Vous êtes sur la page 1sur 874

APEX

Version .

Gregory Hartman, Ph.D.


Department of Applied Mathema cs
Virginia Military Ins tute

Contribu ng Authors
Troy Siemers, Ph.D.
Department of Applied Mathema cs
Virginia Military Ins tute

Brian Heinold, Ph.D.


Department of Mathema cs and Computer Science
Mount Saint Marys University

Dimplekumar Chalishajar, Ph.D.


Department of Applied Mathema cs
Virginia Military Ins tute

Editor

Jennifer Bowen, Ph.D.


Department of Mathema cs and Computer Science
The College of Wooster

Copyright
Gregory Hartman
Licensed to the public under Crea ve Commons
A ribu on-Noncommercial . Interna onal Public
License

Contents
Table of Contents

iii

Preface

vii

Limits
.
An Introduc on To Limits . . . . .
.
Epsilon-Delta Deni on of a Limit
.
Finding Limits Analy cally . . . . .
.
One Sided Limits . . . . . . . . .
.
Con nuity . . . . . . . . . . . . .
.
Limits Involving Innity . . . . . .

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

Deriva ves
.
Instantaneous Rates of Change: The Deriva ve
.
Interpreta ons of the Deriva ve . . . . . . . .
.
Basic Dieren a on Rules . . . . . . . . . . .
.
The Product and Quo ent Rules . . . . . . . .
.
The Chain Rule . . . . . . . . . . . . . . . . .
.
Implicit Dieren a on . . . . . . . . . . . . .
.
Deriva ves of Inverse Func ons . . . . . . . .

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

The Graphical Behavior of Func ons


.
Extreme Values . . . . . . . . . . .
.
The Mean Value Theorem . . . . . .
.
Increasing and Decreasing Func ons
.
Concavity and the Second Deriva ve
.
Curve Sketching . . . . . . . . . . .

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

Applica ons of the Deriva ve


.
Newtons Method . . . . . . . . . . . . . . . . . . . . . . . . .
.
Related Rates . . . . . . . . . . . . . . . . . . . . . . . . . . .

.
.

Op miza on . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Dieren als . . . . . . . . . . . . . . . . . . . . . . . . . . . .

Integra on
.
An deriva ves and Indenite Integra on
.
The Denite Integral . . . . . . . . . . .
.
Riemann Sums . . . . . . . . . . . . . .
.
The Fundamental Theorem of Calculus . .
.
Numerical Integra on . . . . . . . . . . .

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

Techniques of An dieren a on
.
Subs tu on . . . . . . . . . .
.
Integra on by Parts . . . . . .
.
Trigonometric Integrals . . . .
.
Trigonometric Subs tu on . .
.
Par al Frac on Decomposi on
.
Hyperbolic Func ons . . . . .
.
LHpitals Rule . . . . . . . .
.
Improper Integra on . . . . .

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

Applica ons of Integra on


.
Area Between Curves . . . . . . . . . . . . . . . . . . . .
.
Volume by Cross-Sec onal Area; Disk and Washer Methods
.
The Shell Method . . . . . . . . . . . . . . . . . . . . . .
.
Arc Length and Surface Area . . . . . . . . . . . . . . . .
.
Work . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
.
Fluid Forces . . . . . . . . . . . . . . . . . . . . . . . . .

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

Sequences and Series


.
Sequences . . . . . . . . . . . . . . . . . . .
.
Innite Series . . . . . . . . . . . . . . . . .
.
Integral and Comparison Tests . . . . . . . .
.
Ra o and Root Tests . . . . . . . . . . . . . .
.
Alterna ng Series and Absolute Convergence
.
Power Series . . . . . . . . . . . . . . . . . .
.
Taylor Polynomials . . . . . . . . . . . . . .
.
Taylor Series . . . . . . . . . . . . . . . . . .

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

Curves in the Plane


.
Conic Sec ons . . . . . . . . . . .
.
Parametric Equa ons . . . . . . .
.
Calculus and Parametric Equa ons
.
Introduc on to Polar Coordinates

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

Calculus and Polar Func ons . . . . . . . . . . . . . . . . . . .

Vectors
. Introduc on to Cartesian Coordinates in Space
. An Introduc on to Vectors . . . . . . . . . . .
. The Dot Product . . . . . . . . . . . . . . . . .
. The Cross Product . . . . . . . . . . . . . . . .
. Lines . . . . . . . . . . . . . . . . . . . . . . .
. Planes . . . . . . . . . . . . . . . . . . . . . .

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

Vector Valued Func ons


. VectorValued Func ons . . . . . . . . .
. Calculus and VectorValued Func ons . .
. The Calculus of Mo on . . . . . . . . . .
. Unit Tangent and Normal Vectors . . . . .
. The Arc Length Parameter and Curvature

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

Func
.
.
.
.
.
.
.
.

ons of Several Variables


Introduc on to Mul variable Func ons . . . . .
Limits and Con nuity of Mul variable Func ons .
Par al Deriva ves . . . . . . . . . . . . . . . . .
Dieren ability and the Total Dieren al . . . .
The Mul variable Chain Rule . . . . . . . . . . .
Direc onal Deriva ves . . . . . . . . . . . . . .
Tangent Lines, Normal Lines, and Tangent Planes
Extreme Values . . . . . . . . . . . . . . . . . .

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

Mul
.
.
.
.
.
.

ple Integra on
Iterated Integrals and Area . . . . . . . . .
Double Integra on and Volume . . . . . . .
Double Integra on with Polar Coordinates .
Center of Mass . . . . . . . . . . . . . . .
Surface Area . . . . . . . . . . . . . . . . .
Volume Between Surfaces and Triple Integra

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

A Solu ons To Selected Problems


Index

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

. . .
. . .
. . .
. . .
. . .
on .

A.
A.

P
A Note on Using this Text

Thank you for reading this short preface. Allow us to share a few key points
about the text so that you may be er understand what you will nd beyond this
page.
This text comprises a threevolume series on Calculus. The rst part covers
material taught in many Calc courses: limits, deriva ves, and the basics of
integra on, found in Chapters through . . The second text covers material
o en taught in Calc : integra on and its applica ons, along with an introducon to sequences, series and Taylor Polynomials, found in Chapters through
. The third text covers topics common in Calc or mul variable calc: parametric equa ons, polar coordinates, vectorvalued func ons, and func ons of
more than one variable, found in Chapters through . All three are available
separately for free at www.apexcalculus.com.
Prin ng the en re text as one volume makes for a large, heavy, cumbersome
book. One can certainly only print the pages they currently need, but some prefer to have a nice, bound copy of the text. Therefore this text has been split into
these three manageable parts, each of which can be purchased for under $
at Amazon.com.

For Students: How to Read this Text


Mathema cs textbooks have a reputa on for being hard to read. Highlevel
mathema cal wri ng o en seeks to say much with few words, and this style
o en seeps into texts of lowerlevel topics. This book was wri en with the goal
of being easier to read than many other calculus textbooks, without becoming
too verbose.
Each chapter and sec on starts with an introduc on of the coming material,
hopefully se ng the stage for why you should care, and ends with a look ahead
to see how the justlearned material helps address future problems.
Please read the text; it is wri en to explain the concepts of Calculus. There
are numerous examples to demonstrate the meaning of deni ons, the truth
of theorems, and the applica on of mathema cal techniques. When you encounter a sentence you dont understand, read it again. If it s ll doesnt make
sense, read on anyway, as some mes confusing sentences are explained by later
sentences.
You dont have to read every equa on. The examples generally show all
the steps needed to solve a problem. Some mes reading through each step is
helpful; some mes it is confusing. When the steps are illustra ng a new technique, one probably should follow each step closely to learn the new technique.
When the steps are showing the mathema cs needed to nd a number to be
used later, one can usually skip ahead and see how that number is being used,
instead of ge ng bogged down in reading how the number was found.
Most proofs have been omi ed. In mathema cs, proving something is always true is extremely important, and entails much more than tes ng to see if
it works twice. However, students o en are confused by the details of a proof,
or become concerned that they should have been able to construct this proof

on their own. To alleviate this poten al problem, we do not include the proofs
to most theorems in the text. The interested reader is highly encouraged to nd
proofs online or from their instructor. In most cases, one is very capable of understanding what a theorem means and how to apply it without knowing fully
why it is true.

Interac ve, D Graphics


New to Version . is the addi on of interac ve, D graphics in the .pdf version. Nearly all graphs of objects in space can be rotated, shi ed, and zoomed
in/out so the reader can be er understand the object illustrated.
As of this wri ng, the only pdf viewers that support these D graphics are
Adobe Reader & Acrobat (and only the versions for PC/Mac/Unix/Linux computers, not tablets or smartphones). To ac vate the interac ve mode, click on
the image. Once ac vated, one can click/drag to rotate the object and use the
scroll wheel on a mouse to zoom in/out. (A great way to inves gate an image
is to rst zoom in on the page of the pdf viewer so the graphic itself takes up
much of the screen, then zoom inside the graphic itself.) A CTRL-click/drag pans
the object le /right or up/down. By right-clicking on the graph one can access
a menu of other op ons, such as changing the ligh ng scheme or perspec ve.
One can also revert the graph back to its default view. If you wish to deac ve
the interac vity, one can right-click and choose the Disable Content op on.

Thanks
There are many people who deserve recogni on for the important role they
have played in the development of this text. First, I thank Michelle for her support and encouragement, even as this project from work occupied my me
and a en on at home. Many thanks to Troy Siemers, whose most important
contribu ons extend far beyond the sec ons he wrote or the
gures he
coded in Asymptote for D interac on. He provided incredible support, advice
and encouragement for which I am very grateful. My thanks to Brian Heinold
and Dimplekumar Chalishajar for their contribu ons and to Jennifer Bowen for
reading through so much material and providing great feedback early on. Thanks
to Troy, Lee Dewald, Dan Joseph, Meagan Herald, Bill Lowe, John David, Vonda
Walsh, Geo Cox, Jessica Liber ni and other faculty of VMI who have given me
numerous sugges ons and correc ons based on their experience with teaching
from the text. (Special thanks to Troy, Lee & Dan for their pa ence in teaching
Calc III while I was s ll wri ng the Calc III material.) Thanks to Randy Cone for
encouraging his tutors of VMIs Open Math Lab to read through the text and
check the solu ons, and thanks to the tutors for spending their me doing so.
A very special thanks to Kris Brown and Paul Janiczek who took this opportunity far above & beyond what I expected, me culously checking every solu on
and carefully reading every example. Their comments have been extraordinarily
helpful. I am also thankful for the support provided by Wane Schneiter, who as
my Dean provided me with extra me to work on this project. I am blessed to
have so many people give of their me to make this book be er.

APEX Aordable Print and Electronic teXts


APEX is a consor um of authors who collaborate to produce highquality,
lowcost textbooks. The current textbookwri ng paradigm is facing a potenal revolu on as desktop publishing and electronic formats increase in popularity. However, wri ng a good textbook is no easy task, as the me requirements

alone are substan al. It takes countless hours of work to produce text, write
examples and exercises, edit and publish. Through collabora on, however, the
cost to any individual can be lessened, allowing us to create texts that we freely
distribute electronically and sell in printed form for an incredibly low cost. Having said that, nothing is en rely free; someone always bears some cost. This text
cost the authors of this book their me, and that was not enough. APEX Calculus would not exist had not the Virginia Military Ins tute, through a generous
JacksonHope grant, given the lead author signicant me away from teaching
so he could focus on this text.
Each text is available as a free .pdf, protected by a Crea ve Commons Attribu on - Noncommercial . copyright. That means you can give the .pdf to
anyone you like, print it in any form you like, and even edit the original content
and redistribute it. If you do the la er, you must clearly reference this work and
you cannot sell your edited work for money.
We encourage others to adapt this work to t their own needs. One might
add sec ons that are missing or remove sec ons that your students wont
need. The source les can be found at github.com/APEXCalculus.
You can learn more at www.vmi.edu/APEX.

:L
Calculus means a method of calcula on or reasoning. When one computes
the sales tax on a purchase, one employs a simple calculus. When one nds the
area of a polygonal shape by breaking it up into a set of triangles, one is using
another calculus. Proving a theorem in geometry employs yet another calculus.
Despite the wonderful advances in mathema cs that had taken place into
the rst half of the th century, mathema cians and scien sts were keenly
aware of what they could not do. (This is true even today.) In par cular, two
important concepts eluded mastery by the great thinkers of that me: area and
rates of change.
Area seems innocuous enough; areas of circles, rectangles, parallelograms,
etc., are standard topics of study for students today just as they were then. However, the areas of arbitrary shapes could not be computed, even if the boundary
of the shape could be described exactly.
Rates of change were also important. When an object moves at a constant
rate of change, then distance = rate me. But what if the rate is not constant
can distance s ll be computed? Or, if distance is known, can we discover the
rate of change?
It turns out that these two concepts were related. Two mathema cians, Sir
Isaac Newton and Go ried Leibniz, are credited with independently formula ng
a system of compu ng that solved the above problems and showed how they
were connected. Their system of reasoning was a calculus. However, as the
power and importance of their discovery took hold, it became known to many
as the calculus. Today, we generally shorten this to discuss calculus.
The founda on of the calculus is the limit. It is a tool to describe a parcular behavior of a func on. This chapter begins our study of the limit by approxima ng its value graphically and numerically. A er a formal deni on of
the limit, proper es are established that make nding limits tractable. Once
the limit is understood, then the problems of area and rates of change can be
approached.

An Introduc on To Limits

We begin our study of limits by considering examples that demonstrate key concepts that will be explained as we progress.
Consider the func on y = sinx x . When x is near the value , what value (if
any) is y near?
While our ques on is not precisely formed (what cons tutes near the value

Chapter

Limits
?), the answer does not seem dicult to nd. One might think rst to look at a
graph of this func on to approximate the appropriate y values. Consider Figure
. , where y = sinx x is graphed. For values of x near , it seems that y takes on
values near . . In fact, when x = , then y = sin . , so it makes sense
that when x is near , y will be near . .
Consider this again at a dierent value for x. When x is near , what value (if
any) is y near? By considering Figure . , one can see that it seems that y takes
on values near . But what happens when x = ? We have

y
1

0.8

0.6

.
1

0.5

1.5

sin

The expression / has no value; it is indeterminate. Such an expression gives


no informa on about what is going on with the func on nearby. We cannot nd
out how y behaves near x = for this func on simply by le ng x = .
Finding a limit entails understanding how a func on behaves near a par cular value of x. Before con nuing, it will be useful to establish some nota on. Let
y = f(x); that is, let y be a func on of x for some func on f. The expression the
limit of y as x approaches describes a number, o en referred to as L, that y
nears as x nears . We write all this as

Figure . : sin(x)/x near x = .

y
1

lim y = lim f(x) = L.

x
0.9

This is not a complete deni on (that will come in the next sec on); this is a
pseudo-deni on that will allow us to explore the idea of a limit.
Above, where f(x) = sin(x)/x, we approximated

0.8

.
1

Figure . : sin(x)/x near x = .

.
.

x
.
.

sin(x)/x
.
.
.
.
.
.
.

lim

sin x
.
x

and

lim

sin x
.
x

(We approximated these limits, hence used the symbol, since we are working with the pseudo-deni on of a limit, not the actual deni on.)
Once we have the true deni on of a limit, we will nd limits analy cally;
that is, exactly using a variety of mathema cal tools. For now, we will approximate limits both graphically and numerically. Graphing a func on can provide
a good approxima on, though o en not very precise. Numerical methods can
provide a more accurate approxima on. We have already approximated limits
graphically, so we now turn our a en on to numerical approxima ons.
Consider again limx sin(x)/x. To approximate this limit numerically, we
can create a table of x and f(x) values where x is near . This is done in Figure
. .
No ce that for values of x near , we have sin(x)/x near .
. The x = row
is in bold to highlight the fact that when considering limits, we are not concerned

Figure . : Values of sin(x)/x with x near


.

Notes:

.
with the value of the func on at that par cular x value; we are only concerned
with the values of the func on when x is near .
Now approximate limx sin(x)/x numerically. We already approximated
the value of this limit as graphically in Figure . . The table in Figure . shows
the value of sin(x)/x for values of x near . Ten places a er the decimal point
are shown to highlight how close to the value of sin(x)/x gets as x takes on
values very near . We include the x = row in bold again to stress that we are
not concerned with the value of our func on at x = , only on the behavior of
the func on near .
This numerical method gives condence to say that is a good approximaon of limx sin(x)/x; that is,
lim sin(x)/x .

An Introduc on To Limits

x
- .
- .
- .
.

sin(x)/x
.
.
.
not dened
.
.
.

Figure . : Values of sin(x)/x with x near


.

Later we will be able to prove that the limit is exactly .


We now consider several examples that allow us explore dierent aspects
of the limit concept.
Approxima ng the value of a limit
Example
Use graphical and numerical methods to approximate
lim

0.

To graphically approximate the limit, graph


x+ )

on a small interval that contains . To numerically approximate the limit, create


a table of values where the x values are near . This is done in Figures . and
. , respec vely.
The graph shows that when x is near , the value of y is very near . . By
considering values of x near , we see that y = .
is a be er approxima on.
The graph and the table imply that
x x
lim
x
x x+

This example may bring up a few ques ons about approxima ng limits (and
the nature of limits themselves).
. If a graph does not produce as good an approxima on as a table, why
bother with it?
. How many values of x in a table are enough? In the previous example,
could we have just used x = .
and found a ne approxima on?
Notes:

0.
0.

x x
.
x x+

y = (x x )/( x

0. 8
0.

x
.

Figure . : Graphically approxima ng a


limit in Example .

.
.

x
.
.

x x
x x+

.
.
.
not dened
.
.
.

Figure . : Numerically approxima ng a


limit in Example .

Chapter

Limits
Graphs are useful since they give a visual understanding concerning the behavior of a func on. Some mes a func on may act erra cally near certain
x values which is hard to discern numerically but very plain graphically. Since
graphing u li es are very accessible, it makes sense to make proper use of them.
Since tables and graphs are used only to approximate the value of a limit,
there is not a rm answer to how many data points are enough. Include
enough so that a trend is clear, and use values (when possible) both less than
and greater than the value in ques on. In Example , we used both values less
than and greater than . Had we used just x = .
, we might have been
tempted to conclude that the limit had a value of . . While this is not far o,
we could do be er. Using values on both sides of helps us iden fy trends.

y
1

Approxima ng the value of a limit


Example
Graphically and numerically approximate the limit of f(x) as x approaches ,
where
{
x+
x<
f(x) =
.
x +
x>

0.5

.1

0.5

0.5

Figure . : Graphically approxima ng a


limit in Example .

x
- .
- .
- .
.
.
.

f(x)
.
.
.
.

Figure . : Numerically approxima ng a


limit in Example .

Again we graph f(x) and create a table of its values near x =


S
to approximate the limit. Note that this is a piecewise dened func on, so it
behaves dierently on either side of . Figure . shows a graph of f(x), and on
either side of it seems the y values approach . Note that f( ) is not actually
dened, as indicated in the graph with the open circle.
The table shown in Figure . shows values of f(x) for values of x near . It
is clear that as x takes on values very near , f(x) takes on values very near .
It turns out that if we let x = for either piece of f(x), is returned; this is
signicant and well return to this idea later.
The graph and table allow us to say that limx f(x) ; in fact, we are
probably very sure it equals .

Iden fying When Limits Do Not Exist


A func on may not have a limit for all values of x. That is, we cannot say
limxc f(x) = L for some numbers L for all values of c, for there may not be a
number that f(x) is approaching. There are three ways in which a limit may fail
to exist.
. The func on f(x) may approach dierent values on either side of c.
. The func on may grow without upper or lower bound as x approaches c.
. The func on may oscillate as x approaches c.

Notes:

.
Well explore each of these in turn.

An Introduc on To Limits

Dierent Values Approached From Le and Right


Example
Explore why lim f(x) does not exist, where
x

f(x) =

x x+
x

x
x>

A graph of f(x) around x = and a table are given Figures


. and . , respec vely. It is clear that as x approaches , f(x) does not seem
to approach a single number. Instead, it seems as though f(x) approaches two
dierent numbers. When considering values of x less than (approaching
from the le ), it seems that f(x) is approaching ; when considering values of
x greater than (approaching from the right), it seems that f(x) is approaching . Recognizing this behavior is important; well study this in greater depth
later. Right now, it suces to say that the limit does not exist since f(x) is not
approaching one value as x approaches .
S

Figure . : Observing no limit as x in


Example .

.
.

The Func on Grows Without Bound


Example
Explore why lim /(x ) does not exist.

x
.
.
.

A graph and table of f(x) = /(x ) are given in Figures


, respec vely. Both show that as x approaches , f(x) grows larger

. and .
and larger.
We can deduce this on our own, without the aid of the graph and table. If x
is near , then (x ) is very small, and:
very small number

.5

.5

f(x)
.
.
.

Figure . : Values of f(x) near x =


Example .

in

= very large number.

Since f(x) is not approaching a single number, we conclude that


5

lim

does not exist.

(x )

The Func on Oscillates


Example
Explore why lim sin( /x) does not exist.
x

ure .

Notes:

x
.5

.5

Figure . : Observing no limit as x


in Example .

Two graphs of f(x) = sin( /x) are given in Figures . . Fig(a) shows f(x) on the interval [ , ]; no ce how f(x) seems to oscillate

.
.

x
.
.
.

f(x)
.
.
.

.
.

Figure . : Values of f(x) near x =


Example .

in

Chapter

Limits
near x = . One might think that despite the oscilla on, as x approaches ,
f(x) approaches . However, Figure . (b) zooms in on sin( /x), on the interval [ . , . ]. Here the oscilla on is even more pronounced. Finally, in the
table in Figure . (c), we see sin(x)/x evaluated for values of x near . As x
approaches , f(x) does not appear to approach any value.
It can be shown that in reality, as x approaches , sin( /x) takes on all values
between and innite mes! Because of this oscilla on,
lim sin( /x) does not exist.
x

1
0.5

.5

0.5

0.5

x
.

.5

0.5

x
.
.

(a)
Figure .

.
.
.
.
.

sin( /x)
.
.
.
.
.
.
.

(b)

(c)

: Observing that f(x) = sin( /x) has no limit as x

in Example .

Limits of Dierence Quo ents


We have approximated limits of func ons as x approached a par cular number. We will consider another important kind of limit a er explaining a few key
ideas.
Let f(x) represent the posi on func on, in feet, of some par cle that is moving in a straight line, where x is measured in seconds. Lets say that when x = ,
the par cle is at posi on
., and when x = , the par cle is at
. Another
way of expressing this is to say

f( ) =

and

f( ) =

Since the par cle traveled feet in seconds, we can say the par cles average
velocity was . /s. We write this calcula on using a quo ent of dierences,
or, a dierence quo ent:
f( ) f( )
=

Figure . : Interpre ng a dierence


quo ent as the slope of a secant line.

