. On the third leg, he examines the scents carefully for 50.0 m in the direction 30°
west of north. On the fourth leg, Trooper goes directly south for 80.0 m, picks up a fresh scent and turns 23°
west of south for 150.0 m. Find the scalar components of Trooper’s displacement vectors and his displacement
vectors in vector component form for each leg.
Strategy
Let’s adopt a rectangular coordinate system with the positive x-axis in the direction of geographic east, with the
^
positive y-direction pointed to geographic north. Explicitly, the unit vector i of the x-axis points east and the
^
unit vector j of the y-axis points north. Trooper makes five legs, so there are five displacement vectors. We start
by identifying their magnitudes and direction angles, then we use Equation 6.17 to find the scalar components
of the displacements and Equation 6.12 for the displacement vectors.
Solution
On the first leg, the displacement magnitude is L 1 = 200.0 m and the direction is southeast. For direction
angle θ 1 we can take either 45° measured clockwise from the east direction or 45° + 270° measured
counterclockwise from the east direction. With the first choice, θ 1 = −45° . With the second choice,
θ 1 = + 315° . We can use either one of these two angles. The components are

L 1x = L 1 cos θ 1 = (200.0 m) cos 315° = 141.4 m,


L 1y = L 1 sin θ 1 = (200.0 m) sin 315° = −141.4 m.

The displacement vector of the first leg is


→ ^ ^ ^ ^
L 1 = L 1x i + L 1y j = (141.4 i − 141.4 j ) m.

On the second leg of Trooper’s wanderings, the magnitude of the displacement is L 2 = 300.0 m and the
direction is north. The direction angle is θ 2 = + 90° . We obtain the following results:

L 2x = L 2 cos θ 2 = (300.0 m) cos 90° = 0.0 ,


L 2y = L 2 sin θ 2 = (300.0 m) sin 90° = 300.0 m,
→ ^ ^ ^
L 2 = L 2x i + L 2y j = (300.0 m) j .

On the third leg, the displacement magnitude is L 3 = 50.0 m and the direction is 30° west of north. The
direction angle measured counterclockwise from the eastern direction is θ 3 = 30° + 90° = + 120° . This gives
the following answers:
L 3x = L 3 cos θ 3 = (50.0 m) cos 120° = −25.0 m,
L 3y = L 3 sin θ 3 = (50.0 m) sin 120° = + 43.3 m,
→ ^ ^ ^ ^
L 3 = L 3x i + L 3y j = (−25.0 i + 43.3 j )m.

On the fourth leg of the excursion, the displacement magnitude is L 4 = 80.0 m and the direction is south. The
direction angle can be taken as either θ 4 = −90° or θ 4 = + 270° . We obtain
190 Chapter 6 | Introduction to Vectors

L 4x = L 4 cos θ 4 = (80.0 m) cos (−90°) = 0 ,


L 4y = L 4 sin θ 4 = (80.0 m) sin (−90°) = −80.0 m,
→ ^ ^ ^
L 4 = L 4x i + L 4y j = (−80.0 m) j .

On the last leg, the magnitude is L 5 = 150.0 m and the angle is θ 5 = −23° + 270° = + 247° (23° west of
south), which gives
L 5x = L 5 cos θ 5 = (150.0 m) cos 247° = −58.6 m,
L 5y = L 5 sin θ 5 = (150.0 m) sin 247° = −138.1 m,
→ ^ ^ ^ ^
L 5 = L 5x i + L 5y j = (−58.6 i − 138.1 j )m.

6.6 Check Your Understanding If Trooper runs 20 m west before taking a rest, what is his displacement
vector?

Polar Coordinates
To describe locations of points or vectors in a plane, we need two orthogonal directions. In the Cartesian coordinate system
^ ^
these directions are given by unit vectors i and j along the x-axis and the y-axis, respectively. The Cartesian coordinate
system is very convenient to use in describing displacements and velocities of objects and the forces acting on them.
However, it becomes cumbersome when we need to describe the rotation of objects. When describing rotation, we usually
work in the polar coordinate system.
In the polar coordinate system, the location of point P in a plane is given by two polar coordinates (Figure 6.20). The first
polar coordinate is the radial coordinate r, which is the distance of point P from the origin. The second polar coordinate is
an angle φ that the radial vector makes with some chosen direction, usually the positive x-direction. In polar coordinates,
angles are measured in radians, or rads. The radial vector is attached at the origin and points away from the origin to point
^
P. This radial direction is described by a unit radial vector ^
r . The second unit vector t is a vector orthogonal to the
^
radial direction ^
r . The positive + t direction indicates how the angle φ changes in the counterclockwise direction. In
this way, a point P that has coordinates (x, y) in the rectangular system can be described equivalently in the polar coordinate
system by the two polar coordinates (r, φ) . Equation 6.17 is valid for any vector, so we can use it to express the x-
and y-coordinates of vector →
r . In this way, we obtain the connection between the polar coordinates and rectangular
coordinates of point P:

⎧x = r cos φ (6.18)
⎨ .
⎩ y = r sin φ

This OpenStax book is available for free at https://legacy.cnx.org/content/col12169/1.46


Chapter 6 | Introduction to Vectors 191

^
Figure 6.20 Using polar coordinates, the unit vector r defines
the positive direction along the radius r (radial direction) and,
^
orthogonal to it, the unit vector t defines the positive direction of
rotation by the angle φ .

Example 6.6

Polar Coordinates
A treasure hunter finds one silver coin at a location 20.0 m away from a dry well in the direction 20° north of
east and finds one gold coin at a location 10.0 m away from the well in the direction 20° north of west. What are
the polar and rectangular coordinates of these findings with respect to the well?
Strategy
The well marks the origin of the coordinate system and east is the +x-direction. We identify radial distances from
the locations to the origin, which are r S = 20.0 m (for the silver coin) and r G = 10.0 m (for the gold coin). To
find the angular coordinates, we convert 20° to radians: 20° = π20/180 = π/9 . We use Equation 6.18 to find
the x- and y-coordinates of the coins.
Solution
The angular coordinate of the silver coin is φ S = π/9 , whereas the angular coordinate of the gold coin is
φ G = π − π/9 = 8π/9 . Hence, the polar coordinates of the silver coin are (r S, φ S) = (20.0 m, π/9) and those
of the gold coin are (r G, φ G) = (10.0 m, 8π/9) . We substitute these coordinates into Equation 6.18 to obtain
rectangular coordinates. For the gold coin, the coordinates are
⎧x G = r G cos φ G = (10.0 m) cos 8π/9 = −9.4 m
⎨ ⇒ (x G, y G) = (−9.4 m, 3.4 m).
⎩y G = r G sin φ G = (10.0 m) sin 8π/9 = 3.4 m

For the silver coin, the coordinates are


⎧x S = r S cos φ S = (20.0 m) cos π/9 = 18.9 m
⎨ ⇒ (x S, y S) = (18.9 m, 6.8 m).
⎩y S = r S sin φ S = (20.0 m) sin π/9 = 6.8 m

Vectors in Three Dimensions


To specify the location of a point in space, we need three coordinates (x, y, z), where coordinates x and y specify locations in
a plane, and coordinate z gives a vertical positions above or below the plane. Three-dimensional space has three orthogonal
192 Chapter 6 | Introduction to Vectors

directions, so we need not two but three unit vectors to define a three-dimensional coordinate system. In the Cartesian
^ ^
coordinate system, the first two unit vectors are the unit vector of the x-axis i and the unit vector of the y-axis j . The
^
third unit vector k is the direction of the z-axis (Figure 6.21). The order in which the axes are labeled, which is the
order in which the three unit vectors appear, is important because it defines the orientation of the coordinate system. The
^ ^ ^
order x-y-z, which is equivalent to the order i - j - k , defines the standard right-handed coordinate system (positive
orientation).

Figure 6.21 Three unit vectors define a Cartesian system in


three-dimensional space. The order in which these unit vectors
appear defines the orientation of the coordinate system. The
order shown here defines the right-handed orientation.

→ → ^
In three-dimensional space, vector A has three vector components: the x-component A x = A x i , which is the part
→ → ^ →
of vector A along the x-axis; the y-component A y = A y j , which is the part of A along the y-axis; and the
→ ^
z-component A z = A z k , which is the part of the vector along the z-axis. A vector in three-dimensional space is the
vector sum of its three vector components (Figure 6.22):

→ ^ ^ ^ (6.19)
A = A x i + Ay j + Az k .

If we know the coordinates of its origin b(x b, y b, z b) and of its end e(x e, y e, z e) , its scalar components are obtained by
taking their differences: A x and A y are given by Equation 6.13 and the z-component is given by

A z = z e − z b. (6.20)

Magnitude A is obtained by generalizing Equation 6.15 to three dimensions:

This OpenStax book is available for free at https://legacy.cnx.org/content/col12169/1.46


Chapter 6 | Introduction to Vectors 193

A = A 2x + A 2y + A 2z . (6.21)

This expression for the vector magnitude comes from applying the Pythagorean theorem twice. As seen in Figure 6.22,
the diagonal in the xy-plane has length A 2x + A 2y and its square adds to the square A 2z to give A 2 . Note that when the
z-component is zero, the vector lies entirely in the xy-plane and its description is reduced to two dimensions.

Figure 6.22 A vector in three-dimensional space is the vector


sum of its three vector components.

Example 6.7

Takeoff of a Drone
During a takeoff of IAI Heron (Figure 6.23), its position with respect to a control tower is 100 m above the
ground, 300 m to the east, and 200 m to the north. One minute later, its position is 250 m above the ground, 1200
m to the east, and 2100 m to the north. What is the drone’s displacement vector with respect to the control tower?
What is the magnitude of its displacement vector?

Figure 6.23 The drone IAI Heron in flight. (credit: SSgt


Reynaldo Ramon, USAF)

Strategy
We take the origin of the Cartesian coordinate system as the control tower. The direction of the +x-axis is given
194 Chapter 6 | Introduction to Vectors

^ ^
by unit vector i to the east, the direction of the +y-axis is given by unit vector j to the north, and the direction
^
of the +z-axis is given by unit vector k , which points up from the ground. The drone’s first position is the origin
(or, equivalently, the beginning) of the displacement vector and its second position is the end of the displacement
vector.
Solution
We identify b(300.0 m, 200.0 m, 100.0 m) and e(480.0 m, 370.0 m, 250.0m), and use Equation 6.13 and
Equation 6.20 to find the scalar components of the drone’s displacement vector:
⎧D x = x e − x b = 1200.0 m − 300.0 m = 900.0 m,
⎨D y = y e − y b = 2100.0 m − 200.0 m = 1900.0 m,
⎩D z = z e − z b = 250.0 m − 100.0 m = 150.0 m.
We substitute these components into Equation 6.19 to find the displacement vector:
→ ^ ^ ^ ^ ^ ^ ^ ^ ^
D = D x i + D y j + D z k = 900.0 m i + 1900.0 m j + 150.0 m k = (0.90 i + 1.90 j + 0.15 k ) km.

We substitute into Equation 6.21 to find the magnitude of the displacement”

D = D 2x + D 2y + D 2z = (0.90 km) 2 + (1.90 km) 2 + (0.15 km) 2 = 4.44 km.

6.7 Check Your Understanding If the average velocity vector of the drone in the displacement in Example
^ ^ ^
6.7 is →
u = (15.0 i + 31.7 j + 2.5 k )m/s , what is the magnitude of the drone’s velocity vector?

6.3 | Algebra of Vectors


Learning Objectives
By the end of this section, you will be able to:
• Apply analytical methods of vector algebra to find resultant vectors and to solve vector
equations for unknown vectors.
• Interpret physical situations in terms of vector expressions.

Vectors can be added together and multiplied by scalars. Vector addition is associative (Equation 6.8) and commutative
(Equation 6.7), and vector multiplication by a sum of scalars is distributive (Equation 6.9). Also, scalar multiplication
by a sum of vectors is distributive:

→ → → → (6.22)
α( A + B ) = α A + α B .

→ ^ ^ ^
In this equation, α is any number (a scalar). For example, a vector antiparallel to vector A = A x i + A y j + A z k can

be expressed simply by multiplying A by the scalar α = −1 :

→ ^ ^ ^ (6.23)
− A = −A x i − A y j − A z k .

This OpenStax book is available for free at https://legacy.cnx.org/content/col12169/1.46


Chapter 6 | Introduction to Vectors 195

Example 6.8

Direction of Motion
^ ^ ^
In a Cartesian coordinate system where i denotes geographic east, j denotes geographic north, and k
denotes altitude above sea level, a military convoy advances its position through unknown territory with velocity
→ ^ ^ ^
v = (4.0 i + 3.0 j + 0.1 k )km/h . If the convoy had to retreat, in what geographic direction would it be
moving?
Solution
^
The velocity vector has the third component → v z = ( + 0.1km/h) k , which says the convoy is climbing at a rate
of 100 m/h through mountainous terrain. At the same time, its velocity is 4.0 km/h to the east and 3.0 km/h to
the north, so it moves on the ground in direction tan −1(3 /4) ≈ 37° north of east. If the convoy had to retreat,
its new velocity vector →
u would have to be antiparallel to →
v and be in the form →
u = −α →
v , where
^ ^ ^
α is a positive number. Thus, the velocity of the retreat would be →
u = α(−4.0 i − 3.0 j − 0.1 k )km/h . The
negative sign of the third component indicates the convoy would be descending. The direction angle of the retreat
velocity is tan −1(−3α/ − 4α) ≈ 37° south of west. Therefore, the convoy would be moving on the ground in
direction 37° south of west while descending on its way back.


The generalization of the number zero to vector algebra is called the null vector, denoted by 0 . All components of the
→ ^ ^ ^
null vector are zero, 0 = 0 i + 0 j + 0 k , so the null vector has no length and no direction.
→ →
Two vectors A and B are equal vectors if and only if their difference is the null vector:
→ → → ^ ^ ^ ^ ^ ^ ^ ^ ^
0 = A − B = (A x i + A y j + A z k ) − (B x i + B y j + B z k ) = (A x − B x) i + (A y − B y) j + (A z − B z) k .

This vector equation means we must have simultaneously A x − B x = 0 , A y − B y = 0 , and A z − B z = 0 . Hence, we can
→ → → →
write A = B if and only if the corresponding components of vectors A and B are equal:

⎧A x = B x (6.24)
⎨A y = B y.
→ →
A = B ⇔
⎩A z = B z

Two vectors are equal when their corresponding scalar components are equal.
Resolving vectors into their scalar components (i.e., finding their scalar components) and expressing them analytically in
vector component form (given by Equation 6.19) allows us to use vector algebra to find sums or differences of many
→ →
vectors analytically (i.e., without using graphical methods). For example, to find the resultant of two vectors A and B
, we simply add them component by component, as follows:
→ → → ^ ^ ^ ^ ^ ^ ^ ^ ^
R = A + B = (A x i + A y j + A z k ) + (B x i + B y j + B z k ) = (A x + B x) i + (A y + B y) j + (A z + B z) k .

→ ^ ^ ^
In this way, using Equation 6.24, scalar components of the resultant vector R = R x i + R y j + R z k are the sums of
→ →
corresponding scalar components of vectors A and B :
196 Chapter 6 | Introduction to Vectors

⎧R x = A x + B x,
⎨R y = A y + B y,
⎩ R z = A z + B z.
Analytical methods can be used to find components of a resultant of many vectors. For example, if we are to sum up
→ → → → → ^ ^ ^
N vectors F 1, F 2, F 3, … , F N , where each vector is F k = F kx i + F ky j + F kz k , the resultant vector

F R is

⎛ ^ ^⎞
N N
→ → → → → → ^
F R= F 1+ F 2+ F 3+…+ F N = ∑ F k= ∑ ⎝F kx i + F ky j + F kz k

k=1 k=1
⎛ N ⎞^ ⎛ N ⎞^ ⎛ N ⎞^
= ⎜ ∑ F kx⎟ i + ⎜ ∑ F ky⎟ j + ⎜ ∑ F kz⎟ k .
⎝k = 1 ⎠ ⎝k = 1 ⎠ ⎝k = 1 ⎠

Therefore, scalar components of the resultant vector are

⎧ N (6.25)
⎪F Rx = ∑ F kx = F 1x + F 2x + … + F Nx
⎪ k=1
⎪ N
⎨F Ry = ∑ F ky = F 1y + F 2y + … + F Ny
⎪ k=1
⎪ N
⎪F = ∑ F = F + F + … + F .
⎩ Rz k = 1 kz 1z 2z Nz

Having found the scalar components, we can write the resultant in vector component form:
→ ^ ^ ^
F R = F Rx i + F Ry j + F Rz k .

Analytical methods for finding the resultant and, in general, for solving vector equations are very important in physics
because many physical quantities are vectors. For example, we use this method in kinematics to find resultant displacement
vectors and resultant velocity vectors, in mechanics to find resultant force vectors and the resultants of many derived vector
quantities, and in electricity and magnetism to find resultant electric or magnetic vector fields.

Example 6.9

Analytical Computation of a Resultant


→ → →
Three displacement vectors A , B , and C in a plane (Figure 6.13) are specified by their magnitudes
A = 10.0, B = 7.0, and C = 8.0, respectively, and by their respective direction angles with the horizontal
direction α = 35°, β = −110° , and γ = 30° . The physical units of the magnitudes are centimeters. Resolve
→ → → →
the vectors to their scalar components and find the following vector sums: (a) R = A + B + C , (b)
→ → → → → → →
D = A − B , and (c) S = A − 3 B + C .
Strategy
First, we use Equation 6.17 to find the scalar components of each vector and then we express each vector in its
vector component form given by Equation 6.12. Then, we use analytical methods of vector algebra to find the
resultants.
Solution
We resolve the given vectors to their scalar components:

This OpenStax book is available for free at https://legacy.cnx.org/content/col12169/1.46


Chapter 6 | Introduction to Vectors 197

⎧A x = A cos α = (10.0 cm) cos 35° = 8.19 cm



⎩A y = A sin α = (10.0 cm) sin 35° = 5.73 cm
⎧B x = B cos β = (7.0 cm) cos (−110°) = −2.39 cm
⎨ .
⎩B y = B sin β = (7.0 cm) sin (−110°) = −6.58 cm
⎧C x = C cos γ = (8.0 cm) cos 30° = 6.93 cm

⎩C y = C sin γ = (8.0 cm) sin 30° = 4.00 cm

For (a) we may substitute directly into Equation 6.24 to find the scalar components of the resultant:
⎧R x = A x + B x + C x = 8.19 cm − 2.39 cm + 6.93 cm = 12.73 cm
⎨ .
⎩R y = A y + B y + C y = 5.73 cm − 6.58 cm + 4.00 cm = 3.15 cm

→ ^ ^ ^ ^
Therefore, the resultant vector is R = R x i + R y j = (12.7 i + 3.1 j )cm .

For (b), we may want to write the vector difference as


→ → → ^ ^ ^ ^ ^ ^
D = A − B = (A x i + A y j ) − (B x i + B y j ) = (A x − B x) i + (A y − B y) j .

Then, the scalar components of the vector difference are


⎧D x = A x − B x = 8.19 cm − (−2.39 cm) = 10.58 cm
⎨ .
⎩D y = A y − B y = 5.73 cm − (−6.58 cm) = 12.31 cm

→ ^ ^ ^ ^
Hence, the difference vector is D = D x i + D y j = (10.6 i + 12.3 j )cm .


For (c), we can write vector S in the following explicit form:
→ → → → ^ ^ ^ ^ ^ ^
S = A − 3 B + C = (A x i + A y j ) − 3(B x i + B y j ) + (C x i + C y j )
^ ^
= (A x − 3B x + C x) i + (A y − 3B y + C y) j .

Then, the scalar components of S are
⎧S x = A x − 3B x + C x = 8.19 cm − 3(−2.39 cm) + 6.93 cm = 22.29 cm
⎨ .
⎩S y = A y − 3B y + C y = 5.73 cm − 3(−6.58 cm) + 4.00 cm = 29.47 cm

→ ^ ^ ^ ^
The vector is S = S x i + S y j = (22.3 i + 29.5 j )cm .

Significance
Having found the vector components, we can illustrate the vectors by graphing or we can compute magnitudes
and direction angles, as shown in Figure 6.24. Results for the magnitudes in (b) and (c) can be compared with
results for the same problems obtained with the graphical method, shown in Figure 6.14 and Figure 6.15.
Notice that the analytical method produces exact results and its accuracy is not limited by the resolution of a ruler
or a protractor, as it was with the graphical method used in Example 6.2 for finding this same resultant.
198 Chapter 6 | Introduction to Vectors

Figure 6.24 Graphical illustration of the solutions obtained analytically in Example 6.9.

6.8 → → →
Check Your Understanding Three displacement vectors A , B , and F (Figure 6.13) are
specified by their magnitudes A = 10.00, B = 7.00, and F = 20.00, respectively, and by their respective direction
angles with the horizontal direction α = 35° , β = −110° , and φ = 110° . The physical units of the
→ → → →
magnitudes are centimeters. Use the analytical method to find vector G = A + 2 B − F . Verify that
G = 28.15 cm and that θ G = −68.65° .

Example 6.10

The Tug-of-War Game


Four dogs named Ang, Bing, Chang, and Dong play a tug-of-war game with a toy (Figure 6.25). Ang pulls on
the toy in direction α = 55° south of east, Bing pulls in direction β = 60° east of north, and Chang pulls in
direction γ = 55° west of north. Ang pulls strongly with 160.0 units of force (N), which we abbreviate as A =
160.0 N. Bing pulls even stronger than Ang with a force of magnitude B = 200.0 N, and Chang pulls with a force
of magnitude C = 140.0 N. When Dong pulls on the toy in such a way that his force balances out the resultant of
the other three forces, the toy does not move in any direction. With how big a force and in what direction must
Dong pull on the toy for this to happen?

Figure 6.25 Four dogs play a tug-of-war game with a toy.

This OpenStax book is available for free at https://legacy.cnx.org/content/col12169/1.46


Chapter 6 | Introduction to Vectors 199

Strategy
We assume that east is the direction of the positive x-axis and north is the direction of the positive y-axis. As
→ →
in Example 6.9, we have to resolve the three given forces— A (the pull from Ang), B (the pull from

Bing), and C (the pull from Chang)—into their scalar components and then find the scalar components of the
→ → → → →
resultant vector R = A + B + C . When the pulling force D from Dong balances out this resultant,
→ → → → → → →
the sum of D and R must give the null vector D + R = 0 . This means that D = − R , so the

pull from Dong must be antiparallel to R .
Solution
The direction angles are θ A = −α = −55° , θ B = 90° − β = 30° , and θ C = 90° + γ = 145° , and substituting
them into Equation 6.17 gives the scalar components of the three given forces:
⎧A x = A cos θ A = (160.0 N) cos (−55°) = + 91.8 N

⎩A y = A sin θ A = (160.0 N) sin (−55°) = −131.1 N
⎧B x = B cos θ B = (200.0 N) cos 30° = + 173.2 N
⎨ .
⎩B y = B sin θ B = (200.0 N) sin 30° = + 100.0 N
⎧C x = C cos θ C = (140.0 N) cos 145° = −114.7 N

⎩C y = C sin θ C = (140.0 N) sin 145° = + 80.3 N

→ → → →
Now we compute scalar components of the resultant vector R = A + B + C :
⎧R x = A x + B x + C x = + 91.8 N + 173.2 N − 114.7 N = + 150.3 N
⎨ .
⎩R y = A y + B y + C y = −131.1 N + 100.0 N + 80.3 N = + 49.2 N

The antiparallel vector to the resultant R is
→ → ^ ^ ^ ^
D = − R = −R x i − R y j = (−150.3 i − 49.2 j ) N.

The magnitude of Dong’s pulling force is

D = D 2x + D 2y = (−150.3) 2 + (−49.2) 2 N = 158.1 N.

The direction of Dong’s pulling force is


⎛D y ⎞ −1 ⎛ −49.2 N ⎞ −1 ⎛ 49.2 ⎞
θ = tan −1
⎝D x ⎠ = tan ⎝−150.3 N ⎠ = tan ⎝150.3 ⎠ = 18.1°.
Dong pulls in the direction 18.1° south of west because both components are negative, which means the pull
vector lies in the third quadrant (Figure 6.19).

6.9 Check Your Understanding Suppose that Bing in Example 6.10 leaves the game to attend to more
important matters, but Ang, Chang, and Dong continue playing. Ang and Chang’s pull on the toy does not
change, but Dong runs around and bites on the toy in a different place. With how big a force and in what
direction must Dong pull on the toy now to balance out the combined pulls from Chang and Ang? Illustrate this
situation by drawing a vector diagram indicating all forces involved.
200 Chapter 6 | Introduction to Vectors

Example 6.11

Vector Algebra
→ → → → ^
Find the magnitude of the vector C that satisfies the equation 2 A − 6 B + 3 C = 2 j , where
→ ^ ^ → ^ ^
A = i − 2 k and B = − j + k /2 .

Strategy
→ → →
We first solve the given equation for the unknown vector C . Then we substitute A and B ; group the
^ ^ ^
terms along each of the three directions i , j , and k ; and identify the scalar components C x , C y , and C z .
Finally, we substitute into Equation 6.21 to find magnitude C.
Solution
→ → → ^
2 A −6 B +3 C = 2j
→ ^ → →
3 C = 2j −2 A +6 B
→ ^ → →
C = 2 j − 2 A +2 B
3 3
^ ^ ^ ⎛ ^ ^⎞ ^ ^ ^ ^ ^
= 2 j − 2 ( i − 2 k ) + 2⎜− j + k ⎟ = 2 j − 2 i + 4 k − 2 j + k
3 3 ⎝ 2 ⎠ 3 3 3
^ ⎛ ⎞^ ⎛ ⎞^
= − 2 i + ⎝2 − 2⎠ j + ⎝4 + 1⎠ k
3 3 3
^ ^ ^
= −2 i − 4 j + 7 k.
3 3 3
The components are C x = −2 /3 , C y = −4/3 , and C z = 7 /3 , and substituting into Equation 6.21 gives

C = C x2 + C y2 + C z2 = (−2/3) 2 + (−4 /3) 2 + (7 /3) 2 = 23/3.

Example 6.12

Displacement of a Skier
Starting at a ski lodge, a cross-country skier goes 5.0 km north, then 3.0 km west, and finally 4.0 km southwest
before taking a rest. Find his total displacement vector relative to the lodge when he is at the rest point. How far
and in what direction must he ski from the rest point to return directly to the lodge?
Strategy
^
We assume a rectangular coordinate system with the origin at the ski lodge and with the unit vector i pointing
^ → → →
east and the unit vector j pointing north. There are three displacements: D 1 , D 2 , and D 3 . We
identify their magnitudes as D 1 = 5.0 km , D 2 = 3.0 km , and D 3 = 4.0 km . We identify their directions are
the angles θ 1 = 90° , θ 2 = 180° , and θ 3 = 180° + 45° = 225° . We resolve each displacement vector to its
scalar components and substitute the components into Equation 6.24 to obtain the scalar components of the

resultant displacement D from the lodge to the rest point. On the way back from the rest point to the lodge, the
→ → →
displacement is B = − D . Finally, we find the magnitude and direction of B .
Solution
Scalar components of the displacement vectors are

This OpenStax book is available for free at https://legacy.cnx.org/content/col12169/1.46


Chapter 6 | Introduction to Vectors 201

⎧D 1x = D 1 cos θ 1 = (5.0 km) cos 90° = 0



⎩D 1y = D 1 sin θ 1 = (5.0 km) sin 90° = 5.0 km
⎧D 2x = D 2 cos θ 2 = (3.0 km) cos 180° = −3.0 km
⎨ .
⎩D 2y = D 2 sin θ 2 = (3.0 km) sin 180° = 0
⎧D 3x = D 3 cos θ 3 = (4.0 km) cos 225° = −2.8 km

⎩D 3y = D 3 sin θ 3 = (4.0 km) sin 225° = −2.8 km

Scalar components of the net displacement vector are


⎧D x = D 1x + D 2x + D 3x = (0 − 3.0 − 2.8)km = −5.8 km
⎨ .
⎩D y = D 1y + D 2y + D 3y = (5.0 + 0 − 2.8)km = + 2.2 km

→ ^ ^ ^ ^
Hence, the skier’s net displacement vector is D = D x i + D y j = (−5.8 i + 2.2 j )km . On the way back
→ → ^ ^ ^ ^
to the lodge, his displacement is B = − D = −(−5.8 i + 2.2 j )km = (5.8 i − 2.2 j )km . Its magnitude is

B = B 2x + B 2y = (5.8) 2 + (−2.2) 2 km = 6.2 km and its direction angle is θ = tan −1(−2.2/5.8) = −20.8° .
Therefore, to return to the lodge, he must go 6.2 km in a direction about 21° south of east.
Significance
Notice that no figure is needed to solve this problem by the analytical method. Figures are required when using a
graphical method; however, we can check if our solution makes sense by sketching it, which is a useful final step
in solving any vector problem.

Example 6.13

Displacement of a Jogger
A jogger runs up a flight of 200 identical steps to the top of a hill and then runs along the top of the hill 50.0 m
before he stops at a drinking fountain (Figure 6.26). His displacement vector from point A at the bottom of the
→ ^ ^
steps to point B at the fountain is D AB = (−90.0 i + 30.0 j )m . What is the height and width of each step in
the flight? What is the actual distance the jogger covers? If he makes a loop and returns to point A, what is his net
displacement vector?

Figure 6.26 A jogger runs up a flight of steps.


202 Chapter 6 | Introduction to Vectors

Strategy
→ →
The displacement vector D AB is the vector sum of the jogger’s displacement vector D AT along the stairs

(from point A at the bottom of the stairs to point T at the top of the stairs) and his displacement vector D TB on
the top of the hill (from point T at the top of the stairs to the fountain at point B). We must find the horizontal and
→ →
the vertical components of D TB . If each step has width w and height h, the horizontal component of D TB
must have a length of 200w and the vertical component must have a length of 200h. The actual distance the jogger
covers is the sum of the distance he runs up the stairs and the distance of 50.0 m that he runs along the top of the
hill.
Solution
In the coordinate system indicated in Figure 6.26, the jogger’s displacement vector on the top of the hill is
→ ^
D TB = (−50.0 m) i . His net displacement vector is
→ → →
D AB = D AT + D TB.

Therefore, his displacement vector D TB along the stairs is

→ → → ^ ^ ^ ^ ^
D AT = D AB − D = (−90.0 i + 30.0 j )m − (−50.0 m) i = [(−90.0 + 50.0) i + 30.0 j )]m
TB
^ ^
= (−40.0 i + 30.0 j )m.

Its scalar components are D AT x = −40.0 m and D ATy = 30.0 m . Therefore, we must have

200w = | − 40.0|m and 200h = 30.0 m.


Hence, the step width is w = 40.0 m/200 = 0.2 m = 20 cm, and the step height is w = 30.0 m/200 = 0.15 m = 15
cm. The distance that the jogger covers along the stairs is

D AT = D 2AT x + D 2ATy = (−40.0) 2 + (30.0) 2 m = 50.0 m.

Thus, the actual distance he runs is D AT + D TB = 50.0 m + 50.0 m = 100.0 m . When he makes a loop and
comes back from the fountain to his initial position at point A, the total distance he covers is twice this distance,
or 200.0 m. However, his net displacement vector is zero, because when his final position is the same as his initial
position, the scalar components of his net displacement vector are zero (Equation 6.13).

In many physical situations, we often need to know the direction of a vector. For example, we may want to know the
direction of a magnetic field vector at some point or the direction of motion of an object. We have already said direction is
given by a unit vector, which is a dimensionless entity—that is, it has no physical units associated with it. When the vector
in question lies along one of the axes in a Cartesian system of coordinates, the answer is simple, because then its unit vector
of direction is either parallel or antiparallel to the direction of the unit vector of an axis. For example, the direction of vector
→ ^ ^ ^ ^ →
d = −5 m i is unit vector d = − i . The general rule of finding the unit vector V of direction for any vector V is
to divide it by its magnitude V:

^ → (6.26)
V= V .
V

We see from this expression that the unit vector of direction is indeed dimensionless because the numerator and the
denominator in Equation 6.26 have the same physical unit. In this way, Equation 6.26 allows us to express the unit
vector of direction in terms of unit vectors of the axes. The following example illustrates this principle.

This OpenStax book is available for free at https://legacy.cnx.org/content/col12169/1.46


Chapter 6 | Introduction to Vectors 203

Example 6.14

The Unit Vector of Direction


^ ^ ^
If the velocity vector of the military convoy in Example 6.8 is →
v = (4.000 i + 3.000 j + 0.100 k )km/h ,
what is the unit vector of its direction of motion?
Strategy

The unit vector of the convoy’s direction of motion is the unit vector ^
v that is parallel to the velocity vector. The
unit vector is obtained by dividing a vector by its magnitude, in accordance with Equation 6.26.
Solution
The magnitude of the vector →
v is

v = v 2x + v 2y + v 2z = 4.000 2 + 3.000 2 + 0.100 2km/h = 5.001km/h.

To obtain the unit vector ^


v , divide →
v by its magnitude:

→ ^ ^ ^
^ (4.000 i + 3.000 j + 0.100 k )km/h
v = vv =
5.001km/h
^ ^ ^
(4.000 i + 3.000 j + 0.100 k )
=
5.001
4.000 ^ 3.000 ^ 0.100 ^
= i + j + k
5.001 5.001 5.001
^ ^ ^
= (79.98 i + 59.99 j + 2.00 k ) × 10 −2.

Significance
Note that when using the analytical method with a calculator, it is advisable to carry out your calculations to at
least three decimal places and then round off the final answer to the required number of significant figures, which
is the way we performed calculations in this example. If you round off your partial answer too early, you risk your
final answer having a huge numerical error, and it may be far off from the exact answer or from a value measured
in an experiment.

6.10 Check Your Understanding Verify that vector ^ v obtained in Example 6.14 is indeed a unit vector
by computing its magnitude. If the convoy in Example 6.8 was moving across a desert flatland—that is, if the
third component of its velocity was zero—what is the unit vector of its direction of motion? Which geographic
direction does it represent?
204 Chapter 6 | Introduction to Vectors

CHAPTER 6 REVIEW
KEY TERMS
antiparallel vectors two vectors with directions that differ by 180°

associative terms can be grouped in any fashion


commutative operations can be performed in any order
component form of a vector a vector written as the vector sum of its components in terms of unit vectors
difference of two vectors vector sum of the first vector with the vector antiparallel to the second
direction angle in a plane, an angle between the positive direction of the x-axis and the vector, measured
counterclockwise from the axis to the vector
displacement change in position
distributive multiplication can be distributed over terms in summation
equal vectors two vectors are equal if and only if all their corresponding components are equal; alternately, two parallel
vectors of equal magnitudes
magnitude length of a vector
null vector a vector with all its components equal to zero
orthogonal vectors two vectors with directions that differ by exactly 90° , synonymous with perpendicular vectors

parallel vectors two vectors with exactly the same direction angles
parallelogram rule geometric construction of the vector sum in a plane
polar coordinate system an orthogonal coordinate system where location in a plane is given by polar coordinates
polar coordinates a radial coordinate and an angle
radial coordinate distance to the origin in a polar coordinate system
resultant vector vector sum of two (or more) vectors
scalar a number, synonymous with a scalar quantity in physics
scalar component a number that multiplies a unit vector in a vector component of a vector
scalar equation equation in which the left-hand and right-hand sides are numbers
scalar quantity quantity that can be specified completely by a single number with an appropriate physical unit
tail-to-head geometric construction geometric construction for drawing the resultant vector of many vectors
unit vector vector of a unit magnitude that specifies direction; has no physical unit
unit vectors of the axes unit vectors that define orthogonal directions in a plane or in space
vector mathematical object with magnitude and direction
vector components orthogonal components of a vector; a vector is the vector sum of its vector components.
vector equation equation in which the left-hand and right-hand sides are vectors
vector quantity physical quantity described by a mathematical vector—that is, by specifying both its magnitude and its
direction; synonymous with a vector in physics
vector sum resultant of the combination of two (or more) vectors

KEY EQUATIONS
→ →
Multiplication by a scalar (vector equation) B =α A

This OpenStax book is available for free at https://legacy.cnx.org/content/col12169/1.46


Chapter 6 | Introduction to Vectors 205

Multiplication by a scalar (scalar equation for magnitudes) B = |α| A


→ → →
Resultant of two vectors D AD = D AC + D CD

→ → → →
Commutative law A + B = B + A
→ → → → → →
Associative law ( A + B )+ C = A +( B + C )
→ → →
Distributive law α 1 A + α 2 A = (α 1 + α 2) A

→ ^ ^
The component form of a vector in two dimensions A = Ax i + Ay j

⎧A x = xe − xb
Scalar components of a vector in two dimensions ⎨
⎩A y = ye − yb

Magnitude of a vector in a plane A = A 2x + A 2y

⎛A y⎞
The direction angle of a vector in a plane θ A = tan −1
⎝A x ⎠
⎧A x = A cos θ A
Scalar components of a vector in a plane ⎨
⎩A y = A sin θ A

⎧x = r cos φ
Polar coordinates in a plane ⎨
⎩y = r sin φ

→ ^ ^ ^
The component form of a vector in three dimensions A = A x i + Ay j + Az k

The scalar z-component of a vector in three dimensions Az = ze − zb

Magnitude of a vector in three dimensions A = A 2x + A 2y + A 2z

→ → → →
Distributive property α( A + B ) = α A + α B

→ → ^ ^ ^
Antiparallel vector to A − A = −A x i − A y j − A z k

⎧A x = B x
⎨A y = B y
→ →
A = B ⇔
⎩A z = B z
Equal vectors

⎧ N
⎪F Rx = ∑ F kx = F 1x + F 2x + … + F Nx
⎪ k=1
⎪ N
⎨F Ry = ∑ F ky = F 1y + F 2y + … + F Ny

Components of the resultant of N vectors
k=1
⎪ N
⎪F = ∑ F = F + F + … + F
⎩ Rz k = 1 kz 1z 2z Nz

^ →
General unit vector V= V
V
206 Chapter 6 | Introduction to Vectors

SUMMARY
6.1 Scalars and Vectors
• A vector quantity is any quantity that has magnitude and direction, such as displacement or velocity. Vector
quantities are represented by mathematical objects called vectors.
• Geometrically, vectors are represented by arrows, with the end marked by an arrowhead. The length of the vector is
its magnitude, which is a positive scalar. On a plane, the direction of a vector is given by the angle the vector makes
with a reference direction, often an angle with the horizontal. The direction angle of a vector is a scalar.
• Two vectors are equal if and only if they have the same magnitudes and directions. Parallel vectors have the same
direction angles but may have different magnitudes. Antiparallel vectors have direction angles that differ by 180° .
Orthogonal vectors have direction angles that differ by 90° .
• When a vector is multiplied by a scalar, the result is another vector of a different length than the length of the original
vector. Multiplication by a positive scalar does not change the original direction; only the magnitude is affected.
Multiplication by a negative scalar reverses the original direction. The resulting vector is antiparallel to the original
vector. Multiplication by a scalar is distributive. Vectors can be divided by nonzero scalars but cannot be divided by
vectors.
• Two or more vectors can be added to form another vector. The vector sum is called the resultant vector. We can add
vectors to vectors or scalars to scalars, but we cannot add scalars to vectors. Vector addition is commutative and
associative.
• To construct a resultant vector of two vectors in a plane geometrically, we use the parallelogram rule. To construct
a resultant vector of many vectors in a plane geometrically, we use the tail-to-head method.

6.2 Coordinate Systems and Components of a Vector


• Vectors are described in terms of their components in a coordinate system. In two dimensions (in a plane), vectors
have two components. In three dimensions (in space), vectors have three components.
• A vector component of a vector is its part in an axis direction. The vector component is the product of the unit vector
of an axis with its scalar component along this axis. A vector is the resultant of its vector components.
• Scalar components of a vector are differences of coordinates, where coordinates of the origin are subtracted from
end point coordinates of a vector. In a rectangular system, the magnitude of a vector is the square root of the sum of
the squares of its components.
• In a plane, the direction of a vector is given by an angle the vector has with the positive x-axis. This direction angle
is measured counterclockwise. The scalar x-component of a vector can be expressed as the product of its magnitude
with the cosine of its direction angle, and the scalar y-component can be expressed as the product of its magnitude
with the sine of its direction angle.
• In a plane, there are two equivalent coordinate systems. The Cartesian coordinate system is defined by unit vectors
^ ^
i and j along the x-axis and the y-axis, respectively. The polar coordinate system is defined by the radial unit
^
vector ^r , which gives the direction from the origin, and a unit vector t , which is perpendicular (orthogonal) to
the radial direction.

6.3 Algebra of Vectors


• Analytical methods of vector algebra allow us to find resultants of sums or differences of vectors without having to
draw them. Analytical methods of vector addition are exact, contrary to graphical methods, which are approximate.
• Analytical methods of vector algebra are used routinely in mechanics, electricity, and magnetism. They are
important mathematical tools of physics.

This OpenStax book is available for free at https://legacy.cnx.org/content/col12169/1.46


Chapter 6 | Introduction to Vectors 207

CONCEPTUAL QUESTIONS
11. Can a magnitude of a vector be negative?
6.1 Scalars and Vectors
1. A weather forecast states the temperature is predicted to 12. Can the magnitude of a particle’s displacement be
be −5 °C the following day. Is this temperature a vector or greater that the distance traveled?
a scalar quantity? Explain.
13. If two vectors are equal, what can you say about their
2. Which of the following is a vector: a person’s height, components? What can you say about their magnitudes?
the altitude on Mt. Everest, the velocity of a fly, the age of What can you say about their directions?
Earth, the boiling point of water, the cost of a book, Earth’s
population, or the acceleration of gravity? 14. If three vectors sum up to zero, what geometric
condition do they satisfy?
3. Give a specific example of a vector, stating its
magnitude, units, and direction.
6.2 Coordinate Systems and Components of a
4. What do vectors and scalars have in common? How do Vector
they differ?
15. Give an example of a nonzero vector that has a
component of zero.
→ →
5. Suppose you add two vectors A and B . What
relative direction between them produces the resultant with 16. Explain why a vector cannot have a component greater
the greatest magnitude? What is the maximum magnitude? than its own magnitude.
What relative direction between them produces the
resultant with the smallest magnitude? What is the 17. If two vectors are equal, what can you say about their
minimum magnitude? components?

6. Is it possible to add a scalar quantity to a vector → →


quantity? 18. If vectors A and B are orthogonal, what is the
→ →
component of B along the direction of A ? What is
7. Is it possible for two vectors of different magnitudes → →
the component of A along the direction of B ?
to add to zero? Is it possible for three vectors of different
magnitudes to add to zero? Explain.
19. If one of the two components of a vector is not zero,
8. Does the odometer in an automobile indicate a scalar or can the magnitude of the other vector component of this
a vector quantity? vector be zero?