Notes:

= .

/s.

.
This dierence quo ent can be thought of as the familiar rise over run used
to compute the slopes of lines. In fact, that is essen ally what we are doing:
given two points on the graph of f, we are nding the slope of the secant line
through those two points. See Figure . .
Now consider nding the average speed on another me interval. We again
start at x = , but consider the posi on of the par cle h seconds later. That is,
consider the posi ons of the par cle when x = and when x = + h. The
dierence quo ent is now
f( + h) f( )
f( + h) f( )
=
.
( + h)
h

Let f(x) = . x + . x; note that f( ) =


and f( ) = , as in our
discussion. We can compute this dierence quo ent for all values of h (even
nega ve values!) except h = , for then we get / , the indeterminate form
introduced earlier. For all values h = , the dierence quo ent computes the
average velocity of the par cle over an interval of me of length h star ng at
x= .
For small values of h, i.e., values of h close to , we get average veloci es
over very short me periods and compute secant lines over small intervals. See
Figure . . This leads us to wonder what the limit of the dierence quo ent is
as h approaches . That is,
f( + h) f( )
= ?
lim
h
h

An Introduc on To Limits

(a)
f

(b)
f

As we do not yet have a true deni on of a limit nor an exact method for
compu ng it, we se le for approxima ng the value. While we could graph the
dierence quo ent (where the x-axis would represent h values and the y-axis
would represent values of the dierence quo ent) we se le for making a table.
See Figure . . The table gives us reason to assume the value of the limit is
about . .
Proper understanding of limits is key to understanding calculus. With limits,
we can accomplish seemingly impossible mathema cal things, like adding up an
innite number of numbers (and not get innity) and nding the slope of a line
between two points, where the two points are actually the same point. These
are not just mathema cal curiosi es; they allow us to link posi on, velocity and
accelera on together, connect cross-sec onal areas to volume, nd the work
done by a variable force, and much more.
In the next sec on we give the formal deni on of the limit and begin our
study of nding limits analy cally. In the following exercises, we con nue our
introduc on and approximate the value of limits.

Notes:

(c)
Figure . : Secant lines of f(x) at x =
and x = + h, for shrinking values of h
(i.e., h ).

h
.
.
.
.
.
.

f( +h)f( )
h

.
.

.
.

.
.

Figure . : The dierence quo ent evaluated at values of h near .

Exercises .

Terms and Concepts


. In your own words, what does it mean to nd the limit of
f(x) as x approaches ?
is called

. An expression of the form

. T/F: The limit of f(x) as x approaches is f( ).


. Describe three situa ons where lim f(x) does not exist.
xc

. In your own words, what is a dierence quo ent?

. lim f(x), where


x
{
x x+
f(x) =
x+
. lim f(x), where
x
{
cos x
f(x) =
x + x+
.

x
x>

sin x
cos x

x /
x > /

In Exercises

, a func on f and a value a are


given. Approximate the limit of the dierence quo ent,
f(a + h) f(a)
, using h = . , . .
lim
h
h

. f(x) = x + , a =

. lim x x + x
x

. f(x) = x + .

x+
. lim
x x + x

. lim

. lim
x

. f(x) =

x + x+
x + x+

x+

, a=

. f(x) = x + x , a =

x + x+
x x+

. lim f(x), where


x
{
x+
f(x) =
x

, a=

. f(x) = x + x , a =

x x
. lim
x x x +

. f(x) = ln x,

a=

. f(x) = sin x, a =
x
x>

lim f(x), where

Problems

. lim x + x

x/

f(x) =

In Exercises , approximate the given limits both numerically and graphically.

x
x>

. f(x) = cos x,

a=

Epsilon-Delta Deni on of a Limit

This sec on introduces the formal deni on of a limit. Many refer to this as the
epsilondelta, deni on, referring to the le ers and of the Greek alphabet.
Before we give the actual deni on, lets consider a few informal ways of
describing a limit. Given a func on y = f(x) and an x-value, c, we say that the
limit of the func on f, as x approaches c, is a value L:
. if y tends to L as x tends to c.
. if y approaches L as x approaches c.
. if y is near L whenever x is near c.
The problem with these deni ons is that the words tends, approach,
and especially near are not exact. In what way does the variable x tend to, or
approach, c? How near do x and y have to be to c and L, respec vely?
The deni on we describe in this sec on comes from formalizing . A quick
restatement gets us closer to what we want:

. If x is within a certain tolerance level of c, then the corresponding value y =


f(x) is within a certain tolerance level of L.

The tradi onal nota on for the x-tolerance is the lowercase Greek le er
delta, or , and the y-tolerance is denoted by lowercase epsilon, or . One more
rephrasing of nearly gets us to the actual deni on:

. If x is within units of c, then the corresponding value of y is within units


of L.
We can write x is within units of c mathema cally as
|x c| < ,

which is equivalent to

c < x < c + .

Le ng the symbol represent the word implies, we can rewrite


|x c| < |y L| <

or

as

c < x < c + L < y < L + .

The point is that and , being tolerances, can be any posi ve (but typically
small) values. Finally, we have the formal deni on of the limit with the nota on
seen in the previous sec on.

Notes:

Epsilon-Delta Deni on of a Limit

Chapter

Limits

Deni on

The Limit of a Func on f

Let I be an open interval containing c, and let f be a func on dened on


I, except possibly at c. The limit of f(x), as x approaches c, is L, denoted
by
lim f(x) = L,
xc

means that given any > , there exists >


|x c| < , then |f(x) L| < .

such that for all x = c, if

(Mathema cians o en enjoy wri ng ideas without using any words. Here is
the wordless deni on of the limit:
lim f(x) = L > , >

xc

< |x c| < |f(x) L| < .)

Note the order in which and are given. In the deni on, the y-tolerance
is given rst and then the limit will exist if we can nd an x-tolerance that
works.
An example will help us understand this deni on. Note that the explanaon is long, but it will take one through all steps necessary to understand the
ideas.

= .

= .

Example
Evalua ng a limit using the deni on
Show that lim x = .
x

Choose > . Then ...

Before we use the formal deni on, lets try some numerical
S
tolerances. What if the y tolerance is . , or = . ? How close to does x
have to be so that y is within . units of , i.e., . < y < . ? In this case, we
can proceed as follows:
. < y < .

. < x< .

s.t.

= .

= .

.
.
... choose smaller
than each of these

width
= 1.
}|

{z

width
= 2.2
}|

With = . , we pick any < .


Figure .

So, what is the desired x tolerance? Remember, we want to nd a symmetric


interval of x values, namely < x < + . The lower bound of . is .
units from ; the upper bound of . is . units from . We need the smaller
of these two distances; we must have . . See Figure . .

: Illustra ng the process.

< x < .
< x < .

Notes:

.
Given the y tolerance = . , we have found an x tolerance, . , such
that whenever x is within units of , then y is within units of . Thats what
we were trying to nd.
Lets try another value of .
What if the y tolerance is . , i.e., = . ? How close to does x have to
be in order for y to be within . units of (or . < y < . )? Again, we
just square these values to get .
< x < . , or
.

<x< .

What is the desired x tolerance? In this case we must have .


, which
is the minimum distance from of the two bounds given above.
What we have so far: if = . , then . and if = . , then
.
. A pa ern is not easy to see, so weswitch to general try to determine
symbolically. We start by assuming y = x is within units of :
|y | <

< y <

< x <

< x< +

( ) < x < ( + )
+ <x< + +

( )<x<

+ ( + ).

(Deni on of absolute value)

(y = x)
(Add )
(Square all)
(Expand)
(Rewrite in the desired form)

The desired form in the last step is something < x < + something.
Since we want this last interval to describe an x tolerance around , we have that
either or + , whichever is smaller:
min{ , + }.
Since > , the minimum is . Thats the formula: given an , set
.
We can check this for our previous values. If = . , the formula gives
( . )( . ) = . and when = . , the formula gives ( . )
( . ) = .
.
So given any > , set . Then if |x | < (and x = ), then
|f(x) | < , sa sfying
the deni on of the limit. We have shown formally
(and nally!) that lim x = .
x

Notes:

Epsilon-Delta Deni on of a Limit

Chapter

Limits
The previous example was a li le long in that we sampled a few specic cases
of before handling the general case. Normally
this is not done. The previous

example is also a bit unsa sfying in that


= ; why work so hard to prove
something so obvious? Many - proofs are long and dicult to do. In this secon, we will focus on examples where the answer is, frankly, obvious, because
the nonobvious examples are even harder. In the next sec on we will learn
some theorems that allow us to evaluate limits analy cally, that is, without using the - deni on.
Evalua ng a limit using the deni on
Example
Show that lim x = .
x

Lets do this example symbolically from the start. Let >


S
be given; we want |y | < , i.e., |x | < . How do we nd such that
when |x | < , we are guaranteed that |x | < ?
This is a bit trickier than the previous example, but lets start by no cing that
|x | = |x | |x + |. Consider:
|x | < |x | |x + | < |x | <

.
|x + |

( . )

?
|x + |
We are close to an answer, but the catch is that must be a constant value (so
it cant contain x). There is a way to work around this, but we do have to make an
assump on. Remember that is supposed to be a small number, which implies
that will also be a small value. In par cular, we can (probably) assume that
< . If this is true, then |x | < would imply that |x | < , giving
<x< .

Now, back to the frac on


. If < x < , then < x + < (add
|x + |
to all terms in the inequality). Taking reciprocals, we have
Could we not set =

<

|x + |

which implies

<

which implies

<

|x + |

<
.
|x + |
This suggests that we set

we assume |x | < :

Notes:

( . )

. To see why, let consider what follows when

Epsilon-Delta Deni on of a Limit


y

|x | <

|x | <
|x | |x + | < |x + |
|x | < |x + |
|x | < |x + |

(Our choice of )
(Mul ply by |x + |)

length of
= /5

(Combine le side)
< |x + |

=
|x + |

(Using ( . ) as long as < )

We have arrived at |x | < as desired. Note again, in order to make this


happen we needed to rst be less than . That is a safe assump on; we want
to be arbitrarily small, forcing to also be small.
We have also picked to be smaller than necessary. We could get
. by with
a slightly larger , as shown in Figure . . The dashed outer lines show the
boundaries dened by our choice of . The do ed inner lines show the boundaries dened by se ng = / . Note how these do ed lines are within the
dashed lines. That is perfectly ne; by choosing x within the do ed lines we are
guaranteed that f(x) will be within of .
In summary, given > , set = / . Then |x | < implies |x | <
(i.e. |y | < ) as desired. This shows that lim x = . Figure . gives a
x

visualiza on of this; by restric ng x to values within = / of , we see that


f(x) is within of .

Make note of the general pa ern exhibited in these last two examples. In
some sense, each starts out backwards. That is, while we want to
. start with |x c| < and conclude that
. |f(x) L| < ,
we actually start by assuming
. |f(x) L| < , then perform some algebraic manipula ons to give an
inequality of the form
. |x c| < something.
When we have properly done this, the something on the greater than side of
the inequality becomes our . We can refer to this as the scratchwork phase
of our proof. Once we have , we can formally start with |x c| < and use
algebraic manipula ons to conclude that |f(x) L| < , usually by using the
same steps of our scratchwork in reverse order.

Notes:

length of

z }| {

Figure .
ple .

: Choosing = / in Exam-

Chapter

Limits
We highlight this process in the following example.
Evalua ng a limit using the deni on
Example
Prove that lim x x = .
x

We start our scratchwork by considering |f(x) ( )| < :

|f(x) ( )| <
|x x + | <

(Now factor)

|(x )(x + x )| <


|x | <

.
|x + x |

( . )

We are at the phase of saying that |x | < something, where something=


/|x + x |. We want to turn that something into .
Since x is approaching , we are safe to assume that x is between and .
So
<x<
<x < .

(squared each term)

Since < x < , we can add , x and , respec vely, to each part of the inequality and maintain the inequality.
<x +x<
<x +x

< .

(subtracted from each part)

In Equa on ( . ), we wanted |x | < /|x + x |. The above shows that


given any x in [ , ], we know that
x +x

<

which implies that

<

which implies that

x +x

<
.
x +x

( . )

So we set / . This ends our scratchwork, and we begin the formal proof
(which also helps us understand why this was a good choice of ).
Given , let / . We want to show that when |x | < , then |(x

Notes:

Epsilon-Delta Deni on of a Limit

x) ( )| < . We start with |x | < :


|x | <

|x | <
|x | <

|x | |x + x | <

<

|x + x |

(for x near , from Equa on ( . ))

|x x + | <

|(x x) ( )| < ,
which is what we wanted to show. Thus lim x x = .
x

We illustrate evalua ng limits once more.


Evalua ng a limit using the deni on
Example
Prove that lim ex = .
x

Symbolically, we want to take the equa on |ex | < and


unravel it to the form |x | < . Here is our scratchwork:
S

|ex | <
< ex <
< ex < +

ln( ) < x < ln( + )

(Deni on of absolute value)


(Add )
(Take natural logs)

Making the safe assump on that < ensures the last inequality is valid (i.e.,
so that ln( ) is dened). We can then set to be the minimum of | ln( )|
and ln( + ); i.e.,
= min{| ln( )|, ln( + )} = ln( + ).
Now, we work through the actual the proof:
|x | <

< x <
ln( + ) < x < ln( + ).
ln( ) < x < ln( + ).

Notes:

(Deni on of absolute value)


(since ln( ) < ln( + ))

Note: Recall ln = and ln x < when


< x < . So ln( ) < , hence we
consider its absolute value.

Chapter

Limits
The above line is true by our choice of and by the fact that since | ln( )| >
ln( + ) and ln( ) < , we know ln( ) < ln( + ).
< ex <
< ex

(Exponen ate)

<

(Subtract )

In summary, given > , let = ln( + ). Then |x | < implies


|ex | < as desired. We have shown that lim ex = .
x

We note that we could actually show that limxc ex = ec for any constant c.
We do this by factoring out ec from both sides, leaving us to show limxc exc =
instead. By using the subs tu on u = xc, this reduces to showing limu eu =
which we just did in the last example. As an added benet, this shows that in
fact the func on f(x) = ex is con nuous at all values of x, an important concept
we will dene in Sec on . .
This formal deni on of the limit is not an easy concept grasp. Our examples
are actually easy examples, using simple func ons like polynomials, square
roots and exponen als. It is very dicult to prove, using the techniques given
above, that lim (sin x)/x = , as we approximated in the previous sec on.
x

There is hope. The next sec on shows how one can evaluate complicated
limits using certain basic limits as building blocks. While limits are an incredibly
important part of calculus (and hence much of higher mathema cs), rarely are
limits evaluated using the deni on. Rather, the techniques of the following
sec on are employed.

Notes:

Exercises .

Terms and Concepts

Problems
In Exercises

. What is wrong with the following deni on of a limit?

, prove the given limit using an proof.

. lim x =
x

The limit of f(x), as x approaches a, is K


means that given any > there exists >
such that whenever |f(x) K| < , we have
|x a| < .

. lim x
x

. lim x + x
x

. lim x
x

. Which is given rst in establishing a limit, the xtolerance


or the ytolerance?
. T/F: must always be posi ve.

. lim
x

. lim e x
x

. lim sin x =
. T/F: must always be posi ve.

=
(Hint: use the fact that | sin x| |x|, with

equality only when x = .)

Chapter

Limits

Finding Limits Analy cally

In Sec on . we explored the concept of the limit without a strict deni on,
meaning we could only make approxima ons. In the previous sec on we gave
the deni on of the limit and demonstrated how to use it to verify our approxima ons were correct. Thus far, our method of nding a limit is ) make a really
good approxima on either graphically or numerically, and ) verify our approxima on is correct using a - proof.
Recognizing that - proofs are cumbersome, this sec on gives a series of
theorems which allow us to nd limits much more quickly and intui vely.
Suppose that limx f(x) = and limx g(x) = . What is limx (f(x) +
g(x))? Intui on tells us that the limit should be , as we expect limits to behave
in a nice way. The following theorem states that already established limits do
behave nicely.
Theorem

Basic Limit Proper es

Let b, c, L and K be real numbers, let n be a posi ve integer, and let f and g be
func ons with the following limits:
lim f(x) = L and lim g(x) = K.

xc

xc

The following limits hold.


. Constants:
. Iden ty
. Sums/Dierences:
. Scalar Mul ples:
. Products:
. Quo ents:
. Powers:
. Roots:
. Composi ons:

lim b = b

xc

lim x = c

xc

lim (f(x) g(x)) = L K

xc

lim b f(x) = bL

xc

lim f(x) g(x) = LK

xc

lim f(x)/g(x) = L/K, (K = )

xc

lim f(x)n = Ln

n
lim n f(x) = L
xc

xc

Adjust our previously given limit situa on to:


lim f(x) = L and lim g(x) = K.

xc

Then lim g(f(x)) = K.


xc

Notes:

xL

We make a note about Property # : when n is even, L must be greater than


. If n is odd, then the statement is true for all L.
We apply the theorem to an example.
Using basic limit proper es

Example
Let

lim f(x) = ,

lim g(x) =

and

p(x) = x x + .

Find the following limits:


(
)
. lim f(x) + g(x)

. lim p(x)
x

. lim
x

f(x) + g(x)

(
)
. Using the Sum/Dierence rule, we know that lim f(x) + g(x) = + =
x
.
(
. Using the Scalar Mul ple and Sum/Dierence rules, we nd that lim f(x)+
x
)
g(x) = + = .
. Here we combine the Power, Scalar Mul ple, Sum/Dierence and Constant Rules. We show quite a few steps, but in general these can be omitted:
lim p(x) = lim ( x x + )

= lim x lim x + lim


x

=
=

Part of the previous example demonstrates how the limit of a quadra c


polynomial can be determined using the proper es of Theorem . Not only that,
recognize that
lim p(x) = = p( );
x

i.e., the limit at was found just by plugging into the func on. This holds
true for all polynomials, and also for ra onal func ons (which are quo ents of
polynomials), as stated in the following theorem.

Notes:

Finding Limits Analy cally

Chapter

Limits

Theorem

Limits of Polynomial and Ra onal Func ons

Let p(x) and q(x) be polynomials and c a real number. Then:


. lim p(x) = p(c)
xc

. lim

xc

p(c)
p(x)
=
, where q(c) = .
q(x)
q(c)

Finding a limit of a ra onal func on


Example
Using Theorem , nd
x x+
lim
.
x x x +
S

Using Theorem , we can quickly state that


lim

x x+
x x +

=
=

( ) ( ) +
( ) ( ) +
= .

It was likely frustra ng in Sec on . to do a lot of work to prove that


lim x =

as it seemed fairly obvious. The previous theorems state that many func ons
behave in such an obvious fashion, as demonstrated by the ra onal func on
in Example .
Polynomial and ra onal func ons are not the only func ons to behave in
such a predictable way. The following theorem gives a list of func ons whose
behavior is par cularly nice in terms of limits. In the next sec on, we will give
a formal name to these func ons that behave nicely.
Theorem

Special Limits

Let c be a real number in the domain of the given func on and let n be a posi ve integer. The
following limits hold:
. lim sin x = sin c

. lim csc x = csc c

. lim ax = ac (a > )

. lim cos x = cos c

. lim sec x = sec c

. lim ln x = ln c

. lim tan x = tan c

. lim cot x = cot c

. lim

xc
xc

xc

Notes:

xc

xc

xc

xc

xc

xc

x=

.
Evalua ng limits analy cally
Example
Evaluate the following limits.
. lim cos x

. lim eln x

. lim (sec x tan x)

. lim

lim cos x sin x

sin x
x

x/

. This is a straigh orward applica on of Theorem . lim cos x = cos =


x
.
. We can approach this in at least two ways. First, by directly applying Theorem , we have:
lim (sec x tan x) = sec

tan

Using the Pythagorean Theorem, this last expression is ; therefore


lim (sec x tan x) = .

We can also use the Pythagorean Theorem from the start.


lim (sec x tan x) = lim

= ,

using the Constant limit rule. Either way, we nd the limit is .


. Applying the Product limit rule of Theorem and Theorem gives
lim cos x sin x = cos(/ ) sin(/ ) =

x/

= .

. Again, we can approach this in two ways. First, we can use the exponenal/logarithmic iden ty that eln x = x and evaluate lim eln x = lim x = .
x

We can also use the Composi on limit rule of Theorem . Using Theorem
, we have lim ln x = ln = . Applying the Composi on rule,
x

lim eln x = lim ex = e = .

Both approaches are valid, giving the same result.

Notes:

Finding Limits Analy cally

Chapter

Limits
. We encountered this limit in Sec on . . Applying our theorems, we attempt to nd the limit as

sin x
sin
lim
.

x
x
This, of course, violates a condi on of Theorem , as the limit of the denominator is not allowed to be . Therefore, we are s ll unable to evaluate
this limit with tools we currently have at hand.
The sec on could have been tled Using Known Limits to Find Unknown
Limits. By knowing certain limits of func ons, we can nd limits involving sums,
products, powers, etc., of these func ons. We further the development of such
compara ve tools with the Squeeze Theorem, a clever and intui ve way to nd
the value of some limits.
Before sta ng this theorem formally, suppose we have func ons f, g and h
where g always takes on values between f and h; that is, for all x in an interval,
f(x) g(x) h(x).

If f and h have the same limit at c, and g is always squeezed between them,
then g must have the same limit as well. That is what the Squeeze Theorem
states.
Theorem

Squeeze Theorem

Let f, g and h be func ons on an open interval I containing c such that


for all x in I,
f(x) g(x) h(x).
If

lim f(x) = L = lim h(x),

xc

then

xc

lim g(x) = L.

xc

It can take some work to gure out appropriate func ons by which to squeeze
the given func on of which you are trying to evaluate a limit. However, that is
generally the only place work is necessary; the theorem makes the evalua ng
the limit part very simple.
We use the Squeeze Theorem in the following example to nally prove that
sin x
= .
lim
x
x

Notes:

Finding Limits Analy cally

Using the Squeeze Theorem


Example
Use the Squeeze Theorem to show that
lim

sin x
= .
x

We begin by considering the unit circle. Each point on the


S
unit circle has coordinates (cos , sin ) for some angle as shown in Figure . .
Using similar triangles, we can extend the line from the origin through the point
to the point ( , tan ), as shown. (Here we are assuming that / .
Later we will show that we can also consider .)
Figure . shows three regions have been constructed in the rst quadrant,
two triangles and a sector of a circle, which are also drawn below. The area of
the large triangle is tan ; the area of the sector is / ; the area of the triangle
contained inside the sector is sin . It is then clear from the diagram that

sin

tan

Mul ply all terms by

sin

sin

, giving

cos

.
sin

Taking reciprocals reverses the inequali es, giving


cos

sin
.

(These inequali es hold for all values of near , even nega ve values, since
cos() = cos and sin() = sin .)
Now take limits.
lim cos lim

Notes:

(cos , sin )

Figure .
angles.

tan

(1, tan )

sin
lim

(1, 0)

: The unit circle and related tri-

Chapter

Limits
sin

sin

lim

sin
= .
Clearly this means that lim

cos lim

Two notes about the previous example are worth men oning. First, one
might be discouraged by this applica on, thinking I would never have come up
with that on my own. This is too hard! Dont be discouraged; within this text we
will guide you in your use of the Squeeze Theorem. As one gains mathema cal
maturity, clever proofs like this are easier and easier to create.
Second, this limit tells us more than just that as x approaches , sin(x)/x
approaches . Both x and sin x are approaching , but the ra o of x and sin x
approaches , meaning that they are approaching in essen ally the same way.
Another way of viewing this is: for small x, the func ons y = x and y = sin x are
essen ally indis nguishable.
We include this special limit, along with three others, in the following theorem.
Theorem

Special Limits

. lim

sin x
=
x

. lim

cos x
x

. lim ( + x) x = e
x

. lim
x

ex
x

A short word on how to interpret the la er three limits. We know that as


x goes to , cos x goes to . So, in the second limit, both the numerator and
denominator are approaching . However, since the limit is , we can interpret
this as saying that cos x is approaching faster than x is approaching .
In the third limit, inside the parentheses we have an expression that is approaching (though never equaling ), and we know that raised to any power
is s ll . At the same me, the power is growing toward innity. What happens
to a number near raised to a very large power? In this par cular case, the
result approaches Eulers number, e, approximately .
.
In the fourth limit, we see that as x , ex approaches just as fast as
x , resul ng in a limit of .

Notes:

Finding Limits Analy cally

Our nal theorem for this sec on will be mo vated by the following example.
Using algebra to evaluate a limit
Example
Evaluate the following limit:
lim

x
.
x

We begin by a emp ng to apply Theorem and subs tu ng


S
for x in the quo ent. This gives:
lim

x
x

and indeterminate form. We cannot apply the theorem.


By graphing the func on, as in Figure . , we see that the func on seems
to be linear, implying that the limit should be easy to evaluate. Recognize that
the numerator of our quo ent can be factored:
x
x

(x )(x + )
.
x

The func on is not dened when x = , but for all other x,


x
x

(x )(x + )
(x )(x + )
=
=x+ .
x
x

Clearly lim x + = . Recall that when considering limits, we are not concerned
x

with the value of the func on at , only the value the func on approaches as x
approaches . Since (x )/(x ) and x + are the same at all points except
x = , they both approach the same value as x approaches . Therefore we can
conclude that
x
= .
lim
x x
The key to the above example is that the func ons y = (x )/(x ) and
y = x + are iden cal except at x = . Since limits describe a value the func on
is approaching, not the value the func on actually a ains, the limits of the two
func ons are always equal.

Notes:

.
Figure . : Graphing f in Example
understand a limit.

to

Chapter

Limits

Theorem

Limits of Func ons Equal At All But One Point

Let g(x) = f(x) for all x in an open interval, except possibly at c, and let
lim g(x) = L for some real number L. Then
xc

lim f(x) = L.

xc

The Fundamental Theorem of Algebra tells us that when dealing with a rag(x)
onal func on of the form g(x)/f(x) and directly evalua ng the limit lim
xc f(x)
returns / , then (x c) is a factor of both g(x) and f(x). One can then use
algebra to factor this term out, cancel, then apply Theorem . We demonstrate
this once more.
Evalua ng a limit using Theorem
x x x+
.
Evaluate lim
x
x + x x+
Example

We begin by applying Theorem and subs tu ng for x.


S
This returns the familiar indeterminate form of / . Since the numerator and
denominator are each polynomials, we know that (x ) is factor of each. Using
whatever method is most comfortable to you, factor out (x ) from each (using
polynomial division, synthe c division, a computer algebra system, etc.). We
nd that
x x x+
(x )(x + x )
=
.
x + x x+
(x )( x + x )
We can cancel the (x ) terms as long as x = . Using Theorem we conclude:
lim

x x x+
x + x x+

(x )(x + x )
(x )( x + x )
(x + x )
= lim
x ( x + x )
= lim
x

We end this sec on by revisi ng a limit rst seen in Sec on . , a limit of


a dierence quo ent. Let f(x) = . x + . x; we approximated the limit
f( + h) f( )
. . We formally evaluate this limit in the following exlim
h
h
ample.

Notes:

.
Evalua ng the limit of a dierence quo ent
f( + h) f( )
Let f(x) = . x + . x; nd lim
.
h
h
Example

Since f is a polynomial, our rst a empt should be to emS


ploy Theorem and subs tute for h. However, we see that this gives us / .
Knowing that we have a ra onal func on hints that some algebra will help. Consider the following steps:
(
)
. ( + h) + . ( + h) . ( ) + . ( )
f( + h) f( )
= lim
lim
h
h
h
h
. ( + h+h )+ . + . h
= lim
h
h
. h + . h
= lim
h
h
h( . h + . )
= lim
h
h
= lim ( . h + . ) (using Theorem , as h = )
h

= .

(using Theorem )

This matches our previous approxima on.


This sec on contains several valuable tools for evalua ng limits. One of the
main results of this sec on is Theorem ; it states that many func ons that we
use regularly behave in a very nice, predictable way. In the next sec on we give
a name to this nice behavior; we label such func ons as con nuous. Dening
that term will require us to look again at what a limit is and what causes limits
to not exist.

Notes:

Finding Limits Analy cally

Exercises .

Terms and Concepts

Using:
lim f(x) =

. Explain in your own words, without using - formality, why


lim b = b.
xc

. Explain in your own words, without using - formality, why


lim x = c.
xc

lim g(x) =

lim g(x) =

evaluate the limits given in Exercises


If it is not possible to know, state so.

. Sketch a graph that visually demonstrates the Squeeze Theorem.

(
)
. lim g f(x)

. lim f(x)g(x)
x

In Exercises

, evaluate the given limit.

. lim x x +
x

(b) lim g(x) =

. lim

(c) lim f(x)/g(x) =


x

What can be said about the rela ve sizes of f(x) and g(x)
as x approaches ?

x
x

lim cos x sin x

x/

. lim ln x
x

Problems

lim f(x) =

lim g(x) =

lim g(x) =

evaluate the limits given in Exercises


If it is not possible to know, state so.

lim csc x

x/

. lim ln( + x)
, where possible.

. lim

x + x+
x x

. lim

x+
x

. lim (f(x) + g(x))


x

. lim ( f(x)/g(x))
x

. lim

. lim

f(x) g(x)
g(x)
f(x)
g(x)

x x

. lim

Using:
lim f(x) =

(
)
. lim g f(x)

. lim

x x
x x+

. lim

x + x
x x

. lim

x + x
x x+

. lim

x x+
x x

(
)
. lim f g(x)

(
)
. lim g f(f(x))

. lim

x x
x + x+

. lim f(x)g(x) f (x) + g (x)

. lim

x + x+
x x

, where possible.

(
)
. lim cos g(x)

(a) lim f(x) =

. lim f(x)g(x)

. What does the text mean when it says that certain funcons behavior is nice in terms of limits? What, in parcular, is nice?

. You are given the following informa on:

lim f(x) =

Use the Squeeze Theorem in Exercises


priate, to evaluate the given limit.
. lim x sin
x

Exercises

challenge your understanding of limits but


can be evaluated using the knowledge gained in this sec on.
. lim

sin x
x

. lim

sin x
x

. lim

ln( + x)
x

. lim

sin x
, where x is measured in degrees, not radians.
x

. lim f(x), where x


x

. lim f(x), where x


x +

, where appro-

( )

. lim sin x cos


x

f(x) x .
f(x) x on [ , ].

Chapter

Limits

One Sided Limits

We introduced the concept of a limit gently, approxima ng their values graphically and numerically. Next came the rigorous deni on of the limit, along with
an admi edly tedious method for evalua ng them. The previous sec on gave us
tools (which we call theorems) that allow us to compute limits with greater ease.
Chief among the results were the facts that polynomials and ra onal, trigonometric, exponen al and logarithmic func ons (and their sums, products, etc.) all
behave nicely. In this sec on we rigorously dene what we mean by nicely.
In Sec on . we explored the three ways in which limits of func ons failed
to exist:
. The func on approached dierent values from the le and right,
. The func on grows without bound, and
. The func on oscillates.
In this sec on we explore in depth the concepts behind # by introducing
the one-sided limit. We begin with formal deni ons that are very similar to the
deni on of the limit given in Sec on . , but the nota on is slightly dierent
and x = c is replaced with either x < c or x > c.
Deni on

One Sided Limits

Le -Hand Limit
Let I be an open interval containing c, and let f be a func on dened on
I, except possibly at c. The limit of f(x), as x approaches c from the le ,
is L, or, the le hand limit of f at c is L, denoted by
lim f(x) = L,

xc

means that given any > , there exists >


|x c| < , then |f(x) L| < .

such that for all x < c, if

Right-Hand Limit
Let I be an open interval containing c, and let f be a func on dened on I,
except possibly at c. The limit of f(x), as x approaches c from the right,
is L, or, the righthand limit of f at c is L, denoted by
lim f(x) = L,

xc+

means that given any > , there exists >


|x c| < , then |f(x) L| < .

Notes:

such that for all x > c, if

One Sided Limits

Prac cally speaking, when evalua ng a le -hand limit, we consider only values of x to the le of c, i.e., where x < c. The admi edly imperfect nota on
x c is used to imply that we look at values of x to the le of c. The notaon has nothing to do with posi ve or nega ve values of either x or c. A similar
statement holds for evalua ng right-hand limits; there we consider only values
of x to the right of c, i.e., x > c. We can use the theorems from previous sec ons
to help us evaluate these limits; we just restrict our view to one side of c.
We prac ce evalua ng le and right-hand limits through a series of examples.
Example

Evalua ng one sided limits


x
x
, as shown in Figure .
x
<x<

Let f(x) =
following:

. lim f(x)

. lim f(x)

. lim f(x)

. f( )

. lim f(x)

. lim f(x)

. f( )

. f( )

. Find each of the

Figure .

For these problems, the visual aid of the graph is likely more
S
eec ve in evalua ng the limits than using f itself. Therefore we will refer o en
to the graph.
. As x goes to from the le , we see that f(x) is approaching the value of .
Therefore lim f(x) = .
x

. As x goes to from the right, we see that f(x) is approaching the value of .
Recall that it does not ma er that there is an open circle there; we are
evalua ng a limit, not the value of the func on. Therefore lim f(x) = .
x

. The limit of f as x approaches does not exist, as discussed in the rst


sec on. The func on does not approach one par cular value, but two
dierent values from the le and the right.
. Using the deni on and by looking at the graph we see that f( ) = .
. As x goes to from the right, we see that f(x) is also approaching . Therefore lim f(x) = . Note we cannot consider a le -hand limit at as f is
x

not dened for values of x < .

Notes:

: A graph of f in Example

Chapter

Limits
. Using the deni on and the graph, f( ) = .
. As x goes to from the le , we see that f(x) is approaching the value of
. Therefore lim f(x) = .
x

. The graph and the deni on of the func on show that f( ) is not dened.
Note how the le and right-hand limits were dierent at x = . This, of
course, causes the limit to not exist. The following theorem states what is fairly
intui ve: the limit exists precisely when the le and right-hand limits are equal.
Theorem

Limits and One Sided Limits

Let f be a func on dened on an open interval I containing c. Then


lim f(x) = L

xc

if, and only if,


lim f(x) = L

xc

and

lim f(x) = L.

xc+

The phrase if, and only if means the two statements are equivalent: they
are either both true or both false. If the limit equals L, then the le and right
hand limits both equal L. If the limit is not equal to L, then at least one of the
le and right-hand limits is not equal to L (it may not even exist).
One thing to consider in Examples

is that the value of the func on


may/may not be equal to the value(s) of its le /right-hand limits, even when
these limits agree.
Example

Let f(x) =

following.

. lim f(x)

. lim f(x)

. f( )

. lim f(x)

. lim f(x)

. f( )

. f( )

.
Figure .

: A graph of f from Example

Evalua ng limits of a piecewisedened func on


x
<x<
, as shown in Figure . . Evaluate the
(x )
<x<

. lim f(x)
x

Notes:

One Sided Limits

Again we will evaluate each using both the deni on of f and

its graph.

. As x approaches from the le , we see that f(x) approaches . Therefore


lim f(x) = .
x

. As x approaches from the right, we see that again f(x) approaches .


Therefore lim f(x) = .
x +

. The limit of f as x approaches exists and is , as f approaches from both


the right and le . Therefore lim f(x) = .
x

. f( ) is not dened. Note that is not in the domain of f as dened by the


problem, which is indicated on the graph by an open circle when x = .
. As x goes to from the right, f(x) approaches . So lim f(x) = .
x

. f( ) is not dened as is not in the domain of f.


. As x goes to from the le , f(x) approaches . So lim f(x) = .
x

. f( ) is not dened as is not in the domain of f.


Example
Let f(x) =

the following.

Evalua ng limits of a piecewisedened func on


(x )
x , x =
, as shown in Figure . . Evaluate
x=

. lim f(x)
x

. lim f(x)
x

It is clear by looking at the graph that both the le and rightS


hand limits of f, as x approaches , is . Thus it is also clear that the limit is ;
i.e., lim f(x) = . It is also clearly stated that f( ) = .
x

Example {
Let f(x) =
ing.

Notes:

.5

. f( )

. lim f(x)
x

x
x

Evalua ng limits of a piecewisedened func on


x
, as shown in Figure . . Evaluate the follow<x

.
Figure .

: Graphing f in Example

Chapter

Limits

. lim f(x)

. lim f(x)

. lim f(x)

. f( )

It is clear from the deni on of the func on and its graph


S
that all of the following are equal:

.5

lim f(x) = lim f(x) = lim f(x) = f( ) = .

.
Figure .

: Graphing f in Example

In Examples

we were asked to nd both lim f(x) and f( ). Consider


x

the following table:


lim f(x)

Example
Example
Example
Example

does not exist

f( )
not dened

Only in Example do both the func on and the limit exist and agree. This
seems nice; in fact, it seems normal. This is in fact an important situa on
which we explore in the next sec on, en tled Con nuity. In short, a con nuous func on is one in which when a func on approaches a value as x c (i.e.,
when lim f(x) = L), it actually a ains that value at c. Such func ons behave
xc
nicely as they are very predictable.

Notes:

Exercises .

Terms and Concepts

.5

. What are the three ways in which a limit may fail to exist?

.
.5

. T/F: If lim f(x) = , then lim f(x) =


x

.
.5

.5

. T/F: If lim f(x) = , then lim f(x) =

(a) lim f(x)

(d) f( )

(b) lim f(x)

(e) lim f(x)

(c) lim f(x)

(f) lim f(x)

x +

x +

. T/F: If lim f(x) = , then lim f(x) =


x

x +

Problems

.5

In Exercises
graph of f(x).

, evaluate each expression using the given

.
.5
x

.
.5

.5
y

(a) lim f(x)

(c) lim f(x)

(b) lim f(x)

(d) f( )

x
x +

.5

y
.5
x

.
.5

.5

(a) lim f(x)

(d) f( )

(b) lim f(x)

(e) lim f(x)

(c) lim f(x)

(f) lim f(x)

.5

.5

x +

.5

.5

x +

(a) lim f(x)

(c) lim f(x)

(b) lim f(x)

(d) f( )

x +

.5

.
.

.5

.
.5

(d) f( )

(b) lim f(x)

(e) lim f(x)

(c) lim f(x)

(f) lim f(x)

x +
x

.5

(a) lim f(x)


x

x +

(a) lim f(x)

(c) lim f(x)

(b) lim f(x)

(d) f( )

x
x +

. f(x) =
(a)

lim f(x)

x<
x

(c) lim f(x)


x

(d) f()

x +

cos x
x<a
,
sin x
xa
where a is a real number.

. f(x) =

cos x
sin x

(b) lim f(x)

lim f(x)

(e) lim f(x)

lim f(x)

(f) lim f(x)

(a) lim f(x)

(c) lim f(x)

(c) lim f(x)

(g) lim f(x)

(b) lim f(x)

(d) f(a)

(d) f( )

(h) f( )

(a)
(b)

x +

x +

xa+

. f(x) =

xa

xa

x+

x<
x=
x>

(a) lim f(x)

(c) lim f(x)

(b) lim f(x)

(d) f( )

x +

Let a

. f(x) =

be an integer.
(c) lim f(x)

(b) lim f(x)

(d) f(a)

xa

xa+

(c) lim f(x)

(b) lim f(x)

(d) f( )

x +

(a) lim f(x)

(c) lim f(x)

(b) lim f(x)

(d) f(b)

xb

xb

. f(x) =

x+
x

x
x>

xb+

(a) lim f(x)

(c) lim f(x)

(b) lim f(x)

(d) f( )

x
x +

. f(x) =

x + x
sin x

(c) lim f(x)

(b) lim f(x)

(d) f( )

x +

|x|
x

x =
x=

(a) lim f(x)

(c) lim f(x)

(b) lim f(x)

(d) f( )

x +

(a) lim f(x)


x

. f(x) =

x<
x
x

Review
. Evaluate the limit: lim

x + x+
.
x x

. Evaluate the limit: lim

x
x x

. Evaluate the limit: lim

x x+
x x

x
x +
. f(x) =

x +
(a)

(b)

lim f(x)

lim f(x)

x +

x<
x
x>

(e) lim f(x)


x

(f) lim f(x)

.
.

x +

(c) lim f(x)

(g) lim f(x)

(d) f( )

(h) f( )

a(x b) + c
x<b
,
a(x b) + c
xb
where a, b and c are real numbers.

. f(x) =

In Exercises , evaluate the given limits of the piecewise


dened func ons f.

x<
x=
x>

(a) lim f(x)


x

(a) lim f(x)


xa

x
x+

x + x +

. Approximate the limit numerically: lim

x . x+ .
.
x . x

. Approximate the limit numerically: lim

x + . x .
.
x . x+ .

x .

x .

Con nuity

Con nuity

As we have studied limits, we have gained the intui on that limits measure
where a func on is heading. That is, if lim f(x) = , then as x is close to ,
x

f(x) is close to . We have seen, though, that this is not necessarily a good indicator of what f( ) actually this. This can be problema c; func ons can tend
to one value but a ain another. This sec on focuses on func ons that do not
exhibit such behavior.
Deni on

Con nuous Func on

Let f be a func on dened on an open interval I containing c.


. f is con nuous at c if lim f(x) = f(c).
xc

. f is con nuous on I if f is con nuous at c for all values of c in I. If f


is con nuous on (, ), we say f is con nuous everywhere.
A useful way to establish whether or not a func on f is con nuous at c is to
verify the following three things:
. lim f(x) exists,
xc

. f(c) is dened, and


. lim f(x) = f(c).
xc

Finding intervals of con nuity


Example
Let f be dened as shown in Figure . . Give the interval(s) on which f is connuous.
S

y
.5

We proceed by examining the three criteria for con nuity.

. The limits lim f(x) exists for all c between and .


xc

. f(c) is dened for all c between and , except for c =


immediately that f cannot be con nuous at x = .
. The limit lim f(x) = f(c) for all c between
xc
c= .

. We know

and , except, of course, for

We conclude that f is con nuous at every point of ( , ) except at x = .


Therefore f is con nuous on ( , ) ( , ).

Notes:

.5
x

.
Figure .

: A graph of f in Example

Chapter

Limits
Finding intervals of con nuity
Example
The oor func on, f(x) = x, returns the largest integer smaller than the input
x. (For example, f() = = .) The graph of f in Figure . demonstrates
why this is o en called a step func on.
Give the intervals on which f is con nuous.

. The limits limxc f(x) do not exist at the jumps from one step to the
next, which occur at all integer values of c. Therefore the limits exist for
all c except when c is an integer.
2

We examine the three criteria for con nuity.

Figure . : A graph of the step func on


in Example .

. The func on is dened for all values of c.


. The limit lim f(x) = f(c) for all values of c where the limit exist, since each
xc
step consists of just a line.
We conclude that f is con nuous everywhere except at integer values of c. So
the intervals on which f is con nuous are
. . . , ( , ), ( , ), ( , ), ( , ), . . . .
Our deni on of con nuity on an interval species the interval is an open
interval. We can extend the deni on of con nuity to closed intervals by considering the appropriate one-sided limits at the endpoints.
Deni on

Con nuity on Closed Intervals

Let f be dened on the closed interval [a, b] for some real numbers a, b.
f is con nuous on [a, b] if:
. f is con nuous on (a, b),
. lim f(x) = f(a) and
xa+

. lim f(x) = f(b).


xb

We can make the appropriate adjustments to talk about con nuity on half
open intervals such as [a, b) or (a, b] if necessary.

Notes:

.
Determining intervals on which a func on is con nuous
Example
For each of the following func ons, give the domain of the func on and the
interval(s) on which it is con nuous.

. f(x) = /x

. f(x) =

. f(x) = sin x

. f(x) = x

. f(x) = |x|

We examine each in turn.

. The domain of f(x) = /x is (, ) ( , ). As it is a ra onal func on,


we apply Theorem to recognize that f is con nuous on all of its domain.
. The domain of f(x) = sin x is all real numbers, or (, ). Applying
Theorem shows that sin x is con nuous everywhere.

.
The domain of f(x) = x is [ , ). Applying Theorem shows that f(x) =
x is con nuous on its domain of [ , ).

x is [ , ]. Applying Theorems
. The domain of f(x) =
shows that f is con nuous on all of its domain, [ , ].

and

. The domain of f(x){= |x| is (, ). We can dene the absolute value


x x <
func on as f(x) =
. Each piece of this piecewise dened
x x
func on is con nuous on all of its domain, giving that f is con nuous on
(, ) and [ , ). We cannot assume this implies that f is con nuous
on (, ); we need to check that lim f(x) = f( ), as x = is the point
x

where f transi ons from one piece of its deni on to the other. It is
easy to verify that this is indeed true, hence we conclude that f(x) = |x|
is con nuous everywhere.

Con nuity is inherently ed to the proper es of limits. Because of this, the


proper es of limits found in Theorems and apply to con nuity as well. Further, now knowing the deni on of con nuity we can reread Theorem as
giving a list of func ons that are con nuous on their domains. The following
theorem states how con nuous func ons can be combined to form other connuous func ons, followed by a theorem which formally lists func ons that we
know are con nuous on their domains.