9. When a 10,000-m runner competing on a 400-m track 20. If two vectors have the same magnitude, do their
crosses the finish line, what is the runner’s net components have to be the same?
displacement? Can this displacement be zero? Explain.

10. A vector has zero magnitude. Is it necessary to specify


its direction? Explain.

PROBLEMS
ascends again for 7 m and descends again for 24.0 m,
6.1 Scalars and Vectors where he makes a stop, waiting for his buddy. Assuming the
21. A scuba diver makes a slow descent into the depths positive direction up to the surface, express his net vertical
of the ocean. His vertical position with respect to a boat displacement vector in terms of the unit vector. What is his
on the surface changes several times. He makes the first distance to the boat?
stop 9.0 m from the boat but has a problem with equalizing
the pressure, so he ascends 3.0 m and then continues 22. In a tug-of-war game on one campus, 15 students pull
descending for another 12.0 m to the second stop. From on a rope at both ends in an effort to displace the central
there, he ascends 4 m and then descends for 18.0 m, knot to one side or the other. Two students pull with force
208 Chapter 6 | Introduction to Vectors

196 N each to the right, four students pull with force 98 deal during the day and he is blown along the following
N each to the left, five students pull with force 62 N each directions: 2.50 km and 45.0° north of west, then 4.70 km
to the left, three students pull with force 150 N each to the and 60.0° south of east, then 1.30 km and 25.0° south of
right, and one student pulls with force 250 N to the left. west, then 5.10 km straight east, then 1.70 km and 5.00°
Assuming the positive direction to the right, express the net
east of north, then 7.20 km and 55.0° south of west, and
pull on the knot in terms of the unit vector. How big is the
net pull on the knot? In what direction? finally 2.80 km and 10.0° north of east. Use a graphical
method to find the castaway’s final position relative to the
23. Suppose you walk 18.0 m straight west and then 25.0 island.
m straight north. How far are you from your starting point
and what is the compass direction of a line connecting 28. A small plane flies 40.0 km in a direction 60° north
your starting point to your final position? Use a graphical of east and then flies 30.0 km in a direction 15° north of
method. east. Use a graphical method to find the total distance the
plane covers from the starting point and the direction of the
24. For the vectors given in the following figure, use path to the final position.
a graphical method to find the following resultants: (a)
→ → → → → →
A + B , (b) C + B , (c) D + F , (d) 29. A trapper walks a 5.0-km straight-line distance from
→ → → → → → his cabin to the lake, as shown in the following figure. Use
A − B , (e) D − F , (f) A + 2 F , (g); and (h) a graphical method (the parallelogram rule) to determine
→ → → the trapper’s displacement directly to the east and
A −4 D +2 F .
displacement directly to the north that sum up to his
resultant displacement vector. If the trapper walked only in
directions east and north, zigzagging his way to the lake,
how many kilometers would he have to walk to get to the
lake?

30. A surveyor measures the distance across a river that


flows straight north by the following method. Starting
directly across from a tree on the opposite bank, the
surveyor walks 100 m along the river to establish a
baseline. She then sights across to the tree and reads that
the angle from the baseline to the tree is 35° . How wide is
25. A delivery man starts at the post office, drives 40 km
north, then 20 km west, then 60 km northeast, and finally the river?
50 km north to stop for lunch. Use a graphical method to
find his net displacement vector. 31. A pedestrian walks 6.0 km east and then 13.0 km
north. Use a graphical method to find the pedestrian’s
26. An adventurous dog strays from home, runs three resultant displacement and geographic direction.
blocks east, two blocks north, one block east, one block
north, and two blocks west. Assuming that each block is 32. The magnitudes of two displacement vectors are A =
about 100 m, how far from home and in what direction is 20 m and B = 6 m. What are the largest and the smallest
the dog? Use a graphical method. values of the magnitude of the resultant
→ → →
R = A + B ?
27. In an attempt to escape a desert island, a castaway
builds a raft and sets out to sea. The wind shifts a great

This OpenStax book is available for free at https://legacy.cnx.org/content/col12169/1.46


Chapter 6 | Introduction to Vectors 209

nearest centimeter.
6.2 Coordinate Systems and Components of a
Vector 40. A chameleon is resting quietly on a lanai screen,
waiting for an insect to come by. Assume the origin of a
33. Assuming the +x-axis is horizontal and points to the
Cartesian coordinate system at the lower left-hand corner
right, resolve the vectors given in the following figure to
of the screen and the horizontal direction to the right as the
their scalar components and express them in vector
+x-direction. If its coordinates are (2.000 m, 1.000 m), (a)
component form.
how far is it from the corner of the screen? (b) What is its
location in polar coordinates?
34. Suppose you walk 18.0 m straight west and then 25.0
m straight north. How far are you from your starting point?
41. Two points in the Cartesian plane are A(2.00 m, −4.00
What is your displacement vector? What is the direction of
m) and B(−3.00 m, 3.00 m). Find the distance between
your displacement? Assume the +x-axis is horizontal to the
them and their polar coordinates.
right.

42. A fly enters through an open window and zooms


35. You drive 7.50 km in a straight line in a direction 15°
around the room. In a Cartesian coordinate system with
east of north. (a) Find the distances you would have to drive three axes along three edges of the room, the fly changes
straight east and then straight north to arrive at the same its position from point b(4.0 m, 1.5 m, 2.5 m) to point e(1.0
point. (b) Show that you still arrive at the same point if m, 4.5 m, 0.5 m). Find the scalar components of the fly’s
the east and north legs are reversed in order. Assume the displacement vector and express its displacement vector in
+x-axis is to the east. vector component form. What is its magnitude?

36. A sledge is being pulled by two horses on a flat


terrain. The net force on the sledge can be expressed in 6.3 Algebra of Vectors
the Cartesian coordinate system as vector
→ ^ ^ ^ ^ → ^ ^
F = (−2980.0 i + 8200.0 j )N , where i and j 43. For vectors B =−i −4j and
→ ^ ^ → →
denote directions to the east and north, respectively. Find A = −3 i − 2 j , calculate (a) A + B and its
the magnitude and direction of the pull. → →
magnitude and direction angle, and (b) A − B and its
37. A trapper walks a 5.0-km straight-line distance from magnitude and direction angle.
her cabin to the lake, as shown in the following figure.
Determine the east and north components of her 44. A particle undergoes three consecutive displacements
displacement vector. How many more kilometers would → ^ ^ ^
she have to walk if she walked along the component given by vectors D 1 = (3.0 i − 4.0 j − 2.0 k )mm ,
displacements? What is her displacement vector? → ^ ^ ^
D 2 = (1.0 i − 7.0 j + 4.0 k )mm , and
→ ^ ^ ^
D 3 = (−7.0 i + 4.0 j + 1.0 k )mm . (a) Find the
resultant displacement vector of the particle. (b) What is
the magnitude of the resultant displacement? (c) If all
displacements were along one line, how far would the
particle travel?

45. Given two displacement vectors


→ ^ ^ ^
A = (3.00 i − 4.00 j + 4.00 k )m and
→ ^ ^ ^
B = (2.00 i + 3.00 j − 7.00 k )m , find the
displacements and their magnitudes for (a)
38. The polar coordinates of a point are 4π/3 and 5.50 m. → → → → → →
What are its Cartesian coordinates? C = A + B and (b) D = 2 A − B .

39. Two points in a plane have polar coordinates 46. A small plane flies 40.0 km in a direction 60° north
P 1(2.500 m, π/6) and P 2(3.800 m, 2π/3) . Determine of east and then flies 30.0 km in a direction 15° north
their Cartesian coordinates and the distance between them of east. Use the analytical method to find the total distance
in the Cartesian coordinate system. Round the distance to a the plane covers from the starting point, and the geographic
210 Chapter 6 | Introduction to Vectors

direction of its displacement vector. What is its vector. (b) How far is the restaurant from the post office?
displacement vector? (c) If he returns directly from the restaurant to the post
office, what is his displacement vector on the return trip?
47. In an attempt to escape a desert island, a castaway (d) What is his compass heading on the return trip? Assume
builds a raft and sets out to sea. The wind shifts a great the +x-axis is to the east.
deal during the day, and she is blown along the following
straight lines: 2.50 km and 45.0° north of west, then 4.70 51. An adventurous dog strays from home, runs three
km and 60.0° south of east, then 1.30 km and 25.0° south blocks east, two blocks north, and one block east, one block
of west, then 5.10 km due east, then 1.70 km and 5.00° north, and two blocks west. Assuming that each block is
about a 100 yd, use the analytical method to find the dog’s
east of north, then 7.20 km and 55.0° south of west, and
net displacement vector, its magnitude, and its direction.
finally 2.80 km and 10.0° north of east. Use the analytical Assume the +x-axis is to the east. How would your answer
method to find the resultant vector of all her displacement be affected if each block was about 100 m?
vectors. What is its magnitude and direction?
→ ^ ^
48. Assuming the +x-axis is horizontal to the right for 52. If D = (6.00 i − 8.00 j )m ,
the vectors given in the following figure, use the analytical → ^ ^
→ → B = (−8.00 i + 3.00 j )m , and
method to find the following resultants: (a) A + B ,
→ ^ ^
→ → → → → → A = (26.0 i + 19.0 j )m , find the unknown constants a
(b) C + B , (c) D + F , (d) A − B , (e)
→ → → → → → → → → → →
D − F , (f) A + 2 F , (g) C −2 D +3 F , and b such that a D + b B + A = 0 .
→ → →
and (h) A − 4 D + 2 F .
→ ^ ^
53. Given the displacement vector D = (3 i − 4 j )m,

find the displacement vector R so that
→ → ^
D + R = −4D j .

54. Find the unit vector of direction for the following


→ ^ ^
vector quantities: (a) Force F = (3.0 i − 2.0 j )N , (b)
→ ^ ^
displacement D = (−3.0 i − 4.0 j )m , and (c) velocity
→ ^ ^
v = (−5.00 i + 4.00 j )m/s .

55. At one point in space, the direction of the electric field


vector is given in the Cartesian system by the unit vector
^ ^ ^
E = 1 / 5 i − 2 / 5 j . If the magnitude of the electric
field vector is E = 400.0 V/m, what are the scalar
components E x , E y , and E z of the electric field vector

E at this point? What is the direction angle θ E of the

Figure 6.27 electric field vector at this point?

49. Given the vectors in the preceding figure, find vector 56. A barge is pulled by the two tugboats shown in the
→ → → → following figure. One tugboat pulls on the barge with a
R that solves equations (a) D + R = F and (b) force of magnitude 4000 units of force at 15° above the
→ → → →
C − 2 D + 5 R = 3 F . Assume the +x-axis is line AB (see the figure and the other tugboat pulls on the
horizontal to the right. barge with a force of magnitude 5000 units of force at 12°
below the line AB. Resolve the pulling forces to their scalar
components and find the components of the resultant force
50. A delivery man starts at the post office, drives 40 km
pulling on the barge. What is the magnitude of the resultant
north, then 20 km west, then 60 km northeast, and finally
pull? What is its direction relative to the line AB?
50 km north to stop for lunch. Use the analytical method
to determine the following: (a) Find his net displacement

This OpenStax book is available for free at https://legacy.cnx.org/content/col12169/1.46


Chapter 6 | Introduction to Vectors 211

57. In the control tower at a regional airport, an air traffic


controller monitors two aircraft as their positions change
with respect to the control tower. One plane is a cargo
carrier Boeing 747 and the other plane is a Douglas DC-3.
The Boeing is at an altitude of 2500 m, climbing at 10°
above the horizontal, and moving 30° north of west. The
DC-3 is at an altitude of 3000 m, climbing at 5° above the
horizontal, and cruising directly west. (a) Find the position
vectors of the planes relative to the control tower. (b) What
is the distance between the planes at the moment the air
traffic controller makes a note about their positions?

Figure 6.28

ADDITIONAL PROBLEMS
58. You fly 32.0 km in a straight line in still air in the coordinates of the following points: (a) (−x, y), (b) (−2x,
direction 35.0° south of west. (a) Find the distances you −2y), and (c) (3x, −3y).
would have to fly due south and then due west to arrive
at the same point. (b) Find the distances you would have → →
61. Vectors A and B have identical magnitudes
to fly first in a direction 45.0° south of west and then
of 5.0 units. Find the angle between them if
in a direction 45.0° west of north. Note these are the → → ^
components of the displacement along a different set of A + B =5 2j .
axes—namely, the one rotated by 45° with respect to the
axes in (a). 62. Starting at the island of Moi in an unknown
archipelago, a fishing boat makes a round trip with two
59. Rectangular coordinates of a point are given by (2, y) stops at the islands of Noi and Poi. It sails from Moi for
and its polar coordinates are given by (r, π/6) . Find y and 4.76 nautical miles (nmi) in a direction 37° north of east
r. to Noi. From Noi, it sails 69° west of north to Poi. On
its return leg from Poi, it sails 28° east of south. What
60. If the polar coordinates of a point are (r, φ) and distance does the boat sail between Noi and Poi? What
distance does it sail between Moi and Poi? Express your
its rectangular coordinates are (x, y) , determine the polar
answer both in nautical miles and in kilometers. Note: 1
212 Chapter 6 | Introduction to Vectors

nmi = 1852 m. → ^ ^
u = (−18.0 i − 13.0 j )km/h , how fast and in what
63. An air traffic controller notices two signals from two ^ ^
geographic direction is it heading? Here, i and j are
planes on the radar monitor. One plane is at altitude 800
m and in a 19.2-km horizontal distance to the tower in a directions to geographic east and north, respectively.
direction 25° south of west. The second plane is at altitude
1100 m and its horizontal distance is 17.6 km and 20° 69. Find the scalar components of three-dimensional
→ →
south of west. What is the distance between these planes? vectors G and H in the following figure and write
the vectors in vector component form in terms of the unit
→ → → vectors of the axes.
64. Show that when A + B = C , then
Vector G has magnitude 10.0. Its projection in the x y plane
2 2 2
C = A + B + 2AB cos φ , where φ is the angle is between the positive x and positive y directions, at an
→ → angle of 45 degrees from the positive x direction. The angle
between vectors A and B . between vector G and the positive z direction is 60 degrees.
Vector H has magnitude 15.0. Its projection in the x y
65. Four force vectors each have the same magnitude f. plane is between the negative x and positive y directions,
What is the largest magnitude the resultant force vector at an angle of 30 degrees from the positive y direction. The
may have when these forces are added? What is the angle between vector H and the positive z direction is 450
smallest magnitude of the resultant? Make a graph of both degrees.
situations.
70. A diver explores a shallow reef off the coast of Belize.
66. A skater glides along a circular path of radius 5.00 She initially swims 90.0 m north, makes a turn to the east
m in clockwise direction. When he coasts around one- and continues for 200.0 m, then follows a big grouper for
half of the circle, starting from the west point, find (a) 80.0 m in the direction 30° north of east. In the meantime,
the magnitude of his displacement vector and (b) how far a local current displaces her by 150.0 m south. Assuming
he actually skated. (c) What is the magnitude of his the current is no longer present, in what direction and how
displacement vector when he skates all the way around the far should she now swim to come back to the point where
circle and comes back to the west point? she started?

67. A stubborn dog is being walked on a leash by its →


71. A force vector A has x- and y-components,
owner. At one point, the dog encounters an interesting scent
at some spot on the ground and wants to explore it in detail, respectively, of −8.80 units of force and 15.00 units of

but the owner gets impatient and pulls on the leash with force. The x- and y-components of force vector B are,
→ ^ ^ ^ respectively, 13.20 units of force and −6.60 units of force.
force F = (98.0 i + 132.0 j + 32.0 k )N along the →
leash. (a) What is the magnitude of the pulling force? (b) Find the components of force vector C that satisfies the
What angle does the leash make with the vertical? → → →
vector equation A − B + 3 C = 0 .

68. If the velocity vector of a polar bear is

This OpenStax book is available for free at https://legacy.cnx.org/content/col12169/1.46


Chapter 7 | Kinematics in Two and Three Dimensions 213

7 | KINEMATICS IN TWO
AND THREE DIMENSIONS

Figure 7.1 The Red Arrows is the aerobatics display team of Britain’s Royal Air Force. Based in Lincolnshire, England, they
perform precision flying shows at high speeds, which requires accurate measurement of position, velocity, and acceleration in
three dimensions. (credit: modification of work by Phil Long)

Chapter Outline
7.1 Displacement and Velocity Vectors
7.2 Acceleration Vector
7.3 Projectile Motion
7.4 Uniform Circular Motion
7.5 Relative Motion in One and Two Dimensions

Introduction
To give a complete description of kinematics, we must explore motion in two and three dimensions. After all, most objects
in our universe do not move in straight lines; rather, they follow curved paths. From kicked footballs to the flight paths of
birds to the orbital motions of celestial bodies and down to the flow of blood plasma in your veins, most motion follows
curved trajectories.
Fortunately, the treatment of motion in one dimension in the previous chapter has given us a foundation on which to build,
as the concepts of position, displacement, velocity, and acceleration defined in one dimension can be expanded to two and
three dimensions. Consider the Red Arrows, also known as the Royal Air Force Aerobatic team of the United Kingdom.
Each jet follows a unique curved trajectory in three-dimensional airspace, as well as has a unique velocity and acceleration.
Thus, to describe the motion of any of the jets accurately, we must assign to each jet a unique position vector in three
dimensions as well as a unique velocity and acceleration vector. We can apply the same basic equations for displacement,
velocity, and acceleration we derived in Motion Along a Straight Line to describe the motion of the jets in two and three
dimensions, but with some modifications—in particular, the inclusion of vectors.
In this chapter we also explore two special types of motion in two dimensions: projectile motion and circular motion. Last,
we conclude with a discussion of relative motion. In the chapter-opening picture, each jet has a relative motion with respect
214 Chapter 7 | Kinematics in Two and Three Dimensions

to any other jet in the group or to the people observing the air show on the ground.

7.1 | Displacement and Velocity Vectors


Learning Objectives
By the end of this section, you will be able to:
• Calculate position vectors in a multidimensional displacement problem.
• Solve for the displacement in two or three dimensions.
• Calculate the velocity vector given the position vector as a function of time.
• Calculate the average velocity in multiple dimensions.

Displacement and velocity in two or three dimensions are straightforward extensions of the one-dimensional definitions.
However, now they are vector quantities, so calculations with them have to follow the rules of vector algebra, not scalar
algebra.

Displacement Vector
To describe motion in two and three dimensions, we must first establish a coordinate system and a convention for the axes.
We generally use the coordinates x, y, and z to locate a particle at point P(x, y, z) in three dimensions. If the particle is
moving, the variables x, y, and z are functions of time (t):
x = x(t) y = y(t) z = z(t). (7.1)

The position vector from the origin of the coordinate system to point P is →
r (t). In unit vector notation, introduced in
Coordinate Systems and Components of a Vector, →
r (t) is

→ ^ ^ ^ (7.2)
r (t) = x(t) i + y(t) j + z(t) k .

Figure 7.2 shows the coordinate system and the vector to point P, where a particle could be located at a particular time
t. Note the orientation of the x, y, and z axes. This orientation is called a right-handed coordinate system (Coordinate
Systems and Components of a Vector) and it is used throughout the chapter.

Figure 7.2 A three-dimensional coordinate system with a


particle at position P(x(t), y(t), z(t)).

With our definition of the position of a particle in three-dimensional space, we can formulate the three-dimensional
displacement. Figure 7.3 shows a particle at time t 1 located at P 1 with position vector →
r (t 1). At a later time t 2, the

This OpenStax book is available for free at https://legacy.cnx.org/content/col12169/1.46


Chapter 7 | Kinematics in Two and Three Dimensions 215

particle is located at P 2 with position vector →


r (t 2) . The displacement vector Δ →
r is found by subtracting →
r (t 1)
from →
r (t 2)  :

Δ→
r = →
r (t 2) − →
r (t 1). (7.3)

Vector addition is discussed in Algebra of Vectors. Note that this is the same operation we did in one dimension, but now
the vectors are in three-dimensional space.

→ → →
Figure 7.3 The displacement Δ r = r (t 2) − r (t 1) is
the vector from P 1 to P 2 .

The following examples illustrate the concept of displacement in multiple dimensions.

Example 7.1

Polar Orbiting Satellite


A satellite is in a circular polar orbit around Earth at an altitude of 400 km—meaning, it passes directly overhead
at the North and South Poles. What is the magnitude and direction of the displacement vector from when it is
directly over the North Pole to when it is at −45° latitude?
Strategy
We make a picture of the problem to visualize the solution graphically. This will aid in our understanding of the
displacement. We then use unit vectors to solve for the displacement.
Solution
Figure 7.4 shows the surface of Earth and a circle that represents the orbit of the satellite. Although satellites are
moving in three-dimensional space, they follow trajectories of ellipses, which can be graphed in two dimensions.
The position vectors are drawn from the center of Earth, which we take to be the origin of the coordinate system,
with the y-axis as north and the x-axis as east. The vector between them is the displacement of the satellite. We
take the radius of Earth as 6370 km, so the length of each position vector is 6770 km.
216 Chapter 7 | Kinematics in Two and Three Dimensions

Figure 7.4 Two position vectors are drawn from the center of Earth,
which is the origin of the coordinate system, with the y-axis as north and the
x-axis as east. The vector between them is the displacement of the satellite.

In unit vector notation, the position vectors are

→ ^
r (t 1) = 6770. km j
→ ^ ^
r (t 2) = 6770. km (cos 45°) i + 6770. km (sin(−45°)) j .

Evaluating the sine and cosine, we have

→ ^
r (t 1) = 6770. j
→ ^ ^
r (t 2) = 4787 i − 4787 j .

Now we can find Δ →


r , the displacement of the satellite:
^ ^
Δ→
r = →
r (t 2) − →
r (t 1) = 4787 i − 11,557 j .

The magnitude of the displacement is Δ → | |


r = (4787) 2 + (−11,557) 2 = 12,509 km. The angle the displacement makes
⎛−11,557 ⎞
with the x-axis is θ = tan −1 ⎝ ⎠ = −67.5°.
4787
Significance
Plotting the displacement gives information and meaning to the unit vector solution to the problem. When plotting the
displacement, we need to include its components as well as its magnitude and the angle it makes with a chosen axis—in this
case, the x-axis (Figure 7.5).

This OpenStax book is available for free at https://legacy.cnx.org/content/col12169/1.46


Chapter 7 | Kinematics in Two and Three Dimensions 217

Figure 7.5 Displacement vector with components, angle, and magnitude.

Note that the satellite took a curved path along its circular orbit to get from its initial position to its final position in this
example. It also could have traveled 4787 km east, then 11,557 km south to arrive at the same location. Both of these paths
are longer than the length of the displacement vector. In fact, the displacement vector gives the shortest path between two
points in one, two, or three dimensions.
Many applications in physics can have a series of displacements, as discussed in the previous chapter. The total
displacement is the sum of the individual displacements, only this time, we need to be careful, because we are adding
vectors. We illustrate this concept with an example of Brownian motion.

Example 7.2

Brownian Motion
Brownian motion is a chaotic random motion of particles suspended in a fluid, resulting from collisions with
the molecules of the fluid. This motion is three-dimensional. The displacements in numerical order of a particle
undergoing Brownian motion could look like the following, in micrometers (Figure 7.6):

^ ^ ^
Δ→
r 1 = 2.0 i + j + 3.0 k
^ ^
Δ→
r 2 = − i + 3.0 k
^ ^ ^
Δ→
r 3 = 4.0 i − 2.0 j + k
^ ^ ^
Δ→
r 4 = −3.0 i + j + 2.0 k .

What is the total displacement of the particle from the origin?


218 Chapter 7 | Kinematics in Two and Three Dimensions

Figure 7.6 Trajectory of a particle undergoing random


displacements of Brownian motion. The total displacement is
shown in red.

Solution
We form the sum of the displacements and add them as vectors:

Δ→
r Total =∑Δ→
r i=Δ→
r 1+Δ→
r 2+Δ→
r 3+Δ→
r 4
^ ^ ^
= (2.0 − 1.0 + 4.0 − 3.0) i + (1.0 + 0 − 2.0 + 1.0) j + (3.0 + 3.0 + 1.0 + 2.0) k
^ ^ ^
= 2.0 i + 0 j + 9.0 k μm.

To complete the solution, we express the displacement as a magnitude and direction,


⎛ ⎞
|Δ →r Total| = 2.0 2 + 0 2 + 9.0 2 = 9.2 μm, θ = tan −1 ⎝9 ⎠ = 77°,
2
with respect to the x-axis in the xz-plane.
Significance
From the figure we can see the magnitude of the total displacement is less than the sum of the magnitudes of the
individual displacements.

Velocity Vector
In the previous chapter we found the instantaneous velocity by calculating the derivative of the position function with
respect to time. We can do the same operation in two and three dimensions, but we use vectors. The instantaneous velocity
vector is now


r (t + Δt) − →
r (t) d → (7.4)

v (t) = lim = r .
Δt → 0 Δt dt

Let’s look at the relative orientation of the position vector and velocity vector graphically. In Figure 7.7 we show the
vectors → r (t) and → r (t + Δt), which give the position of a particle moving along a path represented by the gray line.
As Δt goes to zero, the velocity vector, given by Equation 7.4, becomes tangent to the path of the particle at time t.

This OpenStax book is available for free at https://legacy.cnx.org/content/col12169/1.46


Chapter 7 | Kinematics in Two and Three Dimensions 219

Figure 7.7 A particle moves along a path given by the gray


line. In the limit as Δt approaches zero, the velocity vector
becomes tangent to the path of the particle.

Equation 7.4 can also be written in terms of the components of →


v (t). Since

→ ^ ^ ^
r (t) = x(t) i + y(t) j + z(t) k ,

we can write

→ ^ ^ ^ (7.5)
v (t) = v x(t) i + v y(t) j + v z(t) k

where

dx(t) dy(t) dz(t) (7.6)


v x(t) = , v y(t) = , v z(t) = .
dt dt dt

If only the average velocity is of concern, we have the vector equivalent of the one-dimensional average velocity for two
and three dimensions:


r (t 2) − → r (t 1) (7.7)

v = .
avg t2 − t1

Example 7.3

Calculating the Velocity Vector


→ ^ ^ ^
The position function of a particle is r (t) = 2.0t 2 i + (2.0 + 3.0t) j + 5.0t k m. (a) What is the
instantaneous velocity and speed at t = 2.0 s? (b) What is the average velocity between 1.0 s and 3.0 s?
Solution
Using Equation 7.5 and Equation 7.6, and taking the derivative of the position function with respect to time,
we find
220 Chapter 7 | Kinematics in Two and Three Dimensions

dr(t) ^ ^ ^
(a) v(t) = = 4.0t i + 3.0 j + 5.0 k m/s
dt

→ ^ ^ ^
v (2.0s) = 8.0 i + 3.0 j + 5.0 k m/s

Speed → | |
v (2.0 s) = 8 2 + 3 2 + 5 2 = 9.9 m/s.

(b) From Equation 7.7,


→ ^ ^ ^ ^ ^ ^
→ r (t 2) − → r (t 1) →
r (3.0 s) − →r (1.0 s) (18 i + 11 j + 15 k ) m − (2 i + 5 j + 5 k ) m
v avg = t2 − t1 = =
3.0 s − 1.0 s 2.0 s
^ ^ ^
(16 i + 6 j + 10 k ) m ^ ^ ^
= = 8.0 i + 3.0 j + 5.0 k m/s.
2.0 s
Significance
We see the average velocity is the same as the instantaneous velocity at t = 2.0 s, as a result of the velocity function
being linear. This need not be the case in general. In fact, most of the time, instantaneous and average velocities
are not the same.

^ ^
Check Your Understanding The position function of a particle is →
7.1
r (t) = 3.0t 3 i + 4.0 j . (a) What is
the instantaneous velocity at t = 3 s? (b) Is the average velocity between 2 s and 4 s equal to the instantaneous
velocity at t = 3 s?

The Independence of Perpendicular Motions


When we look at the three-dimensional equations for position and velocity written in unit vector notation, Equation 7.2
and Equation 7.5, we see the components of these equations are separate and unique functions of time that do not depend
on one another. Motion along the x direction has no part of its motion along the y and z directions, and similarly for the other
two coordinate axes. Thus, the motion of an object in two or three dimensions can be divided into separate, independent
motions along the perpendicular axes of the coordinate system in which the motion takes place.
To illustrate this concept with respect to displacement, consider a woman walking from point A to point B in a city with
square blocks. The woman taking the path from A to B may walk east for so many blocks and then north (two perpendicular
directions) for another set of blocks to arrive at B. How far she walks east is affected only by her motion eastward. Similarly,
how far she walks north is affected only by her motion northward.

Independence of Motion
In the kinematic description of motion, we are able to treat the horizontal and vertical components of motion separately.
In many cases, motion in the horizontal direction does not affect motion in the vertical direction, and vice versa.

An example illustrating the independence of vertical and horizontal motions is given by two baseballs. One baseball is
dropped from rest. At the same instant, another is thrown horizontally from the same height and it follows a curved path. A
stroboscope captures the positions of the balls at fixed time intervals as they fall (Figure 7.8).

This OpenStax book is available for free at https://legacy.cnx.org/content/col12169/1.46


Chapter 7 | Kinematics in Two and Three Dimensions 221

Figure 7.8 A diagram of the motions of two identical balls:


one falls from rest and the other has an initial horizontal
velocity. Each subsequent position is an equal time interval.
Arrows represent the horizontal and vertical velocities at each
position. The ball on the right has an initial horizontal velocity
whereas the ball on the left has no horizontal velocity. Despite
the difference in horizontal velocities, the vertical velocities and
positions are identical for both balls, which shows the vertical
and horizontal motions are independent.

It is remarkable that for each flash of the strobe, the vertical positions of the two balls are the same. This similarity implies
vertical motion is independent of whether the ball is moving horizontally. (Assuming no air resistance, the vertical motion
of a falling object is influenced by gravity only, not by any horizontal forces.) Careful examination of the ball thrown
horizontally shows it travels the same horizontal distance between flashes. This is because there are no additional forces on
the ball in the horizontal direction after it is thrown. This result means horizontal velocity is constant and is affected neither
by vertical motion nor by gravity (which is vertical). Note this case is true for ideal conditions only. In the real world, air
resistance affects the speed of the balls in both directions.
The two-dimensional curved path of the horizontally thrown ball is composed of two independent one-dimensional
motions (horizontal and vertical). The key to analyzing such motion, called projectile motion, is to resolve it into motions
along perpendicular directions. Resolving two-dimensional motion into perpendicular components is possible because the
components are independent.
222 Chapter 7 | Kinematics in Two and Three Dimensions

7.2 | Acceleration Vector


Learning Objectives
By the end of this section, you will be able to:
• Calculate the acceleration vector given the velocity function in unit vector notation.
• Describe the motion of a particle with a constant acceleration in three dimensions.
• Use the one-dimensional motion equations along perpendicular axes to solve a problem in two
or three dimensions with a constant acceleration.
• Express the acceleration in unit vector notation.

Instantaneous Acceleration
In addition to obtaining the displacement and velocity vectors of an object in motion, we often want to know its acceleration
vector at any point in time along its trajectory. This acceleration vector is the instantaneous acceleration and it can be
obtained from the derivative with respect to time of the velocity function, as we have seen in a previous chapter. The only
difference in two or three dimensions is that these are now vector quantities. Taking the derivative with respect to time
→v (t), we find


v (t + Δt) − →
v (t) d →
v (t) (7.8)

a (t) = lim = .
t→0 Δt dt

The acceleration in terms of components is

→ dv (t) ^ dv y(t) ^ dv z(t) ^ (7.9)


a (t) = x i + j + k.
dt dt dt

Also, since the velocity is the derivative of the position function, we can write the acceleration in terms of the second
derivative of the position function:

→ d 2 x(t) ^ d 2 y(t) ^ d 2 z(t) ^ (7.10)


a (t) = i + j + k.
dt 2 dt 2 dt 2

Example 7.4

Finding an Acceleration Vector


^ ^ ^
A particle has a velocity of →
v (t) = 5.0t i + t 2 j − 2.0t 3 k m/s. (a) What is the acceleration function? (b)
What is the acceleration vector at t = 2.0 s? Find its magnitude and direction.
Solution
(a) We take the first derivative with respect to time of the velocity function to find the acceleration. The derivative
is taken component by component:
→ ^ ^ ^
a (t) = 5.0 i + 2.0t j − 6.0t 2 k m/s 2.

This OpenStax book is available for free at https://legacy.cnx.org/content/col12169/1.46


Chapter 7 | Kinematics in Two and Three Dimensions 223

^ ^ ^
(b) Evaluating →
a (2.0 s) = 5.0 i + 4.0 j − 24.0 k m/s 2 gives us the direction in unit vector notation. The
magnitude of the acceleration is → | |
a (2.0 s) = 5.0 2 + 4.0 2 + (−24.0) 2 = 24.8 m/s 2.

Significance
In this example we find that acceleration has a time dependence and is changing throughout the motion. Let’s
consider a different velocity function for the particle.

Example 7.5

Finding a Particle Acceleration


^ ^ ^
A particle has a position function →
r (t) = (10t − t 2) i + 5t j + 5t k m. (a) What is the velocity? (b) What is
the acceleration? (c) Describe the motion from t = 0 s.
Strategy
We can gain some insight into the problem by looking at the position function. It is linear in y and z, so we know
the acceleration in these directions is zero when we take the second derivative. Also, note that the position in the
x direction is zero for t = 0 s and t = 10 s.
Solution
(a) Taking the derivative with respect to time of the position function, we find
→ ^ ^ ^
v (t) = (10 − 2t) i + 5 j + 5 k m/s.

The velocity function is linear in time in the x direction and is constant in the y and z directions.
(b) Taking the derivative of the velocity function, we find
→ ^
a (t) = −2 i m/s 2.
The acceleration vector is a constant in the negative x-direction.
(c) The trajectory of the particle can be seen in Figure 7.9. Let’s look in the y and z directions first. The particle’s
position increases steadily as a function of time with a constant velocity in these directions. In the x direction,
however, the particle follows a path in positive x until t = 5 s, when it reverses direction. We know this from
looking at the velocity function, which becomes zero at this time and negative thereafter. We also know this
because the acceleration is negative and constant—meaning, the particle is decelerating, or accelerating in the
negative direction. The particle’s position reaches 25 m, where it then reverses direction and begins to accelerate
in the negative x direction. The position reaches zero at t = 10 s.
224 Chapter 7 | Kinematics in Two and Three Dimensions

Figure 7.9 The particle starts at point (x, y, z) = (0, 0, 0) with position vector

r = 0. The projection of the trajectory onto the xy-plane is shown. The
values of y and z increase linearly as a function of time, whereas x has a turning
point at t = 5 s and 25 m, when it reverses direction. At this point, the x
component of the velocity becomes negative. At t = 10 s, the particle is back to 0
m in the x direction.

Significance
By graphing the trajectory of the particle, we can better understand its motion, given by the numerical results of
the kinematic equations.

7.2 Check Your Understanding Suppose the acceleration function has the form
→ ^ ^ ^
a (t) = a i + b j + c k m/s 2, where a, b, and c are constants. What can be said about the functional form of
the velocity function?

Constant Acceleration
Multidimensional motion with constant acceleration can be treated the same way as shown in the previous chapter for
one-dimensional motion. Earlier we showed that three-dimensional motion is equivalent to three one-dimensional motions,
each along an axis perpendicular to the others. To develop the relevant equations in each direction, let’s consider the two-
dimensional problem of a particle moving in the xy plane with constant acceleration, ignoring the z-component for the
moment. The acceleration vector is
→ ^ ^
a = a 0x i + a 0y j .

Each component of the motion has a separate set of equations similar to Equation 3.9–Equation 3.13 of the
Section 3.4. We show only the equations for position and velocity in the x- and y-directions. A similar set of kinematic
equations could be written for motion in the z-direction:

This OpenStax book is available for free at https://legacy.cnx.org/content/col12169/1.46


Chapter 7 | Kinematics in Two and Three Dimensions 225

x(t) = x 0 + (v x) avg t (7.11)


v x(t) = v 0x + a x t (7.12)

x(t) = x 0 + v 0x t + 1 a x t 2 (7.13)
2
v 2x(t) = v 20x + 2a x(x − x 0) (7.14)

y(t) = y 0 + (v y) avg t (7.15)


v y(t) = v 0y + a y t (7.16)

y(t) = y 0 + v 0y t + 1 a y t 2 (7.17)
2
v 2y(t) = v 20y + 2a y(y − y 0). (7.18)

Here the subscript 0 denotes the initial position or velocity. Equation 7.11 to Equation 7.18 can be substituted into
Equation 7.2 and Equation 7.5 without the z-component to obtain the position vector and velocity vector as a function
of time in two dimensions:
→ ^ ^ ^ ^
r (t) = x(t) i + y(t) j and →
v (t) = v x(t) i + v y(t) j .

The following example illustrates a practical use of the kinematic equations in two dimensions.

Example 7.6

A Skier
Figure 7.10 shows a skier moving with an acceleration of 2.1 m/s 2 down a slope of 15° at t = 0. With the
origin of the coordinate system at the front of the lodge, her initial position and velocity are
→ ^ ^
r (0) = (75.0 i − 50.0 j ) m

and
→ ^ ^
v (0) = (4.1 i − 1.1 j ) m/s.

(a) What are the x- and y-components of the skier’s position and velocity as functions of time? (b) What are her
position and velocity at t = 10.0 s?
226 Chapter 7 | Kinematics in Two and Three Dimensions

Figure 7.10 A skier has an acceleration of 2.1 m/s 2 down a slope of 15°. The origin of the coordinate system is at
the ski lodge.

Strategy
Since we are evaluating the components of the motion equations in the x and y directions, we need to find the
components of the acceleration and put them into the kinematic equations. The components of the acceleration
are found by referring to the coordinate system in Figure 7.10. Then, by inserting the components of the initial
position and velocity into the motion equations, we can solve for her position and velocity at a later time t.
Solution
(a) The origin of the coordinate system is at the top of the hill with y-axis vertically upward and the x-axis
horizontal. By looking at the trajectory of the skier, the x-component of the acceleration is positive and the
y-component is negative. Since the angle is 15° down the slope, we find

a x = (2.1 m/s 2) cos(15°) = 2.0 m/s 2


a y = (−2.1 m/s 2) sin 15° = −0.54 m/s 2.

Inserting the initial position and velocity into Equation 7.12 and Equation 7.13 for x, we have

x(t) = 75.0 m + (4.1 m/s)t + 1 (2.0 m/s 2)t 2


2
v x(t) = 4.1 m/s + (2.0 m/s 2)t.
For y, we have

y(t) = −50.0 m + (−1.1 m/s)t + 1 (−0.54 m/s 2)t 2


2
v y(t) = −1.1 m/s + (−0.54 m/s 2)t.

(b) Now that we have the equations of motion for x and y as functions of time, we can evaluate them at t = 10.0 s:

x(10.0 s) = 75.0 m + (4.1 m/s 2)(10.0 s) + 1 (2.0 m/s 2)(10.0 s) 2 = 216.0 m


2
v x(10.0 s) = 4.1 m/s + (2.0 m/s 2)(10.0 s) = 24.1m/s

This OpenStax book is available for free at https://legacy.cnx.org/content/col12169/1.46


Chapter 7 | Kinematics in Two and Three Dimensions 227

y(10.0 s) = −50.0 m + (−1.1 m/s)(10.0 s) + 1 (−0.54 m/s 2)(10.0 s) 2 = −88.0 m


2
v y(10.0 s) = −1.1 m/s + (−0.54 m/s 2)(10.0 s) = −6.5 m/s.

The position and velocity at t = 10.0 s are, finally,


→ ^ ^
r (10.0 s) = (216.0 i − 88.0 j ) m
→ ^ ^
v (10.0 s) = (24.1 i − 6.5 j )m/s.

The magnitude of the velocity of the skier at 10.0 s is 25 m/s, which is 60 mi/h.
Significance
It is useful to know that, given the initial conditions of position, velocity, and acceleration of an object, we can
find the position, velocity, and acceleration at any later time.

With Equation 7.8 through Equation 7.10 we have completed the set of expressions for the position, velocity, and
acceleration of an object moving in two or three dimensions. If the trajectories of the objects look something like the “Red
Arrows” in the opening picture for the chapter, then the expressions for the position, velocity, and acceleration can be quite
complicated. In the sections to follow we examine two special cases of motion in two and three dimensions by looking at
projectile motion and circular motion.
At this University of Colorado Boulder website (https://openstaxcollege.org/l/21phetmotladyb) ,
you can explore the position velocity and acceleration of a ladybug with an interactive simulation that allows you
to change these parameters.

7.3 | Projectile Motion


Learning Objectives
By the end of this section, you will be able to:
• Use one-dimensional motion in perpendicular directions to analyze projectile motion.
• Calculate the range, time of flight, and maximum height of a projectile that is launched and
impacts a flat, horizontal surface.
• Find the time of flight and impact velocity of a projectile that lands at a different height from that
of launch.
• Calculate the trajectory of a projectile.

Projectile motion is the motion of an object thrown or projected into the air, subject only to acceleration as a result of
gravity. The applications of projectile motion in physics and engineering are numerous. Some examples include meteors as
they enter Earth’s atmosphere, fireworks, and the motion of any ball in sports. Such objects are called projectiles and their
path is called a trajectory. The motion of falling objects as discussed in Section 3.5 is a simple one-dimensional type of
projectile motion in which there is no horizontal movement. In this section, we consider two-dimensional projectile motion,
and our treatment neglects the effects of air resistance.
The most important fact to remember here is that motions along perpendicular axes are independent and thus can be
analyzed separately. We discussed this fact in Section 7.1, where we saw that vertical and horizontal motions are
independent. The key to analyzing two-dimensional projectile motion is to break it into two motions: one along the
horizontal axis and the other along the vertical. (This choice of axes is the most sensible because acceleration resulting from
gravity is vertical; thus, there is no acceleration along the horizontal axis when air resistance is negligible.) As is customary,
we call the horizontal axis the x-axis and the vertical axis the y-axis. It is not required that we use this choice of axes; it is
simply convenient in the case of gravitational acceleration. In other cases we may choose a different set of axes. Figure
7.11 illustrates the notation for displacement, where we define → s to be the total displacement, and → x and → y are its
component vectors along the horizontal and vertical axes, respectively. The magnitudes of these vectors are s, x, and y.
228 Chapter 7 | Kinematics in Two and Three Dimensions


s has
Figure 7.11 The total displacement s of a soccer ball at a point along its path. The vector
→ →
components x and y along the horizontal and vertical axes. Its magnitude is s and it makes an angle
θ with the horizontal.

To describe projectile motion completely, we must include velocity and acceleration, as well as displacement. We must find
their components along the x- and y-axes. Let’s assume all forces except gravity (such as air resistance and friction, for
example) are negligible. Defining the positive direction to be upward, the components of acceleration are then very simple:
a y = −g = −9.8 m/s 2 ( − 32 ft/s 2).

Because gravity is vertical, a x = 0. If a x = 0, this means the initial velocity in the x direction is equal to the final velocity
in the x direction, or v x = v 0x. With these conditions on acceleration and velocity, we can write the kinematic Equation
7.11 through Equation 7.18 for motion in a uniform gravitational field, including the rest of the kinematic equations
for a constant acceleration from Section 3.4. The kinematic equations for motion in a uniform gravitational field become
kinematic equations with a y = −g, a x = 0 :

Horizontal Motion
v 0x = v x, x = x 0 + v x t (7.19)

Vertical Motion

y = y 0 + 1 (v 0y + v y)t (7.20)
2
v y = v 0y − gt (7.21)

y = y 0 + v 0y t − 1 gt 2 (7.22)
2
v 2y = v 20y − 2g(y − y 0) (7.23)

Using this set of equations, we can analyze projectile motion, keeping in mind some important points.

Problem-Solving Strategy: Projectile Motion


1. Resolve the motion into horizontal and vertical components along the x- and y-axes. The magnitudes of the
components of displacement → s along these axes are x and y. The magnitudes of the components of velocity

v are v x = vcosθ and v y = vsinθ, where v is the magnitude of the velocity and θ is its direction relative

This OpenStax book is available for free at https://legacy.cnx.org/content/col12169/1.46


Chapter 7 | Kinematics in Two and Three Dimensions 229

to the horizontal, as shown in Figure 7.12.


2. Treat the motion as two independent one-dimensional motions: one horizontal and the other vertical. Use the
kinematic equations for horizontal and vertical motion presented earlier.
3. Solve for the unknowns in the two separate motions: one horizontal and one vertical. Note that the only
common variable between the motions is time t. The problem-solving procedures here are the same as those
for one-dimensional kinematics and are illustrated in the following solved examples.
4. Recombine quantities in the horizontal and vertical directions to find the total displacement →
s and velocity

v . Solve for the magnitude and direction of the displacement and velocity using

s = x 2 + y 2, θ = tan −1(y/x), v = v 2x + v 2y,

where θ is the direction of the displacement →


s .

Figure 7.12 (a) We analyze two-dimensional projectile motion by breaking it into two independent one-dimensional motions
along the vertical and horizontal axes. (b) The horizontal motion is simple, because a x = 0 and v x is a constant. (c) The
velocity in the vertical direction begins to decrease as the object rises. At its highest point, the vertical velocity is zero. As the
object falls toward Earth again, the vertical velocity increases again in magnitude but points in the opposite direction to the initial
vertical velocity. (d) The x and y motions are recombined to give the total velocity at any given point on the trajectory.
230 Chapter 7 | Kinematics in Two and Three Dimensions

Example 7.7

A Fireworks Projectile Explodes High and Away


During a fireworks display, a shell is shot into the air with an initial speed of 70.0 m/s at an angle of 75.0° above
the horizontal, as illustrated in Figure 7.13. The fuse is timed to ignite the shell just as it reaches its highest point
above the ground. (a) Calculate the height at which the shell explodes. (b) How much time passes between the
launch of the shell and the explosion? (c) What is the horizontal displacement of the shell when it explodes? (d)
What is the total displacement from the point of launch to the highest point?

Figure 7.13 The trajectory of a fireworks shell. The fuse is set


to explode the shell at the highest point in its trajectory, which is
found to be at a height of 233 m and 125 m away horizontally.

Strategy
The motion can be broken into horizontal and vertical motions in which a x = 0 and a y = −g. We can then
define x 0 and y 0 to be zero and solve for the desired quantities.

Solution
(a) By “height” we mean the altitude or vertical position y above the starting point. The highest point in any
trajectory, called the apex, is reached when v y = 0. Since we know the initial and final velocities, as well as the
initial position, we use the following equation to find y:
v 2y = v 20y − 2g(y − y 0).

Because y 0 and v y are both zero, the equation simplifies to

0 = v 20y − 2gy.

Solving for y gives


v 20y
y= .
2g

Now we must find v 0y, the component of the initial velocity in the y direction. It is given by v 0y = v 0 sinθ 0,
where v 0 is the initial velocity of 70.0 m/s and θ 0 = 75° is the initial angle. Thus,

v 0y = v 0 sinθ = (70.0 m/s)sin 75° = 67.6 m/s

This OpenStax book is available for free at https://legacy.cnx.org/content/col12169/1.46


Chapter 7 | Kinematics in Two and Three Dimensions 231

and y is
(67.6 m/s) 2
y= .
2(9.80 m/s 2)
Thus, we have
y = 233 m.

Note that because up is positive, the initial vertical velocity is positive, as is the maximum height, but the
acceleration resulting from gravity is negative. Note also that the maximum height depends only on the vertical
component of the initial velocity, so that any projectile with a 67.6-m/s initial vertical component of velocity
reaches a maximum height of 233 m (neglecting air resistance). The numbers in this example are reasonable for
large fireworks displays, the shells of which do reach such heights before exploding. In practice, air resistance is
not completely negligible, so the initial velocity would have to be somewhat larger than that given to reach the
same height.
(b) As in many physics problems, there is more than one way to solve for the time the projectile reaches its highest
point. In this case, the easiest method is to use v y = v 0y − gt. Because v y = 0 at the apex, this equation reduces
to simply
0 = v 0y − gt

or
v 0y
t = g = 67.6 m/s2 = 6.90s.
9.80 m/s
This time is also reasonable for large fireworks. If you are able to see the launch of fireworks, notice that several
seconds pass before the shell explodes. Another way of finding the time is by using y = y 0 + 1 (v 0y + v y)t. This
2
is left for you as an exercise to complete.
(c) Because air resistance is negligible, a x = 0 and the horizontal velocity is constant, as discussed earlier. The
horizontal displacement is the horizontal velocity multiplied by time as given by x = x 0 + v x t, where x 0 is
equal to zero. Thus,
x = v x t,
where v x is the x-component of the velocity, which is given by

v x = v 0 cosθ = (70.0 m/s)cos75° = 18.1 m/s.

Time t for both motions is the same, so x is


x = (18.1 m/s)6.90 s = 125 m.
Horizontal motion is a constant velocity in the absence of air resistance. The horizontal displacement found
here could be useful in keeping the fireworks fragments from falling on spectators. When the shell explodes, air
resistance has a major effect, and many fragments land directly below.
(d) The horizontal and vertical components of the displacement were just calculated, so all that is needed here is
to find the magnitude and direction of the displacement at the highest point:
→ ^ ^
s = 125 i + 233 j

| →s | = 125 2 + 233 2 = 264 m


⎛ ⎞
θ = tan −1 ⎝233 ⎠ = 61.8°.
125
Note that the angle for the displacement vector is less than the initial angle of launch. To see why this is, review
Figure 7.11, which shows the curvature of the trajectory toward the ground level.
232 Chapter 7 | Kinematics in Two and Three Dimensions

When solving Example 7.7(a), the expression we found for y is valid for any projectile motion when air resistance is
negligible. Call the maximum height y = h. Then,
v 20y
h= .
2g

This equation defines the maximum height of a projectile above its launch position and it depends only on the vertical
component of the initial velocity.
7.3 Check Your Understanding A rock is thrown horizontally off a cliff 100.0 m high with a velocity of
15.0 m/s. (a) Define the origin of the coordinate system. (b) Which equation describes the horizontal motion?
(c) Which equations describe the vertical motion? (d) What is the rock’s velocity at the point of impact?

Example 7.8

Calculating Projectile Motion: Tennis Player


A tennis player wins a match at Arthur Ashe stadium and hits a ball into the stands at 30 m/s and at an angle
45° above the horizontal (Figure 7.14). On its way down, the ball is caught by a spectator 10 m above the
point where the ball was hit. (a) Calculate the time it takes the tennis ball to reach the spectator. (b) What are the
magnitude and direction of the ball’s velocity at impact?

Figure 7.14 The trajectory of a tennis ball hit into the stands.

Strategy
Again, resolving this two-dimensional motion into two independent one-dimensional motions allows us to solve
for the desired quantities. The time a projectile is in the air is governed by its vertical motion alone. Thus, we solve
for t first. While the ball is rising and falling vertically, the horizontal motion continues at a constant velocity.
This example asks for the final velocity. Thus, we recombine the vertical and horizontal results to obtain → v at
final time t, determined in the first part of the example.
Solution
(a) While the ball is in the air, it rises and then falls to a final position 10.0 m higher than its starting altitude. We
can find the time for this by using Equation 7.22:

y = y 0 + v 0y t − 1 gt 2.
2
If we take the initial position y 0 to be zero, then the final position is y = 10 m. The initial vertical velocity is the
vertical component of the initial velocity:

This OpenStax book is available for free at https://legacy.cnx.org/content/col12169/1.46


Chapter 7 | Kinematics in Two and Three Dimensions 233

v 0y = v 0 sinθ 0 = (30.0 m/s)sin 45° = 21.2 m/s.

Substituting into Equation 7.22 for y gives us


10.0 m = (21.2 m/s)t − (4.90 m/s 2)t 2.
Rearranging terms gives a quadratic equation in t:
(4.90 m/s 2)t 2 − (21.2 m/s)t + 10.0 m = 0.
Use of the quadratic formula yields t = 3.79 s and t = 0.54 s. Since the ball is at a height of 10 m at two times
during its trajectory—once on the way up and once on the way down—we take the longer solution for the time it
takes the ball to reach the spectator:
t = 3.79 s.
The time for projectile motion is determined completely by the vertical motion. Thus, any projectile that has an
initial vertical velocity of 21.2 m/s and lands 10.0 m below its starting altitude spends 3.79 s in the air.
(b) We can find the final horizontal and vertical velocities v x and v y with the use of the result from (a). Then,

we can combine them to find the magnitude of the total velocity vector → v and the angle θ it makes with
the horizontal. Since v x is constant, we can solve for it at any horizontal location. We choose the starting point
because we know both the initial velocity and the initial angle. Therefore,
v x = v 0 cosθ 0 = (30 m/s)cos 45° = 21.2 m/s.

The final vertical velocity is given by Equation 7.21:


v y = v 0y − gt.

Since v 0y was found in part (a) to be 21.2 m/s, we have

v y = 21.2 m/s − 9.8 m/s 2(3.79 s) = −15.9 m/s.

The magnitude of the final velocity →


v is

v = v 2x + v 2y = (21.2 m/s) 2 + (− 15.9 m/s) 2 = 26.5 m/s.

The direction θ v is found using the inverse tangent:

⎛v y ⎞ ⎛ ⎞
θ v = tan −1⎝v ⎠ = tan −1 ⎝ 21.2 ⎠ = −53.1°.
x −15.9
Significance
(a) As mentioned earlier, the time for projectile motion is determined completely by the vertical motion. Thus,
any projectile that has an initial vertical velocity of 21.2 m/s and lands 10.0 m below its starting altitude spends
3.79 s in the air. (b) The negative angle means the velocity is 53.1° below the horizontal at the point of impact.
This result is consistent with the fact that the ball is impacting at a point on the other side of the apex of the
trajectory and therefore has a negative y component of the velocity. The magnitude of the velocity is less than the
magnitude of the initial velocity we expect since it is impacting 10.0 m above the launch elevation.

Time of Flight, Trajectory, and Range


Of interest are the time of flight, trajectory, and range for a projectile launched on a flat horizontal surface and impacting
on the same surface. In this case, kinematic equations give useful expressions for these quantities, which are derived in the
following sections.
Time of flight
We can solve for the time of flight of a projectile that is both launched and impacts on a flat horizontal surface by performing
some manipulations of the kinematic equations. We note the position and displacement in y must be zero at launch and at
impact on an even surface. Thus, we set the displacement in y equal to zero and find
234 Chapter 7 | Kinematics in Two and Three Dimensions

y − y 0 = v 0y t − 1 gt 2 = (v 0 sinθ 0)t − 1 gt 2 = 0.
2 2
Factoring, we have
⎛ gt ⎞
t⎝v 0 sinθ 0 − = 0.
2⎠
Solving for t gives us

2(v 0 sinθ 0) (7.24)


T tof = g .

This is the time of flight for a projectile both launched and impacting on a flat horizontal surface. Equation 7.24 does not
apply when the projectile lands at a different elevation than it was launched, as we saw in Example 7.8 of the tennis player
hitting the ball into the stands. The other solution, t = 0, corresponds to the time at launch. The time of flight is linearly
proportional to the initial velocity in the y direction and inversely proportional to g. Thus, on the Moon, where gravity is
one-sixth that of Earth, a projectile launched with the same velocity as on Earth would be airborne six times as long.
Trajectory
The trajectory of a projectile can be found by eliminating the time variable t from the kinematic equations for arbitrary t and
solving for y(x). We take x 0 = y 0 = 0 so the projectile is launched from the origin. The kinematic equation for x gives

x = v 0x t ⇒ t = vx = x .
0x v 0 cosθ 0

Substituting the expression for t into the equation for the position y = (v 0 sinθ 0)t − 1 gt 2 gives
2

⎛ x ⎞ 1 ⎛ x ⎞
2
y = (v 0 sinθ 0)
⎝v 0 cosθ 0 ⎠ − 2 g⎝v 0 cosθ 0 ⎠ .
Rearranging terms, we have

⎡ g ⎤ (7.25)
y = (tanθ 0)x − ⎢ ⎥x 2.
⎣2(v 0 cosθ 0) ⎦
2

This trajectory equation is of the form y = ax + bx 2, which is an equation of a parabola with coefficients
g
a = tanθ 0, b = − .
2(v 0 cosθ 0) 2
Range
From the trajectory equation we can also find the range, or the horizontal distance traveled by the projectile. Factoring
Equation 7.25, we have
⎡ g ⎤
y = x⎢tanθ 0 − x ⎥.
⎣ 2(v 0 cosθ 0) 2 ⎦

The position y is zero for both the launch point and the impact point, since we are again considering only a flat horizontal
surface. Setting y = 0 in this equation gives solutions x = 0, corresponding to the launch point, and
2v 20 sinθ 0 cosθ 0
x= g ,

corresponding to the impact point. Using the trigonometric identity 2sinθcosθ = sin2θ and setting x = R for range, we
find

This OpenStax book is available for free at https://legacy.cnx.org/content/col12169/1.46


Chapter 7 | Kinematics in Two and Three Dimensions 235

v 20 sin2θ 0 (7.26)
R= g .

Note particularly that Equation 7.26 is valid only for launch and impact on a horizontal surface. We see the range is
directly proportional to the square of the initial speed v 0 and sin2θ 0 , and it is inversely proportional to the acceleration of
gravity. Thus, on the Moon, the range would be six times greater than on Earth for the same initial velocity. Furthermore, we
see from the factor sin2θ 0 that the range is maximum at 45°. These results are shown in Figure 7.15. In (a) we see that
the greater the initial velocity, the greater the range. In (b), we see that the range is maximum at 45°. This is true only for
conditions neglecting air resistance. If air resistance is considered, the maximum angle is somewhat smaller. It is interesting
that the same range is found for two initial launch angles that sum to 90°. The projectile launched with the smaller angle
has a lower apex than the higher angle, but they both have the same range.

Figure 7.15 Trajectories of projectiles on level ground. (a) The


greater the initial speed v 0, the greater the range for a given
initial angle. (b) The effect of initial angle θ 0 on the range of a
projectile with a given initial speed. Note that the range is the same
for initial angles of 15° and 75°, although the maximum
heights of those paths are different.

Example 7.9

Comparing Golf Shots


A golfer finds himself in two different situations on different holes. On the second hole he is 120 m from the
green and wants to hit the ball 90 m and let it run onto the green. He angles the shot low to the ground at 30°
to the horizontal to let the ball roll after impact. On the fourth hole he is 90 m from the green and wants to let
236 Chapter 7 | Kinematics in Two and Three Dimensions

the ball drop with a minimum amount of rolling after impact. Here, he angles the shot at 70° to the horizontal to
minimize rolling after impact. Both shots are hit and impacted on a level surface.
(a) What is the initial speed of the ball at the second hole?
(b) What is the initial speed of the ball at the fourth hole?
(c) Write the trajectory equation for both cases.
(d) Graph the trajectories.
Strategy
We see that the range equation has the initial speed and angle, so we can solve for the initial speed for both (a)
and (b). When we have the initial speed, we can use this value to write the trajectory equation.
Solution
2
v sin2θ Rg 90.0 m(9.8 m/s 2)
(a) R = 0 g 0 ⇒ v 0 = = = 37.0 m/s
sin2θ 0 sin(2(70°))
2
v sin2θ Rg 90.0 m(9.8 m/s 2)
(b) R = 0 g 0 ⇒ v 0 = = = 31.9 m/s
sin2θ 0 sin(2(30°))

(c)
⎡ g ⎤
y = x⎢tanθ 0 − x ⎥
⎣ 2(v 0 cosθ 0) 2 ⎦
⎡ ⎤
Second hole: y = x⎢tan 70° − 9.8 m/s 2 x⎥ = 2.75x − 0.0306x 2
⎣ 2[(37.0 m/s)(cos 70°)] 2 ⎦
⎡ ⎤
Fourth hole: y = x⎢tan 30° − 9.8 m/s 2 x⎥ = 0.58x − 0.0064x 2
⎣ 2[(31.9 m/s)(cos30°)] 2 ⎦

(d) Using a graphing utility, we can compare the two trajectories, which are shown in Figure 7.16.

Figure 7.16 Two trajectories of a golf ball with a range of 90 m. The


impact points of both are at the same level as the launch point.

This OpenStax book is available for free at https://legacy.cnx.org/content/col12169/1.46


Chapter 7 | Kinematics in Two and Three Dimensions 237

Significance
The initial speed for the shot at 70° is greater than the initial speed of the shot at 30°. Note from Figure 7.16
that two projectiles launched at the same speed but at different angles have the same range if the launch angles
add to 90°. The launch angles in this example add to give a number greater than 90°. Thus, the shot at 70° has
to have a greater launch speed to reach 90 m, otherwise it would land at a shorter distance.

7.4 Check Your Understanding If the two golf shots in Example 7.9 were launched at the same speed,
which shot would have the greatest range?

When we speak of the range of a projectile on level ground, we assume R is very small compared with the circumference
of Earth. If, however, the range is large, Earth curves away below the projectile and the acceleration resulting from gravity
changes direction along the path. The range is larger than predicted by the range equation given earlier because the projectile
has farther to fall than it would on level ground, as shown in Figure 7.17, which is based on a drawing in Newton’s
Principia . If the initial speed is great enough, the projectile goes into orbit. Earth’s surface drops 5 m every 8000 m. In 1 s
an object falls 5 m without air resistance. Thus, if an object is given a horizontal velocity of 8000 m/s (or 18,000mi/hr)
near Earth’s surface, it will go into orbit around the planet because the surface continuously falls away from the object. This
is roughly the speed of the Space Shuttle in a low Earth orbit when it was operational, or any satellite in a low Earth orbit.
These and other aspects of orbital motion, such as Earth’s rotation, are covered in greater depth in Newton's Synthesis
(https://legacy.cnx.org/content/m58344/latest/) .

Figure 7.17 Projectile to satellite. In each case shown here, a


projectile is launched from a very high tower to avoid air
resistance. With increasing initial speed, the range increases and
becomes longer than it would be on level ground because Earth
curves away beneath its path. With a speed of 8000 m/s, orbit is
achieved.

At PhET Explorations: Projectile Motion (https://openstaxcollege.org/l/21phetpromot) , learn about


projectile motion in terms of the launch angle and initial velocity.
238 Chapter 7 | Kinematics in Two and Three Dimensions

7.4 | Uniform Circular Motion


Learning Objectives
By the end of this section, you will be able to:
• Solve for the centripetal acceleration of an object moving on a circular path.
• Use the equations of circular motion to find the position, velocity, and acceleration of a particle
executing circular motion.
• Explain the differences between centripetal acceleration and tangential acceleration resulting
from nonuniform circular motion.
• Evaluate centripetal and tangential acceleration in nonuniform circular motion, and find the total
acceleration vector.

Uniform Circular Motion Revisited


Uniform circular motion is a specific type of motion in which an object travels in a circle with a constant speed. For
example, any point on a propeller spinning at a constant rate is executing uniform circular motion. Other examples are
the second, minute, and hour hands of a watch. In Rotation with Constant Angular Acceleration we discussed the
tangential acceleration of such a point. But, even in the absence of any tangential acceleration, when the motion is taking
place at a constant angular speed, the points on these rotating objects are actually accelerating. To see this, we must analyze
the circular motion in terms of vectors.

Centripetal Acceleration
In one-dimensional kinematics, objects with a constant speed have zero acceleration. However, in two- and three-
dimensional kinematics, even if the speed is a constant, a particle can have acceleration if it moves along a curved trajectory
such as a circle. In this case the velocity vector is changing, or d →
v /dt ≠ 0. This is shown in Figure 7.18. As the particle
moves counterclockwise in time Δt on the circular path, its position vector moves from →
r (t) to →
r (t + Δt). The
velocity vector has constant magnitude and is tangent to the path as it changes from →
v (t) to →
v (t + Δt), changing its
v (t) is perpendicular to the position vector →
direction only. Since the velocity vector → r (t), the triangles formed by the
position vectors and Δ → v are similar. Furthermore, since →
r , and the velocity vectors and Δ → r (t) = →
r (t + Δt) | | | |
and → |
v (t) = → | | |
v (t + Δt) , the two triangles are isosceles. From these facts we can make the assertion

Δv = Δr or Δv = v Δr.
v r r

This OpenStax book is available for free at https://legacy.cnx.org/content/col12169/1.46


Chapter 7 | Kinematics in Two and Three Dimensions 239

Figure 7.18 (a) A particle is moving in a circle at a constant speed, with position and velocity vectors at times t

and t + Δt. (b) Velocity vectors forming a triangle. The two triangles in the figure are similar. The vector Δ v
points toward the center of the circle in the limit Δt → 0.

We can find the magnitude of the acceleration from


⎛ ⎞ ⎛ ⎞ 2
a = lim ⎝Δv ⎠ = vr ⎝ lim Δr ⎠ = vr .
Δt → 0 Δt Δt → 0 Δt

The direction of the acceleration can also be found by noting that as Δt and therefore Δθ approach zero, the vector Δ →
v
approaches a direction perpendicular to →
v . In the limit Δt → 0, Δ →
v is perpendicular to →
v . Since →
v is tangent
to the circle, the acceleration d →
v /dt points toward the center of the circle. Summarizing, a particle moving in a circle at
a constant speed has an acceleration with magnitude

2 (7.27)
a C = vr .

The direction of the acceleration vector is toward the center of the circle (Figure 7.19). This is a radial acceleration and
is called the centripetal acceleration, which is why we give it the subscript c. The word centripetal comes from the Latin
words centrum (meaning “center”) and petere (meaning to seek”), and thus takes the meaning “center seeking.”
240 Chapter 7 | Kinematics in Two and Three Dimensions

Figure 7.19 The centripetal acceleration vector points toward


the center of the circular path of motion and is an acceleration in
the radial direction. The velocity vector is also shown and is
tangent to the circle.

Let’s investigate some examples that illustrate the relative magnitudes of the velocity, radius, and centripetal acceleration.

Example 7.10

Creating an Acceleration of 1 g
A jet is flying at 134.1 m/s along a straight line and makes a turn along a circular path level with the ground. What
does the radius of the circle have to be to produce a centripetal acceleration of 1 g on the pilot and jet toward the
center of the circular trajectory?
Strategy
Given the speed of the jet, we can solve for the radius of the circle in the expression for the centripetal
acceleration.
Solution
Set the centripetal acceleration equal to the acceleration of gravity: 9.8 m/s 2 = v 2/r.
Solving for the radius, we find
(134.1 m/s) 2
r= = 1835 m = 1.835 km.
9.8 m/s 2
Significance
To create a greater acceleration than g on the pilot, the jet would either have to decrease the radius of its circular
trajectory or increase its speed on its existing trajectory or both.

7.5 Check Your Understanding A flywheel has a radius of 20.0 cm. What is the speed of a point on the edge
of the flywheel if it experiences a centripetal acceleration of 900.0 cm/s 2 ?

Centripetal acceleration can have a wide range of values, depending on the speed and radius of curvature of the circular
path. Typical centripetal accelerations are given in the following table.

Object Centripetal Acceleration (m/s2 or factors


of g)
Earth around the Sun 5.93 × 10 −3

Table 7.1 Typical Centripetal Accelerations

This OpenStax book is available for free at https://legacy.cnx.org/content/col12169/1.46


Chapter 7 | Kinematics in Two and Three Dimensions 241

Object Centripetal Acceleration (m/s2 or factors


of g)
Moon around the Earth 2.73 × 10 −3
Satellite in geosynchronous orbit 0.233
Outer edge of a CD when playing 5.78
Jet in a barrel roll (2–3 g)
Roller coaster (5 g)
Electron orbiting a proton in a simple Bohr model of the 9.0 × 10 22
atom

Table 7.1 Typical Centripetal Accelerations

Equations of Motion for Uniform Circular Motion


A particle executing circular motion can be described by its position vector →
r (t). Figure 7.20 shows a particle executing
circular motion in a counterclockwise direction. As the particle moves on the circle, its position vector sweeps out the angle
θ with the x-axis. Vector → r (t) making an angle θ with the x-axis is shown with its components along the x- and y-axes.
The magnitude of the position vector is A = → | |
r (t) and is also the radius of the circle, so that in terms of its components,

→ ^ ^ (7.28)
r (t) = Acosωt i + Asin ωt j .

Here, ω is a constant called the angular frequency of the particle. The angular frequency has units of radians (rad) per
second and is simply the number of radians of angular measure through which the particle passes per second. The angle θ
that the position vector has at any particular time is ωt .
If T is the period of motion, or the time to complete one revolution ( 2π rad), then

ω = 2π .
T

Figure 7.20 The position vector for a particle in circular


motion with its components along the x- and y-axes. The particle
moves counterclockwise. Angle θ is the angular frequency ω
in radians per second multiplied by t.

Velocity and acceleration can be obtained from the position function by differentiation:
242 Chapter 7 | Kinematics in Two and Three Dimensions

→ d→
r (t) ^ ^ (7.29)
v (t) = = −Aωsin ωt i + Aωcos ωt j .
dt

It can be shown from Figure 7.20 that the velocity vector is tangential to the circle at the location of the particle, with
magnitude Aω. Similarly, the acceleration vector is found by differentiating the velocity:

→ d→
v (t) ^ ^ (7.30)
a (t) = = −Aω 2 cos ωt i − Aω 2 sin ωt j .
dt

From this equation we see that the acceleration vector has magnitude Aω 2 and is directed opposite the position vector,
toward the origin, because →
a (t) = −ω 2 →
r (t).

Example 7.11

Circular Motion of a Proton


A proton has speed 5 × 10 6 m/s and is moving in a circle in the xy plane of radius r = 0.175 m. What is its
^
position in the xy plane at time t = 2.0 × 10 −7 s = 200 ns? At t = 0, the position of the proton is 0.175 m i
and it circles counterclockwise. Sketch the trajectory.
Solution
From the given data, the proton has period and angular frequency:
2π(0.175 m)
T = 2πr
v = = 2.20 × 10 −7 s
5.0 × 10 6 m/s
ω = 2π = 2π = 2.856 × 10 7 rad/s.
T 2.20 × 10 −7 s

The position of the particle at t = 2.0 × 10 −7 s with A = 0.175 m is

→ ^ ^
r (2.0 × 10 −7 s) = A cos ω(2.0 × 10 −7 s) i + A sin ω(2.0 × 10 −7 s) j m
^
= 0.175cos[(2.856 × 10 7 rad/s)(2.0 × 10 −7 s)] i
^
+0.175sin[(2.856 × 10 7 rad/s)(2.0 × 10 −7 s)] j m
^ ^ ^ ^
= 0.175cos(5.712 rad) i + 0.175sin(5.712 rad) j = 0.147 i − 0.095 j m.

From this result we see that the proton is located slightly below the x-axis. This is shown in Figure 7.21.

This OpenStax book is available for free at https://legacy.cnx.org/content/col12169/1.46


Chapter 7 | Kinematics in Two and Three Dimensions 243

Figure 7.21 Position vector of the proton at


t = 2.0 × 10 −7 s = 200 ns. The trajectory of the proton is
shown. The angle through which the proton travels along the
circle is 5.712 rad, which a little less than one complete
revolution.

Significance
We picked the initial position of the particle to be on the x-axis. This was completely arbitrary. If a different
starting position were given, we would have a different final position at t = 200 ns.

Nonuniform Circular Motion


Circular motion does not have to be at a constant speed. A particle can travel in a circle and speed up or slow down, showing
an acceleration in the direction of the motion.
In uniform circular motion, the particle executing circular motion has a constant speed and the circle is at a fixed radius. If
the speed of the particle is changing as well, then we introduce an additional acceleration in the direction tangential to the
circle. Such accelerations occur at a point on a top that is changing its spin rate, or any accelerating rotor. In Section 7.1
we showed that centripetal acceleration is the time rate of change of the direction of the velocity vector. If the speed of the
particle is changing, then it has a tangential acceleration that is the time rate of change of the magnitude of the velocity:

aT = | |
d →v
.
(7.31)
dt

The direction of tangential acceleration is tangent to the circle whereas the direction of centripetal acceleration is radially
inward toward the center of the circle. Thus, a particle in circular motion with a tangential acceleration has a total
acceleration that is the vector sum of the centripetal and tangential accelerations:


a = →
a → (7.32)
C+ a T.
244 Chapter 7 | Kinematics in Two and Three Dimensions

The acceleration vectors are shown in Figure 7.22. Note that the two acceleration vectors →
a C and →
a T are
perpendicular to each other, with →
a C in the radial direction and →
a T in the tangential direction. The total acceleration

a points at an angle between →
a →
a T.
C and

Figure 7.22 The centripetal acceleration points toward the


center of the circle. The tangential acceleration is tangential to
the circle at the particle’s position. The total acceleration is the
vector sum of the tangential and centripetal accelerations, which
are perpendicular.

Example 7.12

Total Acceleration during Circular Motion


A particle moves in a circle of radius r = 2.0 m. During the time interval from t = 1.5 s to t = 4.0 s its speed varies
with time according to
c2
v(t) = c 1 − , c 1 = 4.0 m/s, c 2 = 6.0 m · s.
t2
What is the total acceleration of the particle at t = 2.0 s?
Strategy
We are given the speed of the particle and the radius of the circle, so we can calculate centripetal acceleration
easily. The direction of the centripetal acceleration is toward the center of the circle. We find the magnitude of the
tangential acceleration by taking the derivative with respect to time of |v(t)| using Equation 7.31 and evaluating
it at t = 2.0 s. We use this and the magnitude of the centripetal acceleration to find the total acceleration.
Solution
Centripetal acceleration is
⎛ ⎞
v(2.0s) = ⎜4.0 − 6.0 2 ⎟m/s = 2.5 m/s
⎝ (2.0) ⎠
2 (2.5 m/s) 2
a C = vr = = 3.1 m/s 2
2.0 m
directed toward the center of the circle. Tangential acceleration is

| |
→ 2c
a T = d v = 32 = 12.03 m/s 2 = 1.5 m/s 2.
dt t (2.0)
Total acceleration is

| →a | = 3.1 2 + 1.5 2m/s 2 = 3.44 m/s 2

This OpenStax book is available for free at https://legacy.cnx.org/content/col12169/1.46


Chapter 7 | Kinematics in Two and Three Dimensions 245

and θ = tan −1 3.1 = 64° from the tangent to the circle. See Figure 7.23.
1.5

Figure 7.23 The tangential and centripetal acceleration


vectors. The net acceleration

a is the vector sum of the two
accelerations.

Significance
The directions of centripetal and tangential accelerations can be described more conveniently in terms of a polar
coordinate system, with unit vectors in the radial and tangential directions. This coordinate system, which is used
for motion along curved paths, is discussed in detail later in the book.

7.5 | Relative Motion in One and Two Dimensions


Learning Objectives
By the end of this section, you will be able to:
• Explain the concept of reference frames.
• Write the position and velocity vector equations for relative motion.
• Draw the position and velocity vectors for relative motion.
• Analyze one-dimensional and two-dimensional relative motion problems using the position and
velocity vector equations.

We have already discussed (in Section 5.6) how velocities transform non-intuitively in the world of relativity. The
discussion in this chapter will assume that the velocities of all the objects involved are much, much less than the speed
of light, so that the relativistic corrections may be ignored. However, from the extensive discussion in the chapter on
Relativistic Kinematics in One Dimension, we already understand the importance in kinematics of asking the
question, "Who is the observer?"
Motion Relative to What?
Motion does not happen in isolation. If you’re riding in a train moving at 10 m/s east, this velocity is measured relative to
the ground on which you’re traveling. However, if another train passes you at 15 m/s east, your velocity relative to this other
train is different from your velocity relative to the ground. Your velocity relative to the other train is 5 m/s west. To explore
this idea further, we first need to establish some terminology.

Reference Frames
To discuss relative motion in one or more dimensions, we return to the concept of reference frames. When we say an object
has a certain velocity, we must state it has a velocity with respect to a given reference frame. In most examples we have
examined so far, this reference frame has been Earth. If you say a person is sitting in a train moving at 10 m/s east, then you
246 Chapter 7 | Kinematics in Two and Three Dimensions

imply the person on the train is moving relative to the surface of Earth at this velocity, and Earth is the reference frame. We
can expand our view of the motion of the person on the train and say Earth is spinning in its orbit around the Sun, in which
case the motion becomes more complicated. In this case, the solar system is the reference frame. In summary, all discussion
of relative motion must define the reference frames involved. We now develop a method to refer to reference frames in
relative motion.

Relative Motion in One Dimension


We introduce relative motion in one dimension first, because the velocity vectors simplify to having only two possible
directions. Take the example of the person sitting in a train moving east. If we choose east as the positive direction and
^
Earth as the reference frame, then we can write the velocity of the train with respect to the Earth as → v TE = 10 m/s i
east, where the subscripts TE refer to train and Earth. Let’s now say the person gets up out of /her seat and walks toward
the back of the train at 2 m/s. This tells us she has a velocity relative to the reference frame of the train. Since the person is
^
walking west, in the negative direction, we write her velocity with respect to the train as → v PT = −2 m/s i . We can add
the two velocity vectors to find the velocity of the person with respect to Earth. This relative velocity is written as

v = →
v → (7.33)
PE PT + v TE.

Note the ordering of the subscripts for the various reference frames in Equation 7.33. The subscripts for the coupling
reference frame, which is the train, appear consecutively in the right-hand side of the equation. Figure 7.24 shows the
correct order of subscripts when forming the vector equation.

Figure 7.24 When constructing the vector equation, the


subscripts for the coupling reference frame appear consecutively
on the inside. The subscripts on the left-hand side of the
equation are the same as the two outside subscripts on the right-
hand side of the equation.

^
Adding the vectors, we find →
v PE = 8 m/s i , so the person is moving 8 m/s east with respect to Earth. Graphically, this
is shown in Figure 7.25.

Figure 7.25 Velocity vectors of the train with respect to Earth, person with respect to the
train, and person with respect to Earth.

Relative Velocity in Two Dimensions


We can now apply these concepts to describing motion in two dimensions. Consider a particle P and reference frames S and
S′, as shown in Figure 7.26. The position of the origin of S′ as measured in S is →
r S′ S, the position of P as measured
in S′ is →
r PS′, and the position of P as measured in S is →
r PS.

This OpenStax book is available for free at https://legacy.cnx.org/content/col12169/1.46


Chapter 7 | Kinematics in Two and Three Dimensions 247

Figure 7.26 The positions of particle P relative to frames S


and S′ are

r →
r
PS and PS′, respectively.

From Figure 7.26 we see that


r = →
r → (7.34)
PS PS′ + r S′ S.

The relative velocities are the time derivatives of the position vectors. Therefore,


v = →
v → (7.35)
PS PS′ + v S′ S.

The velocity of a particle relative to S is equal to its velocity relative to S′ plus the velocity of S′ relative to S.
We can extend Equation 7.35 to any number of reference frames. For particle P with velocities

v PA, →
v PB , and →
v PC in frames A, B, and C,


v = →
v → → (7.36)
PC PA + v AB + v BC.

We can also see how the accelerations are related as observed in two reference frames by differentiating Equation 7.35:


a = →
a → (7.37)
PS PS′ + a S′ S.

We see that if the velocity of S′ relative to S is a constant, then →


a S′ S = 0 and

a = →
a (7.38)
PS PS′.

This says the acceleration of a particle is the same as measured by two observers moving at a constant velocity relative to
each other.
248 Chapter 7 | Kinematics in Two and Three Dimensions

Example 7.13

Motion of a Car Relative to a Truck


A truck is traveling south at a speed of 70 km/h toward an intersection. A car is traveling east toward the
intersection at a speed of 80 km/h (Figure 7.27). What is the velocity of the car relative to the truck?

Figure 7.27 A car travels east toward an intersection while a


truck travels south toward the same intersection.

Strategy
First, we must establish the reference frame common to both vehicles, which is Earth. Then, we write the
velocities of each with respect to the reference frame of Earth, which enables us to form a vector equation that
links the car, the truck, and Earth to solve for the velocity of the car with respect to the truck.
Solution
^
The velocity of the car with respect to Earth is →
v CE = 80 km/h i . The velocity of the truck with respect to
^
Earth is →
v TE = −70 km/h j . Using the velocity addition rule, the relative motion equation we are seeking is

v = →
v →
CT CE + v ET.

Here, → v CT is the velocity of the car with respect to the truck, and Earth is the connecting reference frame.
Since we have the velocity of the truck with respect to Earth, the negative of this vector is the velocity of Earth
with respect to the truck: →
v ET = − → v TE. The vector diagram of this equation is shown in Figure 7.28.

This OpenStax book is available for free at https://legacy.cnx.org/content/col12169/1.46


Chapter 7 | Kinematics in Two and Three Dimensions 249

Figure 7.28 Vector diagram of the vector equation


→v CT = → v CE + → v ET .

We can now solve for the velocity of the car with respect to the truck:

| →v CT| = (80.0 km/h) 2 + (70.0 km/h) 2 = 106. km/h

and
⎛ ⎞
θ = tan −1 ⎝70.0 ⎠ = 41.2° north of east.
80.0
Significance
Drawing a vector diagram showing the velocity vectors can help in understanding the relative velocity of the two
objects.

7.6 Check Your Understanding A boat heads north in still water at 4.5 m/s directly across a river that is
running east at 3.0 m/s. What is the velocity of the boat with respect to Earth?

Example 7.14

Flying a Plane in a Wind


A pilot must fly his plane due north to reach his destination. The plane can fly at 300 km/h in still air. A wind
is blowing out of the northeast at 90 km/h. (a) What is the speed of the plane relative to the ground? (b) In what
direction must the pilot head her plane to fly due north?
Strategy
The pilot must point her plane somewhat east of north to compensate for the wind velocity. We need to construct
a vector equation that contains the velocity of the plane with respect to the ground, the velocity of the plane
with respect to the air, and the velocity of the air with respect to the ground. Since these last two quantities are
known, we can solve for the velocity of the plane with respect to the ground. We can graph the vectors and use
this diagram to evaluate the magnitude of the plane’s velocity with respect to the ground. The diagram will also
tell us the angle the plane’s velocity makes with north with respect to the air, which is the direction the pilot must
head her plane.
Solution
The vector equation is → v PG = → v PA + → v AG, where P = plane, A = air, and G = ground. From the
geometry in Figure 7.29, we can solve easily for the magnitude of the velocity of the plane with respect to the
ground and the angle of the plane’s heading, θ.
250 Chapter 7 | Kinematics in Two and Three Dimensions

Figure 7.29 Vector diagram for Equation 7.34 showing the


→ → →
vectors v PA , v AG, and v PG.

(a) Known quantities:

| →v PA| = 300 km/h


| →v AG| = 90 km/h
Substituting into the equation of motion, we obtain | →
v PG| = 230 km/h.

(b) The angle θ = tan −1 63.64 = 12° east of north.


300

This OpenStax book is available for free at https://legacy.cnx.org/content/col12169/1.46


Chapter 7 | Kinematics in Two and Three Dimensions 251

CHAPTER 7 REVIEW
KEY TERMS
acceleration vector instantaneous acceleration found by taking the derivative of the velocity function with respect to
time in unit vector notation
angular frequency ω, rate of change of an angle with which an object that is moving on a circular path

centripetal acceleration component of acceleration of an object moving in a circle that is directed radially inward
toward the center of the circle
displacement vector vector from the initial position to a final position on a trajectory of a particle
position vector vector from the origin of a chosen coordinate system to the position of a particle in two- or three-
dimensional space
projectile motion motion of an object subject only to the acceleration of gravity
range maximum horizontal distance a projectile travels
reference frame coordinate system in which the position, velocity, and acceleration of an object at rest or moving is
measured
relative velocity velocity of an object as observed from a particular reference frame, or the velocity of one reference
frame with respect to another reference frame
tangential acceleration magnitude of which is the time rate of change of speed. Its direction is tangent to the circle.
time of flight elapsed time a projectile is in the air
total acceleration vector sum of centripetal and tangential accelerations
trajectory path of a projectile through the air
velocity vector vector that gives the instantaneous speed and direction of a particle; tangent to the trajectory

KEY EQUATIONS
→ ^ ^ ^
Position vector r (t) = x(t) i + y(t) j + z(t) k

Displacement vector Δ→
r = →
r (t 2) − →
r (t 1)

r (t + Δt) − →
r (t) d →

v (t) = lim = r
Velocity vector
Δt → 0 Δt dt

→ ^ ^ ^
Velocity in terms of components v (t) = v x(t) i + v y(t) j + v z(t) k

dx(t) dy(t) dz(t)


Velocity components v x(t) = v y(t) = v z(t) =
dt dt dt

r (t 2) − → r (t 1)
Average velocity

v =
avg t2 − t1

v (t + Δt) − →
v (t) d →
v (t)

a (t) = lim =
Instantaneous acceleration
t→0 Δt dt

→ dv (t) ^ dv y(t) ^ dv z(t) ^


Instantaneous acceleration, component form a (t) = x i + j + k
dt dt dt

Instantaneous acceleration as second → d 2 x(t) ^ d 2 y(t) ^ d 2 z(t) ^


a (t) = i + j + k
derivatives of position dt 2 dt 2 dt 2
252 Chapter 7 | Kinematics in Two and Three Dimensions

2(v 0 sinθ)
Time of flight T tof = g

⎡ g ⎤
Trajectory y = (tanθ 0)x − ⎢ ⎥x 2
⎣2(v 0 cosθ 0) 2 ⎦

v 20 sin 2θ 0
Range R= g
2
Centripetal acceleration a C = vr

→ ^ ^
Position vector, uniform circular motion r (t) = Acos ωt i + A sin ωt j

→ d→
r (t) ^ ^
Velocity vector, uniform circular motion v (t) = = −Aω sin ωt i + Aω cos ωt j
dt

→ d→
v (t) ^ ^
Acceleration vector, uniform circular motion a (t) = = −Aω 2 cos ωt i − Aω 2 sin ωt j
dt

Tangential acceleration aT =
d →v | |
dt

a = →
a →
Total acceleration C+ a T

Position vector in frame


S is the position →
r = →
r →
vector in frame S′ plus the vector from the PS PS′ + r S′ S
origin of S to the origin of S′

Relative velocity equation connecting two →


v = →
v →
PS PS′ + v S′ S
reference frames
Relative velocity equation connecting more →
v = →
v → →
PC PA + v AB + v BC
than two reference frames

a = →
a →
Relative acceleration equation PS PS′ + a S′ S

SUMMARY
7.1 Displacement and Velocity Vectors
• The position function →
r (t) gives the position as a function of time of a particle moving in two or three
dimensions. Graphically, it is a vector from the origin of a chosen coordinate system to the point where the particle
is located at a specific time.
• The displacement vector Δ →
r gives the shortest distance between any two points on the trajectory of a particle in
two or three dimensions.
• Instantaneous velocity gives the speed and direction of a particle at a specific time on its trajectory in two or three
dimensions, and is a vector in two and three dimensions.
• The velocity vector is tangent to the trajectory of the particle.
• Displacement →
r (t) can be written as a vector sum of the one-dimensional displacements →
x (t), →
y (t), →
z (t)
along the x, y, and z directions.
• Velocity →
v (t) can be written as a vector sum of the one-dimensional velocities v x(t), v y(t), v z(t) along the x, y,
and z directions.