Notes:

Con nuity

Chapter

Limits

Theorem

Proper es of Con nuous Func ons

Let f and g be con nuous func ons on an interval I, let c be a real number
and let n be a posi ve integer. The following func ons are con nuous on
I.
. Sums/Dierences: f g
. Constant Mul ples: c f
. Products:

fg

. Quo ents:

f/g

. Powers:

fn

n
f

. Roots:
. Composi ons:

Theorem

(as long as g =

on I)

(if n is even then f on I; if n is odd,


then true for all values of f on I.)

Adjust the deni ons of f and g to: Let f be


con nuous on I, where the range of f on I is J,
and let g be con nuous on J. Then g f, i.e.,
g(f(x)), is con nuous on I.

Con nuous Func ons

The following func ons are con nuous on their domains.


. f(x) = sin x

. f(x) = cos x

. f(x) = tan x

. f(x) = cot x

. f(x) = sec x

. f(x) = csc x

. f(x) = n x,

. f(x) = ln x
. f(x) = ax (a > )

(where n is a posi ve integer)

We apply these theorems in the following Example.

Notes:

Con nuity

Determining intervals on which a func on is con nuous


Example
State the interval(s) on which each of the following func ons is con nuous.

. f(x) = tan x
. f(x) = x +
x

. f(x) = ln x
. f(x) = x sin x
S

propriate.

We examine each in turn, applying Theorems and as apy

. The squareroot terms are con nuous on the intervals [ , ) and (, ],


respec vely. As f is con nuous only where each term is con nuous, f is
con nuous on [ , ], the intersec on of these two intervals. A graph of f
is given in Figure . .
. The func ons y = x and y = sin x are each con nuous everywhere, hence
their product is, too.
. Theorem states that f(x) = tan x is con nuous on its domain. Its domain includes all real numbers except odd mul ples of / . Thus f(x) =
tan x is con nuous on
) (
)
(
(
)

, ,
,...,
,
,
... ,
or, equivalently, on D = {x R | x = n , n is an odd integer}.

. The domain of y = x is [ , ). The range of y = ln x is (, ), but if


we restrict its domain to [ , ) its range is [ , ). So restric ng y =ln x
to the domain of [ , ) restricts its
output is [ , ), on which y = x is
dened. Thus the domain of f(x) = ln x is [ , ).
A common way of thinking of a con nuous func on is that its graph can
be sketched without li ing your pencil. That is, its graph forms a con nuous
curve, without holes, breaks or jumps. While beyond the scope of this text,
this pseudodeni on glosses over some of the ner points of con nuity. Very
strange func ons are con nuous that one would be hard pressed to actually
sketch by hand.
This intui ve no on of con nuity does help us understand another important concept as follows. Suppose f is dened on [ , ] and f( ) = and
f( ) = . If f is con nuous on [ , ] (i.e., its graph can be sketched as a con nuous curve from ( , ) to ( , )) then we know intui vely that somewhere on
[ , ] f must be equal to , and , and , , . . . , , / , etc. In short, f

Notes:

.
Figure .

: A graph of f in Example

(a).

Chapter

Limits
takes on all intermediate values between and . It may take on more values; f may actually equal at some me, for instance, but we are guaranteed all
values between and .
While this no on seems intui ve, it is not trivial to prove and its importance
is profound. Therefore the concept is stated in the form of a theorem.
Theorem

Intermediate Value Theorem

Let f be a con nuous func on on [a, b] and, without loss of generality,


let f(a) < f(b). Then for every value y, where f(a) < y < f(b), there is a
value c in [a, b] such that f(c) = y.
One important applica on of the Intermediate Value Theorem is root nding. Given a func on f, we are o en interested in nding values of x where
f(x) = . These roots may be very dicult to nd exactly. Good approxima ons
can be found through successive applica ons of this theorem. Suppose through
direct computa on we nd that f(a) < and f(b) > , where a < b. The Intermediate Value Theorem states that there is a c in [a, b] such that f(c) = . The
theorem does not give us any clue as to where that value is in the interval [a, b],
just that it exists.
There is a technique that produces a good approxima on of c. Let d be the
midpoint of the interval [a, b] and consider f(d). There are three possibili es:
. f(d) = we got lucky and stumbled on the actual value. We stop as we
found a root.
. f(d) < Then we know there is a root of f on the interval [d, b] we have
halved the size of our interval, hence are closer to a good approxima on
of the root.
. f(d) > Then we know there is a root of f on the interval [a, d] again,we
have halved the size of our interval, hence are closer to a good approxima on of the root.
Successively applying this technique is called the Bisec on Method of root
nding. We con nue un l the interval is suciently small. We demonstrate this
in the following example.
Using the Bisec on Method
Example
Approximate the root of f(x) = x cos x, accurate to three places a er the
decimal.
Consider the graph of f(x) = xcos x, shown in Figure . .
S
It is clear that the graph crosses the x-axis somewhere near x = . . To start the
Notes:

.
Bisec on Method, pick an interval that contains . . We choose [ . , . ]. Note
that all we care about are signs of f(x), not their actual value, so this is all we
display.

y
0.5

Itera on : f( . ) < , f( . ) > , and f( . ) > . So replace . with . and


repeat.
Itera on : f( . ) < , f( . ) > , and at the midpoint, . , we have f( . ) >
. So replace . with . and repeat. Note that we dont need to connue to check the endpoints, just the midpoint. Thus we put the rest of
the itera ons in Table . .
No ce that in the th itera on we have the endpoints of the interval each
star ng with .
. Thus we have narrowed the zero down to an accuracy of
the rst three places a er the decimal. Using a computer, we have
f( .

)= .

f( .

)= .

0.5

0.5

Figure .
cos x.

: Graphing a root of f(x) = x

Either endpoint of the interval gives a good approxima on of where f is . The


Intermediate Value Theorem states that the actual zero is s ll within this interval.
While we do not know its exact value, we know it starts with .
.
This type of exercise is rarely done by hand. Rather, it is simple to program
a computer to run such an algorithm and stop when the endpoints dier by a
preset small amount. One of the authors did write such a program and found
the zero of f, accurate to places a er the decimal, to be .
. While
it took a few minutes to write the program, it took less than a thousandth of a
second for the program to run the necessary
itera ons. In less than hundredths of a second, the zero was calculated to
decimal places (with less
than
itera ons).
It is a simple ma er to extend the Bisec on Method to solve problems similar
to Find x, where f(x) = . For instance, we can nd x, where f(x) = . It
actually works very well to dene a new func on g where g(x) = f(x) . Then
use the Bisec on Method to solve g(x) = .
Similarly, given two func ons f and g, we can use the Bisec on Method to
solve f(x) = g(x). Once again, create a new func on h where h(x) = f(x) g(x)
and solve h(x) = .
In Sec on . another equa on solving method will be introduced, called
Newtons Method. In many cases, Newtons Method is much faster. It relies on
more advanced mathema cs, though, so we will wait before introducing it.
This sec on formally dened what it means to be a con nuous func on.
Most func ons that we deal with are con nuous, so o en it feels odd to have
to formally dene this concept. Regardless, it is important, and forms the basis
of the next chapter.
In the next sec on we examine one more aspect of limits: limits that involve
innity.
Notes:

Con nuity

Itera on #

Interval
[ . , . ]
[ . , . ]
[ . , . ]
[ .
, . ]
[ .
, . ]
[ .
, .
]
[ .
, .
]
[ .
, .
]
[ .
, .
]
[ .
, .
]
[ .
, .
]
[ .
, .
]

Midpoint Sign
f( . ) >
f( . ) >
f( .
)<
f( .
)<
f( .
)>
f( .
)>
f( .
)>
f( .
)<
f( .
)<
f( .
)<
f( .
)<

Figure . : Itera ons of the Bisec on


Method of Root Finding

Exercises .

Terms and Concepts

. a=
y

. In your own words, describe what it means for a func on


to be con nuous.

.5

. In your own words, describe what the Intermediate Value


Theorem states.
.5

. What is a root of a func on?

. Given func ons f and g on an interval I, how can the Bisecon Method be used to nd a value c where f(c) = g(c)?

.5

.5

.5

.5

.5

.5

. a=
y

. T/F: If f is dened on an open interval containing c, and


lim f(x) exists, then f is con nuous at c.
xc

.5

. T/F: If f is con nuous at c, then lim f(x) exists.


xc

.5

. T/F: If f is con nuous at c, then lim f(x) = f(c).


xc+

. T/F: If f is con nuous on [a, b], then lim f(x) = f(a).


xa

. T/F: If f is con nuous on [ , ) and [ , ), then f is con nuous on [ , ).

. a=
y

.5

. T/F: The sum of con nuous func ons is also con nuous.

Problems

.5

In Exercises , a graph of a func on f is given along with


a value a. Determine if f is con nuous at a; if it is not, state
why it is not.

. a=
y

. a=
y

.5

.5
x

.
.5

.5

(a) a =

(b) a =

. a=

(c) a =

.5

.5

.
.5

.5

In Exercises
, determine if f is con nuous at the indicated values. If not, explain why.
{

. f(x) =

sin x
x

. Let f be con nuous on [ , ] where f( ) = and f( ) =


. Does a value < c < exist such that f(c) = ?
Why/why not?

x=
x>

(a) x =
(b) x =
x x
x

. f(x) =

(a) x =

x<
x

(b) x =
{

. f(x) =

x + x+
x + x+

x =
x=

(a) x =

(b) x =
{

. f(x) =

x
x x+

x =
x=

(a) x =
(b) x =

. g(x) =

. h(k) =

k+
t

. g(t) =

+x

In Exercises
, use the Bisec on Method to approximate, accurate to two decimal places, the value of the root
of the given func on in the given interval.
. f(x) = x + x on [ , . ].

. f(x) = ex on [ .

k+

. Let f(x) =

x
x

(a) lim f(x)

, . ].

x<
x

(b) lim f(x)


x +

ek

.
(c) lim f(x)
x

(d) f( )

. Numerically approximate the following limits:

(b)

. h(t) = cos t

Review

(a)

. g(s) = ln s

. Let h be a func on on [ , ] where h( ) = and


h( ) =
. Does a value < c < exist such that
h(c) = ? Why/why not?

. f(x) = ex

. f(k) =

. Let f be con nuous on [ , ] where f( ) = and


f( ) =
. Does a value < c < exist such that
f(c) = ? Why/why not?

. f(x) = cos x sin x on [ . , . ].

. f(x) = x x +

. g(x) =

. Let g be con nuous on [ , ] where g( ) = and g( ) =


. Does a value < c < exist such that g(c) = ?
Why/why not?

. f(x) = sin x / on [ . , .

In Exercises
, give the intervals on which the given
func on is con nuous.

. f(t) =

. f(x) = sin(ex + x )

lim

x . x .
x + . x+

lim

x . x .
x + . x+

x / +

x /

. Give an example of func on f(x) for which lim f(x) does not
x
exist.

Chapter

Limits

y
100

In Deni on we stated that in the equa on lim f(x) = L, both c and L were
xc
numbers. In this sec on we relax that deni on a bit by considering situa ons
when it makes sense to let c and/or L be innity.
As a mo va ng example, consider f(x) = /x , as shown in Figure . .
Note how, as x approaches , f(x) grows very, very large. It seems appropriate,
and descrip ve, to state that

50

.
1

0.5

Limits Involving Innity

0.5

lim

Figure . : Graphing f(x) = /x for values of x near .

= .

Also note that as x gets very large, f(x) gets very, very small. We could represent
this concept with nota on such as
lim

= .

We explore both types of use of in turn.


Deni on

Limit of Innity,

We say lim f(x) = if for every M >


xc

there exists >

such that for

all x = c, if |x c| < , then f(x) M.


This is just like the deni on from Sec on . . In that deni on, given
any (small) value , if we let x get close enough to c (within units of c) then f(x)
is guaranteed to be within of f(c). Here, given any (large) value M, if we let x
get close enough to c (within units of c), then f(x) will be at least as large as
M. In other words, if we get close enough to c, then we can make f(x) as large
as we want. We can dene limits equal to in a similar way.
It is important to note that by saying lim f(x) = we are implicitly sta ng
xc

that the limit of f(x), as x approaches c, does not exist. A limit only exists when
f(x) approaches an actual numeric value. We use the concept of limits that approach innity because it is helpful and descrip ve.

Evalua ng limits involving innity

Example
Find lim
x

(x )

as shown in Figure .

In Example of Sec on . , by inspec ng values of x close


S
to we concluded that this limit does not exist. That is, it cannot equal any real

x
.5

.5

Figure . : Observing innite limit as


x in Example .

Notes:

Limits Involving Innity

number. But the limit could be innite. And in fact, we see that the func on
does appear to be growing larger and larger, as f(. ) =
, f(.
) =
,
f(.
) =
. A similar thing happens onthe other side of . In general,
let a large
value M be given. Let = / M. If x is within of , i.e., if
|x | < / M, then:
|x | <
(x ) <
(x )

> M,

y
50

which is what we wanted to show. So we may say lim /(x ) = .


x

Evalua ng limits involving innity

Example
Find lim
x

, as shown in Figure .

.
1

0.5

0.5

It is easy to see that the func on grows without bound near


S
, but it does so in dierent ways on dierent sides of . Since its behavior is not
= . However, we can make a statement
x
about onesided limits. We can state that lim
= and lim
= .
x + x
x x
consistent, we cannot say that lim
x

50

Figure .

: Evalua ng lim
x

Ver cal asymptotes


If the limit of f(x) as x approaches c from either the le or right (or both) is
or , we say the func on has a ver cal asymptote at c.

10

Finding ver cal asymptotes


x
Find the ver cal asymptotes of f(x) =
.
x
Example

Ver cal asymptotes occur where the func on grows without


S
bound; this can occur at values of c where the denominator is . When x is
near c, the denominator is small, which in turn can make the func on take on
large values. In the case of the given func on, the denominator is at x = .
Subs tu ng in values of x close to and seems to indicate that the func on
tends toward or at those points. We can graphically conrm this by
looking at Figure . . Thus the ver cal asymptotes are at x = .

Notes:

10

.
Figure .

: Graphing f(x) =

x
.
x

Chapter

Limits
y

.
Figure

When a ra onal func on has a ver cal asymptote at x = c, we can conclude


that the denominator is at x = c. However, just because the denominator
is at a certain point does not mean there is a ver cal asymptote there. For
instance, f(x) = (x )/(x ) does not have a ver cal asymptote at x = ,
as shown in Figure . . While the denominator does get small near x = ,
the numerator gets small too, matching the denominator step for step. In fact,
factoring the numerator, we get
f(x) =

. : Graphically showing that


x
does not have an asympf(x) =
x
tote at x = .

(x )(x + )
.
x

Canceling the common term, we get that f(x) = x + for x =

. So there is
clearly no asymptote, rather a hole exists in the graph at x = .
The above example may seem a li le contrived. Another example demonstra ng this important concept is f(x) = (sin x)/x. We have considered this
sin x
func on several mes in the previous sec ons. We found that lim
= ;
x
x
i.e., there is no ver cal asymptote. No simple algebraic cancella on makes this
fact obvious; we used the Squeeze Theorem in Sec on . to prove this.
If the denominator is at a certain point but the numerator is not, then
there will usually be a ver cal asymptote at that point. On the other hand, if the
numerator and denominator are both zero at that point, then there may or may
not be a ver cal asymptote at that point. This case where the numerator and
denominator are both zero returns us to an important topic.
Indeterminate Forms
We have seen how the limits
lim

sin x
x

and

lim

x
x

each return the indeterminate form / when we blindly plug in x = and


x = , respec vely. However, / is not a valid arithme cal expression. It gives
no indica on that the respec ve limits are and .
With a li le cleverness, one can come up / expressions which have a limit
of , , or any other real number. That is why this expression is called indeterminate.
A key concept to understand is that such limits do not really return / .
Rather, keep in mind that we are taking limits. What is really happening is that
the numerator is shrinking to while the denominator is also shrinking to .
The respec ve rates at which they do this are very important and determine the
actual value of the limit.

Notes:

.
An indeterminate form indicates that one needs to do more work in order to
compute the limit. That work may be algebraic (such as factoring and canceling)
or it may require a tool such as the Squeeze Theorem. In a later sec on we will
learn a technique called lHospitals Rule that provides another way to handle
indeterminate forms.
Some other common indeterminate forms are , , /, ,
and . Again, keep in mind that these are the blind results of evalua ng a
limit, and each, in and of itself, has no meaning. The expression does
not really mean subtract innity from innity. Rather, it means One quan ty
is subtracted from the other, but both are growing without bound. What is the
result? It is possible to get every value between and
Note that / and / are not indeterminate forms, though they are not
exactly valid mathema cal expressions, either. In each, the func on is growing
without bound, indica ng that the limit will be , , or simply not exist if the
le - and right-hand limits do not match.
Limits at Innity and Horizontal Asymptotes
At the beginning of this sec on we briey considered what happens to f(x) =
/x as x grew very large. Graphically, it concerns the behavior of the func on to
the far right of the graph. We make this no on more explicit in the following
deni on.
Deni on

Limits at Innity and Horizontal Asymptote

. We say lim f(x) = L if for every >


x

there exists M >

such

there exists M <

such

that if x M, then |f(x) L| < .


. We say lim f(x) = L if for every >
x

that if x M, then |f(x) L| < .


. If lim f(x) = L or lim f(x) = L, we say that y = L is a horizontal
x

asymptote of f.

We can also dene limits such as lim f(x) = by combining this deni on
x
with Deni on .

Notes:

Limits Involving Innity

Chapter

Limits
Approxima ng horizontal asymptotes
x
Approximate the horizontal asymptote(s) of f(x) =
.
x +
Example

ma ng the limits

.5

We will approximate the horizontal asymptotes by approxi-

lim

x
x +

and

lim

x
.
x +

Figure . (a) shows a sketch of f, and part (b) gives values of f(x) for large magnitude values of x. It seems reasonable to conclude from both of these sources
that f has a horizontal asymptote at y = .
Later, we will show how to determine this analy cally.

(a)
x

f(x)

(b)

.
.
.
.

Figure . : Using a graph and a table


to approximate a horizontal asymptote in
Example .

Horizontal asymptotes can take on a variety of forms. Figure . (a) shows


that f(x) = x/(x + ) has a horizontal asymptote of y = , where is approached from both above and below.
Figure . (b) shows that f(x) = x/ x + has two horizontal asymptotes;
one at y = and the other at y = .
Figure . (c) shows that f(x) = (sin x)/x has even more interes ng behavior than at just x = ; as x approaches , f(x) approaches , but oscillates as
it does this.

.5

.5
x

.5

.5

.5

(a)

(b)
Figure .

(c)

: Considering dierent types of horizontal asymptotes.

We can analy cally evaluate limits at innity for ra onal func ons once we
understand lim /x. As x gets larger and larger, the /x gets smaller and smaller,
x

approaching . We can, in fact, make /x as small as we want by choosing a large

Notes:

.
enough value of x. Given , we can make /x < by choosing x > /. Thus
we have limx /x = .
It is now not much of a jump to conclude the following:
lim

=
and
lim
=
x xn
xn
Now suppose we need to compute the following limit:
x

x + x+
.
x x x +
A good way of approaching this is to divide through the numerator and denominator by x (hence dividing by ), which is the largest power of x to appear in
the func on. Doing this, we get
lim

lim

x + x+
x x +

= lim

= lim

= lim

x + x+
/x

/x
x x +
x /x + x/x + /x
x /x x /x + /x
+ /x + /x
.
/x + /x

Then using the rules for limits (which also hold for limits at innity), as well as
the fact about limits of /xn , we see that the limit becomes
+ +
= .
+
This procedure works for any ra onal func on. In fact, it gives us the following theorem.
Theorem

Limits of Ra onal Func ons at Innity

Let f(x) be a ra onal func on of the following form:


f(x) =

an xn + an xn + + a x + a
,
bm xm + bm xm + + b x + b

where any of the coecients may be except for an and bm .


. If n = m, then lim f(x) = lim f(x) =
x

an
.
bm

. If n < m, then lim f(x) = lim f(x) = .


x

. If n > m, then lim f(x) and limx f(x) are both innite.
x

Notes:

Limits Involving Innity

Chapter

Limits

We can see why this is true. If the highest power of x is the same in both
the numerator and denominator (i.e. n = m), we will be in a situa on like the
example above, where we will divide by xn and in the limit all the terms will
approach except for an xn /xn and bm xm /xn . Since n = m, this will leave us with
the limit an /bm . If n < m, then a er dividing through by xm , all the terms in the
numerator will approach in the limit, leaving us with /bm or . If n > m, and
we try dividing through by xn , we end up with all the terms in the denominator
tending toward , while the xn term in the numerator does not approach . This
is indica ve of some sort of innite limit.
Intui vely, as x gets very large, all the terms in the numerator are small in
comparison to an xn , and likewise all the terms in the denominator are small
compared to bn xm . If n = m, looking only at these two important terms, we
have (an xn )/(bn xm ). This reduces to an /bm . If n < m, the func on behaves
like an /(bm xmn ), which tends toward . If n > m, the func on behaves like
an xnm /bm , which will tend to either or depending on the values of n,
m, an , bm and whether you are looking for limx f(x) or limx f(x).
With care, we can quickly evaluate limits at innity for a large number of
func ons by considering the largest powers of x. For instance, consider again
x
, graphed in Figure . (b). When x is very large, x + x .
lim
x
x +
Thus

x +

x = |x|,

and

x
x +

x
.
|x|

This expression is when x is posi ve and when x is nega ve. Hence we get
asymptotes of y = and y = , respec vely.
Example

Finding a limit of a ra onal func on

Conrm analy cally that y =


approximated in Example

is the horizontal asymptote of f(x) =

x
, as
x +

Before using Theorem , lets use the technique of evaluS


a ng limits at innity of ra onal func ons that led to that theorem. The largest
power of x in f is , so divide the numerator and denominator of f by x , then

Notes:

Limits Involving Innity

take limits.
lim

x
x +

= lim

= lim

x /x
x /x + /x

+ /x

= .

We can also use Theorem


directly; in this case n = m so the limit is the
ra o of the leading coecients of the numerator and denominator, i.e., / = .
Example
Use Theorem
.

lim

. lim

Finding limits of ra onal func ons


to evaluate each of the following limits.

x + x
x +

x + x
x x

. lim

x
x

(a)
y
.

. The highest power of x is in the denominator. Therefore, the limit is ; see


Figure . (a).

. The highest power of x is x , which occurs in both the numerator and denominator. The limit is therefore the ra o of the coecients of x , which
is / . See Figure . (b).
. The highest power of x is in the numerator so the limit will be or .
To see which, consider only the dominant terms from the numerator and
denominator, which are x and x. The expression in the limit will behave
like x /(x) = x for large values of x. Therefore, the limit is . See
Figure . (c).

.
(b)

y
20

40

20

40

.
(c)
Figure .
Example

Notes:

: Visualizing the func ons in


.

Chapter

Limits

Chapter Summary
In this chapter we:
dened the limit,
found accessible ways to approximate their values numerically and graphically,
developed a notsoeasy method of proving the value of a limit (- proofs),
explored when limits do not exist,
dened con nuity and explored proper es of con nuous func ons, and
considered limits that involved innity.
Why? Mathema cs is famous for building on itself and calculus proves to be
no excep on. In the next chapter we will be interested in dividing by . That
is, we will want to divide a quan ty by a smaller and smaller number and see
what value the quo ent approaches. In other words, we will want to nd a limit.
These limits will enable us to, among other things, determine exactly how fast
something is moving when we are only given posi on informa on.
Later, we will want to add up an innite list of numbers. We will do so by
rst adding up a nite list of numbers, then take a limit as the number of things
we are adding approaches innity. Surprisingly, this sum o en is nite; that is,
we can add up an innite list of numbers and get, for instance, .
These are just two quick examples of why we are interested in limits. Many
students dislike this topic when they are rst introduced to it, but over me an
apprecia on is o en formed based on the scope of its applicability.

Notes:

Exercises .

Terms and Concepts

. f(x) =

. T/F: If lim f(x) = , then we are implicitly sta ng that the

(x )(x )

(a) lim f(x)

(d) lim f(x)

(b) lim f(x)

(e) lim f(x)

(c) lim f(x)

(f) lim f(x)

limit exists.

x +

. T/F: If lim f(x) = , then we are implicitly sta ng that the

x +

limit exists.

y
0

. T/F: If lim f(x) = , then lim f(x) =


x

x +

. T/F: If lim f(x) = , then f has a ver cal asymptote at


x
x= .
0

. T/F: / is not an indeterminate form.


. f(x) =

. List indeterminate forms.

(a)

. Construct a func on with a ver cal asymptote at x = and


a horizontal asymptote at y = .

ex +
lim f(x)

(b) lim f(x)

(d) lim f(x)

x +

. Let lim f(x) = . Explain how we know that f is/is not


x
con nuous at x = .

1
0.5

Problems

(c) lim f(x)

10

10

10

0.5

In Exercises
of the func on.

, evaluate the given limits using the graph

. f(x) = x sin(x)
. f(x) =

(a)
(b)

(x + )
(a)

lim f(x)

lim f(x)

(b) lim f(x)


x

lim f(x)

y
100
50

5
10

5
50

..

100

. f(x) = cos(x)
(a)

In Exercises
, iden fy the horizontal and ver cal
asymptotes, if any, of the given func on.

lim f(x)

(b) lim f(x)

. f(x) =

x x
x +x

. f(x) =

x x
x x

. f(x) =

x +x
x x

. f(x) =

x
x

. f(x) =

x
x+

. f(x) =

x
x

. f(x) =
(a)

lim f(x)

(b) lim f(x)

In Exercises

100

x + x +
x

. lim

x + x +
x

50

.
.

10

x + x +
x

lim

x + x +
x

, numerically approximate the following

(a) lim f(x)


x

(b) lim f(x)


x

lim

.
In Exercises
limits:

(c) lim f(x)


x

, evaluate the given limit.

. lim

150

Review
. Use an proof to show that
lim x = .
x

. Let lim f(x) = and lim g(x) = . Evaluate the following


x

x
. f(x) =
x x
. f(x) =

x + x
x x + x+

x x+
. f(x) =
x x x+
. f(x) =

x x+
x x

limits.

(a) lim (f + g)(x)

(c) lim (f/g)(x)

(b) lim (fg)(x)

(d) lim f(x)g(x)

x
x<
x+
x
Is f con nuous everywhere?

. Let f(x) =

. Evaluate the limit: lim ln x.


xe

:D

The previous chapter introduced the most fundamental of calculus topics: the
limit. This chapter introduces the second most fundamental of calculus topics:
the deriva ve. Limits describe where a func on is going; deriva ves describe
how fast the func on is going.

Instantaneous Rates of Change: The Deriva ve

A common amusement park ride li s riders to a height then allows them to


freefall a certain distance before safely stopping them. Suppose such a ride
drops riders from a height of
feet. Student of physics may recall that the
height (in feet) of the riders, t seconds a er freefall (and ignoring air resistance,
etc.) can be accurately modeled by f(t) = t +
.
Using this formula, it is easy
to
verify
that,
without
interven on, the riders

. . seconds. Suppose the designers of


will hit the ground at t = .
the ride decide to begin slowing the riders fall a er seconds (corresponding
to a height of
.). How fast will the riders be traveling at that me?
We have been given a posi on func on, but what we want to compute is a
velocity at a specic point in me, i.e., we want an instantaneous velocity. We
do not currently know how to calculate this.
However, we do know from common experience how to calculate an average
velocity. (If we travel
miles in hours, we know we had an average velocity
of
mph.) We looked at this concept in Sec on . when we introduced the
dierence quo ent. We have
rise
change in distance
=
= average velocity.
change in me
run
We can approximate the instantaneous velocity at t = by considering the
average velocity over some me period containing t = . If we make the me
interval small, we will get a good approxima on. (This fact is commonly used.
For instance, high speed cameras are used to track fast moving objects. Distances are measured over a xed number of frames to generate an accurate
approxima on of the velocity.)
Consider the interval from t = to t = (just before the riders hit the
ground). On that interval, the average velocity is
f( ) f( )
f( ) f( )
=
=

/s,

Chapter

Deriva ves
where the minus sign indicates that the riders are moving down. By narrowing
the interval we consider, we will likely get a be er approxima on of the instantaneous velocity. On [ , . ] we have
f( . ) f( )
f( . ) f( )
=
=
.
.

/s.

We can do this for smaller and smaller intervals of me. For instance, over
a me span of / th of a second, i.e., on [ , . ], we have
f( . ) f( )
f( . ) f( )
=
=
.
.

Over a me span of /
f( .

th

of a second, on [ , .

f( . ) f( )
) f( )
=
=
.
.

/s.

], the average velocity is


.

/s.

What we are really compu ng is the average velocity on the interval [ , +h]
for small values of h. That is, we are compu ng
f( + h) f( )
h
h
.
.
.
.

Average Velocity
/s

.
.
.

Figure . : Approxima ng the instantaneous velocity with average veloci es


over a small me period h.

where h is small.
What we really want is for h = , but this, of course, returns the familiar
/ indeterminate form. So we employ a limit, as we did in Sec on . .
We can approximate the value of this limit numerically with small values of
h as seen in Figure . . It looks as though the velocity is approaching
/s.
Compu ng the limit directly gives
lim

f( + h) f( )
= lim
h
h

= lim
h

= lim
h

( + h) +
h h
h
h

(
h

( ) +

Graphically, we can view the average veloci es we computed numerically as


the slopes of secant lines on the graph of f going through the points ( , f( )) and
( +h, f( +h)). In Figure . , the secant line corresponding to h = is shown in
three contexts. Figure . (a) shows a zoomed out version of f with its secant
line. In (b), we zoom in around the points of intersec on between f and the
secant line. No ce how well this secant line approximates f between those two
points it is a common prac ce to approximate func ons with straight lines.

Notes:

Instantaneous Rates of Change: The Deriva ve

As h , these secant lines approach the tangent line, a line that goes
through the point ( , f( )) with the special slope of . In parts (c) and (d) of
Figure . , we zoom in around the point ( , ). In (c) we see the secant line,
which approximates f well, but not as well the tangent line shown in (d).

5
x

.
5

.
(a)

.5

(b)

x
.5

.5

(c)

x
.5

.5

(d)

Figure . : Parts (a), (b) and (c) show the secant line to f(x) with h =
dierent amounts. Part (d) shows the tangent line to f at x = .

, zoomed in

We have just introduced a number of important concepts that we will esh


out more within this sec on. First, we formally dene two of them.

Notes:

Chapter

Deriva ves

Deni on

Deriva ve at a Point

Let f be a con nuous func on on an open interval I and let c be in I. The


deriva ve of f at c, denoted f (c), is
lim

f(c + h) f(c)
,
h

provided the limit exists. If the limit exists, we say that f is dieren able
at c; if the limit does not exist, then f is not dieren able at c. If f is
dieren able at every point in I, then f is dieren able on I.

Deni on

Tangent Line

Let f be con nuous on an open interval I and dieren able at c, for some
c in I. The line with equa on (x) = f (c)(x c) + f(c) is the tangent line
to the graph of f at c; that is, it is the line through (c, f(c)) whose slope
is the deriva ve of f at c.
Some examples will help us understand these deni ons.
Finding deriva ves and tangent lines
Example
Let f(x) = x + x . Find:
. f ( )

. f ( )

. The equa on of the tangent line


to the graph of f at x = .

. The equa on of the tangent line


to the graph f at x = .

. We compute this directly using Deni on .


f( + h) f( )
h
( + h) + ( + h) ( ( ) + ( ) )
= lim
h
h
h + h
= lim
h
h
= lim h +
= .

f ( ) = lim
h

Notes:

Instantaneous Rates of Change: The Deriva ve

. The tangent line at x = has slope f ( ) and goes through the point
( , f( )) = ( , ). Thus the tangent line has equa on, in point-slope form,
y = (x ) + . In slope-intercept form we have y = x .
. Again, using the deni on,
f( + h) f( )
h
( + h) + ( + h) ( ( ) + ( ) )
= lim
h
h
h + h
= lim
h
h
= lim h +

f ( ) = lim
h

y
6

. The tangent line at x = has slope and goes through the point ( , f( )) =
( , ). Thus the tangent line has equa on y = (x )+ = x .
A graph of f is given in Figure . along with the tangent lines at x =
x= .

and

Another important line that can be created using informa on from the derivave is the normal line. It is perpendicular to the tangent line, hence its slope is
the oppositereciprocal of the tangent lines slope.
Deni on

Normal Line

Let f be con nuous on an open interval I and dieren able at c, for some
c in I. The normal line to the graph of f at c is the line with equa on
n(x) =

(x c) + f(c),
f (c)

where f ((c) = ) . When f (c) = , the normal line is the ver cal line
through c, f(c) ; that is, x = c.
Finding equa ons of normal lines
Example
Let f(x) = x + x , as in Example . Find the equa ons of the normal lines
to the graph of f at x = and x = .
S

Notes:

In Example

, we found that f ( ) =

. Hence at x = ,

.
Figure . : A graph of f(x) = x + x
and its tangent lines at x = and x = .

Chapter

Deriva ves
the normal line will have slope /

n(x) =

Figure . : A graph of f(x) = x + x ,


along with its normal line at x = .

. An equa on for the normal line is

(x ) + .

The normal line is plo ed with y = f(x) in Figure . . Note how the line looks
perpendicular to f. (A key word here is looks. Mathema cally, we say that the
normal line is perpendicular to f at x = as the slope of the normal line is the
oppositereciprocal of the slope of the tangent line. However, normal lines may
not always look perpendicular. The aspect ra o of the picture of the graph plays
a big role in this.)
We also found that f ( ) = , so the normal line to the graph of f at x =
will have slope / . An equa on for the normal line is
n(x) =

(x ) +

Linear func ons are easy to work with; many func ons that arise in the
course of solving real problems are not easy to work with. A common prac ce
in mathema cal problem solving is to approximate dicult func ons with not
sodicult func ons. Lines are a common choice. It turns out that at any given
point on the graph of a dieren able func on f, the best linear approxima on
to f is its tangent line. That is one reason well spend considerable me nding
tangent lines to func ons.
One type of func on that does not benet from a tangentline approximaon is a line; it is rather simple to recognize that the tangent line to a line is the
line itself. We look at this in the following example.
Finding the Deriva ve of a Line
Example
Consider f(x) = x + . Find the equa on of the tangent line to f at x =
x= .
S

We nd the slope of the tangent line by using Deni on .


f( + h) f( )
h
h
( + h) + ( + )
= lim
h
h
h
= lim
h h
= lim

f ( ) = lim

= .

Notes:

and

Instantaneous Rates of Change: The Deriva ve

We just found that f ( ) = . That is, we found the instantaneous rate of


change of f(x) = x + is . This is not surprising; lines are characterized by
being the only func ons with a constant rate of change. That rate of change
is called the slope of the line. Since their rates of change are constant, their
instantaneous rates of change are always the same; they are all the slope.
So given a line f(x) = ax + b, the deriva ve at any point x will be a; that is,
f (x) = a.
It is now easy to see that the tangent line to the graph of f at x = is just f,
with the same being true for x = .
We o en desire to nd the tangent line to the graph of a func on without
knowing the actual deriva ve of the func on. In these cases, the best we may
be able to do is approximate the tangent line. We demonstrate this in the next
example.
Numerical Approxima on of the Tangent Line
Example
Approximate the equa on of the tangent line to the graph of f(x) = sin x at
x= .
In order to nd the equa on of the tangent line, we need a
S
slope and a point. The point is given to us: ( , sin ) = ( , ). To compute the
slope, we need the deriva ve. This is where we will make an approxima on.
Recall that
sin( + h) sin
f ( )
h
for a small value of h. We choose (somewhat arbitrarily) to let h = . . Thus
f ( )

sin( . ) sin
.

Thus our approxima on of the equa on of the tangent line is y = .


(x
)+ = .
x; it is graphed in Figure . . The graph seems to imply the
approxima on is rather good.
sin x
= , meaning for values of x near ,
x
sin x x. Since the slope of the line y = x is at x = , it should seem reasonable that the slope of f(x) = sin x is near at x = . In fact, since we
approximated the value of the slope to be .
, we might guess the actual
value is . Well come back to this later.
Recall from Sec on . that lim
x

Consider again Example . To nd the deriva ve of f at x = , we needed to


evaluate a limit. To nd the deriva ve of f at x = , we needed to again evaluate
a limit. We have this process:
Notes:

.5

.5

Figure . : f(x) = sin x graphed with an


approxima on to its tangent line at x = .

Chapter

Deriva ves
do something
to f and c

input specic
number c

return
number f (c)

This process describes a func on; given one input (the value of c), we return
exactly one output (the value of f (c)). The do something box is where the
tedious work (taking limits) of this func on occurs.
Instead of applying this func on repeatedly for dierent values of c, let us
apply it just once to the variable x. We then take a limit just once. The process
now looks like:
do something
to f and x

input variable x

return
func on f (x)

The output is the deriva ve func on, f (x). The f (x) func on will take a
number c as input and return the deriva ve of f at c. This calls for a deni on.
Deni on

Deriva ve Func on

Let f be a dieren able func on on an open interval I. The func on


f (x) = lim
h

f(x + h) f(x)
h

is the deriva ve of f.
Nota on:
Let y = f(x). The following nota ons all represent the deriva ve:
f (x) = y =

df
d
d
dy
=
=
(f) =
(y).
dx
dx
dx
dx

dy
is one symbol; it is not the frac on dy/dx. The
Important: The nota on
dx
nota on, while somewhat confusing at rst, was chosen with care. A frac on
looking symbol was chosen because the deriva ve has many frac onlike proper es. Among other places, we see these proper es at work when we talk about
the units of the deriva ve, when we discuss the Chain Rule, and when we learn
about integra on (topics that appear in later sec ons and chapters).
Examples will help us understand this deni on.
Finding the deriva ve of a func on
Example
Let f(x) = x + x as in Example . Find f (x).

Notes:

.
We apply Deni on

Instantaneous Rates of Change: The Deriva ve

f(x + h) f(x)
h
h
(x + h) + (x + h) ( x + x )
= lim
x
h
h + xh + h
= lim
x
h
= lim h + x +

f (x) = lim

= x+
So f (x) = x + . Recall earlier we found that f ( ) =
our new computa on of f (x) arm these facts.

. Note

Finding the deriva ve of a func on

Example
Let f(x) =
S

and f ( ) =

x+

. Find f (x).
We apply Deni on

f (x) = lim
h

= lim
h

f(x + h) f(x)
h

x+h+
x+
h

Now nd common denominator then subtract; pull /h out front to facilitate


reading.
= lim
h

= lim
h

= lim
h

(
(
(

(x +
x+
(x +

x+
x+h+

)(x + h + ) (x + )(x + h + )
)
(x + h + )
)(x + h + )
)
h
)(x + h + )

(x +

= lim
h (x + )(x + h + )

=
(x + )(x + )

=
(x + )

Notes:

Chapter

Deriva ves
So f (x) =

. To prac ce using our nota on, we could also state


(x + )
(
)
d

=
.
dx x +
(x + )

Finding the deriva ve of a func on


Example
Find the deriva ve of f(x) = sin x.
Before applying Deni on , note that once this is found,
S
we can nd the actual tangent line to f(x) = sin x at x = , whereas we se led
for an approxima on in Example .
sin(x + h) sin x
h
h
sin x cos h + cos x sin h sin x
= lim
h
h
sin x(cos h ) + cos x sin h
= lim
h
h
)
(
sin x(cos h )
cos x sin h
+
= lim
h
h
h
= sin x + cos x

f (x) = lim

Use trig iden ty


sin(x + h) = sin x cos h + cos x sin h

(regroup)

(split into two frac ons)

use lim
h

cos h
h

and lim
h

sin h
=
h

= cos x !

We have found that when f(x) = sin x, f (x) = cos x. This should be somewhat
surprising; the result of a tedious limit process and the sine func on is a nice
func on. Then again, perhaps this is not en rely surprising. The sine func on
is periodic it repeats itself on regular intervals. Therefore its rate of change
also repeats itself on the same regular intervals. We should have known the
deriva ve would be periodic; we now know exactly which periodic func on it is.
Thinking back to Example , we can nd the slope of the tangent line to
f(x) = sin x at x = using our deriva ve. We approximated the slope as .
;
we now know the slope is exactly cos = .

y
1

Finding the deriva ve of a piecewise dened func on


Example
Find the deriva ve of the absolute value func on,
{
x x <
f(x) = |x| =
.
x x

0.5

0.5

0.5

Figure . : The absolute value func on,


f(x) = |x|. No ce how the slope of
the lines (and hence the tangent lines)
abruptly changes at x = .

See Figure . .

f(x + h) f(x)
. As f is piecewise
h
dened, we need to consider separately the limits when x < and when x > .
S

Notes:

We need to evaluate lim


h

Instantaneous Rates of Change: The Deriva ve

When x < :
)
d(
(x + h) (x)
x = lim
h
dx
h
h
= lim
h h
= lim
h

= .
When x > , a similar computa on shows that

d( )
x = .
dx

We need to also nd the deriva ve at x = . By the deni on of the derivave at a point, we have
f ( ) = lim
h

f( + h) f( )
.
h

Since x = is the point where our func ons deni on switches from one piece
to other, we need to consider le and right-hand limits. Consider the following,
where we compute the le and right hand limits side by side.
lim

f( + h) f( )
=
h
h
lim
=

h
h
lim =
h

lim

f( + h) f( )
=
h
h
lim
=
+
h
h
lim =
h

At x = , f (x) does not exist; there is a jump discon nuity at ; see Figure . .
So f(x) = |x| is dieren able everywhere except at .

The point of non-dieren ability came where the piecewise dened funcon switched from one piece to the other. Our next example shows that this
does not always cause trouble.
Finding the deriva ve
{of a piecewise dened func on
sin x x /
Find the deriva ve of f(x), where f(x) =
. See Figure . .
x > /

Notes:

The last lines of each column tell the story: the le and right hand limits are not
equal. Therefore the limit does not exist at , and f is not dieren able at . So
we have
{

x<
f (x) =
.
x>

Example

0.5

0.5

.
Figure . : A graph of the deriva ve of
f(x) = |x|.

Chapter

Deriva ves
Using Example , we know that when x < / , f (x) =
cos x. It is easy to verify that when x > / , f (x) = ; consider:
S

lim

f(x + h) f(x)
= lim
x
h

So far we have

.5

f (x) =

.
Figure . : A graph of f(x) as dened in
Example .

= lim

cos x x < /
x > /

= .

We s ll need to nd f (/ ). No ce at x = / that both pieces of f are ,


meaning we can state that f (/ ) = .
Being more rigorous, we can again evaluate the dierence quo ent limit at
x = / , u lizing again le and righthand limits:
f(/ + h) f(/ )
h
h
sin(/ + h) sin(/ )
lim
h
h
sin( ) cos(h) + sin(h) cos( ) sin( )
lim
h
h
cos(h) + sin(h)
lim
h
h
lim

=
=
=

lim

f(/ + h) f(/ )
=
h

lim
=
h
h +
lim

h +

.5

.
Figure . : A graph of f (x) in Example

Since both the le and right hand limits are at x = / , the limit exists and
f (/ ) exists (and is ). Therefore we can fully write f as
{
cos x x /
f (x) =
.
x > /
See Figure . for a graph of this func on.
Recall we pseudodened a con nuous func on as one in which we could
sketch its graph without li ing our pencil. We can give a pseudodeni on for
dieren ability as well: it is a con nuous func on that does not have any sharp
corners. One such sharp corner is shown in Figure . . Even though the funcon f in Example is piecewisedened, the transi on is smooth hence it is
dieren able. Note how in the graph of f in Figure . it is dicult to tell when
f switches from one piece to the other; there is no corner.
This sec on dened the deriva ve; in some sense, it answers the ques on of
What is the deriva ve? The next sec on addresses the ques on What does
the deriva ve mean?

Notes:

Exercises .

Terms and Concepts


. T/F: Let f be a posi on func on. The average rate of change
on [a, b] is the slope of the line through the points (a, f(a))
and (b, f(b)).

. r(x) =

, at x = .

In Exercises
, a func on f and an xvalue a are given.
Approximate the equa on of the tangent line to the graph of
f at x = a by numerically approxima ng f (a), using h = . .

. T/F: The deni on of the deriva ve of a func on at a point


involves taking a limit.

. f(x) = x + x + , x =

. In your own words, explain the dierence between the average rate of change and instantaneous rate of change.

. f(x) =

. In your own words, explain the dierence between Denions and .

. f(x) = ex , x =

. Let y = f(x). Give three dierent nota ons equivalent to


f (x).

. f(x) = cos x, x =

,x=

. The graph of f(x) = x is shown.

Problems

(a) Use the graph to approximate the slope of the tangent line to f at the following points: ( , ), ( , )
and ( , ).

In Exercises , use the deni on of the deriva ve to compute the deriva ve of the given func on.

(b) Using the deni on, nd f (x).

. f(x) =

(c) Find the slope of the tangent line at the points


( , ), ( , ) and ( , ).

. f(x) = x
. f(t) =

x+

. g(x) = x
. f(x) = x x +
. r(x) =
. r(s) =

In Exercises
, a func on and an xvalue c are given.
(Note: these func ons are the same as those given in Exercises through .)
(a) Find the tangent line to the graph of the func on at c.
(b) Find the normal line to the graph of the func on at c.