This OpenStax book is available for free at https://legacy.cnx.org/content/col12169/1.46


Chapter 7 | Kinematics in Two and Three Dimensions 253

• Motion in any given direction is independent of motion in a perpendicular direction.

7.2 Acceleration Vector


• In two and three dimensions, the acceleration vector can have an arbitrary direction and does not necessarily point
along a given component of the velocity.
• The instantaneous acceleration is produced by a change in velocity taken over a very short (infinitesimal) time
period. Instantaneous acceleration is a vector in two or three dimensions. It is found by taking the derivative of the
velocity function with respect to time.
• In three dimensions, acceleration →
a (t) can be written as a vector sum of the one-dimensional accelerations
a x(t), a y(t), and a z(t) along the x-, y-, and z-axes.

• The kinematic equations for constant acceleration can be written as the vector sum of the constant acceleration
equations in the x, y, and z directions.

7.3 Projectile Motion


• Projectile motion is the motion of an object subject only to the acceleration of gravity, where the acceleration is
constant, as near the surface of Earth.
• To solve projectile motion problems, we analyze the motion of the projectile in the horizontal and vertical directions
using the one-dimensional kinematic equations for x and y.
• The time of flight of a projectile launched with initial vertical velocity v 0y on an even surface is given by

2(v 0 sinθ)
T to f = g .

This equation is valid only when the projectile lands at the same elevation from which it was launched.
• The maximum horizontal distance traveled by a projectile is called the range. Again, the equation for range is valid
only when the projectile lands at the same elevation from which it was launched.

7.4 Uniform Circular Motion


• Uniform circular motion is motion in a circle at constant speed.
• Centripetal acceleration →
a C is the acceleration a particle must have to follow a circular path. Centripetal
acceleration always points toward the center of rotation and has magnitude a C = v 2/r.

• Nonuniform circular motion occurs when there is tangential acceleration of an object executing circular motion such
that the speed of the object is changing. This acceleration is called tangential acceleration →
a T. The magnitude of
tangential acceleration is the time rate of change of the magnitude of the velocity. The tangential acceleration vector
is tangential to the circle, whereas the centripetal acceleration vector points radially inward toward the center of the
circle. The total acceleration is the vector sum of tangential and centripetal accelerations.
• An object executing uniform circular motion can be described with equations of motion. The position vector of the
^ ^
object is →
r (t) = A cos ωt i + A sin ωt j , where A is the magnitude | →r (t)|, which is also the radius of the
circle, and ω is the angular frequency.

7.5 Relative Motion in One and Two Dimensions


• When analyzing motion of an object, the reference frame in terms of position, velocity, and acceleration needs to be
specified.
• Relative velocity is the velocity of an object as observed from a particular reference frame, and it varies with the
choice of reference frame.
254 Chapter 7 | Kinematics in Two and Three Dimensions

• If S and S′ are two reference frames moving relative to each other at a constant velocity, then the velocity of an
object relative to S is equal to its velocity relative to S′ plus the velocity of S′ relative to S.
• If two reference frames are moving relative to each other at a constant velocity, then the accelerations of an object
as observed in both reference frames are equal.

CONCEPTUAL QUESTIONS
9. A dime is placed at the edge of a table so it hangs over
7.1 Displacement and Velocity Vectors slightly. A quarter is slid horizontally on the table surface
1. What form does the trajectory of a particle have if perpendicular to the edge and hits the dime head on. Which
the distance from any point A to point B is equal to the coin hits the ground first?
magnitude of the displacement from A to B?

2. Give an example of a trajectory in two or three 7.4 Uniform Circular Motion


dimensions caused by independent perpendicular motions. 10. Can centripetal acceleration change the speed of a
particle undergoing circular motion?
3. If the instantaneous velocity is zero, what can be said
about the slope of the position function? 11. Can tangential acceleration change the speed of a
particle undergoing circular motion?

7.2 Acceleration Vector


4. If the position function of a particle is a linear function 7.5 Relative Motion in One and Two
of time, what can be said about its acceleration? Dimensions
12. What frame or frames of reference do you use
5. If an object has a constant x-component of the velocity instinctively when driving a car? When flying in a
and suddenly experiences an acceleration in the y direction, commercial jet?
does the x-component of its velocity change?
13. A basketball player dribbling down the court usually
6. If an object has a constant x-component of velocity and keeps his eyes fixed on the players around him. He is
suddenly experiences an acceleration at an angle of 70° in moving fast. Why doesn’t he need to keep his eyes on the
the x direction, does the x-component of velocity change? ball?

14. If someone is riding in the back of a pickup truck and


7.3 Projectile Motion throws a softball straight backward, is it possible for the
7. Answer the following questions for projectile motion ball to fall straight down as viewed by a person standing
on level ground assuming negligible air resistance, with at the side of the road? Under what condition would this
the initial angle being neither 0° nor 90° : (a) Is the occur? How would the motion of the ball appear to the
person who threw it?
velocity ever zero? (b) When is the velocity a minimum?
A maximum? (c) Can the velocity ever be the same as the
initial velocity at a time other than at t = 0? (d) Can the 15. The hat of a jogger running at constant velocity falls
speed ever be the same as the initial speed at a time other off the back of his head. Draw a sketch showing the path of
than at t = 0? the hat in the jogger’s frame of reference. Draw its path as
viewed by a stationary observer. Neglect air resistance.
8. Answer the following questions for projectile motion
on level ground assuming negligible air resistance, with 16. A clod of dirt falls from the bed of a moving truck. It
the initial angle being neither 0° nor 90° : (a) Is the strikes the ground directly below the end of the truck. (a)
What is the direction of its velocity relative to the truck just
acceleration ever zero? (b) Is the acceleration ever in the
before it hits? (b) Is this the same as the direction of its
same direction as a component of velocity? (c) Is the
velocity relative to ground just before it hits? Explain your
acceleration ever opposite in direction to a component of
answers.
velocity?

This OpenStax book is available for free at https://legacy.cnx.org/content/col12169/1.46


Chapter 7 | Kinematics in Two and Three Dimensions 255

PROBLEMS
23. The position of a particle is
7.1 Displacement and Velocity Vectors → ^ ^ ^
r (t) = 4.0t 2 i − 3.0 j + 2.0t 3 k m. (a) What is the
17. The coordinates of a particle in a rectangular
coordinate system are (1.0, –4.0, 6.0). What is the position velocity of the particle at 0 s and at 1.0 s? (b) What is the
vector of the particle? average velocity between 0 s and 1.0 s?

18. The position of a particle changes from 24. Clay Matthews, a linebacker for the Green Bay
→ ^ ^ Packers, can reach a speed of 10.0 m/s. At the start of a
r 1 = (2.0 i + 3.0 j )cm to play, Matthews runs downfield at 45° with respect to the
→ ^ ^ 50-yard line and covers 8.0 m in 1 s. He then runs straight
r 2 = (−4.0 i + 3.0 j ) cm. What is the particle’s
down the field at 90° with respect to the 50-yard line for
displacement? 12 m, with an elapsed time of 1.2 s. (a) What is Matthews’
final displacement from the start of the play? (b) What is
19. The 18th hole at Pebble Beach Golf Course is a dogleg his average velocity?
to the left of length 496.0 m. The fairway off the tee is
taken to be the x direction. A golfer hits his tee shot a 25. The F-35B Lighting II is a short-takeoff and vertical
distance of 300.0 m, corresponding to a displacement landing fighter jet. If it does a vertical takeoff to 20.00-m
^
Δ→
r 1 = 300.0 m i , and hits his second shot 189.0 m height above the ground and then follows a flight path
angled at 30° with respect to the ground for 20.00 km,
^ ^
with a displacement Δ → r 2 = 172.0 m i + 80.3 m j . what is the final displacement?
What is the final displacement of the golf ball from the tee?

7.2 Acceleration Vector


20. A bird flies straight northeast a distance of 95.0 km
for 3.0 h. With the x-axis due east and the y-axis due north, 26. The position of a particle is
what is the displacement in unit vector notation for the → ^ ^ ^
r (t) = (3.0t 2 i + 5.0 j − 6.0t k ) m. (a) Determine its
bird? What is the average velocity for the trip?
velocity and acceleration as functions of time. (b) What are
its velocity and acceleration at time t = 0?
21. A cyclist rides 5.0 km due east, then 10.0 km 20°
west of north. From this point she rides 8.0 km due west.
^ ^
What is the final displacement from where the cyclist 27. A particle’s acceleration is (4.0 i + 3.0 j )m/s 2. At
started?
t = 0, its position and velocity are zero. (a) What are the
particle’s position and velocity as functions of time? (b)
22. New York Rangers defenseman Daniel Girardi stands Find the equation of the path of the particle. Draw the x-
at the goal and passes a hockey puck 20 m and 45° from and y-axes and sketch the trajectory of the particle.
straight down the ice to left wing Chris Kreider waiting at
the blue line. Kreider waits for Girardi to reach the blue 28. A boat leaves the dock at t = 0 and heads out into
line and passes the puck directly across the ice to him 10 m ^
away. What is the final displacement of the puck? See the a lake with an acceleration of 2.0 m/s 2 i . A strong wind
following figure. is pushing the boat, giving it an additional velocity of
^ ^
2.0 m/s i + 1.0 m/s j . (a) What is the velocity of the boat
at t = 10 s? (b) What is the position of the boat at t = 10s?
Draw a sketch of the boat’s trajectory and position at t = 10
s, showing the x- and y-axes.

29. The position of a particle for t > 0 is given by


→ ^ ^ ^
r (t) = (3.0t 2 i − 7.0t 3 j − 5.0t −2 k ) m. (a) What is
the velocity as a function of time? (b) What is the
acceleration as a function of time? (c) What is the particle’s
velocity at t = 2.0 s? (d) What is its speed at t = 1.0 s and t
= 3.0 s? (e) What is the average velocity between t = 1.0 s
and t = 2.0 s?
256 Chapter 7 | Kinematics in Two and Three Dimensions

30. The acceleration of a particle is a constant. At t =


^ ^
0 the velocity of the particle is (10 i + 20 j )m/s. At t
^
= 4 s the velocity is 10 j m/s. (a) What is the particle’s
acceleration? (b) How do the position and velocity vary
with time? Assume the particle is initially at the origin.

31. A particle has a position function


→ ^ ^ ^
r (t) = cos(1.0t) i + sin(1.0t) j + t k , where the
arguments of the cosine and sine functions are in radians.
(a) What is the velocity vector? (b) What is the acceleration
vector?
37. Suppose the airplane in the preceding problem fires a
32. A Lockheed Martin F-35 II Lighting jet takes off from projectile horizontally in its direction of motion at a speed
an aircraft carrier with a runway length of 90 m and a of 300 m/s relative to the plane. (a) How far in front of the
takeoff speed 70 m/s at the end of the runway. Jets are release point does the projectile hit the ground? (b) What is
catapulted into airspace from the deck of an aircraft carrier its speed when it hits the ground?
with two sources of propulsion: the jet propulsion and the
catapult. At the point of leaving the deck of the aircraft 38. A fastball pitcher can throw a baseball at a speed of 40
carrier, the F-35’s acceleration decreases to a constant m/s (90 mi/h). (a) Assuming the pitcher can release the ball
acceleration of 5.0 m/s 2 at 30° with respect to the 16.7 m from home plate so the ball is moving horizontally,
horizontal. (a) What is the initial acceleration of the F-35 on how long does it take the ball to reach home plate? (b) How
the deck of the aircraft carrier to make it airborne? (b) Write far does the ball drop between the pitcher’s hand and home
the position and velocity of the F-35 in unit vector notation plate?
from the point it leaves the deck of the aircraft carrier. (c)
At what altitude is the fighter 5.0 s after it leaves the deck 39. A projectile is launched at an angle of 30° and lands
of the aircraft carrier? (d) What is its velocity and speed at 20 s later at the same height as it was launched. (a) What is
this time? (e) How far has it traveled horizontally? the initial speed of the projectile? (b) What is the maximum
altitude? (c) What is the range? (d) Calculate the
displacement from the point of launch to the position on its
7.3 Projectile Motion trajectory at 15 s.
33. A bullet is shot horizontally from shoulder height (1.5
m) with and initial speed 200 m/s. (a) How much time 40. A basketball player shoots toward a basket 6.1 m away
elapses before the bullet hits the ground? (b) How far does and 3.0 m above the floor. If the ball is released 1.8 m above
the bullet travel horizontally? the floor at an angle of 60° above the horizontal, what
must the initial speed be if it were to go through the basket?
34. A marble rolls off a tabletop 1.0 m high and hits the
floor at a point 3.0 m away from the table’s edge in the 41. At a particular instant, a hot air balloon is 100 m in
horizontal direction. (a) How long is the marble in the air? the air and descending at a constant speed of 2.0 m/s. At
(b) What is the speed of the marble when it leaves the this exact instant, a girl throws a ball horizontally, relative
table’s edge? (c) What is its speed when it hits the floor? to herself, with an initial speed of 20 m/s. When she lands,
where will she find the ball? Ignore air resistance.
35. A dart is thrown horizontally at a speed of 10 m/
s at the bull’s-eye of a dartboard 2.4 m away, as in the 42. A man on a motorcycle traveling at a uniform speed
following figure. (a) How far below the intended target of 10 m/s throws an empty can straight upward relative to
does the dart hit? (b) What does your answer tell you about himself with an initial speed of 3.0 m/s. Find the equation
how proficient dart players throw their darts? of the trajectory as seen by a police officer on the side of
the road. Assume the initial position of the can is the point
36. An airplane flying horizontally with a speed of 500 where it is thrown. Ignore air resistance.
km/h at a height of 800 m drops a crate of supplies (see the
following figure). If the parachute fails to open, how far in 43. An athlete can jump a distance of 8.0 m in the broad
front of the release point does the crate hit the ground? jump. What is the maximum distance the athlete can jump
on the Moon, where the gravitational acceleration is one-
sixth that of Earth?

This OpenStax book is available for free at https://legacy.cnx.org/content/col12169/1.46


Chapter 7 | Kinematics in Two and Three Dimensions 257

44. The maximum horizontal distance a boy can throw a


ball is 50 m. Assume he can throw with the same initial
speed at all angles. How high does he throw the ball when
he throws it straight upward?

45. A rock is thrown off a cliff at an angle of 53° with


respect to the horizontal. The cliff is 100 m high. The initial
speed of the rock is 30 m/s. (a) How high above the edge
of the cliff does the rock rise? (b) How far has it moved
horizontally when it is at maximum altitude? (c) How long
after the release does it hit the ground? (d) What is the
49. An astronaut on Mars kicks a soccer ball at an angle of
range of the rock? (e) What are the horizontal and vertical
positions of the rock relative to the edge of the cliff at t =
45° with an initial velocity of 15 m/s. If the acceleration
2.0 s, t = 4.0 s, and t = 6.0 s? of gravity on Mars is 3.7 m/s, (a) what is the range of the
soccer kick on a flat surface? (b) What would be the range
of the same kick on the Moon, where gravity is one-sixth
46. Trying to escape his pursuers, a secret agent skis off a
that of Earth?
slope inclined at 30° below the horizontal at 60 km/h. To
survive and land on the snow 100 m below, he must clear a
50. Mike Powell holds the record for the long jump of 8.95
gorge 60 m wide. Does he make it? Ignore air resistance.
m, established in 1991. If he left the ground at an angle of
15°, what was his initial speed?

51. MIT’s robot cheetah can jump over obstacles 46 cm


high and has speed of 12.0 km/h. (a) If the robot launches
itself at an angle of 60° at this speed, what is its maximum
height? (b) What would the launch angle have to be to reach
a height of 46 cm?

52. Mt. Asama, Japan, is an active volcano. In 2009,


an eruption threw solid volcanic rocks that landed 1 km
horizontally from the crater. If the volcanic rocks were
launched at an angle of 40° with respect to the horizontal
and landed 900 m below the crater, (a) what would be their
initial velocity and (b) what is their time of flight?

53. Drew Brees of the New Orleans Saints can throw a


football 23.0 m/s (50 mph). If he angles the throw at 10°
from the horizontal, what distance does it go if it is to be
47. A golfer on a fairway is 70 m away from the green, caught at the same elevation as it was thrown?
which sits below the level of the fairway by 20 m. If the
golfer hits the ball at an angle of 40° with an initial speed
54. The Lunar Roving Vehicle used in NASA’s late Apollo
of 20 m/s, how close to the green does she come? missions reached an unofficial lunar land speed of 5.0 m/
s by astronaut Eugene Cernan. If the rover was moving at
48. A projectile is shot at a hill, the base of which is 300 this speed on a flat lunar surface and hit a small bump that
m away. The projectile is shot at 60° above the horizontal projected it off the surface at an angle of 20°, how long
with an initial speed of 75 m/s. The hill can be would it be “airborne” on the Moon?
approximated by a plane sloped at 20° to the horizontal.
Relative to the coordinate system shown in the following 55. A soccer goal is 2.44 m high. A player kicks the ball
figure, the equation of this straight line is at a distance 10 m from the goal at an angle of 25°. What
y = (tan20°)x − 109. Where on the hill does the
is the initial speed of the soccer ball?
projectile land?
56. Olympus Mons on Mars is the largest volcano in the
solar system, at a height of 25 km and with a radius of 312
km. If you are standing on the summit, with what initial
velocity would you have to fire a projectile from a cannon
horizontally to clear the volcano and land on the surface
258 Chapter 7 | Kinematics in Two and Three Dimensions

of Mars? Note that Mars has an acceleration of gravity of its equator just above its surface. At what speed must the jet
2 travel if the magnitude of its acceleration is g?
3.7 m/s .

57. In 1999, Robbie Knievel was the first to jump the 67. A fan is rotating at a constant 360.0 rev/min. What is
Grand Canyon on a motorcycle. At a narrow part of the the magnitude of the acceleration of a point on one of its
canyon (69.0 m wide) and traveling 35.8 m/s off the takeoff blades 10.0 cm from the axis of rotation?
ramp, he reached the other side. What was his launch
angle? 68. A point located on the second hand of a large clock
has a radial acceleration of 0.1cm/s 2. How far is the point
58. You throw a baseball at an initial speed of 15.0 m/s at from the axis of rotation of the second hand?
an angle of 30° with respect to the horizontal. What would
the ball’s initial speed have to be at 30° on a planet that
has twice the acceleration of gravity as Earth to achieve the 7.5 Relative Motion in One and Two
same range? Consider launch and impact on a horizontal Dimensions
surface.
69. The coordinate axes of the reference frame S′ remain
59. Aaron Rogers throws a football at 20.0 m/s to his wide parallel to those of S, as S′ moves away from S at a
receiver, who runs straight down the field at 9.4 m/s for → ^ ^ ^
constant velocity v S′ = (4.0 i + 3.0 j + 5.0 k ) m/s.
20.0 m. If Aaron throws the football when the wide receiver
has reached 10.0 m, what angle does Aaron have to launch (a) If at time t = 0 the origins coincide, what is the position
the ball so the receiver catches it at the 20.0 m mark? of the origin O′ in the S frame as a function of time?
(b) How is particle position for →
r (t) and →
r ′(t), as
measured in S and S′, respectively, related? (c) What
7.4 Uniform Circular Motion
is the relationship between particle velocities
60. A flywheel is rotating at 30 rev/s. What is the total →
v (t) and →
v ′(t) ? (d) How are accelerations
angle, in radians, through which a point on the flywheel
rotates in 40 s? →
a (t) and →
a ′ (t) related?

61. A particle travels in a circle of radius 10 m at a 70. The coordinate axes of the reference frame S′ remain
constant speed of 20 m/s. What is the magnitude of the
parallel to those of S, as S′ moves away from S at a
acceleration?
^ ^ ^
constant velocity → v S′ S = (1.0 i + 2.0 j + 3.0 k )t m/s
62. Cam Newton of the Carolina Panthers throws a perfect
. (a) If at time t = 0 the origins coincide, what is the
football spiral at 8.0 rev/s. The radius of a pro football is 8.5
position of origin O′ in the S frame as a function of time?
cm at the middle of the short side. What is the centripetal
acceleration of the laces on the football? (b) How is particle position for →
r (t) and →
r ′(t) , as
measured in S and S′, respectively, related? (c) What
63. A fairground ride spins its occupants inside a flying is the relationship between particle velocities
saucer-shaped container. If the horizontal circular path the →
v (t) and →
v ′(t) ? (d) How are accelerations
riders follow has an 8.00-m radius, at how many
revolutions per minute are the riders subjected to a

a (t) and →
a ′ (t) related?
centripetal acceleration equal to that of gravity?
71. The velocity of a particle in reference frame A is
64. A runner taking part in the 200-m dash must run ^ ^
around the end of a track that has a circular arc with a (2.0 i + 3.0 j ) m/s. The velocity of reference frame A
radius of curvature of 30.0 m. The runner starts the race at ^
a constant speed. If she completes the 200-m dash in 23.2 with respect to reference frame B is 4.0 k m/s, and the
s and runs at constant speed throughout the race, what is velocity of reference frame B with respect to C is
her centripetal acceleration as she runs the curved portion ^
of the track? 2.0 j m/s. What is the velocity of the particle in reference
frame C?
65. What is the acceleration of Venus toward the Sun,
assuming a circular orbit? 72. Raindrops fall vertically at 4.5 m/s relative to the
earth. What does an observer in a car moving at 22.0 m/s in
66. An experimental jet rocket travels around Earth along a straight line measure as the velocity of the raindrops?

This OpenStax book is available for free at https://legacy.cnx.org/content/col12169/1.46


Chapter 7 | Kinematics in Two and Three Dimensions 259

73. A seagull can fly at a velocity of 9.00 m/s in still air. to get across and how far downstream is the boat when it
(a) If it takes the bird 20.0 min to travel 6.00 km straight reaches the opposite shore?
into an oncoming wind, what is the velocity of the wind?
(b) If the bird turns around and flies with the wind, how 76. A small plane flies at 200 km/h in still air. If the
long will it take the bird to return 6.00 km? wind blows directly out of the west at 50 km/h, (a) in what
direction must the pilot head her plane to move directly
74. A ship sets sail from Rotterdam, heading due north at north across land and (b) how long does it take her to reach
7.00 m/s relative to the water. The local ocean current is a point 300 km directly north of her starting point?
1.50 m/s in a direction 40.0° north of east. What is the
velocity of the ship relative to Earth? 77. A cyclist traveling southeast along a road at 15 km/h
feels a wind blowing from the southwest at 25 km/h. To a
75. A boat can be rowed at 8.0 km/h in still water. (a) How stationary observer, what are the speed and direction of the
much time is required to row 1.5 km downstream in a river wind?
moving 3.0 km/h relative to the shore? (b) How much time
is required for the return trip? (c) In what direction must the 78. A river is moving east at 4 m/s. A boat starts from
boat be aimed to row straight across the river? (d) Suppose the dock heading 30° north of west at 7 m/s. If the river
the river is 0.8 km wide. What is the velocity of the boat is 1800 m wide, (a) what is the velocity of the boat with
with respect to Earth and how much time is required to respect to Earth and (b) how long does it take the boat to
get to the opposite shore? (e) Suppose, instead, the boat is cross the river?
aimed straight across the river. How much time is required

ADDITIONAL PROBLEMS
79. A Formula One race car is traveling at 89.0 m/s along
a straight track enters a turn on the race track with radius
of curvature of 200.0 m. What centripetal acceleration must
the car have to stay on the track?

80. A particle travels in a circular orbit of radius 10 m.


Its speed is changing at a rate of 15.0 m/s 2 at an instant
when its speed is 40.0 m/s. What is the magnitude of the
acceleration of the particle?

81. The driver of a car moving at 90.0 km/h presses down


on the brake as the car enters a circular curve of radius
150.0 m. If the speed of the car is decreasing at a rate
of 9.0 km/h each second, what is the magnitude of the
acceleration of the car at the instant its speed is 60.0 km/h?

82. A race car entering the curved part of the track at


the Daytona 500 drops its speed from 85.0 m/s to 80.0 m/
s in 2.0 s. If the radius of the curved part of the track is
316.0 m, calculate the total acceleration of the race car at
the beginning and ending of reduction of speed.

83. An elephant is located on Earth’s surface at a latitude


λ. Calculate the centripetal acceleration of the elephant
84. A proton in a synchrotron is moving in a circle of
resulting from the rotation of Earth around its polar axis.
radius 1 km and increasing its speed by
Express your answer in terms of λ, the radius R E of
v(t) = c 1 + c 2 t 2, where c 1 = 2.0 × 10 5 m/s,
Earth, and time T for one rotation of Earth. Compare your
answer with g for λ = 40°. c 2 = 10 5 m/s 3. (a) What is the proton’s total acceleration
at t = 5.0 s? (b) At what time does the expression for the
velocity become unphysical?
260 Chapter 7 | Kinematics in Two and Three Dimensions

85. A propeller blade at rest starts to rotate from t = 0 s to t velocity 5 s after the rockets fire?
= 5.0 s with a tangential acceleration of the tip of the blade
at 3.00 m/s 2. The tip of the blade is 1.5 m from the axis of 92. A crossbow is aimed horizontally at a target 40 m
rotation. At t = 5.0 s, what is the total acceleration of the tip away. The arrow hits 30 cm below the spot at which it was
of the blade? aimed. What is the initial velocity of the arrow?

86. A particle is executing circular motion with a constant 93. A long jumper can jump a distance of 8.0 m when he
angular frequency of ω = 4.00 rad/s. If time t = 0 takes off at an angle of 45° with respect to the horizontal.
corresponds to the position of the particle being located at Assuming he can jump with the same initial speed at all
y = 0 m and x = 5 m, (a) what is the position of the particle angles, how much distance does he lose by taking off at
at t = 10 s? (b) What is its velocity at this time? (c) What is 30° ?
its acceleration?
94. On planet Arcon, the maximum horizontal range of
87. A particle’s centripetal acceleration is a C = 4.0 m/s 2 a projectile launched at 10 m/s is 20 m. What is the
acceleration of gravity on this planet?
at t = 0 s. It is executing uniform circular motion about an
axis at a distance of 5.0 m. What is its velocity at t = 10 s?
95. A mountain biker encounters a jump on a race course
that sends him into the air at 60° to the horizontal. If he
88. A rod 3.0 m in length is rotating at 2.0 rev/s about
an axis at one end. Compare the centripetal accelerations at lands at a horizontal distance of 45.0 m and 20 m below his
radii of (a) 1.0 m, (b) 2.0 m, and (c) 3.0 m. launch point, what is his initial speed?

^ ^ 96. Which has the greater centripetal acceleration, a car


89. A particle located initially at (1.5 j + 4.0 k )m with a speed of 15.0 m/s along a circular track of radius
^ ^ ^ 100.0 m or a car with a speed of 12.0 m/s along a circular
undergoes a displacement of (2.5 i + 3.2 j − 1.2 k ) m. track of radius 75.0 m?
What is the final position of the particle?
97. A geosynchronous satellite orbits Earth at a distance
90. The position of a particle is given by of 42,250.0 km and has a period of 1 day. What is the
→ ^ ^ centripetal acceleration of the satellite?
r (t) = (50 m/s)t i − (4.9 m/s 2)t 2 j . (a) What are the
particle’s velocity and acceleration as functions of time? (b) 98. Two speedboats are traveling at the same speed
What are the initial conditions to produce the motion? relative to the water in opposite directions in a moving
river. An observer on the riverbank sees the boats moving
91. A spaceship is traveling at a constant velocity of at 4.0 m/s and 5.0 m/s. (a) What is the speed of the boats
→ ^ relative to the river? (b) How fast is the river moving
v (t) = 250.0 i m/s when its rockets fire, giving it an relative to the shore?
^ ^
acceleration of →
a (t) = (3.0 i + 4.0 k )m/s 2. What is its

CHALLENGE PROBLEMS
99. World’s Longest Par 3. The tee of the world’s longest (b) In addition to clearing the crossbar, the football must
par 3 sits atop South Africa’s Hanglip Mountain at 400.0 be high enough in the air early during its flight to clear the
m above the green and can only be reached by helicopter. reach of the onrushing defensive lineman. If the lineman is
The horizontal distance to the green is 359.0 m. Neglect 4.6 m away and has a vertical reach of 2.5 m, can he block
air resistance and answer the following questions. (a) If the 45.7-m field goal attempt? (c) What if the lineman is 1.0
a golfer launches a shot that is 40° with respect to the m away?
horizontal, what initial velocity must she give the ball? (b)
What is the time to reach the green?

100. When a field goal kicker kicks a football as hard as


he can at 45° to the horizontal, the ball just clears the 3-m-
high crossbar of the goalposts 45.7 m away. (a) What is
the maximum speed the kicker can impart to the football?

This OpenStax book is available for free at https://legacy.cnx.org/content/col12169/1.46


Chapter 7 | Kinematics in Two and Three Dimensions 261

101. A truck is traveling east at 80 km/h. At an


intersection 32 km ahead, a car is traveling north at 50
km/h. (a) How long after this moment will the vehicles be
closest to each other? (b) How far apart will they be at that
point?
262 Chapter 7 | Kinematics in Two and Three Dimensions

This OpenStax book is available for free at https://legacy.cnx.org/content/col12169/1.46


Chapter 8 | Overview of the Solar System 263

8 | OVERVIEW OF THE
SOLAR SYSTEM
Chapter Outline
8.1 Overview of Our Planetary System
8.2 Composition and Structure of Planets
8.3 Origin of the Solar System
8.4 Kepler's Laws of Planetary Motion

Introduction

Figure 8.1 This picture was taken by the Curiosity Rover on Mars in 2012. The image is reconstructed digitally from 55
different images taken by a camera on the rover’s extended mast, so that the many positions of the mast (which acted like a selfie
stick) are edited out. (credit: modification of work by NASA/JPL-Caltech/MSSS)

Surrounding the Sun is a complex system of worlds with a wide range of conditions: eight major planets, many dwarf
planets, hundreds of moons, and countless smaller objects. Thanks largely to visits by spacecraft, we can now envision the
members of the solar system as other worlds like our own, each with its own chemical and geological history, and unique
sights that interplanetary tourists may someday visit. Some have called these past few decades the “golden age of planetary
exploration,” comparable to the golden age of exploration in the fifteenth century, when great sailing ships plied Earth’s
oceans and humanity became familiar with our own planet’s surface.
In this chapter, we discuss our planetary system and introduce the idea of comparative planetology—studying how the
planets work by comparing them with one another. We want to get to know the planets not only for what we can learn
about them, but also to see what they can tell us about the origin and evolution of the entire solar system. In the upcoming
chapters, we describe the better-known members of the solar system and begin to compare them to the thousands of planets
that have been discovered recently, orbiting other stars.
264 Chapter 8 | Overview of the Solar System

8.1 | Overview of Our Planetary System


Learning Objectives
By the end of this section you will be able to:
• Describe how the objects in our solar system are identified, explored, and characterized
• Describe the types of small bodies in our solar system, their locations, and how they formed
• Model the solar system with distances from everyday life to better comprehend distances in
space

The solar system[1] consists of the Sun and many smaller objects: the planets, their moons and rings, and such “debris”
as asteroids, comets, and dust. Decades of observation and spacecraft exploration have revealed that most of these objects
formed together with the Sun about 4.5 billion years ago. They represent clumps of material that condensed from an
enormous cloud of gas and dust. The central part of this cloud became the Sun, and a small fraction of the material in the
outer parts eventually formed the other objects.
During the past 50 years, we have learned more about the solar system than anyone imagined before the space age. In
addition to gathering information with powerful new telescopes, we have sent spacecraft directly to many members of the
planetary system. (Planetary astronomy is the only branch of our science in which we can, at least vicariously, travel to the
objects we want to study.) With evocative names such as Voyager, Pioneer, Curiosity, and Pathfinder, our robot explorers
have flown past, orbited, or landed on every planet, returning images and data that have dazzled both astronomers and the
public. In the process, we have also investigated two dwarf planets, hundreds of fascinating moons, four ring systems, a
dozen asteroids, and several comets (smaller members of our solar system that we will discuss later).
Our probes have penetrated the atmosphere of Jupiter and landed on the surfaces of Venus, Mars, our Moon, Saturn’s moon
Titan, the asteroids Eros and Itokawa, and the Comet Churyumov-Gerasimenko (usually referred to as 67P). Humans have
set foot on the Moon and returned samples of its surface soil for laboratory analysis (Figure 8.2). We have even discovered
other places in our solar system that might be able to support some kind of life.

Figure 8.2 The lunar lander and surface rover from the Apollo 15 mission are seen in this view of the one place beyond Earth
that has been explored directly by humans. (credit: modification of work by David R. Scott, NASA)

View this gallery of NASA images (https://openstaxcollege.org/l/30projapolloarc) that trace the history of

1. The generic term for a group of planets and other bodies circling a star is planetary system. Ours is called the solar
system because our Sun is sometimes called Sol. Strictly speaking, then, there is only one solar system; planets orbiting
other stars are in planetary systems.

This OpenStax book is available for free at https://legacy.cnx.org/content/col12169/1.46


Chapter 8 | Overview of the Solar System 265

the Apollo mission.

An Inventory
The Sun, a star that is brighter than about 80% of the stars in the Galaxy, is by far the most massive member of the solar
system, as shown in Table 8.1. It is an enormous ball about 1.4 million kilometers in diameter, with surface layers of
incandescent gas and an interior temperature of millions of degrees. The Sun will be discussed in later chapters as our first,
and best-studied, example of a star.

Mass of Members of the Solar System


Object Percentage of Total Mass of Solar System
Sun 99.80
Jupiter 0.10
Comets 0.0005–0.03 (estimate)
All other planets and dwarf planets 0.04
Moons and rings 0.00005
Asteroids 0.000002 (estimate)
Cosmic dust 0.0000001 (estimate)

Table 8.1

Table 8.1 also shows that most of the material of the planets is actually concentrated in the largest one, Jupiter, which
is more massive than all the rest of the planets combined. Astronomers were able to determine the masses of the planets
centuries ago using Kepler’s laws of planetary motion and Newton’s law of gravity to measure the planets’ gravitational
effects on one another or on moons that orbit them (see Orbits and Gravity (https://legacy.cnx.org/content/
m59773/latest/) ). Today, we make even more precise measurements of their masses by tracking their gravitational effects
on the motion of spacecraft that pass near them.
Beside Earth, five other planets were known to the ancients—Mercury, Venus, Mars, Jupiter, and Saturn—and two were
discovered after the invention of the telescope: Uranus and Neptune. The eight planets all revolve in the same direction
around the Sun. They orbit in approximately the same plane, like cars traveling on concentric tracks on a giant, flat
racecourse. Each planet stays in its own “traffic lane,” following a nearly circular orbit about the Sun and obeying the
“traffic” laws discovered by Galileo, Kepler, and Newton. Besides these planets, we have also been discovering smaller
worlds beyond Neptune that are called trans-Neptunian objects or TNOs (see Figure 8.3). The first to be found, in 1930,
was Pluto, but others have been discovered during the twenty-first century. One of them, Eris, is about the same size as
Pluto and has at least one moon (Pluto has five known moons.) The largest TNOs are also classed as dwarf planets, as
is the largest asteroid, Ceres. (Dwarf planets will be discussed further in the chapter on Rings, Moons, and Pluto
(https://legacy.cnx.org/content/m59860/latest/) ). To date, more than 1750 of these TNOs have been discovered.
266 Chapter 8 | Overview of the Solar System

Figure 8.3 All eight major planets orbit the Sun in roughly the same plane. The five currently known dwarf planets are also
shown: Eris, Haumea, Pluto, Ceres, and Makemake. Note that Pluto’s orbit is not in the plane of the planets.

Each of the planets and dwarf planets also rotates (spins) about an axis running through it, and in most cases the direction
of rotation is the same as the direction of revolution about the Sun. The exceptions are Venus, which rotates backward very
slowly (that is, in a retrograde direction), and Uranus and Pluto, which also have strange rotations, each spinning about an
axis tipped nearly on its side. We do not yet know the spin orientations of Eris, Haumea, and Makemake.
The four planets closest to the Sun (Mercury through Mars) are called the inner or terrestrial planets. Often, the Moon is
also discussed as a part of this group, bringing the total of terrestrial objects to five. (We generally call Earth’s satellite “the
Moon,” with a capital M, and the other satellites “moons,” with lowercase m’s.) The terrestrial planets are relatively small
worlds, composed primarily of rock and metal. All of them have solid surfaces that bear the records of their geological
history in the forms of craters, mountains, and volcanoes (Figure 8.4).

Figure 8.4 The pockmarked face of the terrestrial world of Mercury is more typical of the inner planets than the watery surface
of Earth. This black-and-white image, taken with the Mariner 10 spacecraft, shows a region more than 400 kilometers wide.
(credit: modification of work by NASA/John Hopkins University Applied Physics Laboratory/Carnegie Institution of
Washington)

This OpenStax book is available for free at https://legacy.cnx.org/content/col12169/1.46


Chapter 8 | Overview of the Solar System 267

The next four planets (Jupiter through Neptune) are much larger and are composed primarily of lighter ices, liquids, and
gases. We call these four the jovian planets (after “Jove,” another name for Jupiter in mythology) or giant planets—a name
they richly deserve (Figure 8.5). More than 1400 Earths could fit inside Jupiter, for example. These planets do not have
solid surfaces on which future explorers might land. They are more like vast, spherical oceans with much smaller, dense
cores.

Figure 8.5 This montage shows the four giant planets: Jupiter, Saturn, Uranus, and Neptune. Below them, Earth is shown to
scale. (credit: modification of work by NASA, Solar System Exploration)

Near the outer edge of the system lies Pluto, which was the first of the distant icy worlds to be discovered beyond Neptune
(Pluto was visited by a spacecraft, the NASA New Horizons mission, in 2015 [see Figure 8.6]). Table 8.2 summarizes
some of the main facts about the planets.

Figure 8.6 This intriguing image from the New Horizons spacecraft, taken when it flew by the dwarf planet in July 2015,
shows some of its complex surface features. The rounded white area is temporarily being called the Sputnik Plain, after
humanity’s first spacecraft. (credit: modification of work by NASA/Johns Hopkins University Applied Physics Laboratory/
Southwest Research Institute)

The Planets
Name Distance from Sun Revolution Period Diameter Mass Density
(AU)[2] (y) (km) (1023 kg) (g/cm3)[3]
Mercury 0.39 0.24 4,878 3.3 5.4

Table 8.2

2. An AU (or astronomical unit) is the distance from Earth to the Sun.


3. We give densities in units where the density of water is 1 g/cm3. To get densities in units of kg/m3,
multiply the given value by 1000.
268 Chapter 8 | Overview of the Solar System

The Planets
Name Distance from Sun Revolution Period Diameter Mass Density
(AU) (y) (km) (1023 kg) (g/cm3)
Venus 0.72 0.62 12,120 48.7 5.2
Earth 1.00 1.00 12,756 59.8 5.5
Mars 1.52 1.88 6,787 6.4 3.9
Jupiter 5.20 11.86 142,984 18,991 1.3
Saturn 9.54 29.46 120,536 5686 0.7
Uranus 19.18 84.07 51,118 866 1.3
Neptune 30.06 164.82 49,660 1030 1.6

Table 8.2

Example 8.1

Comparing Densities
Let’s compare the densities of several members of the solar system. The density of an object equals its mass
divided by its volume. The volume (V) of a sphere (like a planet) is calculated using the equation

V = 4 πR 3
3
where π (the Greek letter pi) has a value of approximately 3.14. Although planets are not perfect spheres, this
equation works well enough. The masses and diameters of the planets are given in Table 8.2. For data on selected
moons, see Appendix D. Let’s use Saturn’s moon Mimas as our example, with a mass of 4 × 1019 kg and a
diameter of approximately 400 km (radius, 200 km = 2 × 105 m).
Solution
The volume of Mimas is
3
4 × 3.14 × ⎛2 × 10 5 m⎞ = 3.3 × 10 16 m 3.
3 ⎝ ⎠

Density is mass divided by volume:


4 × 10 19 kg
= 1.2 × 10 3 kg/m 3.
3.3 × 10 16 m 3
Note that the density of water in these units is 1000 kg/m3, so Mimas must be made mainly of ice, not rock. (Note
that the density of Mimas given in Section D.1 is 1.2, but the units used there are different. In that table, we give
density in units of g/cm3, for which the density of water equals 1. Can you show, by converting units, that 1 g/
cm3 is the same as 1000 kg/m3?)
Check Your Learning
Calculate the average density of our own planet, Earth. Show your work. How does it compare to the density of
an ice moon like Mimas? See Table 8.2 for data.

Answer:
For a sphere,
density = ⎛mass3⎞ kg/m 3.
4
⎝ 3 πR ⎠

This OpenStax book is available for free at https://legacy.cnx.org/content/col12169/1.46


Chapter 8 | Overview of the Solar System 269

For Earth, then,


6 × 10 24 kg
density = = 5.5 × 10 3 kg/m 3.
4.2 × 2.6 × 10 20 m 3
This density is four to five times greater than Mimas’. In fact, Earth is the densest of the planets.

Learn more about NASA’s mission to Pluto (https://openstaxcollege.org/l/30NASAmisspluto) and see


high-resolution images of Pluto’s moon Charon.