. The graph of f(x) =

x+

is shown.

(a) Use the graph to approximate the slope of the tangent line to f at the following points: ( , ) and
( , . ).
(b) Using the deni on, nd f (x).
(c) Find the slope of the tangent line at the points ( , )
and ( , . ).

. f(x) = , at x = .
. f(x) = x, at x = .
. f(x) =

x, at x = .

. g(x) = x , at x = .
. f(x) = x x + , at x = .
. r(x) =

, at x = .

In Exercises , a graph of a func on f(x) is given. Using


the graph, sketch f (x).

Review
. Approximate lim

x + x
x . x+

. Use the Bisec on Method to approximate, accurate to two


decimal places, the root of g(x) = x + x + x on
[ . , . ].

. Give intervals on which each of the following func ons are


con nuous.

(a)

(b)

.
6

ex

(a)

(b)
y

(d) Where is f con nuous?

x +

.
y

.5

.5

. Using the graph of g(x) below, answer the following quesons.


(a) Where is g(x) > ?

(c) Where is g (x) < ?

(b) Where is g(x) < ?

(d) Where is g (x) > ?

(c) Where is g(x) = ?

(e) Where is g (x) = ?

y
5

lim f(x) =?

(c) lim f(x) =?

(d)

lim f(x) =?

. Use the graph of f(x) provided to answer the following.

(c)

Interpreta ons of the Deriva ve

The previous sec on dened the deriva ve of a func on and gave examples of
how to compute it using its deni on (i.e., using limits). The sec on also started
with a brief mo va on for this deni on, that is, nding the instantaneous velocity of a falling object given its posi on func on. The next sec on will give us
more accessible tools for compu ng the deriva ve, tools that are easier to use
than repeated use of limits.
This sec on falls in between the What is the deni on of the deriva ve?
and How do I compute the deriva ve? sec ons. Here we are concerned with
What does the deriva ve mean?, or perhaps, when read with the right emphasis, What is the deriva ve? We oer two interconnected interpreta ons
of the deriva ve, hopefully explaining why we care about it and why it is worthy
of study.

Interpreta on of the Deriva ve # : Instantaneous Rate of Change


The previous sec on started with an example of using the posi on of an
object (in this case, a falling amusementpark rider) to nd the objects velocity. This type of example is o en used when introducing the deriva ve because
we tend to readily recognize that velocity is the instantaneous rate of change
of posi on. In general, if f is a func on of x, then f (x) measures the instantaneous rate of change of f with respect to x. Put another way, the derivave answers When x changes, at what rate does f change? Thinking back to
the amusementpark ride, we asked When me changed, at what rate did the
height change? and found the answer to be By feet per second.
Now imagine driving a car and looking at the speedometer, which reads
mph. Five minutes later, you wonder how far you have traveled. Certainly, lots
of things could have happened in those minutes; you could have inten onally
sped up signicantly, you might have come to a complete stop, you might have
slowed to
mph as you passed through construc on. But suppose that you
know, as the driver, none of these things happened. You know you maintained
a fairly consistent speed over those minutes. What is a good approxima on of
the distance traveled?
One could argue the only good approxima on, given the informa on provided, would be based on distance = rate me. In this case, we assume a
constant rate of
mph with a me of / hours. Hence we would approximate the distance traveled as miles.
Referring back to the falling amusementpark ride, knowing that at t = the
velocity was
/s, we could reasonably assume that second later the rid-

Notes:

Interpreta ons of the Deriva ve

Chapter

Deriva ves
ers height would have dropped by about feet. Knowing that the riders were
accelera ng as they fell would inform us that this is an underapproxima on. If
all we knew was that f( ) =
and f ( ) = , wed know that wed have to
stop the riders quickly otherwise they would hit the ground!
Units of the Deriva ve
It is useful to recognize the units of the deriva ve func on. If y is a func on
of x, i.e., y = f(x) for some func on f, and y is measured in feet and x in seconds,
then the units of y = f are feet per second, commonly wri en as /s. In
general, if y is measured in units P and x is measured in units Q, then y will be
measured in units P per Q, or P/Q. Here we see the frac onlike behavior
of the deriva ve in the nota on:
the units of

dy
dx

are

units of y
.
units of x

The meaning of the deriva ve: World Popula on


Example
Let P(t) represent the world popula on t minutes a er : a.m., January ,
. It is fairly accurate to say that P( ) = ,
,
,
(www.prb.org). It
is also fairly accurate to state that P ( ) =
; that is, at midnight on January ,
, the popula on of the world was growing by about
people per minute
(note the units). Twenty days later (or, ,
minutes later) we could reasonably assume the popula on grew by about ,

= ,
,
people.
The meaning of the deriva ve: Manufacturing

Example

The term widget is an economic term for a generic unit of manufacturing


output. Suppose a company produces widgets and knows that the market supports a price of $ per widget. Let P(n) give the prot, in dollars, earned by
manufacturing and selling n widgets. The company likely cannot make a (posi ve) prot making just one widget; the startup costs will likely exceed $ .
Mathema cally, we would write this as P( ) < .
What do P(
)=
and P (
) = . mean? Approximate P(
).
The equa on P(
)=
means that selling ,
widS
gets returns a prot of $
. We interpret P (
) = . as meaning that
the prot is increasing at rate of $ . per widget (the units are dollars per
widget.) Since we have no other informa on to use, our best approxima on
for P(
) is:
P(

) P(

) + P (

We approximate that selling ,

Notes:

=$

widgets returns a prot of $

=$
.

.
The previous examples made use of an important approxima on tool that
we rst used in our previous driving a car at
mph example at the beginning of this sec on. Five minutes a er looking at the speedometer, our best
approxima on for distance traveled assumed the rate of change was constant.
In Examples and we made similar approxima ons. We were given rate of
change informa on which we used to approximate total change. Nota onally,
we would say that
f(c + h) f(c) + f (c) h.
This approxima on is best when h is small. Small is a rela ve term; when
dealing with the world popula on, h =
days = ,
minutes is small in
comparison to years. When manufacturing widgets,
widgets is small when
one plans to manufacture thousands.

The Deriva ve and Mo on


One of the most fundamental applica ons of the deriva ve is the study of
mo on. Let s(t) be a posi on func on, where t is me and s(t) is distance. For
instance, s could measure the height of a projec le or the distance an object has
traveled.
Lets let s(t) measure the distance traveled, in feet, of an object a er t seconds of travel. Then s (t) has units feet per second, and s (t) measures the
instantaneous rate of distance change it measures velocity.
Now consider v(t), a velocity func on. That is, at me t, v(t) gives the velocity of an object. The deriva ve of v, v (t), gives the instantaneous rate of
velocity change accelera on. (We o en think of accelera on in terms of cars:
a car may go from to
in . seconds. This is an average accelera on, a
measurement of how quickly the velocity changed.) If velocity is measured in
feet per second, and me is measured in seconds, then the units of accelera on
(i.e., the units of v (t)) are feet per second per second, or ( /s)/s. We o en
shorten this to feet per second squared, or /s , but this tends to obscure the
meaning of the units.
Perhaps the most well known accelera on is that of gravity. In this text, we
use g =
/s or g = . m/s . What do these numbers mean?
A constant accelera on of ( /s)/s means that the velocity changes by
/s each second. For instance, let v(t) measures the velocity of a ball thrown
straight up into the air, where v has units /s and t is measured in seconds. The
ball will have a posi ve velocity while traveling upwards and a nega ve velocity
while falling down. The accelera on is thus
/s . If v( ) =
/s, then
when t = , the velocity will have decreased by
/s; that is, v( ) =
/s.
We can con nue: v( ) =
/s, and we can also gure that v( ) =
/s.
These ideas are so important we write them out as a Key Idea.

Notes:

Interpreta ons of the Deriva ve

Chapter

Deriva ves

Key Idea

The Deriva ve and Mo on

. Let s(t) be the posi on func on of an object. Then s (t) is the


velocity func on of the object.
. Let v(t) be the velocity func on of an object. Then v (t) is the
accelera on func on of the object.

We now consider the second interpreta on of the deriva ve given in this


sec on. This interpreta on is not independent from the rst by any means;
many of the same concepts will be stressed, just from a slightly dierent perspec ve.

y
6

Interpreta on of the Deriva ve # : The Slope of the Tangent Line


8

.
Figure .

: A graph of f(x) = x .

f(c + h) f(c)
Given a func on y = f(x), the dierence quo ent
gives a
h
change in y values divided by a change in x values; i.e., it is a measure of the
rise(over run,
) or(slope, of the) line that goes through two points on the graph
of f: c, f(c) and c+h, f(c+h) . As h shrinks to , these two points come close
together; in the limit we nd f (c), the slope of a special line called the tangent
line that intersects f only once near x = c.
Lines have a constant rate of change, their slope. Nonlinear func ons do not
have a constant rate of change, but we can measure their instantaneous rate of
change at a given x value c by compu ng f (c). We can get an idea of how f is
behaving by looking at the slopes of its tangent lines. We explore this idea in the
following example.

Understanding the deriva ve: the rate of change


Example
Consider f(x) = x as shown in Figure . . It is clear that at x = the func on
is growing faster than at x = , as it is steeper at x = . How much faster is it
growing?

..

Figure . : A graph of f(x) = x and tangent lines.

We can answer this directly a er the following sec on, where


S
we learn to quickly compute deriva ves. For now, we will answer graphically,
by considering the slopes of the respec ve tangent lines.
With prac ce, one can fairly eec vely sketch tangent lines to a curve at a
par cular point. In Figure . , we have sketched the tangent lines to f at x =
and x = , along with a grid to help us measure the slopes of these lines. At

Notes:

Interpreta ons of the Deriva ve

x = , the slope is ; at x = , the slope is . Thus we can say not only is f


growing faster at x = than at x = , it is growing three mes as fast.

y
f(x)

Understanding the graph of the deriva ve


Example
Consider the graph of f(x) and its deriva ve, f (x), in Figure . (a). Use these
graphs to nd the slopes of the tangent lines to the graph of f at x = , x = ,
and x = .
To nd the appropriate slopes of tangent lines to the graph
of f, we need to look at the corresponding values of f .
The slope of the tangent line to f at x = is f ( ); this looks to be about .
The slope of the tangent line to f at x = is f ( ); this looks to be about .
The slope of the tangent line to f at x = is f ( ); this looks to be about .
Using these slopes, the tangent lines to f are sketched in Figure . (b). Included on the graph of f in this gure are lled circles where x = , x = and
x = to help be er visualize the y value of f at those points.
S

Approxima on with the deriva ve


Example
Consider again the graph of f(x) and its deriva ve f (x) in Example
tangent line to f at x = to approximate the value of f( . ).

=. +

(a)

y
f(x)

5
f (x)

. Use the

Figure . shows the graph of f along with its tangent line,


S
zoomed in at x = . No ce that near x = , the tangent line makes an excellent
approxima on of f. Since lines are easy to deal with, o en it works well to approximate a func on with its tangent line. (This is especially true when you dont
actually know much about the func on at hand, as we dont in this example.)
While the tangent line to f was drawn in Example , it was not explicitly
computed. Recall that the tangent line to f at x = c is y = f (c)(x c) + f(c).
While f is not explicitly given, by the graph it looks like f( ) = . Recalling that
f ( ) = , we can compute the tangent line to be approximately y = (x )+ .
It is o en useful to leave the tangent line in pointslope form.
To use the tangent line to approximate f( . ), we simply evaluate y at .
instead of f.
f( . ) y( . ) = ( . ) +

f (x)

(b)

Figure . : Graphs of f and f in Example


, along with tangent lines in (b).

= . .

We approximate f( . ) . .
To demonstrate the accuracy of the tangent line approxima on, we now
state that in Example , f(x) = x + x x + . We can evaluate f( . ) =
.
. Had we known f all along, certainly we could have just made this computa on. In reality, we o en only know two things:

Notes:

x
.8

Figure . : Zooming in on f at x = for


the func on given in Examples and .

Chapter

Deriva ves
. What f(c) is, for some value of c, and
. what f (c) is.
For instance, we can easily observe the loca on of an object and its instantaneous velocity at a par cular point in me. We do not have a func on f
for the loca on, just an observa on. This is enough to create an approxima ng
func on for f.
This last example has a direct connec on to our approxima on method explained above a er Example . We stated there that
f(c + h) f(c) + f (c) h.
If we know f(c) and f (c) for some value x = c, then compu ng the tangent
line at (c, f(c)) is easy: y(x) = f (c)(x c) + f(c). In Example , we used the
tangent line to approximate a value of f. Lets use the tangent line at x = c
to approximate a value of f near x = c; i.e., compute y(c + h) to approximate
f(c + h), assuming again that h is small. Note:
(
)
y(c + h) = f (c) (c + h) c + f(c) = f (c) h + f(c).
This is the exact same approxima on method used above! Not only does it make
intui ve sense, as explained above, it makes analy cal sense, as this approximaon method is simply using a tangent line to approximate a func ons value.

The importance of understanding the deriva ve cannot be understated. When


f is a func on of x, f (x) measures the instantaneous rate of change of f with respect to x and gives the slope of the tangent line to f at x.

Notes:

Exercises .

Terms and Concepts

In Exercises
, graphs of func ons f(x) and g(x) are
given. Iden fy which func on is the deriva ve of the other.)

. What is the instantaneous rate of change of posi on


called?
. Given a func on y = f(x), in your own words describe how
to nd the units of f (x).

y
g(x)

f(x)

.
4

. What func ons have a constant rate of change?

2
4

Problems

y
4

. Given f( ) =
. Given P(
P(
).
. Given z(

and f ( ) = , approximate f( ).

) =

and P (

) =

g(x)

f(x)

, approximate

2
2

)=

and z (

)=

, approximate z(

).

. Knowing f( ) =
and f ( ) = and the methods described in this sec on, which approxima on is likely to be
most accurate: f( . ), f( ), or f( )? Explain your reasoning.
. Given f( ) =
. Given H( ) =

and f( ) =
and H( ) =

y
5

f(x)

. .

, approximate f ( ).

g(x)

, approximate H ( ).

. Let V(x) measure the volume, in decibels, measured inside


a restaurant with x customers. What are the units of V (x)?
. Let v(t) measure the velocity, in /s, of a car moving in a
straight line t seconds a er star ng. What are the units of
v (t)?
. The height H, in feet, of a river is recorded t hours a er midnight, April . What are the units of H (t)?
. P is the prot, in thousands of dollars, of producing and selling c cars.
(a) What are the units of P (c)?
(b) What is likely true of P( )?
. T is the temperature in degrees Fahrenheit, h hours a er
midnight on July in Sidney, NE.
(a) What are the units of T (h)?
(b) Is T ( ) likely greater than or less than ? Why?
(c) Is T( ) likely greater than or less than ? Why?

y
1

f(x)

.
4

g(x)

Review
In Exercises , use the deni on to compute the derivaves of the following func ons.
. f(x) = x
. f(x) = (x )
In Exercises
, numerically approximate the value of
f (x) at the indicated x value.
. f(x) = cos x at x = .
. f(x) =

x at x = .

Chapter

Deriva ves

Basic Dieren a on Rules

The deriva ve is a powerful tool but is admi edly awkward given its reliance on
limits. Fortunately, one thing mathema cians are good at is abstrac on. For
instance, instead of con nually nding deriva ves at a point, we abstracted and
found the deriva ve func on.
Lets prac ce abstrac on on linear func ons, y = mx + b. What is y ? Without limits, recognize that linear func on are characterized by being func ons
with a constant rate of change (the slope). The deriva ve, y , gives the instantaneous rate of change; with a linear func on, this is constant, m. Thus y = m.
Lets abstract once more. Lets nd the deriva ve of the general quadra c
func on, f(x) = ax + bx + c. Using the deni on of the deriva ve, we have:
a(x + h) + b(x + h) + c (ax + bx + c)
h
ah + ahx + bh
= lim
h
h
= lim ah + ax + b

f (x) = lim
h

= ax + b.
So if y = x +

, we can immediately compute y =

x+

In this sec on (and in some sec ons to follow) we will learn some of what
mathema cians have already discovered about the deriva ves of certain funcons and how deriva ves interact with arithme c opera ons. We start with a
theorem.
Theorem

Deriva ves of Common Func ons

. Constant Rule:
d( )
c = , where c is a constant.
dx

. Power Rule:
d n
(x ) = nxn , where n is an integer, n > .
dx

d
(sin x) = cos x
dx

d
(cos x) = sin x
dx

d x
(e ) = ex
dx

d
(ln x) =
dx
x

This theorem starts by sta ng an intui ve fact: constant func ons have no
rate of change as they are constant. Therefore their deriva ve is (they change

Notes:

Basic Dieren a on Rules

at the rate of ). The theorem then states some fairly amazing things. The Power
Rule states that the deriva ves of Power Func ons (of the form y = xn ) are very
straigh orward: mul ply by the power, then subtract from the power. We see
something incredible about the func on y = ex : it is its own deriva ve. We also
see a new connec on between the sine and cosine func ons.
One special case of the Power Rule is when n = , i.e., when f(x) = x. What
is f (x)? According to the Power Rule,
f (x) =

d( )
d( )
x =
x =
dx
dx

x = .

In words, we are asking At what rate does f change with respect to x? Since f
is x, we are asking At what rate does x change with respect to x? The answer
is: . They change at the same rate.
Lets prac ce using this theorem.
Example
Let f(x) = x .

Using Theorem

to nd, and use, deriva ves

. Find f (x).

. Find the equa on of the line tangent to the graph of f at x = .

. Use the tangent line to approximate ( . ) .

. Sketch f, f and the found tangent line on the same axis.


2

. The Power Rule states that if f(x) = x , then f (x) = x .


. To nd the equa on of the line tangent to the graph of f at x = , we
need a point and the slope. The point is ( , f( )) = ( , ). The
slope is f ( ) = . Thus the tangent line has equa on y = (x( ))+
( ) = x + .
. We can use the tangent line to approximate ( . ) as . is close to
. We have
( . ) ( . ) + = . .
We can easily nd the actual answer; ( . ) = .

. See Figure .

Notes:

Figure . : A graph of f(x) = x , along

with
. its deriva ve f (x) = x and its tangent line at x = .

Chapter

Deriva ves
Theorem
gives useful informa on, but we will need much more. For instance, using the theorem, we can easily nd the deriva ve of y = x , but it does
not tell how to compute the deriva ve of y = x , y = x +sin x nor y = x sin x.
The following theorem helps with the rst two of these examples (the third is
answered in the next sec on).
Theorem

Proper es of the Deriva ve

Let f and g be dieren able on an open interval I and let c be a real


number. Then:
. Sum/Dierence Rule:
)
)
)
d(
d(
d(
f(x) g(x) =
f(x)
g(x) = f (x) g (x)
dx
dx
dx
. Constant Mul ple Rule:
)
)
d(
d(
c f(x) = c
f(x) = c f (x).
dx
dx

Theorem
allows us to nd the deriva ves of a wide variety of func ons.
It can be used in conjunc on with the Power Rule to nd the deriva ves of any
polynomial. Recall in Example
that we found, using the limit deni on, the
deriva ve of f(x) = x + x . We can now nd its deriva ve without expressly
using limits:
)
d( )
d( )
d( )
d(
x + x+
=
x +
x +
dx
dx
dx
dx
= x+ +
= x+ .
We were a bit pedan c here, showing every step. Normally we would do all
)
d(
the arithme c and steps in our head and readily nd
x + x+ = x+ .
dx

Using the tangent line to approximate a func on value


Example
Let f(x) = sin x + x + . Approximate f( ) using an appropriate tangent line.

This problem is inten onally ambiguous; we are to approxiS


mate using an appropriate tangent line. How good of an approxima on are we
seeking? What does appropriate mean?
In the real world, people solving problems deal with these issues all me.
One must make a judgment using whatever seems reasonable. In this example,
the actual answer is f( ) = sin + , where the real problem spot is sin . What
is sin ?

Notes:

Basic Dieren a on Rules

Since is close to , we can assume sin sin = . Thus one guess is


f( ) . Can we do be er? Lets use a tangent line as instructed and examine
the results; it seems best to nd the tangent line at x = .
Using Theorem we nd f (x) = cos x + . The slope of the tangent line is
thus f () = cos + = . Also, f() = + . . So the tangent line to
the graph of f at x = is y = (x)+ + = x++ x+ . . Evaluated
at x = , our tangent line gives y = + . = . . Using the tangent line,
our nal approxima on is that f( ) . .
Using a calculator, we get an answer accurate to places a er the decimal:
f( ) = .
. Our ini al guess was ; our tangent line approxima on was more
accurate, at . .
The point is not Heres a cool way to do some math without a calculator.
Sure, that might be handy some me, but your phone could probably give you
the answer. Rather, the point is to say that tangent lines are a good way of
approxima ng, and many scien sts, engineers and mathema cians o en face
problems too hard to solve directly. So they approximate.

Higher Order Deriva ves


The deriva ve of a func on f is itself a func on, therefore we can take its
deriva ve. The following deni on gives a name to this concept and introduces
its nota on.
Deni on

Higher Order Deriva ves

Let y = f(x) be a dieren able func on on I.


. The second deriva ve of f is:
f (x) =

d( )
d
f (x) =
dx
dx

dy
dx

d y
= y .
dx

. The third deriva ve of f is:


d
d ( )
f (x) =
f (x) =
dx
dx

d y
dx

d y
= y .
dx

. The nth deriva ve of f is:


f

Notes:

(n)

d ( (n ) )
d
(x) =
f
(x) =
dx
dx

d n y
dxn

d ny
= y(n) .
dxn

Note: Deni on
comes with the
caveat Where the corresponding limits
exist. With f dieren able on I, it is possible that f is not dieren able on all of
I, and so on.

Chapter

Deriva ves
In general, when nding the fourth deriva ve and on, we resort to the f ( ) (x)
nota on, not f (x); a er a while, too many cks is too confusing.
Lets prac ce using this new concept.
Finding higher order deriva ves
Example
Find the rst four deriva ves of the following func ons:
. f(x) = ex

. f(x) = x
. f(x) = sin x
S

. Using the Power and Constant Mul ple Rules, we have: f (x) = x. Connuing on, we have
f (x) =

d( )
x = ;
dx

f (x) = ;

f ( ) (x) = .

No ce how all successive deriva ves will also be .


. We employ Theorem
f (x) = cos x;

repeatedly.

f (x) = sin x;

f (x) = cos x;

f ( ) (x) = sin x.

Note how we have come right back to f(x) again. (Can you quickly gure
what f ( ) (x) is?)
. Employing Theorem

and the Constant Mul ple Rule, we can see that

f (x) = f (x) = f (x) = f ( ) (x) = ex .

Interpre ng Higher Order Deriva ves


What do higher order deriva ves mean? What is the prac cal interpretaon?
Our rst answer is a bit wordy, but is technically correct and benecial to
understand. That is,
The second deriva ve of a func on f is the rate of change of the rate
of change of f.

Notes:

.
One way to grasp this concept is to let f describe a posi on func on. Then, as
stated in Key Idea , f describes the rate of posi on change: velocity. We now
consider f , which describes the rate of velocity change. Sports car enthusiasts
talk of how fast a car can go from to
mph; they are bragging about the
accelera on of the car.
We started this chapter with amusementpark riders freefalling with posion func on f(t) = t +
. It is easy to compute f (t) = t /s and

f (t) = ( /s)/s. We may recognize this la er constant; it is the acceleraon due to gravity. In keeping with the unit nota on introduced in the previous
sec on, we say the units are feet per second per second. This is usually shortened to feet per second squared, wri en as /s .
It can be dicult to consider the meaning of the third, and higher order,
deriva ves. The third deriva ve is the rate of change of the rate of change of
the rate of change of f. That is essen ally meaningless to the unini ated. In
the context of our posi on/velocity/accelera on example, the third deriva ve
is the rate of change of accelera on, commonly referred to as jerk.
Make no mistake: higher order deriva ves have great importance even if
their prac cal interpreta ons are hard (or impossible) to understand. The
mathema cal topic of series makes extensive use of higher order deriva ves.

Notes:

Basic Dieren a on Rules

Exercises .

Terms and Concepts

. f(x) = ln( x )

. What is the name of the rule which states that


nxn , where n >
. What is

is an integer?

d ( n)
x =
dx

)
d(
ln x ?
dx

. f(t) = ln(

) + e + sin /

. g(t) = ( + t)
. g(x) = ( x )

. Give an example of a func on f(x) where f (x) = f(x).

. f(x) = ( x)

. Give an example of a func on f(x) where f (x) = .

. f(x) = ( x)

. The deriva ve rules introduced in this sec on explain how


to compute the deriva ve of which of the following funcons?

. A property of logarithms is that loga x =

f(x) =

j(x) = sin x cos x

g(x) = x x +
h(x) = ln x

k(x) = ex

m(x) = x

. Explain in your own words how to nd the third deriva ve


of a func on f(x).
. Give an example of a func on where f (x) = and f (x) =
.
. Explain in your own words what the second deriva ve
means.
. If f(x) describes a posi on func on, then f (x) describes
what kind of func on? What kind of func on is f (x)?
. Let f(x) be a func on measured in pounds, where x is measured in feet. What are the units of f (x)?

(a) Rewrite this iden ty when b = e, i.e., using loge x =


ln x.
(b) Use part (a) to nd the deriva ve of y = loga x.
(c) Give the deriva ve of y = log x.

In Exercises
given func on.

, compute the rst four deriva ves of the

. f(x) = x
. g(x) = cos x
. h(t) = t et
. p() =
. f() = sin cos
. f(x) = ,
In Exercises
, nd the equa ons of the tangent and
normal lines to the graph of the func on at the given point.

Problems
In Exercises
on.

bases a, b > , = .

logb x
, for all
logb a

, compute the deriva ve of the given func-

. f(x) = x x at x =
. f(t) = et + at t =

. f(x) = x x +
. g(x) =

. g(x) = ln x at x =

x + x +

. m(t) = t t + t
. f() = sin +

cos

. f(r) = er
. g(t) =

t cos t + sin t

. f(x) = ln x x
. p(s) = s + s + s + s +
t

. h(t) = e sin t cos t

. f(x) = sin x at x = /
. f(x) = cos x at x = /
. f(x) = x + at x =

Review
. Given that e = , approximate the value of e
tangent line to f(x) = ex at x = .
. Approximate the value of ( .
f(x) = x at x = .

using the

) using the tangent line to

The Product and Quo ent Rules

The Product and Quo ent Rules

The previous sec on showed that, in some ways, deriva ves behave nicely. The
Constant Mul ple and Sum/Dierence Rules established that the deriva ve of
f(x) = x + sin x was not complicated. We neglected compu ng the deriva ve
of things like g(x) = x sin x and h(x) = sinx x on purpose; their deriva ves are
not as straigh orward. (If you had to guess what their respec ve deriva ves are,
you would probably guess wrong.) For these, we need the Product and Quo ent
Rules, respec vely, which are dened in this sec on.
We begin with the Product Rule.
Theorem

Product Rule

Let f and g be dieren able func ons on an open interval I. Then fg is a


dieren able func on on I, and
)
d(
f(x)g(x) = f(x)g (x) + f (x)g(x).
dx
)
d(
Important:
f(x)g(x) = f (x)g (x)! While this answer is simpler than
dx
the Product Rule, it is wrong.
We prac ce using this new rule in an example, followed by an example that
demonstrates why this theorem is true.
Using the Product Rule
Example
Use the Product Rule to compute the deriva ve of y = x sin x. Evaluate the
deriva ve at x = / .
To make our use of the Product Rule explicit, lets set f(x) =
x and g(x) = sin x. We easily compute/recall that f (x) = x and g (x) =
cos x. Employing the rule, we have
)
d(
x sin x = x cos x + x sin x.
dx
At x = / , we have
()
()
()

y (/ ) =
cos
sin
+
= .
S

We graph y and its tangent line at x = / , which has a slope of , in Figure


. . While this does not prove that the Produce Rule is the correct way to handle deriva ves of products, it helps validate its truth.

Notes:

Figure . : A graph of y = x sin x and


its tangent line at x = / .

Chapter

Deriva ves
We now inves gate why the Product Rule is true.
A proof of the Product Rule
Example
Use the deni on of the deriva ve to prove Theorem
S

By the limit deni on, we have


)
d(
f(x + h)g(x + h) f(x)g(x)
.
f(x)g(x) = lim
h
dx
h

We now do something a bit unexpected; add to the numerator (so that nothing
is changed) in the form of f(x+h)g(x)+f(x+h)g(x), then do some regrouping
as shown.
)
f(x + h)g(x + h) f(x)g(x)
d(
f(x)g(x) = lim
(now add to the numerator)
h
dx
h
f(x + h)g(x + h) f(x + h)g(x) + f(x + h)g(x) f(x)g(x)
= lim
(regroup)
h
h
(
) (
)
f(x + h)g(x + h) f(x + h)g(x) + f(x + h)g(x) f(x)g(x)
= lim
h
h
f(x + h)g(x) f(x)g(x)
f(x + h)g(x + h) f(x + h)g(x)
+ lim
= lim
(factor)
h
h
h
h
g(x + h) g(x)
f(x + h) f(x)
= lim f(x + h)
+ lim
g(x) (apply limits)
h
h
h
h
= f(x)g (x) + f (x)g(x)

It is o en true that we can recognize that a theorem is true through its proof
yet somehow doubt its applicability to real problems. In the following example,
we compute the deriva ve of a product of func ons in two ways to verify that
the Product Rule is indeed right.
Exploring alternate deriva ve methods
Example
Let y = (x + x + )( x x + ). Find y two ways: rst, by expanding
the given product and then taking the deriva ve, and second, by applying the
Product Rule. Verify that both methods give the same answer.
We rst expand the expression for y; a li le algebra shows
S
that y = x + x x + . It is easy to compute y ;
y = x + x

Notes:

x.

.
Now apply the Product Rule.
y = (x + x + )( x ) + ( x + )( x x + )
(
) (
)
= x + x x + x x+
= x + x

x.

The uninformed usually assume that the deriva ve of the product is the
product of the deriva ves. Thus we are tempted to say that y = ( x + )( x
) = x + x . Obviously this is not correct.
Using the Product Rule with a product of three func ons
Example
Let y = x ln x cos x. Find y .
We have a product of three func ons while the Product Rule
S
only species how to handle a product of two func ons. Our method of handling
this problem
is to) simply group the la er two func ons together, and consider
(
y = x ln x cos x . Following the Product Rule, we have
(
)
(
)
y = (x ) ln x cos x + x ln x cos x

(
)
To evaluate ln x cos x , we apply the Product Rule again:
(
= (x ) ln x( sin x) +
= x ln x( sin x) + x

(
)
)
cos x + x ln x cos x

cos x + x ln x cos x

Recognize the pa ern in our answer above: when applying the Product Rule to
a product of three func ons, there are three terms added together in the nal
deriva ve. Each terms contains only one deriva ve of one of the original funcons, and each func ons deriva ve shows up in only one term. It is straigh orward to extend this pa ern to nding the deriva ve of a product of or more
func ons.
We consider one more example before discussing another deriva ve rule.
Using the Product Rule
Example
Find the deriva ves of the following func ons.
. f(x) = x ln x
. g(x) = x ln x x.

Notes:

The Product and Quo ent Rules

Chapter

Deriva ves
Recalling that the deriva ve of ln x is /x, we use the Product
Rule to nd our answers.
)
d(
.
x ln x = x /x + ln x = + ln x.
dx
. Using the result from above, we compute
)
d(
x ln x x = + ln x = ln x.
dx

This seems signicant; if the natural log func on ln x is an important func on (it
is), it seems worthwhile to know a func on whose deriva ve is ln x. We have
found one. (We leave it to the reader to nd another; a correct answer will be
very similar to this one.)
We have learned how to compute the deriva ves of sums, dierences, and
products of func ons. We now learn how to nd the deriva ve of a quo ent of
func ons.

Theorem

Quo ent Rule

Let f and g be func ons dened on an open interval I, where g(x) =


on I. Then f/g is dieren able on I, and
(
)
g(x)f (x) f(x)g (x)
d f(x)
=
.
dx g(x)
g(x)
The Quo ent Rule is not hard to use, although it might be a bit tricky to remember. A useful mnemonic works as follows. Consider a frac ons numerator
and denominator as HI and LO, respec vely. Then
( )
d HI
LO dHI HI dLO
=
,
dx LO
LOLO
read low dee high minus high dee low, over low low. Said fast, that phrase can
roll o the tongue, making it easy to memorize. The dee high and dee low
parts refer to the deriva ves of the numerator and denominator, respec vely.
Lets prac ce using the Quo ent Rule.
Using the Quo ent Rule
x
Let f(x) =
. Find f (x).
sin x
Example

Notes:

.
S

The Product and Quo ent Rules

Directly applying the Quo ent Rule gives:


d
dx

x
sin x

=
=

sin x

x x cos x
sin x
x sin x x cos x
.
sin x

The Quo ent Rule allows us to ll in holes in our understanding of deriva ves
of the common trigonometric func ons. We start with nding the deriva ve of
the tangent func on.
Using the Quo ent Rule to nd
Example
Find the deriva ve of y = tan x.

d
dx

)
tan x .

At rst, one might feel unequipped to answer this ques on.


S
But recall that tan x = sin x/ cos x, so we can apply the Quo ent Rule.
y

(
)
)
d sin x
d(
tan x =
dx
dx cos x
cos x cos x sin x( sin x)
=
cos x
cos x + sin x
=
cos x

5
x

cos x
= sec x.
This is beau ful result. To conrm its truth, we can nd the equa on of the tangent line to y = tan x at x = / . The slope is sec (/ ) = ; y = tan x, along
with its tangent line, is graphed in Figure . .
We include this result in the following theorem about the deriva ves of the
trigonometric func ons. Recall we found the deriva ve of y = sin x in Example
and stated the deriva ve of the cosine func on in Theorem . The derivaves of the cotangent, cosecant and secant func ons can all be computed directly using Theorem and the Quo ent Rule.

Notes:

.
.
Figure
. : A graph of y = tan x along

with its tangent line at x = / .

Chapter

Deriva ves

Theorem

Deriva ves of Trigonometric Func ons

)
d(
sin x = cos x
dx
)
d(
tan x = sec x
.
dx
)
d(
.
sec x = sec x tan x
dx

)
d(
cos x = sin x
dx
)
d(
cot x = csc x
.
dx
)
d(
.
csc x = csc x cot x
dx
.

To remember the above, it may be helpful to keep in mind that the derivaves of the trigonometric func ons that start with c have a minus sign in them.
Exploring alternate deriva ve methods
x
was found using the Quo ent Rule.
In Example the deriva ve of f(x) =
sin x

Rewri ng f as f(x) = x csc x, nd f using Theorem


and verify the two
answers are the same.
Example

x sin x x cos x
.
sin x

We now nd f using the Product Rule, considering f as f(x) = x csc x.


)
d(
f (x) =
x csc x
dx
= x ( csc x cot x) + x csc x
(now rewrite trig func ons)
cos x
x

+
= x
sin x sin x
sin x
x cos x
x
(get common denominator)
=
+
sin x
sin x
x sin x x cos x
=
sin x
S

We found in Example

that the f (x) =

Finding f using either method returned the same result. At rst, the answers
looked dierent, but some algebra veried they are the same. In general, there
is not one nal form that we seek; the immediate result from the Product Rule
is ne. Work to simplify your results into a form that is most readable and
useful to you.
The Quo ent Rule gives other useful results, as show in the next example.

Notes:

.
Using the Quo ent Rule to expand the Power Rule
Example
Find the deriva ves of the following func ons.
. f(x) =
. f(x) =

x
xn

, where n >

is an integer.

We employ the Quo ent Rule.

. f (x) =

x
x

. f (x) =

xn nxn
nxn
n
=

= n+ .
n
n
(x )
x
x

The deriva ve of y = n turned out to be rather nice. It gets be er. Conx


sider:
( )
d
d ( n )
=
(apply result from Example )
x
dx xn
dx
n
= n+
(rewrite algebraically)
x
= nx(n+ )
= nxn .

This is reminiscent of the Power Rule: mul ply by the power, then subtract
from the power. We now add to our previous Power Rule, which had the restric on of n > .
Theorem

Power Rule with Integer Exponents

Let f(x) = xn , where n =

is an integer. Then
f (x) = n xn .

Taking the deriva ve of many func ons is rela vely straigh orward. It is
clear (with prac ce) what rules apply and in what order they should be applied.
Other func ons present mul ple paths; dierent rules may be applied depending on how the func on is treated. One of the beau ful things about calculus
is that there is not the right way; each path, when applied correctly, leads to

Notes:

The Product and Quo ent Rules

Chapter

Deriva ves
the same result, the deriva ve. We demonstrate this concept in an example.
Exploring alternate deriva ve methods
x x+
. Find f (x) in each of the following ways:
Let f(x) =
x

Example

. By applying the Quo ent Rule,


(
)
. by viewing f as f(x) = x x + x and applying the Product and
Power Rules, and
. by simplifying rst through division.
Verify that all three methods give the same result.
S

. Applying the Quo ent Rule gives:


(
) (
)
x x x x+

f (x) =
x

x
x

. By rewri ng f, we can apply the Product and Power Rules as follows:


(
)
(
)
f (x) = x x + ( )x + x x
x
x x+
+
x
x
x x+
x x
=
+
x
x
x
= ,
=
x
x
=

the same result as above.


. As x = , we can divide through by x rst, giving f(x) = x + . Now
x
apply the Power Rule.
f (x) =

the same result as before.


Example demonstrates three methods of nding f . One is hard pressed
to argue for a best method as all three gave the same result without too much
diculty, although it is clear that using the Product Rule required more steps.
Ul mately, the important principle to take away from this is: reduce the answer

Notes:

.
to a form that seems simple and easy to interpret. In that example, we saw
dierent expressions for f , including:

(
)
x x+
x

(
)
(
)
= x x + ( )x + x x .

They are equal; they are all correct; only the rst is clear. Work to make answers clear.
In the next sec on we con nue to learn rules that allow us to more easily
compute deriva ves than using the limit deni on directly. We have to memorize the deriva ves of a certain set of func ons, such as the deriva ve of sin x
is cos x. The Sum/Dierence, Constant Mul ple, Power, Product and Quo ent
Rules show us how to nd the deriva ves of certain combina ons of these funcons. The next sec on shows how to nd the deriva ves when we compose
these func ons together.

Notes:

The Product and Quo ent Rules

Exercises .

Terms and Concepts


. T/F: The Product Rule states that

In Exercises
on.
)
d(
x sin x = x cos x.
dx

. T/F: The Quo ent Rule states that

d
dx

x
sin x

cos x
.
x

, compute the deriva ve of the given func-

. f(x) = x sin x
. f(t) =

(csc t )

. T/F: The deriva ves of the trigonometric func ons that


start with c have minus signs in them.

. g(x) =

x+
x

. What deriva ve rule is used to extend the Power Rule to


include nega ve integer exponents?

. g(t) =

t
cos t t

. T/F: Regardless of the func on, there is always exactly one


right way of compu ng its deriva ve.
. In your own words, explain what it means to make your answers clear.

. h(x) = cot x ex
. h(t) = t + t
. f(x) =

x + x
x+

. f(x) = (

Problems
In Exercises

x +

sin x
cos x +

. f(x) =

(b) Manipulate the func on algebraically and dieren ate without the Product Rule.

. g(x) = e

(c) Show that the answers from (a) and (b) are equivalent.
. f(x) = x(x + x)

. h(s) = ( s )(s + )
. f(x) = (x + )( x )
:

(a) Use the Quo ent Rule to dieren ate the func on.
(b) Manipulate the func on algebraically and dieren ate without the Quo ent Rule.

. g(x) =
. h(s) =
. f(t) =

x x
x
s
t
t+

sin(/ )

t sin t +
t cos t +

. f(x) = x ex tan x
. g(x) = x sin x sec x
In Exercises
, nd the equa ons of the tangent and
normal lines to the graph of g at the indicated point.
. g(s) = es (s + ) at ( , ).
. g(t) = t sin t at (

(c) Show that the answers from (a) and (b) are equivalent.
x +
. f(x) =
x

. g(t) = t et sin t cos t


. h(t) =

. g(x) = x ( x )

x
x + x + x

. f(t) = t (sec t + et )

(a) Use the Product Rule to dieren ate the func on.

In Exercises

x + x)

. g(x) =

x
x

. g() =

cos
at ( , )
+

at ( , )

In Exercises , nd the xvalues where the graph of the


func on has a horizontal tangent line.
. f(x) = x

. f(x) = x sin x on [ , ]

. f(x) =

x
x+

. f(x) =

x
x+

y
6

In Exercises

, nd the requested deriva ve.

. f(x) = x sin x; nd f (x).

. f(x) = x sin x; nd f ( ) (x).


. f(x) = csc x; nd f (x).

. f(x) = (x x + )(x + x ); nd f ( ) (x).

In Exercises

, use the graph of f(x) to sketch f (x).

10
5

.
5

.
.

10

Chapter

Deriva ves

The Chain Rule

We have covered almost all of the deriva ve rules that deal with combina ons
of two (or more) func ons. The opera ons of addi on, subtrac on, mul plicaon (including by a constant) and division led to the Sum and Dierence rules,
the Constant Mul ple Rule, the Power Rule, the Product Rule and the Quo ent
Rule. To complete the list of dieren a on rules, we look at the last way two (or
more) func ons can be combined: the process of composi on (i.e. one func on
inside another).
One example of a composi on of func ons is f(x) = cos(x ). We currently
do not know how to compute this deriva ve. If forced to guess, one would likely
guess f (x) = sin( x), where we recognize sin x as the deriva ve of cos x
and x as the deriva ve of x . However, this is not the case; f (x) = sin( x).
In Example
well see the correct answer, which employs the new rule this
sec on introduces, the Chain Rule.
Before we dene this new rule, recall the nota on for composi on of funcons. We write (f g)(x) or f(g(x)), read as f of g of x, to denote composing f
with g. In shorthand, we simply write f g or f(g) and read it as f of g. Before
giving the corresponding dieren a on rule, we note that the rule extends to
mul ple composi ons like f(g(h(x))) or f(g(h(j(x)))), etc.
To mo vate the rule, lets look at three deriva ves we can already compute.
Exploring similar deriva ves
Example
Find the deriva ves of F (x) = ( x) , F (x) = ( x) , and F (x) = (
x) . (Well see later why we are using subscripts for dierent func ons and an
uppercase F.)
In order to use the rules we already have, we must rst exS
pand each func on as F (x) = x + x , F (x) = x + x x and
F (x) = x + x x + x .
It is not hard to see that:
F (x) = + x,

F (x) = + x x and
F (x) = +

x + x .

An interes ng fact is that these can be rewri en as


F (x) = ( x),

F (x) = ( x) and F (x) = ( x) .

A pa ern might jump out at you. Recognize that each of these func ons is a

Notes:

.
composi on, le ng g(x) =

x:

F (x) = f (g(x)),

where f (x) = x ,

F (x) = f (g(x)),
F (x) = f (g(x)),

where f (x) = x ,
where f (x) = x .

Well come back to this example a er giving the formal statements of the
Chain Rule; for now, we are just illustra ng a pa ern.

Theorem

The Chain Rule

Let y = f(u) be a dieren able func on of u and let u = g(x) be a


dieren able func on of x. Then y = f(g(x)) is a dieren able func on
of x, and
y = f (g(x)) g (x).
To help understand the Chain Rule, we return to Example

Using the Chain Rule


Example
Use the Chain Rule to nd the deriva ves of the following func ons, as given in
Example .
Example ended with the recogni on that each of the given
S
func ons was actually a composi on of func ons. To avoid confusion, we ignore
most of the subscripts here.
F (x) = ( x) :
We found that
y = ( x) = f(g(x)), where f(x) = x and g(x) =

x.

To nd y , we apply the Chain Rule. We need f (x) = x and g (x) = .


Part of the Chain Rule uses f (g(x)). This means subs tute g(x) for x in the
equa on for f (x). That is, f (x) = ( x). Finishing out the Chain Rule we
have
y = f (g(x)) g (x) = ( x) ( ) = ( x) = x .
F (x) = ( x) :

Notes:

The Chain Rule

Chapter

Deriva ves
Let y = ( x) = f(g(x)), where f(x) = x and g(x) = ( x). We have
f (x) = x , so f (g(x)) = ( x) . The Chain Rule then states
y = f (g(x)) g (x) = ( x) ( ) = ( x) .
F (x) = ( x) :
Finally, when y = ( x) , we have f(x) = x and g(x) = ( x). Thus
f (x) = x and f (g(x)) = ( x) . Thus
y = f (g(x)) g (x) = ( x) ( ) = ( x) .
Example
demonstrated a par cular pa ern: when f(x) = xn , then y =
n

n (g(x))
g (x). This is called the Generalized Power Rule.
Theorem

Generalized Power Rule

Let g(x) be a dieren able func on and let n =

be an integer. Then

)
(
)n
d(
g(x)n = n g(x)
g (x).
dx
This allows us to quickly nd the deriva ve of func ons like y = ( x x +
+ sin x) . While it may look in mida ng, the Generalized Power Rule states
that
y = ( x x + + sin x) ( x + cos x).