Smaller Members of the Solar System


Most of the planets are accompanied by one or more moons; only Mercury and Venus move through space alone. There
are more than 180 known moons orbiting planets and dwarf planets (see Appendix D for a listing of the larger ones), and
undoubtedly many other small ones remain undiscovered. The largest of the moons are as big as small planets and just as
interesting. In addition to our Moon, they include the four largest moons of Jupiter (called the Galilean moons, after their
discoverer) and the largest moons of Saturn and Neptune (confusingly named Titan and Triton).
Each of the giant planets also has rings made up of countless small bodies ranging in size from mountains to mere grains
of dust, all in orbit about the equator of the planet. The bright rings of Saturn are, by far, the easiest to see. They are among
the most beautiful sights in the solar system (Figure 8.7). But, all four ring systems are interesting to scientists because of
their complicated forms, influenced by the pull of the moons that also orbit these giant planets.

Figure 8.7 This 2007 Cassini image shows Saturn and its complex system of rings, taken from a distance of about 1.2 million
kilometers. This natural-color image is a composite of 36 images taken over the course of 2.5 hours. (credit: modification of work
by NASA/JPL/Space Science Institute)

The solar system has many other less-conspicuous members. Another group is the asteroids, rocky bodies that orbit the
Sun like miniature planets, mostly in the space between Mars and Jupiter (although some do cross the orbits of planets like
Earth—see Figure 8.8). Most asteroids are remnants of the initial population of the solar system that existed before the
planets themselves formed. Some of the smallest moons of the planets, such as the moons of Mars, are very likely captured
asteroids.
270 Chapter 8 | Overview of the Solar System

Figure 8.8 This small Earth-crossing asteroid image was taken by the NEAR-Shoemaker spacecraft from an altitude of about
100 kilometers. This view of the heavily cratered surface is about 10 kilometers wide. The spacecraft orbited Eros for a year
before landing gently on its surface. (credit: modification of work by NASA/JHUAPL)

Another class of small bodies is composed mostly of ice, made of frozen gases such as water, carbon dioxide, and carbon
monoxide; these objects are called comets (see Figure 8.9). Comets also are remnants from the formation of the solar
system, but they were formed and continue (with rare exceptions) to orbit the Sun in distant, cooler regions—stored in a
sort of cosmic deep freeze. This is also the realm of the larger icy worlds, called dwarf planets.

Figure 8.9 This image shows Comet Churyumov-Gerasimenko, also known as 67P, near its closest approach to the Sun in
2015, as seen from the Rosetta spacecraft. Note the jets of gas escaping from the solid surface. (credit: modification of work by
ESA/Rosetta/NAVACAM, CC BY-SA IGO 3.0 (http://creativecommons.org/licenses/by-sa/3.0/igo/) )

Finally, there are countless grains of broken rock, which we call cosmic dust, scattered throughout the solar system. When
these particles enter Earth’s atmosphere (as millions do each day) they burn up, producing a brief flash of light in the
night sky known as a meteor (meteors are often referred to as shooting stars). Occasionally, some larger chunk of rocky or
metallic material survives its passage through the atmosphere and lands on Earth. Any piece that strikes the ground is known
as a meteorite. (You can see meteorites on display in many natural history museums and can sometimes even purchase
pieces of them from gem and mineral dealers.)

Carl Sagan: Solar System Advocate


The best-known astronomer in the world during the 1970s and 1980s, Carl Sagan devoted most of his professional
career to studying the planets and considerable energy to raising public awareness of what we can learn from exploring
the solar system (see Figure 8.10). Born in Brooklyn, New York, in 1934, Sagan became interested in astronomy as
a youngster; he also credits science fiction stories for sustaining his fascination with what’s out in the universe.

This OpenStax book is available for free at https://legacy.cnx.org/content/col12169/1.46


Chapter 8 | Overview of the Solar System 271

Figure 8.10 Sagan was Tyson’s inspiration to become a scientist. (credit “Sagan”:
modification of work by NASA, JPL; credit “Tyson”: modification of work by Bruce F. Press)

In the early 1960s, when many scientists still thought Venus might turn out to be a hospitable place, Sagan calculated
that the thick atmosphere of Venus could act like a giant greenhouse, keeping the heat in and raising the temperature
enormously. He showed that the seasonal changes astronomers had seen on Mars were caused, not by vegetation, but
by wind-blown dust. He was a member of the scientific teams for many of the robotic missions that explored the solar
system and was instrumental in getting NASA to put a message-bearing plaque aboard the Pioneer spacecraft, as well
as audio-video records on the Voyager spacecraft—all of them destined to leave our solar system entirely and send
these little bits of Earth technology out among the stars.
To encourage public interest and public support of planetary exploration, Sagan helped found The Planetary Society,
now the largest space-interest organization in the world. He was a tireless and eloquent advocate of the need to study
the solar system close-up and the value of learning about other worlds in order to take better care of our own.
Sagan simulated conditions on early Earth to demonstrate how some of life’s fundamental building blocks might have
formed from the “primordial soup” of natural compounds on our planet. In addition, he and his colleagues developed
computer models showing the consequences of nuclear war for Earth would be even more devastating than anyone
had thought (this is now called the nuclear winter hypothesis) and demonstrating some of the serious consequences of
continued pollution of our atmosphere.
Sagan was perhaps best known, however, as a brilliant popularizer of astronomy and the author of many books on
science, including the best-selling Cosmos, and several evocative tributes to solar system exploration such as The
Cosmic Connection and Pale Blue Dot. His book The Demon Haunted World, completed just before his death in
1996, is perhaps the best antidote to fuzzy thinking about pseudo-science and irrationality in print today. An intriguing
science fiction novel he wrote, titled Contact, which became a successful film as well, is still recommended by many
science instructors as a scenario for making contact with life elsewhere that is much more reasonable than most science
fiction.
Sagan was a master, too, of the television medium. His 13-part public television series, Cosmos, was seen by an
estimated 500 million people in 60 countries and has become one of the most-watched series in the history of public
broadcasting. A few astronomers scoffed at a scientist who spent so much time in the public eye, but it is probably fair
to say that Sagan’s enthusiasm and skill as an explainer won more friends for the science of astronomy than anyone or
anything else in the second half of the twentieth century.
In the two decades since Sagan’s death, no other scientist has achieved the same level of public recognition. Perhaps
closest is the director of the Hayden Planetarium, Neil deGrasse Tyson, who followed in Sagan’s footsteps by making
an updated version of the Cosmos program in 2014. Tyson is quick to point out that Sagan was his inspiration to
become a scientist, telling how Sagan invited him to visit for a day at Cornell when he was a high school student
looking for a career. However, the media environment has fragmented a great deal since Sagan’s time. It is interesting
to speculate whether Sagan could have adapted his communication style to the world of cable television, Twitter,
Facebook, and podcasts.
272 Chapter 8 | Overview of the Solar System

Two imaginative videos provide a tour of the solar system objects we have been discussing. Shane Gellert’s I
Need Some Space (https://openstaxcollege.org/l/30needsomespace) uses NASA photography and models
to show the various worlds with which we share our system. In the more science fiction-oriented Wanderers
(https://openstaxcollege.org/l/30wanderers) video, we see some of the planets and moons as tourist
destinations for future explorers, with commentary taken from recordings by Carl Sagan.

A Scale Model of the Solar System


Astronomy often deals with dimensions and distances that far exceed our ordinary experience. What does 1.4 billion
kilometers—the distance from the Sun to Saturn—really mean to anyone? It can be helpful to visualize such large systems
in terms of a scale model.
In our imaginations, let us build a scale model of the solar system, adopting a scale factor of 1 billion (109)—that is, reducing
the actual solar system by dividing every dimension by a factor of 109. Earth, then, has a diameter of 1.3 centimeters, about
the size of a grape. The Moon is a pea orbiting this at a distance of 40 centimeters, or a little more than a foot away. The
Earth-Moon system fits into a standard backpack.
In this model, the Sun is nearly 1.5 meters in diameter, about the average height of an adult, and our Earth is at a distance
of 150 meters—about one city block—from the Sun. Jupiter is five blocks away from the Sun, and its diameter is 15
centimeters, about the size of a very large grapefruit. Saturn is 10 blocks from the Sun; Uranus, 20 blocks; and Neptune, 30
blocks. Pluto, with a distance that varies quite a bit during its 249-year orbit, is currently just beyond 30 blocks and getting
farther with time. Most of the moons of the outer solar system are the sizes of various kinds of seeds orbiting the grapefruit,
oranges, and lemons that represent the outer planets.
In our scale model, a human is reduced to the dimensions of a single atom, and cars and spacecraft to the size of molecules.
Sending the Voyager spacecraft to Neptune involves navigating a single molecule from the Earth–grape toward a lemon 5
kilometers away with an accuracy equivalent to the width of a thread in a spider’s web.
If that model represents the solar system, where would the nearest stars be? If we keep the same scale, the closest stars
would be tens of thousands of kilometers away. If you built this scale model in the city where you live, you would have to
place the representations of these stars on the other side of Earth or beyond.
By the way, model solar systems like the one we just presented have been built in cities throughout the world. In Sweden,
for example, Stockholm’s huge Globe Arena has become a model for the Sun, and Pluto is represented by a 12-centimeter
sculpture in the small town of Delsbo, 300 kilometers away. Another model solar system is in Washington on the Mall
between the White House and Congress (perhaps proving they are worlds apart?).

Names in the Solar System


We humans just don’t feel comfortable until something has a name. Types of butterflies, new elements, and the
mountains of Venus all need names for us to feel we are acquainted with them. How do we give names to objects and
features in the solar system?
Planets and moons are named after gods and heroes in Greek and Roman mythology (with a few exceptions among the
moons of Uranus, which have names drawn from English literature). When William Herschel, a German immigrant to
England, first discovered the planet we now call Uranus, he wanted to name it Georgium Sidus (George’s star) after
King George III of his adopted country. This caused such an outcry among astronomers in other nations, however, that
the classic tradition was upheld—and has been maintained ever since. Luckily, there were a lot of minor gods in the
ancient pantheon, so plenty of names are left for the many small moons we are discovering around the giant planets.
(Appendix G lists the larger moons).
Comets are often named after their discoverers (offering an extra incentive to comet hunters). Asteroids are named by
their discoverers after just about anyone or anything they want. Recently, asteroid names have been used to recognize
people who have made significant contributions to astronomy, including the three original authors of this book.
That was pretty much all the naming that was needed while our study of the solar system was confined to Earth.
But now, our spacecraft have surveyed and photographed many worlds in great detail, and each world has a host of
features that also need names. To make sure that naming things in space remains multinational, rational, and somewhat
dignified, astronomers have given the responsibility of approving names to a special committee of the International
Astronomical Union (IAU), the body that includes scientists from every country that does astronomy.
This IAU committee has developed a set of rules for naming features on other worlds. For example, craters on Venus

This OpenStax book is available for free at https://legacy.cnx.org/content/col12169/1.46


Chapter 8 | Overview of the Solar System 273

are named for women who have made significant contributions to human knowledge and welfare. Volcanic features
on Jupiter’s moon Io, which is in a constant state of volcanic activity, are named after gods of fire and thunder
from the mythologies of many cultures. Craters on Mercury commemorate famous novelists, playwrights, artists, and
composers. On Saturn’s moon Tethys, all the features are named after characters and places in Homer’s great epic
poem, The Odyssey. As we explore further, it may well turn out that more places in the solar system need names than
Earth history can provide. Perhaps by then, explorers and settlers on these worlds will be ready to develop their own
names for the places they may (if but for a while) call home.
You may be surprised to know that the meaning of the word planet has recently become controversial because we
have discovered many other planetary systems that don’t look very much like our own. Even within our solar system,
the planets differ greatly in size and chemical properties. The biggest dispute concerns Pluto, which is much smaller
than the other eight major planets. The category of dwarf planet was invented to include Pluto and similar icy objects
beyond Neptune. But is a dwarf planet also a planet? Logically, it should be, but even this simple issue of grammar has
been the subject of heated debate among both astronomers and the general public.

8.2 | Composition and Structure of Planets


Learning Objectives
By the end of this section you will be able to:
• Describe the characteristics of the giant planets, terrestrial planets, and small bodies in the
solar system
• Explain what influences the temperature of a planet’s surface
• Explain why there is geological activity on some planets and not on others

The fact that there are two distinct kinds of planets—the rocky terrestrial planets and the gas-rich jovian planets—leads us
to believe that they formed under different conditions. Certainly their compositions are dominated by different elements.
Let us look at each type in more detail.

The Giant Planets


The two largest planets, Jupiter and Saturn, have nearly the same chemical makeup as the Sun; they are composed primarily
of the two elements hydrogen and helium, with 75% of their mass being hydrogen and 25% helium. On Earth, both
hydrogen and helium are gases, so Jupiter and Saturn are sometimes called gas planets. But, this name is misleading. Jupiter
and Saturn are so large that the gas is compressed in their interior until the hydrogen becomes a liquid. Because the bulk of
both planets consists of compressed, liquefied hydrogen, we should really call them liquid planets.
Under the force of gravity, the heavier elements sink toward the inner parts of a liquid or gaseous planet. Both Jupiter and
Saturn, therefore, have cores composed of heavier rock, metal, and ice, but we cannot see these regions directly. In fact,
when we look down from above, all we see is the atmosphere with its swirling clouds (Figure 8.11). We must infer the
existence of the denser core inside these planets from studies of each planet’s gravity.
274 Chapter 8 | Overview of the Solar System

Figure 8.11 This true-color image of Jupiter was taken from the Cassini spacecraft in 2000. (credit: modification of work by
NASA/JPL/University of Arizona)

Uranus and Neptune are much smaller than Jupiter and Saturn, but each also has a core of rock, metal, and ice. Uranus and
Neptune were less efficient at attracting hydrogen and helium gas, so they have much smaller atmospheres in proportion to
their cores.
Chemically, each giant planet is dominated by hydrogen and its many compounds. Nearly all the oxygen present is
combined chemically with hydrogen to form water (H2O). Chemists call such a hydrogen-dominated composition reduced.
Throughout the outer solar system, we find abundant water (mostly in the form of ice) and reducing chemistry.

The Terrestrial Planets


The terrestrial planets are quite different from the giants. In addition to being much smaller, they are composed primarily
of rocks and metals. These, in turn, are made of elements that are less common in the universe as a whole. The most
abundant rocks, called silicates, are made of silicon and oxygen, and the most common metal is iron. We can tell from their
densities (see m59816 (https://legacy.cnx.org/content/m59816/latest/#fs-id1170326125105) ) that Mercury has
the greatest proportion of metals (which are denser) and the Moon has the lowest. Earth, Venus, and Mars all have roughly
similar bulk compositions: about one third of their mass consists of iron-nickel or iron-sulfur combinations; two thirds is
made of silicates. Because these planets are largely composed of oxygen compounds (such as the silicate minerals of their
crusts), their chemistry is said to be oxidized.
When we look at the internal structure of each of the terrestrial planets, we find that the densest metals are in a central core,
with the lighter silicates near the surface. If these planets were liquid, like the giant planets, we could understand this effect
as the result the sinking of heavier elements due to the pull of gravity. This leads us to conclude that, although the terrestrial
planets are solid today, at one time they must have been hot enough to melt.
Differentiation is the process by which gravity helps separate a planet’s interior into layers of different compositions and
densities. The heavier metals sink to form a core, while the lightest minerals float to the surface to form a crust. Later, when
the planet cools, this layered structure is preserved. In order for a rocky planet to differentiate, it must be heated to the
melting point of rocks, which is typically more than 1300 K.

Moons, Asteroids, and Comets


Chemically and structurally, Earth’s Moon is like the terrestrial planets, but most moons are in the outer solar system,
and they have compositions similar to the cores of the giant planets around which they orbit. The three largest
moons—Ganymede and Callisto in the jovian system, and Titan in the saturnian system—are composed half of frozen water,
and half of rocks and metals. Most of these moons differentiated during formation, and today they have cores of rock and
metal, with upper layers and crusts of very cold and—thus very hard—ice (Figure 8.12).

This OpenStax book is available for free at https://legacy.cnx.org/content/col12169/1.46


Chapter 8 | Overview of the Solar System 275

Figure 8.12 This view of Jupiter’s moon Ganymede was taken in June 1996 by the Galileo spacecraft. The brownish gray
color of the surface indicates a dusty mixture of rocky material and ice. The bright spots are places where recent impacts have
uncovered fresh ice from underneath. (credit: modification of work by NASA/JPL)

Most of the asteroids and comets, as well as the smallest moons, were probably never heated to the melting point. However,
some of the largest asteroids, such as Vesta, appear to be differentiated; others are fragments from differentiated bodies.
Because most asteroids and comets retain their original composition, they represent relatively unmodified material dating
back to the time of the formation of the solar system. In a sense, they act as chemical fossils, helping us to learn about a
time long ago whose traces have been erased on larger worlds.

Temperatures: Going to Extremes


Generally speaking, the farther a planet or moon is from the Sun, the cooler its surface. The planets are heated by the radiant
energy of the Sun, which gets weaker with the square of the distance. You know how rapidly the heating effect of a fireplace
or an outdoor radiant heater diminishes as you walk away from it; the same effect applies to the Sun. Mercury, the closest
planet to the Sun, has a blistering surface temperature that ranges from 280–430 °C on its sunlit side, whereas the surface
temperature on Pluto is only about –220 °C, colder than liquid air.
Mathematically, the temperatures decrease approximately in proportion to the square root of the distance from the Sun.
Pluto is about 30 AU at its closest to the Sun (or 100 times the distance of Mercury) and about 49 AU at its farthest from
the Sun. Thus, Pluto’s temperature is less than that of Mercury by the square root of 100, or a factor of 10: from 500 K to
50 K.
In addition to its distance from the Sun, the surface temperature of a planet can be influenced strongly by its atmosphere.
Without our atmospheric insulation (the greenhouse effect, which keeps the heat in), the oceans of Earth would be
permanently frozen. Conversely, if Mars once had a larger atmosphere in the past, it could have supported a more temperate
climate than it has today. Venus is an even more extreme example, where its thick atmosphere of carbon dioxide acts as
insulation, reducing the escape of heat built up at the surface, resulting in temperatures greater than those on Mercury.
Today, Earth is the only planet where surface temperatures generally lie between the freezing and boiling points of water.
As far as we know, Earth is the only planet to support life.

There’s No Place Like Home


In the classic film The Wizard of Oz, Dorothy, the heroine, concludes after her many adventures in “alien”
environments that “there’s no place like home.” The same can be said of the other worlds in our solar system. There
are many fascinating places, large and small, that we might like to visit, but humans could not survive on any without
a great deal of artificial assistance.
A thick carbon dioxide atmosphere keeps the surface temperature on our neighbor Venus at a sizzling 700 K (near 900
°F). Mars, on the other hand, has temperatures generally below freezing, with air (also mostly carbon dioxide) so thin
that it resembles that found at an altitude of 30 kilometers (100,000 feet) in Earth’s atmosphere. And the red planet is
so dry that it has not had any rain for billions of years.
The outer layers of the jovian planets are neither warm enough nor solid enough for human habitation. Any bases
276 Chapter 8 | Overview of the Solar System

we build in the systems of the giant planets may well have to be in space or one of their moons—none of which is
particularly hospitable to a luxury hotel with a swimming pool and palm trees. Perhaps we will find warmer havens
deep inside the clouds of Jupiter or in the ocean under the frozen ice of its moon Europa.
All of this suggests that we had better take good care of Earth because it is the only site where life as we know it could
survive. Recent human activity may be reducing the habitability of our planet by adding pollutants to the atmosphere,
especially the potent greenhouse gas carbon dioxide. Human civilization is changing our planet dramatically, and
these changes are not necessarily for the better. In a solar system that seems unready to receive us, making Earth less
hospitable to life may be a grave mistake.

Geological Activity
The crusts of all of the terrestrial planets, as well as of the larger moons, have been modified over their histories by
both internal and external forces. Externally, each has been battered by a slow rain of projectiles from space, leaving
their surfaces pockmarked by impact craters of all sizes (see m59816 (https://legacy.cnx.org/content/m59816/
latest/#OSC_Astro_07_01_Mercury) ). We have good evidence that this bombardment was far greater in the early
history of the solar system, but it certainly continues to this day, even if at a lower rate. The collision of more than 20 large
pieces of Comet Shoemaker–Levy 9 with Jupiter in the summer of 1994 (see Figure 8.13) is one dramatic example of this
process.

Figure 8.13 In this image of Comet Shoemaker–Levy 9 taken on May 17, 1994, by NASA’s Hubble Space Telescope, you can
see about 20 icy fragments into which the comet broke. The comet was approximately 660 million kilometers from Earth,
heading on a collision course with Jupiter. (credit: modification of work by NASA, ESA, H. Weaver (STScl), E. Smith (STScl))

Figure 8.14 shows the aftermath of these collisions, when debris clouds larger than Earth could be seen in Jupiter’s
atmosphere.

Figure 8.14 The Hubble Space Telescope took this sequence of images of Jupiter in summer 1994, when fragments of Comet
Shoemaker–Levy 9 collided with the giant planet. Here we see the site hit by fragment G, from five minutes to five days after
impact. Several of the dust clouds generated by the collisions became larger than Earth. (credit: modification of work by H.
Hammel, NASA)

This OpenStax book is available for free at https://legacy.cnx.org/content/col12169/1.46


Chapter 8 | Overview of the Solar System 277

During the time all the planets have been subject to such impacts, internal forces on the terrestrial planets have buckled and
twisted their crusts, built up mountain ranges, erupted as volcanoes, and generally reshaped the surfaces in what we call
geological activity. (The prefix geo means “Earth,” so this is a bit of an “Earth-chauvinist” term, but it is so widely used
that we bow to tradition.) Among the terrestrial planets, Earth and Venus have experienced the most geological activity over
their histories, although some of the moons in the outer solar system are also surprisingly active. In contrast, our own Moon
is a dead world where geological activity ceased billions of years ago.
Geological activity on a planet is the result of a hot interior. The forces of volcanism and mountain building are driven by
heat escaping from the interiors of planets. As we will see, each of the planets was heated at the time of its birth, and this
primordial heat initially powered extensive volcanic activity, even on our Moon. But, small objects such as the Moon soon
cooled off. The larger the planet or moon, the longer it retains its internal heat, and therefore the more we expect to see
surface evidence of continuing geological activity. The effect is similar to our own experience with a hot baked potato: the
larger the potato, the more slowly it cools. If we want a potato to cool quickly, we cut it into small pieces.
For the most part, the history of volcanic activity on the terrestrial planets conforms to the predictions of this simple theory.
The Moon, the smallest of these objects, is a geologically dead world. Although we know less about Mercury, it seems
likely that this planet, too, ceased most volcanic activity about the same time the Moon did. Mars represents an intermediate
case. It has been much more active than the Moon, but less so than Earth. Earth and Venus, the largest terrestrial planets,
still have molten interiors even today, some 4.5 billion years after their birth.

8.3 | Origin of the Solar System


Learning Objectives
By the end of this section you will be able to:
• Describe the characteristics of planets that are used to create formation models of the solar
system.
• Describe how the characteristics of extrasolar systems help us to model our own solar system.
• Explain the importance of collisions in the formation of the solar system.

Much of astronomy is motivated by a desire to understand the origin of things: to find at least partial answers to age-old
questions of where the universe, the Sun, Earth, and we ourselves came from. Each planet and moon is a fascinating place
that may stimulate our imagination as we try to picture what it would be like to visit. Taken together, the members of the
solar system preserve patterns that can tell us about the formation of the entire system. As we begin our exploration of the
planets, we want to introduce our modern picture of how the solar system formed.
The recent discovery of hundreds of planets in orbit around other stars has shown astronomers that many exoplanetary
systems can be quite different from our own solar system. For example, it is common for these systems to include planets
intermediate in size between our terrestrial and giant planets. These are often called superearths. Some exoplanet systems
even have giant planets close to the star, reversing the order we see in our system. In The Birth of Stars and the
Discovery of Planets outside the Solar System (https://legacy.cnx.org/content/m59915/latest/) , we will
look at these exoplanet systems. But for now, let us focus on theories of how our own particular system has formed and
evolved.

Looking for Patterns


One way to approach our question of origin is to look for regularities among the planets. We found, for example, that all
the planets lie in nearly the same plane and revolve in the same direction around the Sun. The Sun also spins in the same
direction about its own axis. Astronomers interpret this pattern as evidence that the Sun and planets formed together from a
spinning cloud of gas and dust that we call the solar nebula (Figure 8.15).
278 Chapter 8 | Overview of the Solar System

Figure 8.15 This artist’s conception of the solar nebula shows the flattened
cloud of gas and dust from which our planetary system formed. Icy and rocky
planetesimals (precursors of the planets) can be seen in the foreground. The bright
center is where the Sun is forming. (credit: William K. Hartmann, Planetary
Science Institute)

The composition of the planets gives another clue about origins. Spectroscopic analysis allows us to determine which
elements are present in the Sun and the planets. The Sun has the same hydrogen-dominated composition as Jupiter and
Saturn, and therefore appears to have been formed from the same reservoir of material. In comparison, the terrestrial planets
and our Moon are relatively deficient in the light gases and the various ices that form from the common elements oxygen,
carbon, and nitrogen. Instead, on Earth and its neighbors, we see mostly the rarer heavy elements such as iron and silicon.
This pattern suggests that the processes that led to planet formation in the inner solar system must somehow have excluded
much of the lighter materials that are common elsewhere. These lighter materials must have escaped, leaving a residue of
heavy stuff.
The reason for this is not hard to guess, bearing in mind the heat of the Sun. The inner planets and most of the asteroids are
made of rock and metal, which can survive heat, but they contain very little ice or gas, which evaporate when temperatures
are high. (To see what we mean, just compare how long a rock and an ice cube survive when they are placed in the sunlight.)
In the outer solar system, where it has always been cooler, the planets and their moons, as well as icy dwarf planets and
comets, are composed mostly of ice and gas.

The Evidence from Far Away


A second approach to understanding the origins of the solar system is to look outward for evidence that other systems
of planets are forming elsewhere. We cannot look back in time to the formation of our own system, but many stars in
space are much younger than the Sun. In these systems, the processes of planet formation might still be accessible to direct
observation. We observe that there are many other “solar nebulas” or circumstellar disks—flattened, spinning clouds of gas
and dust surrounding young stars. These disks resemble our own solar system’s initial stages of formation billions of years
ago (Figure 8.16).

This OpenStax book is available for free at https://legacy.cnx.org/content/col12169/1.46


Chapter 8 | Overview of the Solar System 279

Figure 8.16 These Hubble Space Telescope photos show sections of the Orion Nebula, a relatively close-by region where stars
are currently forming. Each image shows an embedded circumstellar disk orbiting a very young star. Seen from different angles,
some are energized to glow by the light of a nearby star while others are dark and seen in silhouette against the bright glowing
gas of the Orion Nebula. Each is a contemporary analog of our own solar nebula—a location where planets are probably being
formed today. (credit: modification of work by NASA/ESA, L. Ricci (ESO))

Building Planets
Circumstellar disks are a common occurrence around very young stars, suggesting that disks and stars form together.
Astronomers can use theoretical calculations to see how solid bodies might form from the gas and dust in these disks as they
cool. These models show that material begins to coalesce first by forming smaller objects, precursors of the planets, which
we call planetesimals.
Today’s fast computers can simulate the way millions of planetesimals, probably no larger than 100 kilometers in diameter,
might gather together under their mutual gravity to form the planets we see today. We are beginning to understand that this
process was a violent one, with planetesimals crashing into each other and sometimes even disrupting the growing planets
themselves. As a consequence of those violent impacts (and the heat from radioactive elements in them), all the planets
were heated until they were liquid and gas, and therefore differentiated, which helps explain their present internal structures.
The process of impacts and collisions in the early solar system was complex and, apparently, often random. The solar nebula
model can explain many of the regularities we find in the solar system, but the random collisions of massive collections
of planetesimals could be the reason for some exceptions to the “rules” of solar system behavior. For example, why do the
planets Uranus and Pluto spin on their sides? Why does Venus spin slowly and in the opposite direction from the other
planets? Why does the composition of the Moon resemble Earth in many ways and yet exhibit substantial differences? The
answers to such questions probably lie in enormous collisions that took place in the solar system long before life on Earth
began.
Today, some 4.5 billion years after its origin, the solar system is—thank goodness—a much less violent place. As we will
see, however, some planetesimals have continued to interact and collide, and their fragments move about the solar system
as roving “transients” that can make trouble for the established members of the Sun’s family, such as our own Earth. (We
discuss this “troublemaking” in Comets and Asteroids: Debris of the Solar System (https://legacy.cnx.org/
content/m59865/latest/) .)

A great variety of infographics (https://openstaxcollege.org/l/30worldsinsolar) at space.com let you explore


what it would be like to live on various worlds in the solar system.

8.4 | Kepler's Laws of Planetary Motion


Learning Objectives
By the end of this section, you will be able to:
• Describe the conic sections and how they relate to orbital motion.
• Describe how orbital velocity is related to orbital distance.
• Determine the period of an elliptical orbit from its semimajor axis and vice versa.
280 Chapter 8 | Overview of the Solar System

Figure 8.17 This space habitat and laboratory orbits Earth once every 90 minutes. (credit: modification of work by NASA)

How would you find a new planet at the outskirts of our solar system that is too dim to be seen with the unaided eye and is
so far away that it moves very slowly among the stars? This was the problem confronting astronomers during the nineteenth
century as they tried to pin down a full inventory of our solar system.
If we could look down on the solar system from somewhere out in space, interpreting planetary motions would be much
simpler. But the fact is, we must observe the positions of all the other planets from our own moving planet. Scientists of
the Renaissance did not know the details of Earth’s motions any better than the motions of the other planets. Their problem
was that they had to deduce the nature of all planetary motion using only their earthbound observations of the other planets’
positions in the sky. To solve this complex problem more fully, better observations and better models of the planetary system
were needed.
At about the time that Galileo was beginning his experiments with falling bodies, the efforts of two other scientists
dramatically advanced our understanding of the motions of the planets. These two astronomers were the observer Tycho
Brahe and the mathematician Johannes Kepler. Together, they placed the speculations of Copernicus on a sound
mathematical basis and paved the way for the work of Isaac Newton in the next century.

Tycho Brahe’s Observatory


Three years after the publication of Copernicus’ De Revolutionibus, Tycho Brahe was born to a family of Danish nobility.
He developed an early interest in astronomy and, as a young man, made significant astronomical observations. Among
these was a careful study of what we now know was an exploding star that flared up to great brilliance in the night sky.
His growing reputation gained him the patronage of the Danish King Frederick II, and at the age of 30, Brahe was able to
establish a fine astronomical observatory on the North Sea island of Hven (Figure 8.18). Brahe was the last and greatest
of the pre-telescopic observers in Europe.

This OpenStax book is available for free at https://legacy.cnx.org/content/col12169/1.46


Chapter 8 | Overview of the Solar System 281

Figure 8.18 (a) A stylized engraving shows Tycho Brahe using his instruments to measure the altitude of celestial objects
above the horizon. The large curved instrument in the foreground allowed him to measure precise angles in the sky. Note that the
scene includes hints of the grandeur of Brahe’s observatory at Hven. (b) Kepler was a German mathematician and astronomer.
His discovery of the basic laws that describe planetary motion placed the heliocentric cosmology of Copernicus on a firm
mathematical basis.

At Hven, Brahe made a continuous record of the positions of the Sun, Moon, and planets for almost 20 years. His extensive
and precise observations enabled him to note that the positions of the planets varied from those given in published tables,
which were based on the work of Ptolemy. These data were extremely valuable, but Brahe didn’t have the ability to analyze
them and develop a better model than what Ptolemy had published. He was further inhibited because he was an extravagant
and cantankerous fellow, and he accumulated enemies among government officials. When his patron, Frederick II, died
in 1597, Brahe lost his political base and decided to leave Denmark. He took up residence in Prague, where he became
court astronomer to Emperor Rudolf of Bohemia. There, in the year before his death, Brahe found a most able young
mathematician, Johannes Kepler, to assist him in analyzing his extensive planetary data.

Johannes Kepler
Johannes Kepler was born into a poor family in the German province of Württemberg and lived much of his life amid the
turmoil of the Thirty Years’ War (see Figure 8.18). He attended university at Tubingen and studied for a theological career.
There, he learned the principles of the Copernican system and became converted to the heliocentric hypothesis. Eventually,
Kepler went to Prague to serve as an assistant to Brahe, who set him to work trying to find a satisfactory theory of planetary
motion—one that was compatible with the long series of observations made at Hven. Brahe was reluctant to provide Kepler
with much material at any one time for fear that Kepler would discover the secrets of the universal motion by himself,
thereby robbing Brahe of some of the glory. Only after Brahe’s death in 1601 did Kepler get full possession of the priceless
records. Their study occupied most of Kepler’s time for more than 20 years.
Through his analysis of the motions of the planets, Kepler developed a series of principles, now known as Kepler’s three
laws, which described the behavior of planets based on their paths through space. The first two laws of planetary motion
were published in 1609 in The New Astronomy. Their discovery was a profound step in the development of modern science.

The Laws of Planetary Motion


The path of an object through space is called its orbit. Kepler initially assumed that the orbits of planets were circles, but
doing so did not allow him to find orbits that were consistent with Brahe’s observations. Working with the data for Mars, he
282 Chapter 8 | Overview of the Solar System

eventually discovered that the orbit of that planet had the shape of a somewhat flattened circle, or ellipse. Next to the circle,
the ellipse is the simplest kind of closed curve, belonging to a family of curves known as conic sections (Figure 8.19).

Figure 8.19 The circle, ellipse, parabola, and hyperbola are


all formed by the intersection of a plane with a cone. This is why
such curves are called conic sections.

Using the precise data collected by Tycho Brahe, Johannes Kepler carefully analyzed the positions in the sky of all the
known planets and the Moon, plotting their positions at regular intervals of time. From this analysis, he formulated three
laws, which we address in this section.

Kepler’s First Law


The prevailing view during the time of Kepler was that all planetary orbits were circular. The data for Mars presented the
greatest challenge to this view and that eventually encouraged Kepler to give up the popular idea. Kepler’s first law states
that every planet moves along an ellipse, with the Sun located at a focus of the ellipse. An ellipse is defined as the set of
all points such that the sum of the distance from each point to two foci is a constant. Figure 8.20 shows an ellipse and
describes a simple way to create it.

Figure 8.20 (a) An ellipse is a curve in which the sum of the distances from a point on the curve to two foci
( f 1 and f 2) is a constant. From this definition, you can see that an ellipse can be created in the following way. Place a
pin at each focus, then place a loop of string around a pencil and the pins. Keeping the string taught, move the pencil
around in a complete circuit. If the two foci occupy the same place, the result is a circle—a special case of an ellipse. (b)
For an elliptical orbit, if m ≪ M , then m follows an elliptical path with M at one focus. More exactly, both m and M
move in their own ellipse about the common center of mass.

For elliptical orbits, the point of closest approach of a planet to the Sun is called the perihelion. It is labeled point A in

This OpenStax book is available for free at https://legacy.cnx.org/content/col12169/1.46


Chapter 8 | Overview of the Solar System 283

Figure 8.20. The farthest point is the aphelion and is labeled point B in the figure. For the Moon’s orbit about Earth, those
points are called the perigee and apogee, respectively.
An ellipse has several mathematical forms, but all are a specific case of the more general equation for conic sections. There
are four different conic sections, all given by the equation
α = 1 + ecosθ. (8.1)
r
The variables r and θ are shown in Figure 8.21 in the case of an ellipse. The values of α and e determine which of the
four conic sections (hyperbola, parabola, ellipse or circle) represents the path of the satellite. For an ellipse, 0 ≤ e < 1,
with an eccentricity of zero meaning a circular orbit.

Figure 8.21 As before, the distance between the planet and


the Sun is r, and the angle measured from the x-axis, which is
along the major axis of the ellipse, is θ .

You can see an animation of two interacting objects at the My Solar System page at Phet
(https://openstaxcollege.org/l/21mysolarsys) . Choose the Sun and Planet preset option. You can also view
the more complicated multiple body problems as well. You may find the actual path of the Moon quite surprising,
yet is obeying Newton’s simple laws of motion.

Kepler’s Second Law


Kepler’s second law states that a planet sweeps out equal areas in equal times, that is, the area divided by time, called the
areal velocity, is constant. Consider Figure 8.22. The time it takes a planet to move from position A to B, sweeping out
area A 1 , is exactly the time taken to move from position C to D, sweeping area A 2 , and to move from E to F, sweeping
out area A 3 . These areas are the same: A 1 = A 2 = A 3 .
284 Chapter 8 | Overview of the Solar System

Figure 8.22 The shaded regions shown have equal areas and
represent the same time interval.

Comparing the areas in the figure and the distance traveled along the ellipse in each case, we can see that in order for the
areas to be equal, the planet must speed up as it gets closer to the Sun and slow down as it moves away.
Now consider Figure 8.23. A small triangular area ΔA is swept out in time Δt . The velocity is along the path and it
makes an angle θ with the radial direction. Hence, the perpendicular velocity is given by v perp = vsinθ . The planet moves
a distance Δs = vΔtsinθ projected along the direction perpendicular to r. Since the area of a triangle is one-half the base
(r) times the height (Δs) , for a small displacement, the area is given by ΔA = 1 rΔs .
2

Figure 8.23 The element of area ΔA swept out in time Δt as the planet moves
through angle Δϕ . The angle between the radial direction and

v is θ .

The areal velocity is simply the rate of change of area with time, so we have
1 r Δs (8.2)
areal velocity = ΔA = 2 = 1r v
Δt Δt 2
The fact that the areal velocity remains constant, then, implies that the product of the planet's distance from the Sun ( r ) and
its instantaneous speed ( v ) is a constant of the motion. Hence, the closer it is to the Sun, the faster it moves, and vice versa.
You can view an animated version (https://openstaxcollege.org/l/21animationgrav) of Figure 8.22,
and many other interesting animations as well, at the School of Physics (University of New South Wales) site.

This OpenStax book is available for free at https://legacy.cnx.org/content/col12169/1.46


Chapter 8 | Overview of the Solar System 285

Kepler’s Third Law


Kepler’s first two laws of planetary motion describe the shape of a planet’s orbit and allow us to calculate the speed of its
motion at any point in the orbit. Kepler was pleased to have discovered such fundamental rules, but they did not satisfy his
quest to fully understand planetary motions. He wanted to know why the orbits of the planets were spaced as they are and
to find a mathematical pattern in their movements—a “harmony of the spheres” as he called it. For many years he worked
to discover mathematical relationships governing planetary spacing and the time each planet took to go around the Sun.
In 1619, Kepler discovered a basic relationship to relate the planets’ orbits to their relative distances from the Sun. We
define a planet’s orbital period, (T), as the time it takes a planet to travel once around the Sun. Also, recall that a planet’s
semimajor axis, a, is equal to its average distance from the Sun. The relationship, now known as Kepler’s third law, says
that a planet’s orbital period squared is proportional to the semimajor axis of its orbit cubed, or
T 2 ∝ a3
When T (the orbital period) is measured in years, and a is expressed in a quantity known as an astronomical unit (AU),
the two sides of the formula are not only proportional but equal. One AU is the average distance between Earth and the Sun
and is approximately equal to 1.5 × 108 kilometers. In these units,
Kepler's Third Law in its Original Form
T 2 = a3 (8.3)

Kepler’s third law applies to all objects orbiting the Sun, including Earth, and provides a means for calculating their relative
distances from the Sun from the time they take to orbit. Let’s look at a specific example to illustrate how useful Kepler’s
third law is.
For instance, suppose you time how long Mars takes to go around the Sun (in Earth years). Kepler’s third law can then
be used to calculate Mars’ average distance from the Sun. Mars’ orbital period (1.88 Earth years) squared, or T2, is 1.882
= 3.53, and according to the equation for Kepler’s third law, this equals the cube of its semimajor axis, or a3. So what
number must be cubed to give 3.53? The answer is 1.52 (since 1.52 × 1.52 × 1.52 = 3.53). Thus, Mars’ semimajor axis in
astronomical units must be 1.52 AU. In other words, to go around the Sun in a little less than two years, Mars must be about
50% (half again) as far from the Sun as Earth is.

Example 8.2

Calculating Periods
Imagine an object is traveling around the Sun. What would be the orbital period of the object if its orbit has a
semimajor axis of 50 AU?
Solution
From Kepler’s third law, we know that (when we use units of years and AU)
T 2 = a3
If the object’s orbit has a semimajor axis of 50 AU (a = 50), we can cube 50 and then take the square root of the
result to get T:

T = a3
T = 50 × 50 × 50 = 125,000 = 353.6 years

Therefore, the orbital period of the object is about 350 years. This would place our hypothetical object beyond
the orbit of Pluto.

Kepler’s third law states that the square of the period is proportional to the cube of the semi-major axis of the orbit. When
written in the form of Equation 8.3, it is simple and easy to use for objects orbiting our Sun. The distances are measured
in AU and the periods in years. These units came naturally out of Kepler's work on the solar system. But, in fact, Kepler's
Third Law is much more general, and can be applied to bodies orbiting any large object. To do so, it makes the most sense to
measure quantities in SI units. Before we do so, however, we must realize that Kepler's Third Law in the form of Equation
8.3 does in fact contain a missing constant of proportionality, which is the the inverse of the mass of the Sun. The reason
that it does not appear in Equation 8.3 is that it is measured in units of "solar masses", which for the Sun has a value of
exactly 1. Nevertheless, the full equation for Kepler's Third Law should read:
286 Chapter 8 | Overview of the Solar System

Kepler's Third Law in Solar Units


3 (8.4)
T2 = a
M
where M is the mass of the large body (around which the satellite is in orbit) measured in units of solar masses. (Again, T
is measured in years, while a is measured in AU.)
Finally, to express all quantities in SI units, we can rewrite the equation in what is often referred to as "Newton's version":

Newton's Version of Kepler's Third Law


2 (8.5)
T 2 = 4π a 3
GM

In this equation, the period T is measured in seconds, the distance a in meters, and the mass M in kg. The constant G is
Newton's universal gravitational constant, which we will encounter in Chapter 9.

Example 8.3

Orbit of Halley’s Comet


Determine the semi-major axis of the orbit of Halley’s comet, given that it arrives at perihelion every 75.3 years.
If the perihelion is 0.586 AU, what is the aphelion?
Strategy
We are given the period, so we can rearrange Equation 8.5, solving for the semi-major axis. Since we know
the value for the perihelion, we can use the definition of the semi-major axis, given earlier in this section, to find
the aphelion. We note that 1 Astronomical Unit (AU) is the average radius of Earth’s orbit and is defined to be
1 AU = 1.50 × 10 11 m .
Solution
Rearranging Equation 8.5 and inserting the values of the period of Halley’s comet and the mass of the Sun, we
have

⎛ ⎞
1/3
a = GM2 T 2
⎝ 4π ⎠
⎛(6.67 × 10 −11 N · m 2 /kg 2)(2.00 × 10 30 kg) ⎞
1/3
=⎜ (75.3 yr × 365 days/yr × 24 hr/day × 3600 s/hr) 2⎟ .
⎝ 4π 2 ⎠

This yields a value of 2.67 × 10 12 m or 17.8 AU for the semi-major axis.


The semi-major axis is one-half the sum of the aphelion and perihelion, so we have

a = 1 (aphelion + perihelion)
2
aphelion = 2a − perihelion.

Substituting for the values, we found for the semi-major axis and the value given for the perihelion, we find the
value of the aphelion to be 35.0 AU.
Significance
Edmond Halley, a contemporary of Newton, first suspected that three comets, reported in 1531, 1607, and 1682,
were actually the same comet. Before Tycho Brahe made measurements of comets, it was believed that they were
one-time events, perhaps disturbances in the atmosphere, and that they were not affected by the Sun. Halley used
Newton’s new mechanics to predict his namesake comet’s return in 1758.

This OpenStax book is available for free at https://legacy.cnx.org/content/col12169/1.46


Chapter 8 | Overview of the Solar System 287

8.1 Check Your Understanding The nearly circular orbit of Saturn has an average radius of about 9.5 AU
and has a period of 30 years, whereas Uranus averages about 19 AU and has a period of 84 years. Is this
consistent with our results for Halley’s comet?
288 Chapter 8 | Overview of the Solar System

CHAPTER 8 REVIEW
KEY TERMS
aphelion farthest point from the Sun of an orbiting body; the corresponding term for the Moon’s farthest point from Earth
is the apogee
asteroid a stony or metallic object orbiting the Sun that is smaller than a major planet but that shows no evidence of an
atmosphere or of other types of activity associated with comets
astronomical unit (AU) the unit of length defined as the average distance between Earth and the Sun; this distance is
about 1.5 × 108 kilometers
comet a small body of icy and dusty matter that revolves about the Sun; when a comet comes near the Sun, some of its
material vaporizes, forming a large head of tenuous gas and often a tail
differentiation gravitational separation of materials of different density into layers in the interior of a planet or moon
eccentricity in an ellipse, the ratio of the distance between the foci to the major axis
ellipse a closed curve for which the sum of the distances from any point on the ellipse to two points inside (called the
foci) is always the same
focus (plural: foci) one of two fixed points inside an ellipse from which the sum of the distances to any point on the
ellipse is constant
giant planet any of the planets Jupiter, Saturn, Uranus, and Neptune in our solar system, or planets of roughly that mass
and composition in other planetary systems
Kepler’s first law each planet moves around the Sun in an orbit that is an ellipse, with the Sun at one focus of the ellipse
Kepler’s second law the straight line joining a planet and the Sun sweeps out equal areas in space in equal intervals of
time
Kepler’s third law the square of a planet’s orbital period is directly proportional to the cube of the semimajor axis of its
orbit
major axis the maximum diameter of an ellipse
meteor a small piece of solid matter that enters Earth’s atmosphere and burns up, popularly called a shooting star because
it is seen as a small flash of light
meteorite a portion of a meteor that survives passage through an atmosphere and strikes the ground
orbit the path of an object that is in revolution about another object or point
orbital period (P) the time it takes an object to travel once around the Sun
orbital speed the speed at which an object (usually a planet) orbits around the mass of another object; in the case of a
planet, the speed at which each planet moves along its ellipse
perihelion point of closest approach to the Sun of an orbiting body; the corresponding term for the Moon’s closest
approach to Earth is the perigee
planetesimals objects, from tens to hundreds of kilometers in diameter, that formed in the solar nebula as an
intermediate step between tiny grains and the larger planetary objects we see today; the comets and some asteroids
may be leftover planetesimals
semimajor axis half of the major axis of a conic section, such as an ellipse
solar nebula the cloud of gas and dust from which the solar system formed
terrestrial planet any of the planets Mercury, Venus, Earth, or Mars; sometimes the Moon is included in the list

This OpenStax book is available for free at https://legacy.cnx.org/content/col12169/1.46


Chapter 8 | Overview of the Solar System 289

SUMMARY
8.1 Overview of Our Planetary System

Our solar system currently consists of the Sun, eight planets, five dwarf planets, nearly 200 known moons, and a host of
smaller objects. The planets can be divided into two groups: the inner terrestrial planets and the outer giant planets. Pluto,
Eris, Haumea, and Makemake do not fit into either category; as icy dwarf planets, they exist in an ice realm on the fringes
of the main planetary system. The giant planets are composed mostly of liquids and gases. Smaller members of the solar
system include asteroids (including the dwarf planet Ceres), which are rocky and metallic objects found mostly between
Mars and Jupiter; comets, which are made mostly of frozen gases and generally orbit far from the Sun; and countless smaller
grains of cosmic dust. When a meteor survives its passage through our atmosphere and falls to Earth, we call it a meteorite.

8.2 Composition and Structure of Planets

The giant planets have dense cores roughly 10 times the mass of Earth, surrounded by layers of hydrogen and helium. The
terrestrial planets consist mostly of rocks and metals. They were once molten, which allowed their structures to differentiate
(that is, their denser materials sank to the center). The Moon resembles the terrestrial planets in composition, but most of
the other moons—which orbit the giant planets—have larger quantities of frozen ice within them. In general, worlds closer
to the Sun have higher surface temperatures. The surfaces of terrestrial planets have been modified by impacts from space
and by varying degrees of geological activity.

8.3 Origin of the Solar System

Regularities among the planets have led astronomers to hypothesize that the Sun and the planets formed together in a giant,
spinning cloud of gas and dust called the solar nebula. Astronomical observations show tantalizingly similar circumstellar
disks around other stars. Within the solar nebula, material first coalesced into planetesimals; many of these gathered together
to make the planets and moons. The remainder can still be seen as comets and asteroids. Probably all planetary systems have
formed in similar ways, but many exoplanet systems have evolved along quite different paths, as we will see in Cosmic
Samples and the Origin of the Solar System (https://legacy.cnx.org/content/m59870/latest/) .

8.4 Kepler's Laws of Planetary Motion


• All orbital motion follows the path of a conic section. Bound or closed orbits are either a circle or an ellipse;
unbounded or open orbits are either a parabola or a hyperbola.
• The areal velocity of any orbit is constant, meaning that the product of the body's speed and its distance from the
Sun is constant.
• The square of the period of an elliptical orbit is proportional to the cube of the semi-major axis of that orbit.

CONCEPTUAL QUESTIONS
4. Which type of planets have the most moons? Where did
8.3 Origin of the Solar System these moons likely originate?
1. Venus rotates backward and Uranus and Pluto spin
about an axis tipped nearly on its side. Based on what you 5. What is the difference between a meteor and a
learned about the motion of small bodies in the solar system meteorite?
and the surfaces of the planets, what might be the cause of
these strange rotations? 6. Explain our ideas about why the terrestrial planets are
rocky and have less gas than the giant planets.
2. What is the difference between a differentiated body
and an undifferentiated body, and how might that influence 7. Do all planetary systems look the same as our own?
a body’s ability to retain heat for the age of the solar
system?
8. What is comparative planetology and why is it useful to
astronomers?
3. What does a planet need in order to retain an
atmosphere? How does an atmosphere affect the surface of
9. What changed in our understanding of the Moon and
a planet and the ability of life to exist?
290 Chapter 8 | Overview of the Solar System

Moon-Earth system as a result of humans landing on the with Earth at a distance of about one city block from the
Moon’s surface? Sun. If you were to make a model of the distances in the
solar system to match your height, with the Sun at the top
10. If Earth was to be hit by an extraterrestrial object, of your head and Pluto at your feet, which planet would
where in the solar system could it come from and how be near your waist? How far down would the zone of the
would we know its source region? terrestrial planets reach?

11. List some reasons that the study of the planets has 22. Seasons are a result of the inclination of a planet’s
progressed more in the past few decades than any other axial tilt being inclined from the normal of the planet’s
branch of astronomy. orbital plane. For example, Earth has an axis tilt of 23.4°
(Appendix F (https://legacy.cnx.org/content/
m59998/latest/) ). Using information about just the
12. Imagine you are a travel agent in the next century.
inclination alone, which planets might you expect to have
An eccentric billionaire asks you to arrange a “Guinness
seasonal cycles similar to Earth, although different in
Book of Solar System Records” kind of tour. Where would
duration because orbital periods around the Sun are
you direct him to find the following (use this chapter and
different?
Appendix D:
A. the least-dense planet
B. the densest planet 23. Again using Appendix D, which planet(s) might you
C. the largest moon in the solar system expect not to have significant seasonal activity? Why?
D. excluding the jovian planets, the planet where
you would weigh the most on its surface (Hint: 24. Again using Appendix D, which planets might you
Weight is directly proportional to surface gravity.) expect to have extreme seasons? Why?
E. the smallest planet
F. the planet that takes the longest time to rotate 25. Using some of the astronomical resources in your
G. the planet that takes the shortest time to rotate college library or the Internet, find five names of features
H. the planet with a diameter closest to Earth’s on each of three other worlds that are named after real
I. the moon with the thickest atmosphere people. In a sentence or two, describe each of these people
J. the densest moon and what contributions they made to the progress of science
K. the most massive moon or human thought.

13. What characteristics do the worlds in our solar system 26. Explain why the planet Venus is differentiated, but
have in common that lead astronomers to believe that they asteroid Fraknoi, a very boring and small member of the
all formed from the same “mother cloud” (solar nebula)? asteroid belt, is not.

14. How do terrestrial and giant planets differ? List as 27. Would you expect as many impact craters per unit area
many ways as you can think of. on the surface of Venus as on the surface of Mars? Why or
why not?
15. Why are there so many craters on the Moon and so few
on Earth? 28. Interview a sample of 20 people who are not taking
an astronomy class and ask them if they can name a living
16. How do asteroids and comets differ? astronomer. What percentage of those interviewed were
able to name one? Typically, the two living astronomers
17. How and why is Earth’s Moon different from the the public knows these days are Stephen Hawking and Neil
larger moons of the giant planets? deGrasse Tyson. Why are they better known than most
astronomers? How would your result have differed if you
had asked the same people to name a movie star or a
18. Where would you look for some “original”
professional basketball player?
planetesimals left over from the formation of our solar
system?
29. Using Appendix B, complete the following table that
describes the characteristics of the Galilean moons of
19. What was the solar nebula like? Why did the Sun form
Jupiter, starting from Jupiter and moving outward in
at its center?
distance.

20. What can we learn about the formation of our solar


system by studying other stars? Explain.

21. Earlier in this chapter, we modeled the solar system

This OpenStax book is available for free at https://legacy.cnx.org/content/col12169/1.46


Chapter 8 | Overview of the Solar System 291

Moon Semimajor Diameter Density about a much larger mass, indicate where its speed is the
Axis (km3) (g/cm3) greatest and where it is the least. What conservation law
dictates this behavior? Indicate the directions of the force,
Io acceleration, and velocity at these points. Draw vectors for
these same three quantities at the two points where the
Europa
y-axis intersects (along the semi-minor axis) and from this
Ganymede determine whether the speed is increasing decreasing, or at
a max/min.
Callisto

Table A

This system has often been described as a mini solar


system. Why might this be so? If Jupiter were to represent
the Sun and the Galilean moons represented planets, which
moons could be considered more terrestrial in nature and
which ones more like gas/ice giants? Why? (Hint: Use the
values in your table to help explain your categorization.)

8.4 Kepler's Laws of Planetary Motion


30. Are Kepler’s laws purely descriptive, or do they
contain causal information?

31. In the diagram below for a satellite in an elliptical orbit

PROBLEMS
the Sun’s commonly listed value of 1.989 × 10 30 kg .
8.3 Origin of the Solar System
32. Calculate the density of Jupiter. Show your work. Is it
38. Io orbits Jupiter with an average radius of 421,700 km
more or less dense than Earth? Why?
and a period of 1.769 days. Based upon these data, what is
the mass of Jupiter?
33. Calculate the density of Saturn. Show your work. How
does it compare with the density of water? Explain how this
39. The “mean” orbital radius listed for astronomical
can be.
objects orbiting the Sun is typically not an integrated
average but is calculated such that it gives the correct
34. What is the density of Jupiter’s moon Europa (see period when applied to the equation for circular orbits.
Appendix D for data on moons)? Show your work. Given that, what is the mean orbital radius in terms of
aphelion and perihelion?
35. Look at Appendix F (https://legacy.cnx.org/
content/m59998/latest/) and Appendix G 40. The perihelion of Halley’s comet is 0.586 AU and
(https://legacy.cnx.org/content/m59999/latest/) and the aphelion is 17.8 AU. Given that its speed at perihelion
indicate the moon with a diameter that is the largest fraction is 55 km/s, what is the speed at aphelion (
of the diameter of the planet or dwarf planet it orbits.
1 AU = 1.496 × 10 11 m )? (Hint: You may use either
conservation of energy or angular momentum, but the latter
36. Barnard’s Star, the second closest star to us, is about
is much easier.)
56 trillion (5.6 × 1012) km away. Calculate how far it
would be using the scale model of the solar system given in
Overview of Our Planetary System. 41. The perihelion of the comet Lagerkvist is 2.61 AU and
it has a period of 7.36 years. Show that the aphelion for this
comet is 4.95 AU.
8.4 Kepler's Laws of Planetary Motion
42. What is the ratio of the speed at perihelion to that at
37. Calculate the mass of the Sun based on data for aphelion for the comet Lagerkvist in the previous problem?
average Earth’s orbit and compare the value obtained with
292 Chapter 8 | Overview of the Solar System

43. Eros has an elliptical orbit about the Sun, with a 1.78 AU. What is the period of its orbit?
perihelion distance of 1.13 AU and aphelion distance of

FOR FURTHER EXPLORATION


The following sites present introductory information and
8.3 Origin of the Solar System pictures about each of the worlds of our solar system:
Articles • NASA/JPL Solar System Exploration pages:
Davidson, K. “Carl Sagan’s Coming of Age.” Astronomy. http://solarsystem.nasa.gov/index.cfm.
(November 1999): 40. About the noted popularizer of • National Space Science Data Center Lunar and
science and how he developed his interest in astronomy. Planetary Science pages:
Garget, J. “Mysterious Microworlds.” Astronomy. (July http://nssdc.gsfc.nasa.gov/planetary/.
2005): 32. A quick tour of a number of the moons in the • Nine [now 8] Planets Solar System Tour:
solar system. http://www.nineplanets.org/.
Hartmann, W. “The Great Solar System Revision.” • Planetary Society solar system pages:
Astronomy. (August 1998): 40. How our views have http://www.planetary.org/explore/space-topics/
changed over the past 25 years. compare/.
Kross, J. “What’s in a Name?” Sky & Telescope. (May • Views of the Solar System by Calvin J. Hamilton:
1995): 28. How worlds are named. http://www.solarviews.com/eng/homepage.htm.
Rubin, A. “Secrets of Primitive Meteorites.” Scientific Videos
American. (February 2013): 36. What meteorites can teach
Brown Dwarfs and Free Floating Planets: When You Are
us about the environment in which the solar system formed.
Just Too Small to Be a Star: https://www.youtube.com/
Soter, S. “What Is a Planet?” Scientific American. (January watch?v=zXCDsb4n4KU. A nontechnical talk by Gibor
2007): 34. The IAU’s new definition of a planet in our solar Basri of the University of California at Berkeley, discussing
system, and what happened to Pluto as a result. some of the controversies about the meaning of the word
Talcott, R. “How the Solar System Came to Be.” “planet” (1:32:52).
Astronomy. (November 2012): 24. On the formation period In the Land of Enchantment: The Epic Story of the Cassini
of the Sun and the planets. Mission to Saturn: https://www.youtube.com/
Wood, J. “Forging the Planets: The Origin of our Solar watch?v=Vx135n8VFxY. A public lecture by Dr. Carolyn
System.” Sky & Telescope. (January 1999): 36. Good Porco that focuses mainly on the exploration of Saturn and
overview. its moons, but also presents an eloquent explanation of why
we explore the solar system (1:37:52).
Websites
Origins of the Solar System: http://www.pbs.org/wgbh/
Gazetteer of Planetary Nomenclature: nova/space/origins-solar-system.html. A video from PBS
http://planetarynames.wr.usgs.gov/. Outlines the rules for that focuses on the evidence from meteorites, narrated by
naming bodies and features in the solar system. Neil deGrasse Tyson (13:02).
Planetary Photojournal: http://photojournal.jpl.nasa.gov/ To Scale: The Solar System: https://www.youtube.com/
index.html. This NASA site features thousands of the best watch?t=84&v=zR3Igc3Rhfg. Constructing a scale model
images from planetary exploration, with detailed captions of the solar system in the Nevada desert (7:06).
and excellent indexing. You can find images by world,
feature name, or mission, and download them in a number
of formats. And the images are copyright-free because your
tax dollars paid for them.

This OpenStax book is available for free at https://legacy.cnx.org/content/col12169/1.46


Chapter 9 | Newton's Synthesis 293

9 | NEWTON'S SYNTHESIS
294 Chapter 9 | Newton's Synthesis

This OpenStax book is available for free at https://legacy.cnx.org/content/col12169/1.46


Chapter 10 | Newton's Laws for Rotations 295

10 | NEWTON'S LAWS FOR


ROTATIONS
296 Chapter 10 | Newton's Laws for Rotations

This OpenStax book is available for free at https://legacy.cnx.org/content/col12169/1.46


Chapter 11 | Work and Energy 297

11 | WORK AND ENERGY


298 Chapter 11 | Work and Energy

This OpenStax book is available for free at https://legacy.cnx.org/content/col12169/1.46


Chapter 12 | Momentum in One Dimension 299

12 | MOMENTUM IN ONE
DIMENSION
300 Chapter 12 | Momentum in One Dimension

This OpenStax book is available for free at https://legacy.cnx.org/content/col12169/1.46


Chapter 13 | Angular Momentum 301

13 | ANGULAR MOMENTUM
302 Chapter 13 | Angular Momentum

This OpenStax book is available for free at https://legacy.cnx.org/content/col12169/1.46


Chapter 14 | Reflection and Refraction 303

14 | REFLECTION AND
REFRACTION
304 Chapter 14 | Reflection and Refraction

This OpenStax book is available for free at https://legacy.cnx.org/content/col12169/1.46


Chapter 15 | Dispersion, Luminosity and Apparent Brightness 305

15 | DISPERSION,
LUMINOSITY AND
APPARENT BRIGHTNESS
306 Chapter 15 | Dispersion, Luminosity and Apparent Brightness

This OpenStax book is available for free at https://legacy.cnx.org/content/col12169/1.46


Chapter 16 | Image Formation 307

16 | IMAGE FORMATION
308 Chapter 16 | Image Formation

This OpenStax book is available for free at https://legacy.cnx.org/content/col12169/1.46


Chapter 17 | Wave Properties of Light 309

17 | WAVE PROPERTIES OF
LIGHT
310 Chapter 17 | Wave Properties of Light

This OpenStax book is available for free at https://legacy.cnx.org/content/col12169/1.46


Chapter 18 | Interference and Diffraction 311

18 | INTERFERENCE AND
DIFFRACTION
312 Chapter 18 | Interference and Diffraction

This OpenStax book is available for free at https://legacy.cnx.org/content/col12169/1.46


Chapter 19 | Spectroscopy and the Doppler Effect 313

19 | SPECTROSCOPY AND
THE DOPPLER EFFECT
314 Chapter 19 | Spectroscopy and the Doppler Effect

This OpenStax book is available for free at https://legacy.cnx.org/content/col12169/1.46


Chapter 20 | Quantum Optics 315

20 | QUANTUM OPTICS
316 Chapter 20 | Quantum Optics

This OpenStax book is available for free at https://legacy.cnx.org/content/col12169/1.46


Chapter 21 | Temperature 317

21 | TEMPERATURE
318 Chapter 21 | Temperature

This OpenStax book is available for free at https://legacy.cnx.org/content/col12169/1.46


Chapter 22 | The First Law of Thermodynamics 319

22 | THE FIRST LAW OF


THERMODYNAMICS
320 Chapter 22 | The First Law of Thermodynamics

This OpenStax book is available for free at https://legacy.cnx.org/content/col12169/1.46


Chapter 23 | Kinetic Theory 321

23 | KINETIC THEORY
322 Chapter 23 | Kinetic Theory

This OpenStax book is available for free at https://legacy.cnx.org/content/col12169/1.46


Chapter 24 | The Nebular Theory of Solar System Formation 323

24 | THE NEBULAR THEORY


OF SOLAR SYSTEM
FORMATION
324 Chapter 24 | The Nebular Theory of Solar System Formation

This OpenStax book is available for free at https://legacy.cnx.org/content/col12169/1.46


Chapter 25 | Nuclear Physics and Comparative Geology 325

25 | NUCLEAR PHYSICS
AND COMPARATIVE
GEOLOGY
326 Chapter 25 | Nuclear Physics and Comparative Geology

This OpenStax book is available for free at https://legacy.cnx.org/content/col12169/1.46


Chapter 26 | Comparative Atmospheres 327

26 | COMPARATIVE
ATMOSPHERES
328 Chapter 26 | Comparative Atmospheres

This OpenStax book is available for free at https://legacy.cnx.org/content/col12169/1.46


Chapter 27 | Exoplanets 329

27 | EXOPLANETS
330 Chapter 27 | Exoplanets

This OpenStax book is available for free at https://legacy.cnx.org/content/col12169/1.46


Chapter 28 | The Sun 331

28 | THE SUN
332 Chapter 28 | The Sun

This OpenStax book is available for free at https://legacy.cnx.org/content/col12169/1.46


Chapter 29 | Stellar Properties 333

29 | STELLAR PROPERTIES
334 Chapter 29 | Stellar Properties

This OpenStax book is available for free at https://legacy.cnx.org/content/col12169/1.46


Chapter 30 | Stellar Life Cycles 335

30 | STELLAR LIFE CYCLES


336 Chapter 30 | Stellar Life Cycles

This OpenStax book is available for free at https://legacy.cnx.org/content/col12169/1.46


Chapter 31 | The Milky Way as a Prototype Galaxy 337

31 | THE MILKY WAY AS A


PROTOTYPE GALAXY
338 Chapter 31 | The Milky Way as a Prototype Galaxy

This OpenStax book is available for free at https://legacy.cnx.org/content/col12169/1.46


Chapter 32 | Survey of Galaxies 339

32 | SURVEY OF GALAXIES
340 Chapter 32 | Survey of Galaxies

This OpenStax book is available for free at https://legacy.cnx.org/content/col12169/1.46


Chapter 33 | Galactic Evolution 341

33 | GALACTIC EVOLUTION
342 Chapter 33 | Galactic Evolution

This OpenStax book is available for free at https://legacy.cnx.org/content/col12169/1.46


Chapter 34 | Big Bang Cosmology 343

34 | BIG BANG
COSMOLOGY
344 Chapter 34 | Big Bang Cosmology

This OpenStax book is available for free at https://legacy.cnx.org/content/col12169/1.46


Chapter 35 | The Fate of the Universe 345

35 | THE FATE OF THE


UNIVERSE
346 Chapter 35 | The Fate of the Universe

This OpenStax book is available for free at https://legacy.cnx.org/content/col12169/1.46


Appendix A 347

APPENDIX A | UNITS
Quantity Common Unit Unit in Terms of Base SI
Symbol Units
Acceleration →
a m/s2 m/s2

Amount of substance n mole mol


Angle θ, ϕ radian (rad)

Angular acceleration →
α rad/s2 s−2

Angular frequency ω rad/s s−1



Angular momentum L kg · m 2 /s kg · m 2 /s

Angular velocity →
ω rad/s s−1

Area A m2 m2
Atomic number Z
Capacitance C farad (F) A 2 · s 4 /kg · m 2

Charge q, Q, e coulomb (C) A·s


Charge density:
Line λ C/m A · s/m
Surface σ C/m2 A · s/m 2
Volume ρ C/m3 A · s/m 3
Conductivity σ 1/Ω · m A 2 · s 3 /kg · m 3

Current I ampere A
Current density → A/m2 A/m2
J

Density ρ kg/m3 kg/m3

Dielectric constant κ
Electric dipole moment →
p C·m A·s·m


Electric field E N/C kg · m/A · s 3

Electric flux Φ N · m 2 /C kg · m 3 /A · s 3

Electromotive force ε volt (V) kg · m 2 /A · s 3

Energy E,U,K joule (J) kg · m 2 /s 2

Entropy S J/K kg · m 2 /s 2 · K

Table A1 Units Used in Physics (Fundamental units in bold)


348 Appendix A

Quantity Common Unit Unit in Terms of Base SI


Symbol Units

Force F newton (N) kg · m/s 2

Frequency f hertz (Hz) s−1


Heat Q joule (J) kg · m 2 /s 2

Inductance L henry (H) kg · m 2 /A 2 · s 2

Length: ℓ, L meter m

Displacement Δx, Δ →
r

Distance d, h
Position x, y, z, →
r
→ N · J/T
Magnetic dipole μ A · m2
moment

tesla(T) = ⎛⎝Wb/m 2⎞⎠



Magnetic field B kg/A · s 2

Magnetic flux Φm weber (Wb) kg · m 2 /A · s 2

Mass m, M kilogram kg
Molar specific heat C J/mol · K kg · m 2 /s 2 · mol · K

Moment of inertia I kg · m 2 kg · m 2

Momentum →
p kg · m/s kg · m/s

Period T s s
Permeability of free μ0 2
N/A =(H/m) kg · m/A 2 · s 2
space
Permittivity of free ε0 C 2 /N · m 2 =(F/m) A 2 · s 4 /kg · m 3
space
Potential V volt(V) = (J/C) kg · m 2 /A · s 3

Power P watt(W) = (J/s) kg · m 2 /s 3

Pressure p pascal(Pa) = ⎛⎝N/m 2⎞⎠ kg/m · s 2

Resistance R ohm(Ω) = (V/A) kg · m 2 /A 2 · s 3

Specific heat c J/kg · K m 2 /s 2 · K


Speed ν m/s m/s

Temperature T kelvin K
Time t second s
→ N·m
Torque τ kg · m 2 /s 2

Table A1 Units Used in Physics (Fundamental units in bold)

This OpenStax book is available for free at https://legacy.cnx.org/content/col12169/1.46


Appendix A 349

Quantity Common Unit Unit in Terms of Base SI


Symbol Units
Velocity →
v m/s m/s

Volume V m3 m3
Wavelength λ m m

Work W joule(J) = (N · m) kg · m 2 /s 2

Table A1 Units Used in Physics (Fundamental units in bold)


350 Appendix A

This OpenStax book is available for free at https://legacy.cnx.org/content/col12169/1.46


Appendix B 351

APPENDIX B | UNIT
CONVERSIONS
m cm km
1 meter 1 102 10−3
1 centimeter 10−2 1 10−5
1 kilometer 103 105 1
1 inch 2.540 × 10 −2 2.540 2.540 × 10 −5
1 foot 0.3048 30.48 3.048 × 10 −4
1 mile 1609 1.609 × 10 4 1.609

1 angstrom 10−10
1 fermi 10−15
1 light-year 9.461 × 10 15 9.461 × 10 12
1 parsec 3.084 × 10 16 3.084 × 10 13
in. ft mi
1 meter 39.37 3.281 6.214 × 10 −4
1 centimeter 0.3937 3.281 × 10 −2 6.214 × 10 −6
1 kilometer 3.937 × 10 4 3.281 × 10 3 0.6214

1 inch 1 8.333 × 10 −2 1.578 × 10 −5


1 foot 12 1 1.894 × 10 −4
1 mile 6.336 × 10 4 5280 1

Table B1 Length

Area
1 cm 2 = 0.155 in. 2

1 m 2 = 10 4 cm 2 = 10.76 ft 2

1 in. 2 = 6.452 cm 2

1 ft 2 = 144 in. 2 = 0.0929 m 2


Volume
1 liter = 1000 cm 3 = 10 −3 m 3 = 0.03531 ft 3 = 61.02 in. 3

1 ft 3 = 0.02832 m 3 = 28.32 liters = 7.477 gallons

1 gallon = 3.788 liters


352 Appendix B

s min h day yr
1 second 1 1.667 × 10 −2 2.778 × 10 −4 1.157 × 10 −5 3.169 × 10 −8
1 minute 60 1 1.667 × 10 −2 6.944 × 10 −4 1.901 × 10 −6
1 hour 3600 60 1 4.167 × 10 −2 1.141 × 10 −4
1 day 8.640 × 10 4 1440 24 1 2.738 × 10 −3
1 year 3.156 × 10 7 5.259 × 10 5 8.766 × 10 3 365.25 1

Table B2 Time

m/s cm/s ft/s mi/h


1 meter/second 1 102 3.281 2.237

3.281 × 10 −2 2.237 × 10 −2
−2
1 centimeter/second 10 1

1 foot/second 0.3048 30.48 1 0.6818


1 mile/hour 0.4470 44.70 1.467 1

Table B3 Speed

Acceleration
1 m/s 2 = 100 cm/s 2 = 3.281 ft/s 2

1 cm/s 2 = 0.01 m/s 2 = 0.03281 ft/s 2

1 ft/s 2 = 0.3048 m/s 2 = 30.48 cm/s 2

1 mi/h · s = 1.467 ft/s 2

kg g slug u
1 kilogram 1 103 6.852 × 10 −2 6.024 × 10 26
1 gram 10−3 1 6.852 × 10 −5 6.024 × 10 23
1 slug 14.59 1.459 × 10 4 1 8.789 × 10 27
1 atomic mass unit 1.661 × 10 −27 1.661 × 10 −24 1.138 × 10 −28 1

1 metric ton 1000

Table B4 Mass

N dyne lb
1 newton 1 105 0.2248

1 dyne 10−5 1 2.248 × 10 −6


1 pound 4.448 4.448 × 10 5 1

Table B5 Force

This OpenStax book is available for free at https://legacy.cnx.org/content/col12169/1.46


Appendix B 353

Pa dyne/cm2 atm cmHg lb/in.2


1 pascal 1 10 9.869 × 10 −6 7.501 × 10 −4 1.450 × 10 −4
1 dyne/ 10−1 1 9.869 × 10 −7 7.501 × 10 −5 1.450 × 10 −5
centimeter2
1 atmosphere 1.013 × 10 5 1.013 × 10 6 1 76 14.70

1 centimeter 1.333 × 10 3 1.333 × 10 4 1.316 × 10 −2 1 0.1934


mercury*
1 pound/inch2 6.895 × 10 3 6.895 × 10 4 6.805 × 10 −2 5.171 1

1 bar 105
1 torr 1 (mmHg)

*Where the acceleration due to gravity is 9.80665 m/s2 and the temperature is 0°C

Table B6 Pressure

J erg ft.lb
1 joule 1 107 0.7376
1 erg 10−7 1 7.376 × 10 −8
1 foot-pound 1.356 1.356 × 10 7 1

1 electron-volt 1.602 × 10 −19 1.602 × 10 -12 1.182 × 10 −19


1 calorie 4.186 4.186 × 10 7 3.088

1 British thermal unit 1.055 × 10 3 1.055 × 10 10 7.779 × 10 2


1 kilowatt-hour 3.600 × 10 6
eV cal Btu
1 joule 6.242 × 10 18 0.2389 9.481 × 10 −4
1 erg 6.242 × 10 11 2.389 × 10 −8 9.481 × 10 −11
1 foot-pound 8.464 × 10 18 0.3239 1.285 × 10 −3
1 electron-volt 1 3.827 × 10 −20 1.519 × 10 −22
1 calorie 2.613 × 10 19 1 3.968 × 10 −3
1 British thermal unit 6.585 × 10 21 2.520 × 10 2 1

Table B7 Work, Energy, Heat

Power
1 W = 1 J/s
1 hp = 746 W = 550 ft · lb/s

1 Btu/h = 0.293 W
Angle
354 Appendix B

1 rad = 57.30° = 180°/π


1° = 0.01745 rad = π/180 rad
1 revolution = 360° = 2π rad
1 rev/min(rpm) = 0.1047 rad/s

This OpenStax book is available for free at https://legacy.cnx.org/content/col12169/1.46


Appendix C 355

APPENDIX C |
FUNDAMENTAL PHYSICS
CONSTANTS
Quantity Symbol Value
Atomic mass unit u 1.660 538 782 (83) × 10 −27 kg
931.494 028 (23) MeV/c 2

Avogadro’s number NA 6.022 141 79 (30) × 10 23 particles/mol

Bohr magneton μ B = eℏ 9.274 009 15 (23) × 10 −24 J/T


2m e

Bohr radius
a0 = ℏ2 5.291 772 085 9 (36) × 10 −11 m
me e2 ke

Boltzmann’s constant kB = R 1.380 650 4 (24) × 10 −23 J/K


NA

Compton wavelength λ C = mh c 2.426 310 217 5 (33) × 10 −12 m


e

Coulomb constant ke = 1 8.987 551 788... × 10 9 N · m 2 /C 2 (exact)


4πε 0

Deuteron mass md 3.343 583 20 (17) × 10 −27 kg


2.013 553 212 724(78) u
1875.612 859 MeV/c 2

Electron mass me 9.109 382 15 (45) × 10 −31 kg


5.485 799 094 3(23) × 10 −4 u
0.510 998 910 (13) MeV/c 2

Electron volt eV 1.602 176 487 (40) × 10 −19 J

Elementary charge e 1.602 176 487 (40) × 10 −19 C

Gas constant R 8.314 472 (15) J/mol · K

Gravitational constant G 6.674 28 (67) × 10 −11 N · m 2 /kg 2

Hydrogen atom mass mH 1.673 × 10 −27 kg

Table C1 Fundamental Constants Note: These constants are the values recommended in 2006
by CODATA, based on a least-squares adjustment of data from different measurements. The
numbers in parentheses for the values represent the uncertainties of the last two digits.
356 Appendix C

Quantity Symbol Value


Neutron mass mn 1.674 927 211 (84) × 10 −27 kg
1.008 664 915 97 (43) u
939.565 346 (23) MeV/c 2

Nuclear magneton μ n = eℏ 5.050 783 24 (13) × 10 −27 J/T


2m p

Permeability of free space μ0 4π × 10 −7 T · m/A(exact)

Permittivity of free space ε0 = 1 8.854 187 817... × 10 −12 C 2 /N · m 2 (exact)


μ0 c2

Planck’s constant h 6.626 068 96 (33) × 10 −34 J · s


ℏ= h 1.054 571 628 (53) × 10 −34 J · s

Proton mass mp 1.672 621 637 (83) × 10 −27 kg
1.007 276 466 77 (10) u
938.272 013 (23) MeV/c 2

Rydberg constant RH 1.097 373 156 852 7 (73) × 10 7 m −1

Speed of light in vacuum c 2.997 924 58 × 10 8 m/s (exact)

Stefan-Boltzmann constant σ 5.670 × 10 −8 W/m 2 · K 4


Wien's Law constant λ max T 2.898 × 10 −3 m/K

Table C1 Fundamental Constants Note: These constants are the values recommended in 2006
by CODATA, based on a least-squares adjustment of data from different measurements. The
numbers in parentheses for the values represent the uncertainties of the last two digits.

Useful combinations of constants for calculations:


hc = 12,400 eV · Å = 1240 eV · nm = 1240 MeV · fm

ℏc = 1973 eV · Å = 197.3 eV · nm = 197.3 MeV · fm

k e e 2 = 14.40 eV · Å = 1.440 eV · nm = 1.440 MeV · fm


k B T = 0.02585 eV at T = 300 K

This OpenStax book is available for free at https://legacy.cnx.org/content/col12169/1.46


Appendix D 357

APPENDIX D |
ASTRONOMICAL DATA
Astronomical Constants
Name Value
astronomical unit (AU) 1.496 × 1011 m
Light-year (ly) 9.461 × 1015 m
parsec (pc) 3.086 × 1016 m = 3.262 light-years
sidereal year (y) 3.156 × 107 s
mass of Earth (MEarth) 5.974 × 1024 kg
equatorial radius of Earth (REarth) 6.378 × 106 m
obliquity of ecliptic 23.4° 26’
surface gravity of Earth (g) 9.807 m/s2
escape velocity of Earth (vEarth) 1.119 × 104 m/s
mass of Sun (MSun) 1.989 × 1030 kg
equatorial radius of Sun (RSun) 6.960 × 108 m
luminosity of Sun (LSun) 3.85 × 1026 W
solar constant (flux of energy 1.368 × 103 W/m2
received at Earth) (S)
Hubble constant (H0) approximately 20 km/s per million light-years, or approximately
70 km/s per megaparsec

Table D1

Celestial Mean Distance Period of Revolution Period of Eccentricity


Object from Sun (million (d = days) (y = years) Rotation at of Orbit
km) Equator
Sun − − 27 d −
Mercury 57.9 88 d 59 d 0.206
Venus 108.2 224.7 d 243 d 0.007
Earth 149.6 365.26 d 23 h 56 min 4 s 0.017
Mars 227.9 687 d 24 h 37 min 23 s 0.093
Jupiter 778.4 11.9 y 9 h 50 min 30 s 0.048
Saturn 1426.7 29.5 6 10 h 14 min 0.054
Uranus 2871.0 84.0 y 17 h 14 min 0.047
Neptune 4498.3 164.8 y 16 h 0.009

Table D2 Comparative Planetary Data


358 Appendix D

Celestial Mean Distance Period of Revolution Period of Eccentricity


Object from Sun (million (d = days) (y = years) Rotation at of Orbit
km) Equator
Earth’s 149.6 (0.386 from 27.3 d 27.3 d 0.055
Moon Earth)
Celestial Equatorial Diameter Mass (Earth = 1) Density (g/cm3)
Object (km)
Sun 1,392,000 333,000.00 1.4
Mercury 4879 0.06 5.4
Venus 12,104 0.82 5.2
Earth 12,756 1.00 5.5
Mars 6794 0.11 3.9
Jupiter 142,984 317.83 1.3
Saturn 120,536 95.16 0.7
Uranus 51,118 14.54 1.3
Neptune 49,528 17.15 1.6
Earth’s 3476 0.01 3.3
Moon

Table D2 Comparative Planetary Data

D1 Physical and Orbital Data for the Planets


Physical Data for the Major Planets
Major Mean Mean Mass Mean Rotation Inclination Surface Velocity
Planet Diameter Diameter (Earth Density Period of Equator Gravity of
(km) (Earth = = 1) (g/cm3) (d) to Orbit (°) (Earth Escape
1) = 1[g]) (km/s)
Mercury 4879 0.38 0.055 5.43 58. 0.0 0.38 4.3
Venus 12,104 0.95 0.815 5.24 −243. 177 0.90 10.4
Earth 12,756 1.00 1.00 5.51 1.000 23.4 1.00 11.2
Mars 6779 0.53 0.11 3.93 1.026 25.2 0.38 5.0
Jupiter 140,000 10.9 318 1.33 0.414 3.1 2.53 60.
Saturn 117,000 9.13 95.2 0.69 0.440 26.7 1.07 36.
Uranus 50,700 3.98 14.5 1.27 −0.718 97.9 0.89 21.
Neptune 49,200 3.86 17.2 1.64 0.671 29.6 1.14 23.

Table D3

This OpenStax book is available for free at https://legacy.cnx.org/content/col12169/1.46


Appendix D 359

Physical Data for Well-Studied Dwarf Planets


Well- Diameter Diameter Mass Mean Rotation Inclination Surface Velocity
Studied (km) (Earth = (Earth Density Period of Equator Gravity of
Dwarf 1) = 1) (g/cm3) (d) to Orbit (°) (Earth Escape
Planet = 1[g]) (km/s)
Ceres 950 0.07 0.0002 2.2 0.378 3 0.03 0.5
Pluto 2470 0.18 0.0024 1.9 −6.387 122 0.06 1.3
Haumea 1700 0.13 0.0007 3 0.163 — — 0.8
Makemake 1400 0.11 0.0005 2 0.321 — — 0.8
[1]
Eris 2326 0.18 0.0028 2.5 1.25 — — 1.4

Table D4

Orbital Data for the Major Planets


Major Semimajor Semimajor Sidereal Sidereal Mean Orbital Inclination
Planet Axis (AU) Axis (106 Period Period Orbital Eccentricity of Orbit to
km) (y) (d) Speed Ecliptic (°)
(km/s)
Mercury 0.39 58 0.24 88.0 47.9 0.206 7.0
Venus 0.72 108 0.6 224.7 35.0 0.007 3.4
Earth 1.00 149 1.00 365.2 29.8 0.017 0.0
Mars 1.52 228 1.88 687.0 24.1 0.093 1.9
Jupiter 5.20 778 11.86 — 13.1 0.048 1.3
Saturn 9.54 1427 29.46 — 9.6 0.056 2.5
Uranus 19.19 2871 84.01 — 6.8 0.046 0.8
Neptune 30.06 4497 164.82 — 5.4 0.010 1.8

Table D5

Orbital Data for Well-Studied Dwarf Planets


Well- Semimajor Semimajor Sidereal Mean Orbital Inclination
Studied Axis (AU) Axis (106 Period Orbital Eccentricity of Orbit to
Dwarf km) (y) Speed Ecliptic (°)
Planet (km/s)
Ceres 2.77 414.0 4.6 18 0.08 11
Pluto 39.5 5915 248.6 4.7 0.25 17
Haumea 43.1 6452 283.3 4.5 0.19 28
Makemake 45.8 6850 309.9 4.4 0.16 29
Eris 68.0 10,120 560.9 3.4 0.44 44

Table D6

D2 Selected Moons of the Planets


Note: As this book goes to press, nearly two hundred moons are now known in the solar system and more are being

1. This measurement is quite uncertain.


360 Appendix D

discovered on a regular basis. Of the major planets, only Mercury and Venus do not have moons. In addition to moons
of the planets, there are many moons of asteroids. In this appendix, we list only the largest and most interesting objects
that orbit each planet (including dwarf planets). The number given for each planet is discoveries through 2015. For
further information see https://solarsystem.nasa.gov/planets/solarsystem/moons and https://en.wikipedia.org/wiki/
List_of_natural_satellites.

Selected Moons of the Planets


Planet Satellite Discovery Semimajor Period Diameter Mass Density
(moons) Name Axis (km × (d) (km) (1020 (g/cm3)
1000) kg)
Earth (1) Moon — 384 27.32 3476 735 3.3
Mars (2) Phobos Hall (1877) 9.4 0.32 23 1 × 10−4 2.0
Deimos Hall (1877) 23.5 1.26 13 2 × 10−5 1.7
Jupiter Amalthea Barnard 181 0.50 200 — —
(67) (1892)
Thebe Voyager 222 0.67 90 — —
(1979)
Io Galileo 422 1.77 3630 894 3.6
(1610)
Europa Galileo 671 3.55 3138 480 3.0
(1610)
Ganymede Galileo 1070 7.16 5262 1482 1.9
(1610)
Callisto Galileo 1883 16.69 4800 1077 1.9
(1610)
Himalia Perrine 11,460 251 170 — —
(1904)
Saturn Pan Voyager 133.6 0.58 20 3 × 10−5 —
(62) (1985)
Atlas Voyager 137.7 0.60 40 — —
(1980)
Prometheus Voyager 139.4 0.61 80 — —
(1980)
Pandora Voyager 141.7 0.63 100 — —
(1980)
Janus Dollfus 151.4 0.69 190 — —
(1966)
Epimetheus Fountain, 151.4 0.69 120 — —
Larson
(1980)
Mimas Herschel 186 0.94 394 0.4 1.2
(1789)
Enceladus Herschel 238 1.37 502 0.8 1.2
(1789)
Tethys Cassini 295 1.89 1048 7.5 1.3
(1684)

Table D7

This OpenStax book is available for free at https://legacy.cnx.org/content/col12169/1.46


Appendix D 361

Selected Moons of the Planets


Planet Satellite Discovery Semimajor Period Diameter Mass Density
(moons) Name Axis (km × (d) (km) (1020 (g/cm3)
1000) kg)
Dione Cassini 377 2.74 1120 11 1.3
(1684)
Rhea Cassini 527 4.52 1530 25 1.3
(1672)
Titan Huygens 1222 15.95 5150 1346 1.9
(1655)
Hyperion Bond, 1481 21.3 270 — —
Lassell
(1848)
Iapetus Cassini 3561 79.3 1435 19 1.2
(1671)
Phoebe Pickering 12,950 550 220 — —
(1898) (R)[2]
Uranus Puck Voyager 86.0 0.76 170 — —
(27) (1985)
Miranda Kuiper 130 1.41 485 0.8 1.3
(1948)
Ariel Lassell 191 2.52 1160 13 1.6
(1851)
Umbriel Lassell 266 4.14 1190 13 1.4
(1851)
Titania Herschel 436 8.71 1610 35 1.6
(1787)
Oberon Herschel 583 13.5 1550 29 1.5
(1787)
Neptune Despina Voyager 53 0.33 150 — —
(14) (1989)
Galatea Voyager 62 0.40 150 — —
(1989)
Larissa Voyager 118 1.12 400 — —
(1989)
Triton Lassell 355 5.88 2720 220 2.1
(1846) (R)[3]
Nereid Kuiper 5511 360 340 — —
(1949)
Pluto (5) Charon Christy 19.7 6.39 1200 — 1.7
(1978)

Table D7

2. R stands for retrograde rotation (backward from the direction that most objects in the solar system revolve
and rotate).
3. R stands for retrograde rotation (backward from the direction that most objects in the solar system revolve
and rotate).
362 Appendix D

Selected Moons of the Planets


Planet Satellite Discovery Semimajor Period Diameter Mass Density
(moons) Name Axis (km × (d) (km) (1020 (g/cm3)
1000) kg)
Styx Showalter 42 20 20 — —
et al
(2012)
Nix Weaver et 48 24 46 — 2.1
al (2005)
Kerberos Showalter 58 24 28 — 1.4
et al
(2011)
Hydra Weaver et 65 38 61 — 0.8
al (2005)
Eris (1) Dysnomea Brown et al 38 16 684 — —
(2005)
Makemake (MK2) Parker et — — 160 — —
(1) al (2016)
Haumea Hi’iaka Brown et al 50 49 400 — —
(2) (2005)
Namaka Brown et al 39 35 200 — —
(2005)

Table D7

D3 The Nearest Stars, Brown Dwarfs, and White Dwarfs


The Nearest Stars, Brown Dwarfs, and White Dwarfs
Star System Discovery Distance Spectral Location: Location: Luminosity
Name (light- Type RA[4] Dec[5] (Sun = 1)
year)
Sun — G2 V — — 1
1 1 Proxima 4.2 M5.5 V 14 29 −62 40 5 × 10−5
Centauri
2 2 Alpha 4.4 G2 V 14 39 −60 50 1.5
Centauri A
3 Alpha 4.4 K2 IV 14 39 −60 50 0.5
Centauri B
4 3 Barnard’s Star 6.0 M4 V 17 57 +04 42 4.4 × 10−4
5 4 Wolf 359 7.8 M6 V 10 56 +07 00 2 × 10−5
6 5 Lalande 21 8.3 M2 V 11 03 +35 58 5.7 × 10−3
185
7 6 Sirius A 8.6 A1 V 06 45 −16 42 23.1

Table D8

4. Location (right ascension) given for Epoch 2000.0


5. Location (declination) given for Epoch 2000.0

This OpenStax book is available for free at https://legacy.cnx.org/content/col12169/1.46


Appendix D 363

The Nearest Stars, Brown Dwarfs, and White Dwarfs


Star System Discovery Distance Spectral Location: Location: Luminosity
Name (light- Type RA Dec (Sun = 1)
year)
8 Sirius B 8.6 DA2[6] 06 45 −16 43 2.5 × 10−3
9 7 Luyten 726-8 8.7 M5.5 V 01 39 −17 57 6 × 10−5
A
10 Luyten 726-8 8.7 M6 V 01 39 −17 57 4 × 10−5
B (UV Ceti)
11 8 Ross 154 9.7 M.05 V 18 49 −23 50 5 × 10−4
12 9 Ross 248 (HH 10.3 M5.5 V 23 41 +44 10 1.0 × 10−4
Andromedae)
13 10 Epsilon 10.5 K2 V 03 32 −09 27 0.29
Eridani
14 11 Lacaille 9352 10.7 M0.5 V 23 05 −35 51 0.011
15 12 Ross 128 (FI 10.9 M4 V 11 47 +00 48 3.4 × 10−4
Virginis)
16 13 Luyten 789-6 11.3 M5 V 22 38 −15 17 5 × 10−5
A (EZ Aquarii
A)
17 Luyten 789-6 11.3 M5.5 V 22 38 −15 15 5 × 10−5
B (EZ Aquarii
B)
18 Luyten 789-6 11.3 M6.5 V 22 38 −15 17 2 × 10−5
C (EZ Aquarii
C)
19 14 61 Cygni A 11.4 K5 V 21 06 +38 44 0.086
20 61 Cygni B 11.4 K7 V 21 06 +38 44 0.041
21 15 Procyon A 11.4 F51V 07 39 +05 13 7.38
[7]
22 Procyon B 11.4 wd 07 39 +05 13 5.5 × 10−4
23 16 Sigma 2398 A 11.5 M3 V 18 42 +59 37 0.003
24 Sigma 2398 B 11.5 M3.5 V 18 42 +59 37 1.4 × 10−3
25 17 Groombridge 11.6 M1.5 V 00 18 +44 01 6.4 × 10−3
34 A (GX
Andromedae)
26 Groombridge 11.6 M3.5 V 00 18 +44 01 4.1 × 10−4
34 B (GQ
Andromedae)
27 18 Epsilon Indi A 11.8 K5 V 22 03 −56 46 0.150
[8]
28 Epsilon Indi 11.7 T1 22 04 −56 46 —
Ba

Table D8

6. White dwarf stellar remnant


7. White dwarf stellar remnant
364 Appendix D

The Nearest Stars, Brown Dwarfs, and White Dwarfs


Star System Discovery Distance Spectral Location: Location: Luminosity
Name (light- Type RA Dec (Sun = 1)
year)
29 Epsilon Indi 11.7 T6[9] 22 04 −56 46 —
Bb
30 19 G 51-15 (DX 11.8 M6.5 V 08 29 +26 46 1 × 10−5
Cancri)
31 20 Tau Ceti 11.9 G8.5 V 01 44 −15 56 0.458
32 21 Luyten 372-58 12.0 M5 V 03 35 −44 30 7 × 10−5
33 22 Luyten 725-32 12.1 M4.5 V 01 12 −16 59 1.8 × 10−4
(YZ Ceti)
34 23 Luyten’s Star 12.4 M3.5 V 07 27 +05 13 1.4 × 10−3
35 24 SCR 12.6 M8.5 V 18 45 −63 57 1 × 10−6
J184-6357 A
36 SCR 12.7 T6[10] 18 45 −63 57 —
J184-6357 B
37 25 Teegarden’s 12.5 M6 V 02 53 +16 52 1 × 10−5
Star
38 26 Kapteyn’s Star 12.8 M1 V 05 11 −45 01 3.8 × 10−3
39 27 Lacaille 8760 12.9 K7 V 21 17 −38 52 0.029
(AX
Microscopium)

Table D8

D4 Stellar Data
Note: These are the stars that appear the brightest visually, as seen from our vantage point on Earth. They are not the stars
that are intrinsically the most luminous.

8. Brown dwarf
9. Brown dwarf
10. Brown dwarf

This OpenStax book is available for free at https://legacy.cnx.org/content/col12169/1.46


Appendix D 365

Figure D1 The brightest stars typically have names from antiquity. Next to each star’s ancient name, we have added a column
with its name in the system originated by Bayer (see the Naming Stars (https://legacy.cnx.org/content/m59905/
latest/#fs-id1165721974313) feature box.) The distances of the more remote stars are estimated from their spectral types and
apparent brightnesses and are only approximate. The luminosities for those stars are approximate to the same degree. Right
ascension and declination is given for Epoch 2000.0.
366 Appendix D

D5 Galaxies Visible from 40°N Latitude

Figure D2 Source: Roberto Mura (https://commons.wikimedia.org/wiki/User:Roberto_Mura) , Galaxies table


40°N (https://commons.wikimedia.org/wiki/File:Galaxies_table_40°N.png) , file type, CC BY-SA 4.0
(https://creativecommons.org/licenses/by-sa/4.0/legalcode)

This OpenStax book is available for free at https://legacy.cnx.org/content/col12169/1.46


Appendix E 367

APPENDIX E | USEFUL
MATHEMATICAL
FORMULAS
Quadratic formula
2
If ax 2 + bx + c = 0, then x = −b ± b − 4ac
2a

Triangle of base b and height h Area = 1 bh


2

Circle of radius r Circumference = 2πr Area = πr 2

Sphere of radius r Surface area = 4πr 2 Volume = 4 πr 3


3
Cylinder of radius r and height h Area of curved surface = 2πrh Volume = πr 2 h

Table E1 Geometry

Trigonometry
Trigonometric Identities
1. sin θ = 1/csc θ
2. cos θ = 1/sec θ
3. tan θ = 1/cot θ

4. sin⎛⎝90 0 − θ⎞⎠ = cos θ

5. cos⎛⎝90 0 − θ⎞⎠ = sin θ

6. tan⎛⎝90 0 − θ⎞⎠ = cot θ

7. sin 2 θ + cos 2 θ = 1

8. sec 2 θ − tan 2 θ = 1
9. tan θ = sin θ/cos θ
10. sin⎛⎝α ± β⎞⎠ = sin α cos β ± cos α sin β

11. cos⎛⎝α ± β⎞⎠ = cos α cos β ∓ sin α sin β

tan α ± tan β
12. tan⎛⎝α ± β⎞⎠ =
1 ∓ tan α tan β

13. sin 2θ = 2sin θ cos θ

14. cos 2θ = cos 2 θ − sin 2 θ = 2 cos 2 θ − 1 = 1 − 2 sin 2 θ


368 Appendix E

15. sin α + sin β = 2 sin 1 ⎛⎝α + β⎞⎠cos 1 ⎛⎝α − β⎞⎠


2 2

16. cos α + cos β = 2 cos 1 ⎛⎝α + β⎞⎠cos 1 ⎛⎝α − β⎞⎠


2 2
Triangles

1. Law of sines: a = b = c
sin α sin β sin γ

2. Law of cosines: c 2 = a 2 + b 2 − 2ab cos γ

3. Pythagorean theorem: a 2 + b 2 = c 2

Series expansions
n(n − 1)a n − 2 b 2 n(n − 1)(n − 2)a n − 3 b 3
1. Binomial theorem: (a + b) n = a n + na n − 1 b + + + ···
2! 3!

n(n − 1)x 2
2. (1 ± x) n = 1 ± nx + ± ···⎛⎝x 2 < 1⎞⎠
1! 2!

n(n + 1)x 2
3. (1 ± x) −n = 1 ∓ nx + ∓ ···⎛⎝x 2 < 1⎞⎠
1! 2!
3 5
4. sin x = x − x + x − ···
3! 5!
2 4
5. cos x = 1 − x + x − ···
2! 4!
3 5
6. tan x = x + x + 2x + ···
3 15
2
7. e x = 1 + x + x + ···
2!

8. ln(1 + x) = x − 1 x 2 + 1 x 3 − ···(|x| < 1)


2 3

This OpenStax book is available for free at https://legacy.cnx.org/content/col12169/1.46


Appendix E 369

Derivatives

1. d ⎡⎣a f (x)⎤⎦ = a d f (x)


dx dx

2. d ⎡⎣ f (x) + g(x)⎤⎦ = d f (x) + d g(x)


dx dx dx

3. d ⎡⎣ f (x)g(x)⎤⎦ = f (x) d g(x) + g(x) d f (x)


dx dx dx
d f (u) = ⎡ d f (u)⎤du
4.
dx ⎣ du ⎦ dx

5. d x m = mx m − 1
dx

6. d sin x = cos x
dx

7. d cos x = −sin x
dx

8. d tan x = sec 2 x
dx

9. d cot x = −csc 2 x
dx

10. d sec x = tan x sec x


dx

11. d csc x = −cot x csc x


dx

12. d ex = ex
dx

13. d ln x = 1
dx x

14. d sin −1 x = 1
dx 1 − x2

15. d cos −1 x = − 1
dx 1 − x2

16. d tan −1 x = − 1
dx 1 + x2
370 Appendix E

This OpenStax book is available for free at https://legacy.cnx.org/content/col12169/1.46


Appendix F 371

APPENDIX F | PERIODIC
TABLE OF THE ELEMENTS
372 Appendix F

This OpenStax book is available for free at https://legacy.cnx.org/content/col12169/1.46


Appendix G 373

APPENDIX G | THE GREEK


ALPHABET
Name Capital Lowercase Name Capital Lowercase
Alpha A α Nu N ν
Beta B β Xi Ξ ξ

Gamma Γ γ Omicron O ο

Delta Δ δ Pi Π π
Epsilon E ε Rho P ρ

Zeta Z ζ Sigma Σ σ

Eta H η Tau T τ

Theta Θ θ Upsilon ϒ υ
lota I ι Phi Φ ϕ

Kappa K κ Chi X χ

Lambda Λ λ Psi ψ ψ

Mu M μ Omega Ω ω

Table G1 The Greek Alphabet


374 Appendix G

This OpenStax book is available for free at https://legacy.cnx.org/content/col12169/1.46


Answer Key 375

ANSWER KEY
CHAPTER 1
CHECK YOUR UNDERSTANDING
1.1. 4.79 × 10 2 Mg or 479 Mg
1.2. 3 × 10 8 m/s
1.3. 10 8 km 2
1.4. The numbers were too small, by a factor of 4.45.
1.5. 4πr 3 /3
1.6. yes
1.7. 3 × 10 4 m or 30 km. It is probably an underestimate because the density of the atmosphere decreases with altitude. (In fact,
30 km does not even get us out of the stratosphere.)
1.8. No, the coach’s new stopwatch will not be helpful. The uncertainty in the stopwatch is too great to differentiate between the
sprint times effectively.
CONCEPTUAL QUESTIONS
1. Physics is the science concerned with describing the interactions of energy, matter, space, and time to uncover the fundamental
mechanisms that underlie every phenomenon.
3. No, neither of these two theories is more valid than the other. Experimentation is the ultimate decider. If experimental evidence
does not suggest one theory over the other, then both are equally valid. A given physicist might prefer one theory over another
on the grounds that one seems more simple, more natural, or more beautiful than the other, but that physicist would quickly
acknowledge that he or she cannot say the other theory is invalid. Rather, he or she would be honest about the fact that more
experimental evidence is needed to determine which theory is a better description of nature.
5. Probably not. As the saying goes, “Extraordinary claims require extraordinary evidence.”
7. Conversions between units require factors of 10 only, which simplifies calculations. Also, the same basic units can be scaled up
or down using metric prefixes to sizes appropriate for the problem at hand.
9. a. Base units are defined by a particular process of measuring a base quantity whereas derived units are defined as algebraic
combinations of base units. b. A base quantity is chosen by convention and practical considerations. Derived quantities are
expressed as algebraic combinations of base quantities. c. A base unit is a standard for expressing the measurement of a base
quantity within a particular system of units. So, a measurement of a base quantity could be expressed in terms of a base unit in any
system of units using the same base quantities. For example, length is a base quantity in both SI and the English system, but the
meter is a base unit in the SI system only.
11. a. Uncertainty is a quantitative measure of precision. b. Discrepancy is a quantitative measure of accuracy.
13. Check to make sure it makes sense and assess its significance.
PROBLEMS
15. a. 103; b. 105; c. 102; d. 1015; e. 102; f. 1057
17. 102 generations
19. 1011 atoms
21. 103 nerve impulses/s
23. 1026 floating-point operations per human lifetime
25. a. 957 ks; b. 4.5 cs or 45 ms; c. 550 ns; d. 31.6 Ms
27. a. 75.9 Mm; b. 7.4 mm; c. 88 pm; d. 16.3 Tm
29. a. 3.8 cg or 38 mg; b. 230 Eg; c. 24 ng; d. 8 Eg e. 4.2 g
31. a. 27.8 m/s; b. 62 mi/h
33. a. 3.6 km/h; b. 2.2 mi/h
35. 1.05 × 10 5 ft 2
37. 8.847 km
39. a. 1.3 × 10 −9 m; b. 40 km/My
41. 10 6 Mg/μL
43. 62.4 lbm/ft3
45. 0.017 rad
47. 1 light-nanosecond
49. 3.6 × 10 −4 m 3
51. a. Yes, both terms have dimension L2T-2 b. No. c. Yes, both terms have dimension LT-1 d. Yes, both terms have dimension
376 Answer Key

LT-2
⎡ ⎤ ⎡ ⎤ ⎡ ⎤
53. a. [v] = LT–1; b. [a] = LT–2; c. ⎣∫ vdt⎦ = L; d. ⎣∫ adt⎦ = LT ; e. ⎣ da ⎦ = LT
–1 –3
dt
55. a. L; b. L; c. L0 = 1 (that is, it is dimensionless)
57. 1028 atoms
59. 1051 molecules
61. 1016 solar systems
63. a. Volume = 1027 m3, diameter is 109 m.; b. 1011 m
65. a. A reasonable estimate might be one operation per second for a total of 109 in a lifetime.; b. about (109)(10–17 s) = 10–8 s, or
about 10 ns
67. 2 kg
69. 4%
71. 67 mL
73. a. The number 99 has 2 significant figures; 100. has 3 significant figures. b. 1.00%; c. percent uncertainties
75. a. 2%; b. 1 mm Hg
77. 7.557 cm2
79. a. 37.2 lb; because the number of bags is an exact value, it is not considered in the significant figures; b. 1.4 N; because the
value 55 kg has only two significant figures, the final value must also contain two significant figures
ADDITIONAL PROBLEMS
81. a. [s 0] = L and units are meters (m); b. [v 0] = LT −1 and units are meters per second (m/s); c. [a 0] = LT −2 and units

are meters per second squared (m/s2); d. [ j 0] = LT −3 and units are meters per second cubed (m/s3); e. [S 0] = LT −4 and units

are m/s4; f. [c] = LT −5 and units are m/s5.


83. a. 0.059%; b. 0.01%; c. 4.681 m/s; d. 0.07%, 0.003 m/s
85. a. 0.02%; b. 1×104 lbm
87. a. 143.6 cm3; b. 0.2 cm3 or 0.14%
CHALLENGE PROBLEMS
89. Since each term in the power series involves the argument raised to a different power, the only way that every term in the
power series can have the same dimension is if the argument is dimensionless. To see this explicitly, suppose [x] = LaMbTc. Then,
[xn] = [x]n = LanMbnTcn. If we want [x] = [xn], then an = a, bn = b, and cn = c for all n. The only way this can happen is if a = b =
c = 0.

CHAPTER 3
CHECK YOUR UNDERSTANDING
3.1. (a) The rider’s displacement is Δx = x f − x 0 = −1 km . (The displacement is negative because we take east to be positive
and west to be negative.) (b) The distance traveled is 3 km + 2 km = 5 km. (c) The magnitude of the displacement is 1 km.
3.2. Inserting the knowns, we have
–a = Δv = 2.0 × 10 7 m/s − 0 = 2.0 × 10 11 m/s 2.
Δt 10 −4 s − 0
3.3. If we take east to be positive, then the airplane has negative acceleration because it is accelerating toward the west. It is also
decelerating; its acceleration is opposite in direction to its velocity.
3.4. To answer this, choose an equation that allows us to solve for time t, given only a , v0 , and v:
v = v 0 + at.
Rearrange to solve for t:
v − v 0 400 m/s − 0 m/s
t= a = = 20 s.
20 m/s 2
2
3.5. a = m/s .
2
3
3.6. It takes 2.47 s to hit the water. The quantity distance traveled increases faster.
CONCEPTUAL QUESTIONS
1. You drive your car into town and return to drive past your house to a friend’s house.
3. If the bacteria are moving back and forth, then the displacements are canceling each other and the final displacement is small.
5. Distance traveled
7. Average speed is the total distance traveled divided by the elapsed time. If you go for a walk, leaving and returning to your

This OpenStax book is available for free at https://legacy.cnx.org/content/col12169/1.46


Answer Key 377

home, your average speed is a positive number. Since Average velocity = Displacement/Elapsed time, your average velocity is
zero.
9. Average speed. They are the same if the car doesn’t reverse direction.
11. No, in one dimension constant speed requires zero acceleration.
13. A ball is thrown into the air and its velocity is zero at the apex of the throw, but acceleration is not zero.
15. Plus, minus
17. If the acceleration, time, and displacement are the knowns, and the initial and final velocities are the unknowns, then two
kinematic equations must be solved simultaneously. Also if the final velocity, time, and displacement are the knowns then two
kinematic equations must be solved for the initial velocity and acceleration.
19. a. at the top of its trajectory; b. yes, at the top of its trajectory; c. yes
g g
21. Earth v = v 0 − gt = −gt ; Moon v′ = t′ v = v′ − gt = − t′ t′ = 6t ; Earth y = − 1 gt 2 Moon
6 6 2
g ⎛ ⎞
y′ = − 1 (6t) 2 = − 1 g6t 2 = −6⎝1 gt 2⎠ = −6y
26 2 2
PROBLEMS

25. a. →
^ ^
x 1 = (−2.0 m) i , →
x 2 = (5.0 m) i ; b. 7.0 m east
27. a. t = 2.0 s; b. x(6.0) − x(3.0) = −8.0 − (−2.0) = −6.0 m

29. a. 150.0 s, v = 156.7 m/s ; b. 45.7% the speed of sound at sea level

31.

33.
34. a = 4.29m/s 2
378 Answer Key

36.
38. a = 11.1g
40. 150 m
42. a. 525 m;
b. v = 180 m/s
44. a.

b. The acceleration has the greatest positive value at t a


c. The acceleration is zero at t e and t h
d. The acceleration is negative at t i ,t j ,t k ,t l
46. a. a = −1.3 m/s 2 ;
b. v 0 = 18 m/s ;
c. t = 13.8 s
48. v = 502.20 m/s
50. a.

b. Knowns: a = 2.40 m/s 2, t = 12.0 s, v 0 = 0 m/s , and x 0 = 0 m ;

This OpenStax book is available for free at https://legacy.cnx.org/content/col12169/1.46


Answer Key 379

c. x = x 0 + v 0 t + 1 at = 1 at = 2.40 m/s (12.0 s) = 172.80 m , the answer seems reasonable at about 172.8 m; d.
2 2 2 2
2 2
v = 28.8 m/s
52. a.

b. Knowns: v = 30.0 cm/s, x = 1.80 cm ;


c. a = 250 cm/s 2, t = 0.12 s ;
d. yes
54. a. 6.87 s2; b. x = 52.26 m
56. a. a = 8450 m/s 2 ;
b. t = 0.0077 s
58. a. a = 9.18 g;
b. t = 6.67 × 10 −3 s ;

a = −40.0 m/s 2
c.
a = 4.08 g
60. Knowns: x = 3 m, v = 0 m/s, v 0 = 54 m/s . We want a, so we can use this equation: a = −486 m/s 2 .
62. a. a = 32.58 m/s 2 ;
b. v = 161.85 m/s ;
c. v > v max , because the assumption of constant acceleration is not valid for a dragster. A dragster changes gears and would have
a greater acceleration in first gear than second gear than third gear, and so on. The acceleration would be greatest at the beginning,
so it would not be accelerating at 32.6 m/s 2 during the last few meters, but substantially less, and the final velocity would be less
than 162 m/s .

y = −8.23 m
64. a. ;
v 1 = −18.9 m/s

y = −18.9 m
b. ;
v 2 = 23.8 m/s

y = −32.0 m
c. ;
v 3 = −28.7 m/s

y = −47.6 m
d. ;
v 4 = −33.6 m/s

y = −65.6 m
e.
v 5 = −38.5 m/s

66. a. Knowns: a = −9.8 m/s 2 v 0 = −1.4 m/s t = 1.8 s y 0 = 0 m ;

b. y = y 0 + v 0 t − 1 gt y = v 0 t − 1 gt = −1.4 m/s(1.8 sec) − 1 (9.8)(1.8 s) = −18.4 m and the origin is at the rescuers,
2 2
2 2 2
who are 18.4 m above the water.
2 2
v
68. a. v 2 = v 20 − 2g(y − y 0) y 0 = 0 v = 0 y = 0 = (4.0 m/s) = 0.82 m ; b. to the apex v = 0.41 s times 2 to the
2g 2(9.80)
380 Answer Key

board = 0.82 s from the board to the water y = y 0 + v 0 t − 1 gt 2 y = −1.80 m y 0 = 0 v 0 = 4.0 m/s
2
−1.8 = 4.0t − 4.9t 2 4.9t 2 − 4.0t − 1.80 = 0 , solution to quadratic equation gives 1.13 s; c.
2
v = v 20 − 2g(y − y 0) y0 = 0 v 0 = 4.0 m/s y = −1.80 m
v = 7.16 m/s
70. Time to the apex: t = 1.12 s times 2 equals 2.24 s to a height of 2.20 m. To 1.80 m in height is an additional 0.40 m.

y = y 0 + v 0 t − 1 gt 2 y = −0.40 m y 0 = 0 v 0 = −11.0 m/s


2
y = y 0 + v 0 t − 1 gt 2 y = −0.40 m y 0 = 0 v 0 = −11.0 m/s .
2
−0.40 = −11.0t − 4.9t 2 or 4.9t 2 + 11.0t − 0.40 = 0

Take the positive root, so the time to go the additional 0.4 m is 0.04 s. Total time is 2.24 s + 0.04 s = 2.28 s .

2 2
72. a. v = v 0 − 2g(y − y 0) y 0 = 0 v = 0 y = 2.50 m ; b. t = 0.72 s times 2 gives 1.44 s in the air
v 20 = 2gy ⇒ v 0 = 2(9.80)(2.50) = 7.0 m/s
74. a. v = 70.0 m/s ; b. time heard after rock begins to fall: 0.75 s, time to reach the ground: 6.09 s

ADDITIONAL PROBLEMS
76. Take west to be the positive direction.

1st plane: ν = 600 km/h
2nd plane –
ν = 667.0 km/h
v−v −3.4 cm/s − v 0
78. a= t−t0, t = 0, a = = 1.2 cm/s 2 ⇒ v 0 = − 8.2 cm/s v = v 0 + at = − 8.2 + 1.2 t ;
0 4s
v = −7.0 cm/s v = −1.0 cm/s
79. a = −3 m/s 2
81. a.
v = 8.7 × 10 5 m/s ;
b. t = 7.8 × 10 −8 s

83. 1 km = v 0(80.0 s) + 1 a(80.0) ; 2 km = v 0(200.0) + 1 a(200.0) solve simultaneously to get a = −


2 2 0.1 km/s 2
2 2 2400.0
and v 0 = 0.014167 km/s , which is 51.0 km/h . Velocity at the end of the trip is v = 21.0 km/h .
85. a = −0.9 m/s 2
87. Equation for the speeding car: This car has a constant velocity, which is the average velocity, and is not accelerating, so use
– –
the equation for displacement with x 0 = 0 : x = x 0 + v t = v t ; Equation for the police car: This car is accelerating, so use the

equation for displacement with x 0 = 0 and v 0 = 0 , since the police car starts from rest: x = x 0 + v 0 t + 1 at = 1 at ; Now
2 2
2 2
we have an equation of motion for each car with a common parameter, which can be eliminated to find the solution. In this case,
– –
we solve for t . Step 1, eliminating x : x = v t = 1 at ; Step 2, solving for t : t = 2v
2
2 a . The speeding car has a constant velocity
of 40 m/s, which is its average velocity. The acceleration of the police car is 4 m/s2. Evaluating t, the time for the police car to
– 2(40)
reach the speeding car, we have t = 2v
a = = 20 s .
4
−v
89. At this acceleration she comes to a full stop in t = a 0 = 8 = 16 s , but the distance covered is
0.5
x = 8 m/s(16 s) − 1 (0.5)(16 s) 2 = 64 m , which is less than the distance she is away from the finish line, so she never finishes
2
the race.
3
91. x 1 = v 0 t
2
x2 = 5 x1
3
93. v 0 = 7.9 m/s velocity at the bottom of the window.

This OpenStax book is available for free at https://legacy.cnx.org/content/col12169/1.46


Answer Key 381

v = 7.9 m/s
v 0 = 14.1 m/s
95. a. v = 5.42 m/s ;
b. v = 4.64 m/s ;
c. a = 2874.28 m/s 2 ;
d. (x − x 0) = 5.11 × 10 −3 m
97. Consider the players fall from rest at the height 1.0 m and 0.3 m.
0.9 s
0.5 s
99. a. t = 6.37 s taking the positive root;
b. v = 59.5 m/s
101. a. y = 4.9 m ;
b. v = 38.3 m/s ;
c. −33.3 m
103. h = 1 gt , h = total height and time to drop to ground
2
2
2 h = 1 g(t − 1) 2 in t – 1 seconds it drops 2/3h
3 2
2 1 gt 2⎞ = 1 g(t − 1) 2 or t 2 = 1 (t − 1) 2

3 ⎝2 ⎠ 2 3 2
2
0 = t 2 − 6t + 3 t = 6 ± 6 − 4 · 3 = 3 ± 24
2 2
2
t = 5.45 s and h = 145.5 m. Other root is less than 1 s. Check for t = 4.45 s h = 1 gt = 97.0 m = (145.5)
2
2 3
CHALLENGE PROBLEMS

106. 96 km/h = 26.67 m/s, a =


26.67 m/s = 6.67m/s 2 , 295.38 km/h = 82.05 m/s, t = 12.3 s time to accelerate to
4.0 s
maximum speed
x = 504.55 m distance covered during acceleration
7495.44 m at a constant speed
7495.44 m = 91.35 s so total time is 91.35 s + 12.3 s = 103.65 s .
82.05 m/s

CHAPTER 4
CHECK YOUR UNDERSTANDING

4.1. a. 40.0 rev/s = 2π(40.0) rad/s , – = Δω = 2π(40.0) − 0 rad/s = 2π(2.0) = 4.0π rad/s 2 ; b. Since the angular
α
Δt 20.0 s
velocity increases linearly, there has to be a constant acceleration throughout the indicated time. Therefore, the instantaneous
angular acceleration at any time is the solution to 4.0π rad/s 2 .
7000.0(2π rad)
4.2. a. Using Equation 4.22, we have 7000 rpm = = 733.0 rad/s,
60.0 s
ω − ω 0 733.0 rad/s
α= t = = 73.3 rad/s 2 ;
10.0 s
b. Using Equation 4.25, we have
ω 2 − ω 20 0 − (733.0 rad/s) 2
ω 2 = ω 20 + 2αΔθ ⇒ Δθ = = = 3665.2 rad
2α 2(73.3 rad/s 2)
(5.0 − 0)rad/s
4.3. The angular acceleration is α = = 0.25 rad/s 2 . Therefore, the total angle that the boy passes through is
20.0 s
ω 2 − ω 20 (5.0) 2 − 0
Δθ = = = 50 rad .
2α 2(0.25)
Thus, we calculate
382 Answer Key

s = rθ = 5.0 m(50.0 rad) = 250.0 m .

CONCEPTUAL QUESTIONS
1. The second hand rotates clockwise, so by the right-hand rule, the angular velocity vector is into the wall.
3. They have the same angular velocity. Points further out on the bat have greater tangential speeds.
4. straight line, linear in time variable
6. constant
PROBLEMS

9. ω =
2π rad = 0.14 rad/s
45.0 s
s 3.0 m = 2.0 rad ; b. ω = 2.0 rad = 2.0 rad/s
11. a. θ = r =
1.5 m 1.0 s
Δω 0 rad/s − 10.0(2π) rad/s = 31.4 s
13. The propeller takes only Δt = α = to come to rest, when the propeller is at 0 rad/s,
−2.0 rad/s 2
it would start rotating in the opposite direction. This would be impossible due to the magnitude of forces involved in getting the
propeller to stop and start rotating in the opposite direction.
15. a. ω = 25.0(2.0 s) = 50.0 rad/s ; b. α =
dω = 25.0 rad/s 2
dt
16. a. ω = 54.8 rad/s ;
b. t = 11.0 s
18. a. 0.87 rad/s 2 ;
b. θ = 66,264 rad
20. a. ω = 42.0 rad/s ;

v t = 42 m/s
b. θ = 200 rad ; c.
a t = 4.0 m/s 2

22. a. ω = 7.0 rad/s ;


b. θ = 22.5 rad ; c. a t = 0.1 m/s
24. α = 28.6 rad/s 2 .
27. (a) 62.8 m/s
(b) α = −0.314 rad/s 2

CHAPTER 5
CHECK YOUR UNDERSTANDING
5.1. Special relativity applies only to objects moving at constant velocity, whereas general relativity applies to objects that undergo
acceleration.
γ= 1 = 1 = 1.32
5.2. 1− v2 (0.650c) 2
1−
c2 c2

Δt = Δτ = 2.10 × 10 −8 s = 2.71 × 10 −8 s.
5.3. a. 1− v2 8 m/s)
2
c2 1 − (1.90 × 10 2
(3.00 × 10 8 m/s)
5.3. b. Only the relative speed of the two spacecraft matters because there is no absolute motion through space. The signal is
emitted from a fixed location in the frame of reference of A, so the proper time interval of its emission is τ = 1.00 s. The duration
of the signal measured from frame of reference B is then
Δt = Δτ = 1.00 s = 1.01 s.
v2 2
1− (4.00 × 10 7 m/s)
c2 1− 2
(3.00 × 10 8 m/s)

This OpenStax book is available for free at https://legacy.cnx.org/content/col12169/1.46


Answer Key 383

2
5.4. L = L 0 1 − v = (2.50 km) 1 −
(0.750c) 2
2 2
= 1.65 km
c c
5.5. Start with the definition of the proper time increment:
dτ = −(ds) 2 /c 2 = dt 2 − ⎛⎝dx 2 + dx 2 + dx 2⎞⎠/c 2.
where (dx, dy, dx, cdt) are measured in the inertial frame of an observer who does not necessarily see that particle at rest. This
therefore becomes
dτ = −(ds) 2 /c 2 = dt 2 − ⎡⎣(dx) 2 + ⎛⎝dy⎞⎠ 2 + (dz) 2⎤⎦/c 2
⎡ ⎛dy ⎞ 2⎤ 2
⎛ ⎞ ⎛ ⎞
2
= dt 1 − ⎢⎝dx ⎠ + ⎝ ⎠ + ⎝dz ⎠ ⎥/c 2
⎣ dt dt dt ⎦

= dt 1 − v 2 /c 2
dt = γdτ.
5.6. Although displacements perpendicular to the relative motion are the same in both frames of reference, the time interval
between events differ, and differences in dt and dt′ lead to different velocities seen from the two frames.

CONCEPTUAL QUESTIONS
1. the second postulate, involving the speed of light; classical physics already included the idea that the laws of mechanics, at least,
were the same in all inertial frames, but the velocity of a light pulse was different in different frames moving with respect to each
other
3. yes, provided the plane is flying at constant velocity relative to the Earth; in that case, an object with no force acting on it within
the plane has no change in velocity relative to the plane and no change in velocity relative to the Earth; both the plane and the
ground are inertial frames for describing the motion of the object
5. The observer moving with the process sees its interval of proper time, which is the shortest seen by any observer.
7. The length of an object is greatest to an observer who is moving with the object, and therefore measures its proper length.
9. a. No, not within the astronaut’s own frame of reference. b. He sees Earth clocks to be in their rest frame moving by him, and
therefore sees them slowed. c. No, not within the astronaut’s own frame of reference. d. Yes, he measures the distance between the
two stars to be shorter. e. The two observers agree on their relative speed.
PROBLEMS
10. a. 1.0328; b. 1.15
12. 5.96 × 10 −8 s
14. 0.800c
16. 0.140c
18. 48.6 m
20. Using the values given in m58563 (https://legacy.cnx.org/content/m58563/latest/#fs-id1167793912924) : a. 1.39
km; b. 0.433 km; c. 0.433 km
22. a. 10.0c; b. The resulting speed of the canister is greater than c, an impossibility. c. It is unreasonable to assume that the canister
will move toward the earth at 1.20c.
24. The angle α approaches 45°, and the t′- and x′-axes rotate toward the edge of the light cone.
26. 15 m/s east
28. 32 m/s
30. a. The second ball approaches with velocity −v and comes to rest while the other ball continues with velocity −v; b. This
conserves momentum.
t 1′ = 0; x 1′ = 0; t 1′ = 0; x 1′ = 0;
; τ −vτ
32. a.
t 2′ = τ; x 2′ = 0 b. t 2′ = 2 2
; x 2′ =
1 − v /c 1 − v 2 /c 2
34. 0.615c
36. 0.696c
38. (Proof)
ADDITIONAL PROBLEMS
40. a. 0.866c; b. 0.995c
42. a. 4.303 y to four digits to show any effect; b. 0.1434 y; c. 1/ ⎛⎝1 − v 2 /c 2⎞⎠ = 29.88.
44. a. 4.00; b. v = 0.867c
46. a. A sends a radio pulse at each heartbeat to B, who knows their relative velocity and uses the time dilation formula to calculate
384 Answer Key

the proper time interval between heartbeats from the observed signal. b. (66 beats/min) 1 − v 2/c 2 = 57.1 beats/min
48. a. first photon: (0, 0, 0) at t = t′; second photon:
2
t′ = −vx/c 2 = −(c/2)(1.00 m)/c = − 0.577 m = 1.93 × 10 −9 s
0.75 c
1 − v 2 /c 2
x′ = x = 1.00 m = 1.15 m
2 2
1 − v /c 0.75
b. simultaneous in A, not simultaneous in B
⎛ ⎛ 3 m⎞
2 ⎝4.5 × 10 −4 s⎞⎠ − (0.6c)⎝150 × 10
2 ⎠
t′ = t − vx/c = c
1 − v 2 /c 2 1 − (0.6) 2
= 1.88 × 10 −4 s
3 ⎛ 8 ⎞⎛ −4 ⎞
50.
x′ = x − vt = 150 × 10 m − (0.60)⎝3.00 × 10 m/s⎠⎝4.5 × 10 s⎠
1 − v 2 /c 2 1 − (0.6) 2
= −1.01 × 10 5 m = −101 km
y = y′ = 15 km
z = z′ = 1 km
2
Δt = Δt′ + vΔx′/c
2 2
1 − v /c
52.
Δt′ + v(500 m)/c 2
0 = ;
1 − v 2 /c 2
since v ≪ c, we can ignore the term v 2 /c 2 and find
(50 m/s)(500 m)
Δt′ = − 2
= −2.78 × 10 −13 s
⎛ 8 ⎞
⎝3.00 × 10 m/s⎠

The breakdown of Newtonian simultaneity is negligibly small, but not exactly zero, at realistic train speeds of 50 m/s.
2
Δt′ = Δt − vΔx/c
2 2
1 − v /c
⎛ ⎞
(v)⎝2.0 × 10 9 m⎠
(0.30 s) − 2
⎛ 8 ⎞
⎝3.00 × 10 m/s⎠
54. 0 =
1 − v 2 /c 2
(0.30 s) ⎛ 2
v = ⎛ 9 ⎞⎝
3.00 × 10 8 m/s⎞⎠
⎝2.0 × 10 m⎠

v = 1.35 × 10 7 m/s
56. Note that all answers to this problem are reported to five significant figures, to distinguish the results. a. 0.99947c; b.
1.2064 × 10 11 y; c. 1.2058 × 10 11 y
58. a. –0.400c; b. –0.909c
60. a. 1.65 km/s; b. Yes, if the speed of light were this small, speeds that we can achieve in everyday life would be larger than 1%
of the speed of light and we could observe relativistic effects much more often.

CHAPTER 6
CHECK YOUR UNDERSTANDING
6.1. a. not equal because they are orthogonal; b. not equal because they have different magnitudes; c. not equal because they have
different magnitudes and directions; d. not equal because they are antiparallel; e. equal.
6.2. 16 m; → D = −16 m ^ u
6.3. G = 28.2 cm, θ G = 291°

6.4. → ^ ^
D = (−5.0 i − 3.0 j )cm ; the fly moved 5.0 cm to the left and 3.0 cm down from its landing site.

This OpenStax book is available for free at https://legacy.cnx.org/content/col12169/1.46


Answer Key 385

6.5. 5.83 cm, 211°


6.6. → ^
D = (−20 m) j
6.7. 35.1 m/s = 126.4 km/h
6.8. → ^ ^
G = (10.25 i − 26.22 j )cm
6.9. D = 55.7 N; direction 65.7° north of east
6.10. ^ ^ ^
v = 0.8 i + 0.6 j , 36.87° north of east

CONCEPTUAL QUESTIONS
1. scalar
3. answers may vary
5. parallel, sum of magnitudes, antiparallel, zero
7. no, yes
9. zero, yes
11. no
13. equal, equal, the same
15. a unit vector of the x-axis
17. They are equal.
19. yes
PROBLEMS
21. → h = −49 m ^ u , 49 m
23. 30.8 m, 35.7° west of north
25. 134 km, 80°
27. 7.34 km, 63.5° south of east
29. 3.8 km east, 3.2 km north, 7.0 km
31. 14.3 km, 65°
33. a. → ^ ^ → ^ ^ → ^ ^
A = + 8.66 i + 5.00 j , b. B = + 30.09 i + 39.93 j , c. C = + 6.00 i − 10.39 j , d.
→ ^ ^ → ^ ^
D = −15.97 i + 12.04 j , f. F = −17.32 i − 10.00 j

35. a. 1.94 km, 7.24 km; b. proof


386 Answer Key

37. 3.8 km east, 3.2 km north, 2.0 km, → ^ ^


D = (3.8 i + 3.2 j )km
39. P 1(2.165 m, 1.250 m) , P 2(−1.900 m, 3.290 m) , 5.27 m
41. 8.60 m, A(2 5 m, 0.647π) , B(3 2 m, 0.75π)

43. a. → → ^ ^
A + B = −4 i − 6 j , | → →
|
A + B = 7.211, θ = 213.7° ; b. → → ^ ^
A − B =2i −2j ,

| → →
|
A − B = 2 2, θ = −45°

45. a. → ^ ^ ^
C = (5.0 i − 1.0 j − 3.0 k )m, C = 5.92 m ;
b. → ^ ^ ^
D = (4.0 i − 11.0 j + 15.0 k )m, D = 19.03 m
47. → ^ ^ ^
D = (3.3 i − 6.6 j )km , i is to the east, 7.34 km, −63.5°
49. a. → ^ ^ → ^ ^
R = −1.35 i − 22.04 j , b. R = −17.98 i + 0.89 j
51. → ^ ^
D = (200 i + 300 j )yd , D = 360.5 yd, 56.3° north of east; The numerical answers would stay the same but the physical
unit would be meters. The physical meaning and distances would be about the same because 1 yd is comparable with 1 m.
53. → ^ ^
R = −3 i − 16 j
55. → ^ E = −357.8V/m , E z = 0.0V/m , θ E = −tan −1(2)
E = EE , E x = + 178.9V/m , y
→ ^ ^ ^ → ^ ^
57. a. R B = (12.278 i + 7.089 j + 2.500 k )km , R D = (−0.262 i + 3.000 k )km ; b.

| →
R B−

R D| = 14.414 km

ADDITIONAL PROBLEMS
58. a. 18.4 km and 26.2 km, b. 31.5 km and 5.56 km
60. a. (r, φ + π/2) , b. (2r, φ + 2π) , (c) (3r, −φ)
62. d PM = 33.12 nmi = 61.34 km, d NP = 35.47 nmi = 65.69 km
64. proof
66. a. 10.00 m, b. 5π m , c. 0
68. 22.2 km/h, 35.8° south of west
70. 240.2 m, 2.2° south of west

CHAPTER 7
CHECK YOUR UNDERSTANDING
7.1. (a) Taking the derivative with respect to time of the position function, we have
→ ^ ^
v (t) = 9.0t 2 i and →
v (3.0s) = 81.0 i m/s. (b) Since the velocity function is nonlinear, we suspect the average velocity is
not equal to the instantaneous velocity. We check this and find
→ ^ ^
→ r (t 2) − → r (t 1) →
r (4.0 s) − → r (2.0 s) (144.0 i − 36.0 i ) m ^
v avg = t2 − t1 = = = 54.0 i m/s,
4.0 s − 2.0 s 2.0 s
which is different from → ^
v (3.0s) = 81.0 i m/s.
7.2. The acceleration vector is constant and doesn’t change with time. If a, b, and c are not zero, then the velocity function must be
→ ^ ^ ^ ^ ^ ^
linear in time. We have →
v (t) = ∫ a dt = ∫ (a i + b j + c k )dt = (a i + b j + c k )t m/s, since taking the derivative of

the velocity function produces →


a (t). If any of the components of the acceleration are zero, then that component of the velocity
would be a constant.
7.3. (a) Choose the top of the cliff where the rock is thrown from the origin of the coordinate system. Although it is arbitrary, we
typically choose time t = 0 to correspond to the origin. (b) The equation that describes the horizontal motion is x = x 0 + v x t.
With x 0 = 0, this equation becomes x = v x t. (c) m58290 (https://legacy.cnx.org/content/m58290/latest/#fs-

This OpenStax book is available for free at https://legacy.cnx.org/content/col12169/1.46


Answer Key 387

id1165038300036) through m58290 (https://legacy.cnx.org/content/m58290/latest/#fs-id1165038323452) and


Equation 7.46 describe the vertical motion, but since y 0 = 0 and v 0y = 0, these equations simplify greatly to become

y = 1 (v 0y + v y)t = 1 v y t, v y = −gt, y = − 1 gt 2, and v 2y = −2gy. (d) We use the kinematic equations to find the
2 2 2
x and y components of the velocity at the point of impact. Using v 2y = −2gy and noting the point of impact is −100.0 m, we

find the y component of the velocity at impact is v y = 44.3 m/s. We are given the x component, v x = 15.0 m/s, so we can
calculate the total velocity at impact: v = 46.8 m/s and θ = 71.3° below the horizontal.
7.4. The golf shot at 30°.
7.5. 134.0 cm/s
7.6. Labeling subscripts for the vector equation, we have B = boat, R = river, and E = Earth. The vector equation becomes
→v BE = → v BR + → v RE. We have right triangle geometry shown in Figure 04_05_BoatRiv_img. Solving for → v BE , we
have
v BE = v 2BR + v 2RE = 4.5 2 + 3.0 2
⎛ ⎞
v BE = 5.4 m/s, θ = tan −1 ⎝3.0 ⎠ = 33.7°.
4.5

CONCEPTUAL QUESTIONS
1. straight line
3. The slope must be zero because the velocity vector is tangent to the graph of the position function.
5. No, motions in perpendicular directions are independent.
7. a. no; b. minimum at apex of trajectory and maximum at launch and impact; c. no, velocity is a vector; d. yes, where it lands
9. They both hit the ground at the same time.
11. yes
13. If he is going to pass the ball to another player, he needs to keep his eyes on the reference frame in which the other players on
the team are located.
388 Answer Key

15.
PROBLEMS

17. → ^ ^ ^
r = 1.0 i − 4.0 j + 6.0 k
19. Δ →
^ ^
r Total = 472.0 m i + 80.3 m j

21. Sum of displacements = −6.4 km ^i + 9.4 km ^j

23. a. → ^ ^ → → ^ ^
v (t) = 8.0t i + 6.0t 2 k , v (0) = 0, v (1.0) = 8.0 i + 6.0 k m/s ,
^ ^
b. →
v avg = 4.0 i + 2.0 k m/s

25. Δ →
^ ^ ^
r 1 = 20.00 m j , Δ →
r 2 = (2.000 × 10 4 m) (cos30° i + sin 30° j )
^ ^
Δ→
r = 1.700 × 10 4 m i + 1.002 × 10 4 m j
^ ^ → ^ ^
27. a. → r (t) = (2.0t 2 i + 3 t 2 j ) m ,
v (t) = (4.0t i + 3.0t j )m/s, 2
2 3 2 2 x
b. x(t) = 2.0t m, y(t) = t m, t = ⇒ y = x
3
2 2 4

29. a. → ^ ^ ^
v (t) = (6.0t i − 21.0t 2 j + 10.0t −3 k )m/s ,
b. → ^ ^ ^
a (t) = (6.0 i − 42.0t j − 30t −4 k )m/s 2 ,
c. → ^ ^ ^
v (2.0s) = (12.0 i − 84.0 j + 1.25 k )m/s ,
^ ^ ^
d. →
v (1.0 s) = 6.0 i − 21.0 j + 10.0 k m/s, | →v (1.0 s)| = 24.0 m/s
v (3.0 s) = 18.0 i − 189.0 j + 0.37 k m/s, | v (3.0 s)| = 199.0 m/s ,
→ ^ ^ ^ →

e. → ^ ^ ^
r (t) = (3.0t 2 i − 7.0t 3 j − 5.0t −2 k )cm

This OpenStax book is available for free at https://legacy.cnx.org/content/col12169/1.46


Answer Key 389

→ ^ ^ ^
v avg = 9.0 i − 49.0 j − 6.3 k m/s
^ ^ ^ ^ ^
v (t) = −sin(1.0t) i + cos(1.0t) j + k , b. →
31. a. → a (t) = −cos(1.0t) i − sin(1.0t) j
33. a. t = 0.55 s , b. x = 110 m
35. a. t = 0.24s, d = 0.28 m , b. They aim high.

37. a., t = 12.8 s, x = 5619 m b. v y = 125.0 m/s, v x = 439.0 m/s, | →v | = 456.0 m/s
39. a. v y = v 0y − gt, t = 10s, v y = 0, v 0y = 98.0 m/s, v 0 = 196.0 m/s , b. h = 490.0 m,
c. v 0x = 169.7 m/s, x = 3394.0 m,

x = 2545.5 m
d. = 465.5 m
y
→ ^ ^
s = 2545.5 m i + 465.5 m j

41. −100 m = (−2.0 m/s)t − (4.9 m/s 2)t 2, t = 4.3 s, x = 86.0 m


43. R Moon = 48 m
2 2
45. a. v 0y = 24 m/s v y = v 0y − 2gy ⇒ h = 23.4 m ,

b. t = 3 s v 0x = 18 m/s x = 54 m ,

c. y = −100 m y 0 = 0 y − y 0 = v 0y t − 1 gt
2
− 100 = 24t − 4.9t 2 ⇒ t = 7.58 s ,
2
d. x = 136.44 m ,
e. t = 2.0 s y = 28.4 m x = 36 m
t = 4.0 s y = 17.6 m x = 22.4 m
t = 6.0 s y = −32.4 m x = 108 m
47. v 0y = 12.9 m/s y − y 0 = v 0y t − 1 gt
2
− 20.0 = 12.9t − 4.9t 2
2
t = 3.7 s v 0x = 15.3 m/s ⇒ x = 56.7 m
So the golfer’s shot lands 13.3 m short of the green.
49. a. R = 60.8 m ,
390 Answer Key

b. R = 137.8 m
2 2
51. a. v y = v 0y − 2gy ⇒ y = 2.9 m/s
y = 3.3 m/s
v 20y (v 0 sinθ) 2
y= = ⇒ sinθ = 0.91 ⇒ θ = 65.5°
2g 2g
53. R = 18.5 m
⎡ g ⎤
55. y = (tanθ 0)x − ⎢ ⎥x 2 ⇒ v 0 = 16.4 m/s
⎣2(v 0 cosθ 0) ⎦
2

2
v sin 2θ 0
57. R = 0
g ⇒ θ 0 = 15.0°
59. It takes the wide receiver 1.1 s to cover the last 10 m of his run.
2(v 0 sinθ)
T tof = g ⇒ sinθ = 0.27 ⇒ θ = 15.6°

61. a C = 40 m/s 2
2
63. a C = v ⇒ v 2 = r a C = 78.4, v = 8.85 m/s
r
T = 5.68 s, which is 0.176 rev/s = 10.6 rev/min
65. Venus is 108.2 million km from the Sun and has an orbital period of 0.6152 y.
r = 1.082 × 10 11 m T = 1.94 × 10 7 s
v = 3.5 × 10 4 m/s, a C = 1.135 × 10 −2 m/s 2
67. 360 rev/min = 6 rev/s
v = 3.8 m/s a C = 144. m/s 2

69. a. O′(t) = (4.0 ^i + 3.0 ^j + 5.0 ^


k )t m ,
→ = → →   → ^ ^ ^
b. r PS r PS′ + r S′ S, r (t) = →
r ′(t) + (4.0 i + 3.0 j + 5.0 k )t m ,
c. → ^ ^ ^
v (t) = →
v ′(t) + (4.0 i + 3.0 j + 5.0 k ) m/s , d. The accelerations are the same.
71. →
^ ^ ^
v PC = (2.0 i + 5.0 j + 4.0 k )m/s
73. a. A = air, S = seagull, G = ground

v = 9.0 m/s velocity of seagull with respect to still air
SA

v → → → → → = → →
AG = ? v SG = 5 m/s v SG = v SA + v AG ⇒ v AG v SG − v SA
→v AG = −4.0 m/s
→ →
b. v SG = v SA + v
→ ⇒ →
v = −13.0 m/s
AG SG
−6000 m = 7 min 42 s
−13.0 m/s
75. Take the positive direction to be the same direction that the river is flowing, which is east. S = shore/Earth, W = water, and B
= boat.
a.

v = 11 km/h
BS
t = 8.2 min

b. v = −5 km/h
BS
t = 18 min

c. v BS = v
→ + →
v θ = 22° west of north
BW WS

This OpenStax book is available for free at https://legacy.cnx.org/content/col12169/1.46


Answer Key 391

d. | →v BS| = 7.4 km/h t = 6.5 min


e.

v = 8.54 km/h, but only the component of the velocity straight across the river is used to get the time
BS

t = 6.0 min
Downstream = 0.3 km

v = →

v
77. AG v CGAC +

|→
|
v AC = 25 km/h → |
v CG = 15 km/h → | |
v AG = 29.15 km/h → | v AG = → v AC + → v CG
→ →
The angle between v AC and v AG is 31°, so the direction of the wind is 14° north of east.

ADDITIONAL PROBLEMS
79. a C = 39.6 m/s 2
81. 90.0 km/h = 25.0 m/s, 9.0 km/h = 2.5 m/s, 60.0 km/h = 16.7 m/s
2 2 2
a T = −2.5 m/s , a C = 1.86 m/s , a = 3.1 m/s
4π 2 R E cos λ
83. The radius of the circle of revolution at latitude λ is R E cos λ. The velocity of the body is 2πr . a C = for
T 2
T
392 Answer Key

λ = 40°, a C = 0.26% g
85. a T = 3.00 m/s 2

v(5 s) = 15.00 m/s a C = 150.00 m/s 2 θ = 88.8° with respect to the tangent to the circle of revolution directed inward.

| |
→a = 150.03 m/s 2

87. → ^ ^
a (t) = −Aω 2 cos ωt i − Aω 2 sin ωt j
a C = 5.0 mω 2 ω = 0.89 rad/s
→ ^ ^
v (t) = −2.24 m/s i − 3.87 m/s j
89. →
^ ^ ^ ^ ^
r 1 = 1.5 j + 4.0 k →
r 2 =Δ→
r + →
r 1 = 2.5 i + 4.7 j + 2.8 k
91. v x(t) = 265.0 m/s
v y(t) = 20.0 m/s
→ ^ ^
v (5.0 s) = (265.0 i + 20.0 j )m/s
93. R = 1.07 m
95. v 0 = 20.1 m/s
97. v = 3072.5 m/s
a C = 0.223 m/s 2

CHALLENGE PROBLEMS
v 2
99. a. −400.0 m = v 0y t − 4.9t 2 359.0 m = v 0x t t = 359.0 − 400.0 = 359.0 0y − 4.9( 359.0 )
v 0x v 0x v 0x

−400.0 = 359.0 tan 40 − 631,516.9


2
⇒ v 20x = 900.6 v 0x = 30.0 m/s v 0y = v 0x tan 40 = 25.2 m/s
v 0x
v = 39.2 m/s , b. t = 12.0 s
^ ^ 2
101. a. →
r TC = (−32 + 80t) i + 50t j , | →r TC| = (−32 + 80t) 2 + (50t) 2
2(−32 + 80t) + 100t
2r dr = 2(−32 + 80t) + 100t dr = =0
dt dt 2r
260t = 64 ⇒ t = 15 min ,
b.

r| |
= 17 km
TC

CHAPTER 8
CHECK YOUR UNDERSTANDING
8.1. The semi-major axis for the highly elliptical orbit of Halley’s comet is 17.8 AU and is the average of the perihelion and
aphelion. This lies between the 9.5 AU and 19 AU orbital radii for Saturn and Uranus, respectively. The radius for a circular orbit
is the same as the semi-major axis, and since the period increases with an increase of the semi-major axis, the fact that Halley’s
period is between the periods of Saturn and Uranus is expected.
CONCEPTUAL QUESTIONS
31. The speed is greatest where the satellite is closest to the large mass and least where farther away—at the periapsis and apoapsis,
respectively. It is conservation of angular momentum that governs this relationship. But it can also be gleaned from conservation
of energy, the kinetic energy must be greatest where the gravitational potential energy is the least (most negative). The force, and
hence acceleration, is always directed towards M in the diagram, and the velocity is always tangent to the path at all points. The
acceleration vector has a tangential component along the direction of the velocity at the upper location on the y-axis; hence, the
satellite is speeding up. Just the opposite is true at the lower position.
PROBLEMS
37. 1.98 × 10 30 kg ; The values are the same within 0.05%.
39. Compare m58348 (https://legacy.cnx.org/content/m58348/latest/#fs-id1168327874347) and ??? to see that they

This OpenStax book is available for free at https://legacy.cnx.org/content/col12169/1.46


Answer Key 393

differ only in that the circular radius, r, is replaced by the semi-major axis, a. Therefore, the mean radius is one-half the sum of the
aphelion and perihelion, the same as the semi-major axis.
41. The semi-major axis, 3.78 AU is found from the equation for the period. This is one-half the sum of the aphelion and
perihelion, giving an aphelion distance of 4.95 AU.
43. 1.75 years
394 Answer Key

This OpenStax book is available for free at https://legacy.cnx.org/content/col12169/1.46


Index 395

INDEX
A differentiation, 288 International Astronomical
acceleration due to gravity, 91, dimension, 25, 39 Union, 272
97 dimensionally consistent, 26, 39
dimensionless, 26, 39 J
acceleration vector, 222, 251
direction angle, 186, 204 Jupiter, 265, 267, 268, 273, 274,
accuracy, 32, 39
discrepancy, 33, 39 276, 276
air resistance, 91
Albert Einstein, 126 displacement, 64, 97, 171, 204 K
Andromeda galaxy, 56, 57 displacement vector, 215, 251 Kepler, 281, 285
angular acceleration, 111, 121 distance traveled, 66, 97 Kepler’s first law, 282, 288
angular frequency, 241, 251 distributive, 176, 204 Kepler’s second law, 283, 288
angular position, 109, 121 Kepler’s third law, 285, 288
E
angular velocity, 110, 121 kilogram, 20, 39
Earth, 51, 268, 274
antiparallel, 172 kinematics, 63, 97
eccentricity, 288
antiparallel vectors, 204 kinematics of rotational motion,
elapsed time, 66, 97
aphelion, 283, 288 114, 121
ellipse, 282, 288
associative, 176, 204
English units, 18, 39 L
asteroid, 288
equal vectors, 195, 204 Large Magellanic Cloud, 56
asteroids, 269
Eris, 265, 266 law, 17, 39
astronomical unit (AU), 285, 288
estimation, 29, 39 Length contraction, 143
average acceleration, 73, 97
event, 147, 161 length contraction, 161
average speed, 71, 97
average velocity, 66, 97 F light cone, 152
Fermi, 16 Local Group, 58
B Lorentz factor, 134
first postulate, 128
base quantities, 19 Lorentz transformation, 149,
first postulate of special
base quantity, 39 161
relativity, 161
base unit, 19, 39
focus, 288 M
Brahe, 280
frame of reference, 64 magnitude, 171, 204
Brownian motion, 217
free fall, 91, 97 major axis, 288
C Makemake, 266
G
centripetal acceleration, 239, Mars, 10, 263, 268, 274
Galilean relativity, 127, 161
251 Mars Climate Orbiter, 25
Galilean transformation, 148,
Ceres, 108, 265, 266 Mercury, 266, 267, 274, 275
161
classical (Galilean) velocity meteor, 270, 288
Galileo, 90, 280
addition, 158, 161 meteorite, 270, 288
Ganymede, 275
comet, 288 meter, 20, 39
giant planet, 288
Comet Shoemaker–Levy 9, 276, method of adding percents, 39
giant planets, 267
276 metric system, 21, 39
comets, 270 H Michelson-Morley experiment,
commutative, 176, 204 Halley, 286 128, 161
component form of a vector, 204 Halley’s comet, 286 Milky Way, 12
conversion factor, 23, 39 Haumea, 266 Milky Way Galaxy, 52, 54
conversion factors, 23 model, 16, 39
coordinate system, 183 I Moon, 51, 264, 266, 274
Crab Nebula, 11 inertial frame of reference, 127, muons, 135
Curie, 16 161
instantaneous acceleration, 75, N
D 97 Neptune, 267, 268, 274
density, 268 instantaneous angular null vector, 195, 204
derived quantity, 19, 39 acceleration, 111, 121
derived units, 19, 39 instantaneous angular velocity, O
difference of two vectors, 176, 110, 121 orbit, 281, 288
204 instantaneous speed, 72, 97 orbital period, 285
Differentiation, 274 instantaneous velocity, 69, 97 orbital period (P), 288
396 Index

orbital speed, 288 special theory of relativity, 127,


order of magnitude, 14, 39 161 W
Orion Nebula, 48, 279 speed of light, 128 world line, 161
orthogonal vectors, 172, 204 Sun, 52
P T
parallel vectors, 172, 204 tail-to-head geometric
parallelogram rule, 204 construction, 180, 204
percent uncertainty, 33, 39 tangential acceleration, 243,
perihelion, 282, 288 251
physical quantity, 18, 39 tangential speeds, 110
physics, 12, 39 terrestrial planet, 288
planetesimal, 278 terrestrial planets, 266
planetesimals, 279, 288 the component form of a vector,
Pluto, 265, 266, 266, 267, 275 184
polar coordinate system, 190, the parallelogram rule, 179
204 the theory of relativity, 126
polar coordinates, 190, 204 theory, 17, 40
position, 64, 97 thought experiment, 131
position vector, 214, 251 Time dilation, 133
precision, 32, 39 time dilation, 161
projectile motion, 228, 251 time of flight, 234, 251
Proper length, 143 Titan, 274
proper length, 161 total acceleration, 243, 251
proper time, 134, 161 total displacement, 65, 97
Proxima Centauri, 52 trajectory, 227, 234, 251
trans-Neptunian object, 265
R translational motion, 107
radial coordinate, 190, 204 twin paradox, 140, 153
range, 234, 251 two-body pursuit problem, 88,
reference frame, 245, 251 97
relative velocity, 246, 251 two-body pursuit problems, 78
relativistic velocity addition, 158, Tyson, 271
161
rest frame, 127, 161 U
resultant vector, 175, 204 uncertainty, 33, 40
rigid body, 108 unit vector, 176, 204
rotational motion, 107 Unit vectors of the axes, 184
unit vectors of the axes, 204
S units, 18, 40
Sagan, 270 Uranus, 267, 268, 274
Saturn, 267, 268, 269, 269, 273
scalar, 170, 204 V
scalar component, 184, 204 Valles Marineris, 10
scalar equation, 174, 204 vector, 204
scalar quantity, 170, 204 vector components, 183, 204
second, 20, 39 vector equation, 174, 204
second postulate of special vector quantities, 170
relativity, 129, 161 vector quantity, 204
semimajor axis, 288 vector sum, 175, 204
SI units, 18, 39 vectors, 170
significant figures, 35, 39 velocity vector, 218, 251
Simultaneity, 131, 154 Venus, 266, 268, 274
Small Magellanic Cloud, 56 Vesta, 108, 275
solar nebula, 277, 288 Virgo Supercluster, 58
space-time, 151
space-time interval, 152

This OpenStax book is available for free at https://legacy.cnx.org/content/col12169/1.46


Index 397

ATTRIBUTIONS
Collection: The Cosmic Universe
Edited by: Steven Mellema
URL: https://legacy.cnx.org/content/col12169/1.46/
Copyright: Steven Mellema
License: http://creativecommons.org/licenses/by/4.0/

Module: Cosmic Universe Preface


Used here as: Preface
By: Steven Mellema
URL: https://legacy.cnx.org/content/m63936/1.10/
Copyright: Steven Mellema
License: http://creativecommons.org/licenses/by/4.0/
Based on: Preface <http://legacy.cnx.org/content/m62204/1.7> by OpenStax University Physics.

Module: Beginning Your Study of Physics


Used here as: Introducing the Story
By: Steven Mellema
URL: https://legacy.cnx.org/content/m63978/1.14/
Copyright: Steven Mellema
License: http://creativecommons.org/licenses/by/4.0/

Module: Cosmic Universe Introduction


Used here as: Introducing Astrophysics
By: Steven Mellema
URL: https://legacy.cnx.org/content/m63947/1.10/
Copyright: Steven Mellema
License: http://creativecommons.org/licenses/by/4.0/
Based on: Introduction <http://legacy.cnx.org/content/m58268/1.5> by OpenStax.

Module: Cosmic Universe The Scope and Scale of Physics


Used here as: The Scope and Scale of Physics
By: Steven Mellema
URL: https://legacy.cnx.org/content/m64024/1.1/
Copyright: Steven Mellema
License: http://creativecommons.org/licenses/by/4.0/
Based on: The Scope and Scale of Physics <http://legacy.cnx.org/content/m58269/1.6> by OpenStax.

Module: Cosmic Universe Units and Standards


Used here as: Units and Standards
By: Steven Mellema
URL: https://legacy.cnx.org/content/m64027/1.1/
Copyright: Steven Mellema
License: http://creativecommons.org/licenses/by/4.0/
Based on: Units and Standards <http://legacy.cnx.org/content/m58270/1.5> by OpenStax.

Module: Cosmic Universe Unit Conversion


Used here as: Unit Conversion
By: Steven Mellema
URL: https://legacy.cnx.org/content/m63948/1.3/
Copyright: Steven Mellema
License: http://creativecommons.org/licenses/by/4.0/
Based on: Unit Conversion <http://legacy.cnx.org/content/m58271/1.5> by OpenStax.

Module: Dimensional Analysis


By: OpenStax
URL: https://legacy.cnx.org/content/m58272/1.7/
398 Index

Copyright: Rice University


License: http://creativecommons.org/licenses/by/4.0/

Module: Cosmic Universe Estimates and Fermi Calculations


Used here as: Estimates and Fermi Calculations
By: Steven Mellema
URL: https://legacy.cnx.org/content/m64025/1.4/
Copyright: Steven Mellema
License: http://creativecommons.org/licenses/by/4.0/
Based on: Estimates and Fermi Calculations <http://legacy.cnx.org/content/m58273/1.5> by OpenStax.

Module: Cosmic Universe Significant Figures


Used here as: Significant Figures
By: Steven Mellema
URL: https://legacy.cnx.org/content/m63995/1.2/
Copyright: Steven Mellema
License: http://creativecommons.org/licenses/by/4.0/
Based on: Significant Figures <http://legacy.cnx.org/content/m58274/1.5> by OpenStax.

Module: Solving Problems in Physics


By: OpenStax
URL: https://legacy.cnx.org/content/m58275/1.5/
Copyright: Rice University
License: http://creativecommons.org/licenses/by/4.0/

Module: Numbers in the Universe


Used here as: Use of Numbers
By: Steven Mellema
URL: https://legacy.cnx.org/content/m63977/1.4/
Copyright: Steven Mellema
License: http://creativecommons.org/licenses/by/4.0/
Based on: Numbers in Astronomy <http://legacy.cnx.org/content/m59751/1.5> by OpenStax Astronomy.

Module: Derived copy of Consequences of Light Travel Time


Used here as: Consequences of Light Travel Time
By: Steven Mellema
URL: https://legacy.cnx.org/content/m63983/1.1/
Copyright: Steven Mellema
License: http://creativecommons.org/licenses/by/4.0/
Based on: Consequences of Light Travel Time <http://legacy.cnx.org/content/m59752/1.4> by OpenStax
Astronomy.

Module: Cosmic Universe A Tour of the Universe


Used here as: A Tour of the Universe
By: Steven Mellema
URL: https://legacy.cnx.org/content/m63984/1.2/
Copyright: Steven Mellema
License: http://creativecommons.org/licenses/by/4.0/
Based on: A Tour of the Universe <http://legacy.cnx.org/content/m59753/1.4> by OpenStax Astronomy.

Module: Cosmic Universe The Universe on the Large Scale


Used here as: The Universe on the Large Scale
By: Steven Mellema
URL: https://legacy.cnx.org/content/m63985/1.1/
Copyright: Steven Mellema
License: http://creativecommons.org/licenses/by/4.0/
Based on: The Universe on the Large Scale <http://legacy.cnx.org/content/m59754/1.4> by OpenStax Astronomy.

Module: Cosmic Universe The Universe of the Very Small

This OpenStax book is available for free at https://legacy.cnx.org/content/col12169/1.46


Index 399

Used here as: The Universe of the Very Small


By: Steven Mellema
URL: https://legacy.cnx.org/content/m63986/1.2/
Copyright: Steven Mellema
License: http://creativecommons.org/licenses/by/4.0/
Based on: The Universe of the Very Small <http://legacy.cnx.org/content/m59760/1.4> by OpenStax Astronomy.

Module: Cosmic Universe A Conclusion and a Beginning


Used here as: A Conclusion and a Beginning
By: Steven Mellema
URL: https://legacy.cnx.org/content/m63987/1.1/
Copyright: Steven Mellema
License: http://creativecommons.org/licenses/by/4.0/
Based on: A Conclusion and a Beginning <http://legacy.cnx.org/content/m59762/1.4> by OpenStax Astronomy.

Module: Derived copy of 1D Kinematics Introduction


Used here as: Introduction to Kinematics
By: Steven Mellema
URL: https://legacy.cnx.org/content/m63994/1.1/
Copyright: Steven Mellema
License: http://creativecommons.org/licenses/by/4.0/
Based on: Introduction <http://legacy.cnx.org/content/m58281/1.4> by OpenStax.

Module: Cosmic Universe Position, Displacement, and Average Velocity


Used here as: Position, Displacement and Average Velocity
By: Steven Mellema
URL: https://legacy.cnx.org/content/m64026/1.1/
Copyright: Steven Mellema
License: http://creativecommons.org/licenses/by/4.0/
Based on: Position, Displacement, and Average Velocity <http://legacy.cnx.org/content/m58282/1.4> by OpenStax.

Module: Cosmic Universe Instantaneous Velocity and Speed


Used here as: Instantaneous Velocity and Speed
By: Steven Mellema
URL: https://legacy.cnx.org/content/m63980/1.1/
Copyright: Steven Mellema
License: http://creativecommons.org/licenses/by/4.0/
Based on: Instantaneous Velocity and Speed <http://legacy.cnx.org/content/m58283/1.4> by OpenStax.

Module: Cosmic Universe Average and Instantaneous Acceleration


Used here as: Average and Instantaneous Acceleration
By: Steven Mellema
URL: https://legacy.cnx.org/content/m63981/1.3/
Copyright: Steven Mellema
License: http://creativecommons.org/licenses/by/4.0/
Based on: Average and Instantaneous Acceleration <http://legacy.cnx.org/content/m58284/1.4> by OpenStax.

Module: Motion with Constant Acceleration


By: OpenStax
URL: https://legacy.cnx.org/content/m58285/1.4/
Copyright: Rice University
License: http://creativecommons.org/licenses/by/4.0/

Module: Cosmic Universe Free Fall


Used here as: Free Fall
By: Steven Mellema
URL: https://legacy.cnx.org/content/m64004/1.2/
Copyright: Steven Mellema
400 Index

License: http://creativecommons.org/licenses/by/4.0/
Based on: Free Fall <http://legacy.cnx.org/content/m58286/1.4> by OpenStax.

Module: Cosmic Universe Introduction to Rotation


Used here as: Introduction to Rotation
By: Steven Mellema
URL: https://legacy.cnx.org/content/m64031/1.6/
Copyright: Steven Mellema
License: http://creativecommons.org/licenses/by/4.0/
Based on: Introduction <http://legacy.cnx.org/content/m58325/1.3> by OpenStax.

Module: Cosmic Universe Rotational Variables


Used here as: Rotational Variables
By: Steven Mellema
URL: https://legacy.cnx.org/content/m64032/1.5/
Copyright: Steven Mellema
License: http://creativecommons.org/licenses/by/4.0/
Based on: Rotational Variables <http://legacy.cnx.org/content/m58326/1.3> by OpenStax.

Module: Cosmic Universe Rotation with Constant Angular Acceleration


Used here as: Rotation with Constant Angular Acceleration
By: Steven Mellema
URL: https://legacy.cnx.org/content/m64036/1.2/
Copyright: Steven Mellema
License: http://creativecommons.org/licenses/by/4.0/
Based on: Rotation with Constant Angular Acceleration <http://legacy.cnx.org/content/m58327/1.4> by OpenStax.

Module: Cosmic Universe Relating Angular and Translational Quantities


Used here as: Relating Angular and Translational Quantities
By: Steven Mellema
URL: https://legacy.cnx.org/content/m64037/1.4/
Copyright: Steven Mellema
License: http://creativecommons.org/licenses/by/4.0/
Based on: Relating Angular and Translational Quantities <http://legacy.cnx.org/content/m58328/1.3> by OpenStax.

Module: Cosmic Universe Introduction to Relativity


Used here as: Introduction
By: Steven Mellema
URL: https://legacy.cnx.org/content/m64038/1.3/
Copyright: Steven Mellema
License: http://creativecommons.org/licenses/by/4.0/
Based on: Introduction <http://legacy.cnx.org/content/m58555/1.2> by OpenStax.

Module: Invariance of Physical Laws


By: OpenStax
URL: https://legacy.cnx.org/content/m58556/1.2/
Copyright: Rice University
License: http://creativecommons.org/licenses/by/4.0/

Module: Relativity of Simultaneity


By: OpenStax
URL: https://legacy.cnx.org/content/m58656/1.2/
Copyright: Rice University
License: http://creativecommons.org/licenses/by/4.0/

Module: Cosmic Universe Time Dilation


Used here as: Time Dilation
By: Steven Mellema
URL: https://legacy.cnx.org/content/m64161/1.1/

This OpenStax book is available for free at https://legacy.cnx.org/content/col12169/1.46


Index 401

Copyright: Steven Mellema


License: http://creativecommons.org/licenses/by/4.0/
Based on: Time Dilation <http://legacy.cnx.org/content/m58563/1.2> by OpenStax.

Module: Length Contraction


By: OpenStax
URL: https://legacy.cnx.org/content/m58565/1.2/
Copyright: Rice University
License: http://creativecommons.org/licenses/by/4.0/

Module: Cosmic Universe The Lorentz Transformation


Used here as: The Lorentz Transformation
By: Steven Mellema
URL: https://legacy.cnx.org/content/m64039/1.1/
Copyright: Steven Mellema
License: http://creativecommons.org/licenses/by/4.0/
Based on: The Lorentz Transformation <http://legacy.cnx.org/content/m58568/1.2> by OpenStax.

Module: Cosmic Universe Relativistic Velocity Transformation


Used here as: Relativistic Velocity Transformation
By: Steven Mellema
URL: https://legacy.cnx.org/content/m64048/1.4/
Copyright: Steven Mellema
License: http://creativecommons.org/licenses/by/4.0/
Based on: Relativistic Velocity Transformation <http://legacy.cnx.org/content/m58569/1.2> by OpenStax.

Module: Introduction
By: OpenStax
URL: https://legacy.cnx.org/content/m58276/1.4/
Copyright: Rice University
License: http://creativecommons.org/licenses/by/4.0/

Module: Scalars and Vectors


By: OpenStax
URL: https://legacy.cnx.org/content/m58277/1.6/
Copyright: Rice University
License: http://creativecommons.org/licenses/by/4.0/

Module: Coordinate Systems and Components of a Vector


By: OpenStax
URL: https://legacy.cnx.org/content/m58278/1.6/
Copyright: Rice University
License: http://creativecommons.org/licenses/by/4.0/

Module: Cosmic Universe Algebra of Vectors


Used here as: Algebra of Vectors
By: Steven Mellema
URL: https://legacy.cnx.org/content/m64049/1.1/
Copyright: Steven Mellema
License: http://creativecommons.org/licenses/by/4.0/
Based on: Algebra of Vectors <http://legacy.cnx.org/content/m58279/1.6> by OpenStax.

Module: Cosmic Universe Introduction to 2D&3D Kinematics


Used here as: Introduction
By: Steven Mellema
URL: https://legacy.cnx.org/content/m64054/1.1/
Copyright: Steven Mellema
License: http://creativecommons.org/licenses/by/4.0/
Based on: Introduction <http://legacy.cnx.org/content/m58288/1.6> by OpenStax.
402 Index

Module: Cosmic Universe Displacement and Velocity Vectors


Used here as: Displacement and Velocity Vectors
By: Steven Mellema
URL: https://legacy.cnx.org/content/m64162/1.1/
Copyright: Steven Mellema
License: http://creativecommons.org/licenses/by/4.0/
Based on: Displacement and Velocity Vectors <http://legacy.cnx.org/content/m58289/1.4> by OpenStax.

Module: Cosmic Universe Acceleration Vector


Used here as: Acceleration Vector
By: Steven Mellema
URL: https://legacy.cnx.org/content/m64064/1.1/
Copyright: Steven Mellema
License: http://creativecommons.org/licenses/by/4.0/
Based on: Acceleration Vector <http://legacy.cnx.org/content/m58290/1.5> by OpenStax.

Module: Cosmic Universe Projectile Motion


Used here as: Projectile Motion
By: Steven Mellema
URL: https://legacy.cnx.org/content/m64053/1.2/
Copyright: Steven Mellema
License: http://creativecommons.org/licenses/by/4.0/
Based on: Projectile Motion <http://legacy.cnx.org/content/m58291/1.4> by OpenStax.

Module: Cosmic Universe Uniform Circular Motion


Used here as: Uniform Circular Motion
By: Steven Mellema
URL: https://legacy.cnx.org/content/m64051/1.2/
Copyright: Steven Mellema
License: http://creativecommons.org/licenses/by/4.0/
Based on: Uniform Circular Motion <http://legacy.cnx.org/content/m58292/1.4> by OpenStax.

Module: Cosmic Universe Relative Motion in One and Two Dimensions


Used here as: Relative Motion in One and Two Dimensions
By: Steven Mellema
URL: https://legacy.cnx.org/content/m64050/1.2/
Copyright: Steven Mellema
License: http://creativecommons.org/licenses/by/4.0/
Based on: Relative Motion in One and Two Dimensions <http://legacy.cnx.org/content/m58293/1.4> by OpenStax.

Module: Cosmic Universe Introduction to the Solar System


Used here as: Introduction
By: Steven Mellema
URL: https://legacy.cnx.org/content/m64057/1.1/
Copyright: Steven Mellema
License: http://creativecommons.org/licenses/by/4.0/
Based on: Thinking Ahead <http://legacy.cnx.org/content/m59814/1.3> by OpenStax Astronomy.

Module: Cosmic Universe Overview of Our Planetary System


Used here as: Overview of Our Planetary System
By: Steven Mellema
URL: https://legacy.cnx.org/content/m64059/1.1/
Copyright: Steven Mellema
License: http://creativecommons.org/licenses/by/4.0/
Based on: Overview of Our Planetary System <http://legacy.cnx.org/content/m59816/1.3> by OpenStax Astronomy.

Module: Cosmic Universe Composition and Structure of Planets


Used here as: Composition and Structure of Planets

This OpenStax book is available for free at https://legacy.cnx.org/content/col12169/1.46


Index 403

By: Steven Mellema


URL: https://legacy.cnx.org/content/m64060/1.1/
Copyright: Steven Mellema
License: http://creativecommons.org/licenses/by/4.0/
Based on: Composition and Structure of Planets <http://legacy.cnx.org/content/m59819/1.3> by OpenStax
Astronomy.

Module: Cosmic Universe Origin of the Solar System


Used here as: Origin of the Solar System
By: Steven Mellema
URL: https://legacy.cnx.org/content/m64056/1.4/
Copyright: Steven Mellema
License: http://creativecommons.org/licenses/by/4.0/
Based on: Origin of the Solar System <http://legacy.cnx.org/content/m59823/1.4> by OpenStax Astronomy.

Module: Cosmic Universe Kepler's Laws of Planetary Motion


Used here as: Kepler's Laws of Planetary Motion
By: Steven Mellema
URL: https://legacy.cnx.org/content/m64058/1.7/
Copyright: Steven Mellema
License: http://creativecommons.org/licenses/by/4.0/
Based on: Kepler's Laws of Planetary Motion <http://legacy.cnx.org/content/m58350/1.3> by OpenStax.

Module: Cosmic Universe Units


Used here as: Units
By: Steven Mellema
URL: https://legacy.cnx.org/content/m64018/1.1/
Copyright: Steven Mellema
License: http://creativecommons.org/licenses/by/4.0/
Based on: Units <http://legacy.cnx.org/content/m58777/1.5> by OpenStax.

Module: Cosmic Universe Conversion Factors


Used here as: Unit Conversions
By: Steven Mellema
URL: https://legacy.cnx.org/content/m64015/1.1/
Copyright: Steven Mellema
License: http://creativecommons.org/licenses/by/4.0/
Based on: Conversion Factors <http://legacy.cnx.org/content/m58615/1.5> by OpenStax.

Module: Cosmic Universe Fundamental Constants


Used here as: Fundamental Physics Constants
By: Steven Mellema
URL: https://legacy.cnx.org/content/m64016/1.1/
Copyright: Steven Mellema
License: http://creativecommons.org/licenses/by/4.0/
Based on: Fundamental Constants <http://legacy.cnx.org/content/m58616/1.4> by OpenStax.

Module: Cosmic Universe Astronomical Data


Used here as: Astronomical Data
By: Steven Mellema
URL: https://legacy.cnx.org/content/m64014/1.4/
Copyright: Steven Mellema
License: http://creativecommons.org/licenses/by/4.0/
Based on: Astronomical Data <http://legacy.cnx.org/content/m58617/1.5> by OpenStax.

Module: Cosmic Universe Mathematical Formulas


Used here as: Useful Mathematical Formulas
By: Steven Mellema
404 Index

URL: https://legacy.cnx.org/content/m64017/1.1/
Copyright: Steven Mellema
License: http://creativecommons.org/licenses/by/4.0/
Based on: Mathematical Formulas <http://legacy.cnx.org/content/m58618/1.5> by OpenStax.

Module: Chemistry
Used here as: Periodic Table of the Elements
By: OpenStax
URL: https://legacy.cnx.org/content/m58619/1.5/
Copyright: Rice University
License: http://creativecommons.org/licenses/by/4.0/

Module: The Greek Alphabet


By: OpenStax
URL: https://legacy.cnx.org/content/m58620/1.5/
Copyright: Rice University
License: http://creativecommons.org/licenses/by/4.0/

This OpenStax book is available for free at https://legacy.cnx.org/content/col12169/1.46


Index 405

ABOUT CONNEXIONS
Since 1999, Connexions has been pioneering a global system where anyone can create course materials and make them fully
accessible and easily reusable free of charge. We are a Web-based authoring, teaching and learning environment open to anyone
interested in education, including students, teachers, professors and lifelong learners. We connect ideas and facilitate educational
communities. Connexions's modular, interactive courses are in use worldwide by universities, community colleges, K-12 schools,
distance learners, and lifelong learners. Connexions materials are in many languages, including English, Spanish, Chinese,
Japanese, Italian, Vietnamese, French, Portuguese, and Thai.

Anda mungkin juga menyukai