Treat the deriva vetaking process stepbystep. In the example just given,
rst mul ply by , the rewrite the inside of the parentheses, raising it all to
the th power. Then think about the deriva ve of the expression inside the
parentheses, and mul ply by that.
We now consider more examples that employ the Chain Rule.
Using the Chain Rule
Example
Find the deriva ves of the following func ons:
. y = sin x

. y = ln( x x )

. y = ex

. Consider y = sin x. Recognize that this is a composi on of func ons,


where f(x) = sin x and g(x) = x. Thus
y = f (g(x)) g (x) = cos( x)
Notes:

= cos x.

The Chain Rule

. Recognize that y = ln( x x ) is the composi on of f(x) = ln x and


g(x) = x x . Also, recall that
)
d(
ln x = .
dx
x

This leads us to:


y =

x x

x x) =

x x
x( x )
( x )
=
=
.
x x
x( x x)
x x

. Recognize that y = ex is the composi on of f(x) = ex and g(x) = x .


Remembering that f (x) = ex , we have
y = ex ( x) = ( x)ex .
y

Using the Chain Rule to nd a tangent line


Example
Let f(x) = cos x . Find the equa on of the line tangent to the graph of f at x = .
The tangent line goes through the point ( , f( )) ( , . )
with slope f ( ). To nd f , we need the Chain Rule.
f (x) = sin(x ) ( x) = x sin x . Evaluated at x = , we have f ( ) =
sin . . Thus the equa on of the tangent line is

.5

y= .

(x ) + .

.5

The tangent line is sketched along with f in Figure .

The Chain Rule is used o en in taking deriva ves. Because of this, one can
become familiar with the basic process and learn pa erns that facilitate nding
deriva ves quickly. For instance,
)
(anything)
d(
(anything) =
.
ln(anything) =
dx
anything
anything

A concrete example of this is

)
d(
ln( x cos x + ex ) =
dx

x + sin x + ex
.
x cos x + ex

While the deriva ve may look in mida ng at rst, look for the pa ern. The
denominator is the same as what was inside the natural log func on; the numerator is simply its deriva ve.

Notes:

Figure . : f(x) = cos x sketched along


with its tangent line at x = .

Chapter

Deriva ves
This pa ern recogni on process can be applied to lots of func ons. In general, instead of wri ng anything, we use u as a generic func on of x. We then
say
) u
d(
ln u = .
dx
u
The following is a short list of how the Chain Rule can be quickly applied to familiar func ons.
d ( n)
u = n un u .
dx
d ( u)
e = u eu .
.
dx
)
d(
.
sin u = u cos u.
dx

)
d(
cos u = u sin u.
dx
)
d(
tan u = u sec u.
.
dx

Of course, the Chain Rule can be applied in conjunc on with any of the other
rules we have already learned. We prac ce this next.
Using the Product, Quo ent and Chain Rules
Example
Find the deriva ves of the following func ons.
. f(x) = x sin x

. f(x) =

x
.
ex

. We must use the Product and Chain Rules. Do not think that you must be
able to see the whole answer immediately; rather, just proceed step
bystep.
(
)
(
)
f (x) = x x cos x + x sin x = x cos x + x sin x .

. We must employ the Quo ent Rule along with the Chain Rule. Again, proceed stepbystep.
(
)
(
)
(
)
ex
x + x
ex
x x ( x)ex

=
f (x) =
(
)
e x
ex
(
)
= ex
x + x .

A key to correctly working these problems is to break the problem down


into smaller, more manageable pieces. For instance, when using the Product

Notes:

.
and Chain Rules together, just consider the rst part of the Product Rule at rst:
f(x)g (x). Just rewrite f(x), then nd g (x). Then move on to the f (x)g(x) part.
Dont a empt to gure out both parts at once.
Likewise, using the Quo ent Rule, approach the numerator in two steps and
handle the denominator a er comple ng that. Only simplify a erward.
We can also employ the Chain Rule itself several mes, as shown in the next
example.
Using the Chain Rule mul ple mes
Example
Find the deriva ve of y = tan ( x x).
Recognize that we have the g(x) = tan( x x) func on
S
(
)
inside the f(x) = x func on; that is, we have y = tan( x x) . We begin
using the Generalized Power Rule; in this rst step, we do not fully compute the
deriva ve. Rather, we are approaching this stepbystep.
(
)
y = tan( x x) g (x).

We now nd g (x). We again need the Chain Rule;


g (x) = sec ( x x) (

x ).

Combine this with what we found above to give


(
)
y = tan( x x) sec ( x x) (
=(

x )

) sec ( x x) tan ( x x).

This func on is frankly a ridiculous func on, possessing no real prac cal
value. It is very dicult to graph, as the tangent func on has many ver cal
asymptotes and x x grows so very fast. The important thing to learn from
this is that the deriva ve can be found. In fact, it is not hard; one must take
several simple steps and be careful to keep track of how to apply each of these
steps.
It is a tradi onal mathema cal exercise to nd the deriva ves of arbitrarily
complicated func ons just to demonstrate that it can be done. Just break everything down into smaller pieces.
Using the Product, Quo ent and Chain Rules
x cos(x ) sin (e x )
Find the deriva ve of f(x) =
.
ln(x + x )

Example

This func on likely has no prac cal use outside of demonS


stra ng deriva ve skills. The answer is given below without simplica on. It

Notes:

The Chain Rule

Chapter

Deriva ves
employs the Quo ent Rule, the Product Rule, and the Chain Rule three mes.
f (x) =

[(
]
)
ln(x + x ) x ( sin(x )) ( x ) + cos(x ) sin(e x ) cos(e x ) ( e x )

(
)
x
x cos(x ) sin (e x ) xx+
+ x
.
(
)
ln(x + x )

The reader is highly encouraged to look at each term and recognize why it
is there. (I.e., the Quo ent Rule is used; in the numerator, iden fy the LOdHI
term, etc.) This example demonstrates that deriva ves can be computed systema cally, no ma er how arbitrarily complicated the func on is.
The Chain Rule also has theore c value. That is, it can be used to nd the
deriva ves of func ons that we have not yet learned as we do in the following
example.
The Chain Rule and exponen al func ons
Example
Use the Chain Rule to nd the deriva ve of y = ax where a >
constant.

, a =

is

We only know how to nd the deriva ve of one exponen al


S
func on: y = ex ; this problem is asking us to nd the deriva ve of func ons
such as y = x .
This can be accomplished by rewri ng ax in terms of e. Recalling that ex and
ln x are inverse func ons, we can write
a = eln a
get

and so

y = ax = eln(a ) .

By the exponent property of logarithms, we can bring down the power to


y = ax = ex(ln a) .

The func on is now the composi on y = f(g(x)), with f(x) = ex and g(x) =
x(ln a). Since f (x) = ex and g (x) = ln a, the Chain Rule gives
y = ex(ln a) ln a.

Recall that the ex(ln a) term on the right hand side is just ax , our original func on.
Thus, the deriva ve contains the original func on itself. We have
y = y ln a = ax ln a.

Notes:

.
The Chain Rule, coupled with the deriva ve rule of ex , allows us to nd the
deriva ves of all exponen al func ons.
The previous example produced a result worthy of its own box.
Theorem

Deriva ves of Exponen al Func ons

Let f(x) = ax , for a >


numbers and

, a =

. Then f is dieren able for all real

f (x) = ln a ax .

Alternate Chain Rule Nota on


It is instruc ve to understand what the Chain Rule looks like using dy
dx nota on instead of y nota on. Suppose that y = f(u) is a func on of u, where
u = g(x) is a func on of x, as stated in Theorem . Then, through the composi on f g, we can think of y as a func on of x, as y = f(g(x)). Thus the
deriva ve of y with respect to x makes sense; we can talk about dy
dx . This leads
to an interes ng progression of nota on:
y = f (g(x)) g (x)
dy
= y (u) u (x)
(since y = f(u) and u = g(x))
dx
dy
dy du
=
(using frac onal nota on for the deriva ve)

dx
du dx
Here the frac onal aspect of the deriva ve nota on stands out. On the
right hand side, it seems as though the du terms cancel out, leaving
dy
dy
= .
dx
dx
It is important to realize that we are not canceling these terms; the deriva ve
nota on of dy
dx is one symbol. It is equally important to realize that this nota on
was chosen precisely because of this behavior. It makes applying the Chain Rule
easy with mul ple variables. For instance,
dy
dy d d
=

.
dt
d d dt
where and are any variables youd like to use.

Notes:

The Chain Rule

.
Chapter

Deriva ves
One of the most common ways of visualizing the Chain Rule is to consider
a set of gears, as shown in Figure . . The gears have , , and teeth,
respec vely. That means for every revolu on of the x gear, the u gear revolves
twice. That is, the rate at which the u gear makes a revolu on is twice as fast
as the rate at which the x gear makes a revolu on. Using the terminology of
calculus, the rate of u-change, with respect to x, is du
dx = .
dy
= . How does
Likewise, every revolu on of u causes revolu ons of y: du
y change with respect to x? For each revolu on of x, y revolves
mes; that is,
x

dy du
dy
=

=
dx
du dx
dy
=6
dx

du
=2
dx

y
dy
=3
du

Figure . : A series of gears to demon=


strate the Chain Rule. Note how dy
dx
dy
du

du
dx

= .

We can then extend the Chain Rule with more variables by adding more gears
to the picture.
It is dicult to overstate the importance of the Chain Rule. So o en the
func ons that we deal with are composi ons of two or more func ons, requiring
us to use this rule to compute deriva ves. It is o en used in prac ce when actual
dy
and
func ons are unknown. Rather, through measurement, we can calculate du
dy
du
.
With
our
knowledge
of
the
Chain
Rule,
nding
is
straigh
orward.
dx
dx
In the next sec on, we use the Chain Rule to jus fy another dieren a on
technique. There are many curves that we can draw in the plane that fail the
ver cal line test. For instance, consider x + y = , which describes the unit
circle. We may s ll be interested in nding slopes of tangent lines to the circle at
various points. The next sec on shows how we can nd dy
dx without rst solving
for y. While we can in this instance, in many other instances solving for y is
impossible. In these situa ons, implicit dieren a on is indispensable.

Notes:

Exercises .

Terms and Concepts


. T/F: The Chain Rule describes how to evaluate the derivave of a composi on of func ons.
)
d(
g(x)n =
. T/F: The Generalized Power Rule states that
dx
(
)n
n g(x)
.

. T/F:

. T/F:
. T/F:

)
d(
ln(x ) = .
dx
x
d ( x)
.
dx

dx
dx dt
=

dy
dt dy

. T/F: Taking the deriva ve of f(x) = x sin( x) requires the


use of both the Product and Chain Rules.

t
t

. m(w) =

. f(x) =

. g(t) = cos(t + t) sin( t )


. g(t) = cos( t )e

, compute the deriva ve of the given func-

. f(t) = ( t ) at t =
. g() = (sin + cos ) at = /

. Compute

. g() = (sin + cos )


t +t

)
(
. f(x) = x + x
. f(x) = cos( x)

. g(x) = tan( x)
. h(t) = sin ( t)
. p(t) = cos (t + t + )
. f(x) = ln(cos x)

t +t

at t =

)
d(
ln(kx) two ways:
dx

(a) Using the Chain Rule, and

(b) by rst using the logarithm rule ln(ab) = ln a + ln b,


then taking the deriva ve.
. Compute

)
d(
ln(xk ) two ways:
dx

(a) Using the Chain Rule, and

(b) by rst using the logarithm rule ln(ap ) = p ln a, then


taking the deriva ve.

Review
. The wind chill factor is a measurement of how cold it
feels during cold, windy weather. Let W(w) be the wind
chill factor, in degrees Fahrenheit, when it is F outside
with a wind of w mph.
(a) What are the units of W (w)?

. g(r) =

. g(t) =

cos t

(b) What would you expect the sign of W (


. Find the deriva ves of the following func ons.

. g(t) =
. m(w) =

In Exercises , nd the equa ons of tangent and normal


lines to the graph of the func on at the given point. Note: the
func ons here are the same as in Exercises through .

. f(t) = ( t )

. f(x) = ln(x)

+x
x

. f(x) = x sin( x)

. h(t) = e

. f(x) = ln(x )

+
w

. f(x) = ( x x)

. h(t) = e

+
+

. f(x) = ( x x) at x =

Problems
In Exercises
on.

. h(t) =

w
w

(a) f(x) = x ex cot x


(b) g(x) =

x x x

) to be?

Chapter

Deriva ves

.
Figure . : A graph of the implicit funcon sin(y) + y = x .

Implicit Dieren a on

In the previous sec ons we learned to nd the deriva ve, dy


dx , or y , when y is
given explicitly as a func on of x. That is, if we know y = f(x) for some func on
f, we can nd y . For example, given y = x , we can easily nd y = x.
(Here we explicitly state how x and y are related. Knowing x, we can directly nd
y.)
Some mes the rela onship between y and x is not explicit; rather, it is implicit. For instance, we might know that x y = . This equality denes a
rela onship between x and y; if we know x, we could gure out y. Can we s ll
nd y ? In this case, sure; we solve for y to get y = x (hence we now know
y explicitly) and then dieren ate to get y = x.
Some mes the implicit rela onship between x and y is complicated. Suppose we are given sin(y) + y = x . A graph of this implicit func on is given
in Figure . . In this case there is absolutely no way to solve for y in terms of
elementary func ons. The surprising thing is, however, that we can s ll nd y
via a process known as implicit dieren a on.
Implicit dieren a on is a technique based on the Chain Rule that is used to
nd a deriva ve when the rela onship between the variables is given implicitly
rather than explicitly (solved for one variable in terms of the other).

We begin by reviewing the Chain Rule. Let f and g be func ons of x. Then
)
d(
f(g(x)) = f (g(x)) g (x).
dx

Suppose now that y = g(x). We can rewrite the above as


)
d(
f(y)) = f (y)) y ,
dx

or

)
d(
dy
f(y)) = f (y) .
dx
dx

( . )

These equa ons look strange; the key concept to learn here is that we can nd
y even if we dont exactly know how y and x relate.
We demonstrate this process in the following example.
Using Implicit Dieren a on
Example
Find y given that sin(y) + y = x .
We start by taking the deriva ve of both sides (thus mainS
taining the equality.) We have :
)
)
d(
d(
sin(y) + y =
x .
dx
dx

Notes:

.
The right hand side is easy; it returns x .
The le hand side requires more considera on. We take the deriva ve term
byterm. Using the technique derived from Equa on . above, we can see that
)
d(
sin y = cos y y .
dx
We apply the same process to the y term.
)
d( )
d(
y =
(y) = (y) y .
dx
dx

Pu ng this together with the right hand side, we have

Now solve for y .

cos(y)y + y y = x .
cos(y)y + y y = x .
(
)
cos y + y y = x
y =

x
cos y + y

This equa on for y probably seems unusual for it contains both x and y
terms. How is it to be used? Well address that next.
Implicit func ons are generally harder to deal with than explicit func ons.
With an explicit func on, given an x value, we have an explicit formula for compu ng the corresponding y value. With an implicit func on, one o en has to
nd x and y values at the same me that sa sfy the equa on. It is much easier to demonstrate that a given point sa ses the equa on than to actually nd
such a point.

For instance, we can arm easily that the point ( , ) lies on the graph of
the implicit func on siny + y = x . Plugging in for y, we see the le hand
, we see the right hand side is also ; the equa on is
side is . Se ng x =
sa sed. The following example nds the equa on of the tangent line to this
func on at this point.
Using Implicit Dieren a on to nd a tangent line
Example
Find the equa on of the line tangent to
the curve of the implicitly dened funcon sin y + y = x at the point ( , ).
S

In Example

we found that

y =

Notes:

x
.
cos y + y

Implicit Dieren a on

Chapter

Deriva ves
y

.
Figure . : The func on sin y + y =
x and its tangent line at the point
( , ).

We nd the slope of the tangent


line at the point ( , ) by subs tu ng
for
x and for y. Thus at the point ( , ), we have the slope as

( )

=
. .
y =
cos +
Therefore the equa on of thetangent line to the implicitly dened func on
sin y + y = x at the point ( , ) is

(x
)+ . x+ .
y=

The curve and this tangent line are shown in Figure .

This suggests a general method for implicit dieren a on. For the steps below assume y is a func on of x.
. Take the deriva ve of each term in the equa on. Treat the x terms like
normal. When taking the deriva ves of y terms, the usual rules apply
except that, because of the Chain Rule, we need to mul ply each term
by y .
. Get all the y terms on one side of the equal sign and put the remaining
terms on the other side.
. Factor out y ; solve for y by dividing.
Prac cal Note: When working by hand, it may be benecial to use the symbol
dy

dx instead of y , as the la er can be easily confused for y or y .


Using Implicit Dieren a on
Example
Given the implicitly dened func on y + x y = + x, nd y .
We will take the implicit deriva ves term by term. The derivaS
ve of y is y y .
The second term, x y , is a li le tricky. It requires the Product Rule as it is the
product of two func ons of x: x and y . Its deriva ve is x ( y y ) + xy . The
rst part of this expression requires a y because we are taking the deriva ve of a
y term. The second part does not require it because we are taking the deriva ve
of x .
The deriva ve of the right hand side is easily found to be . In all, we get:
y y + x y y + xy = .
Move terms around so that the le side consists only of the y terms and the
right side consists of all the other terms:
y y + x y y =

Notes:

xy .

.
Factor out y from the le side and solve to get
y =

Implicit Dieren a on

xy
.
y + x y

To conrm the validity of our work, lets nd the equa on of a tangent line
to this func on at a point. It is easy to conrm that the point ( , ) lies on the
graph of this func on. At this point, y = / . So the equa on of the tangent
line is y = / (x )+ . The func on and its tangent line are graphed in Figure
. .
No ce how our func on looks much dierent than other func ons we have
seen. For one, it fails the ver cal line test. Such func ons are important in many
areas of mathema cs, so developing tools to deal with them is also important.

10

10
.

Figure . : A graph of the implicitly dened func on y + x y = + x along


with its tangent line at the point ( , ).

Using Implicit Dieren a on


Example
Given the implicitly dened func on sin(x y ) + y = x + y, nd y .
Dieren a ng term by term, we nd the most diculty in
S
the rst term. It requires both the Chain and Product Rules.
)
)
d(
d(
sin(x y ) = cos(x y )
x y
dx
(dx
)
= cos(x y ) x ( yy ) + xy

= (x yy + xy ) cos(x y ).

We leave the deriva ves of the other terms to the reader. A er taking the
deriva ves of both sides, we have
(x yy + xy ) cos(x y ) + y y =

+ y .

We now have to be careful to properly solve for y , par cularly because of


the product on the le . It is best to mul ply out the product. Doing this, we get
x y cos(x y )y + xy cos(x y ) + y y =

+ y .

From here we can safely move around terms to get the following:
x y cos(x y )y + y y y =

xy cos(x y ).

Then we can solve for y to get


y =

Notes:

xy cos(x y )
.
x y cos(x y ) + y

.
Figure . : A graph of the implicitly dened func on sin(x y ) + y = x + y.

Chapter

Deriva ves
y

1
x

1
1

.
Figure . : A graph of the implicitly dened func on sin(x y ) + y = x + y and
certain tangent lines.

A graph of this implicit func on is given in Figure . . It is easy to verify that


the points ( , ), ( , ) and ( , ) all lie on the graph. We can nd the slopes
of the tangent lines at each of these points using our formula for y .
At ( , ), the slope is .
At ( , ), the slope is / .
At ( , ), the slope is also / .
The tangent lines have been added to the graph of the func on in Figure
. .
Quite a few famous curves have equa ons that are given implicitly. We can
use implicit dieren a on to nd the slope at various points on those curves.
We inves gate two such curves in the next examples.
Finding slopes of tangent lines to a circle
Example

/ ).
Find the slope of the tangent line to the circle x +y = at the point ( / ,
S

Taking deriva ves, we get x+ yy = . Solving for y gives:


y =

y
( / ,

This is a clever formula. Recall that the slope of the line through the origin and
the point (x, y) on the circle will be y/x. We have found that the slope of the
tangent line to the circle at that point is the opposite reciprocal of y/x, namely,
x/y. Hence these two
lines are always perpendicular.
/ ), we have the tangent lines slope as
At the point ( / ,

/ )

y =
= .
/

x
.
y

Figure . : The
unit circle with its tan/ ).
gent line at ( / ,

/ ) is given in Figure
A graph of the circle and its tangent line at ( / ,
. , along with a thin dashed line from the origin that is perpendicular to the
tangent line. (It turns out that all normal lines to a circle pass through the center
of the circle.)
This sec on has shown how to nd the deriva ves of implicitly dened funcons, whose graphs include a wide variety of interes ng and unusual shapes.
Implicit dieren a on can also be used to further our understanding of regular dieren a on.
One hole in our current understanding of deriva ves is this: what is the
deriva ve of the square root func on? That is,
d(
d ( )
x =
x
dx
dx
Notes:

=?

.
We allude to a possible solu on, as we can write the square root func on as
a power func on with a ra onal (or, frac onal) power. We are then tempted to
apply the Power Rule and obtain
d(
x
dx

= .
x

The trouble with this is that the Power Rule was ini ally dened only for
posi ve integer powers, n > . While we did not jus fy this at the me, generally the Power Rule is proved using something called the Binomial Theorem,
which deals only with posi ve integers. The Quo ent Rule allowed us to extend
the Power Rule to nega ve integer powers. Implicit Dieren a on allows us to
extend the Power Rule to ra onal powers, as shown below.
Let y = xm/n , where m and n are integers with no common factors (so m =
and n = is ne, but m = and n = is not). We can rewrite this explicit
func on implicitly as yn = xm . Now apply implicit dieren a on.
y = xm/n
yn = xm
d ( n)
d ( m)
y =
x
dx
dx
n yn y = m xm

m xm
(now subs tute xm/n for y)
n yn
m xm
(apply lots of algebra)
=
n (xm/n )n

y =

m (mn)/n
x
n

m m/n
x
.
n

The above deriva on is the key to the proof extending the Power Rule to raonal powers. Using limits, we can extend this once more to include all powers,
including irra onal (even transcendental!) powers, giving the following theorem.
Theorem

Power Rule for Dieren a on


n

Let f(x) = x , where n = is a real number. Then f is a dieren able


func on, and f (x) = n xn .

Notes:

Implicit Dieren a on

Chapter

Deriva ves
y
20

This theorem allows us to say the deriva ve of x is x .


We now apply this nal version of the Power Rule in the next example, the
second inves ga on of a famous curve.
20

20

Using the Power Rule


Example
Find the slope of x / + y / = at the point ( , ).

20

Figure . : An astroid, traced out by a


point on the smaller circle as it rolls inside
the larger circle.

This is a par cularly interes ng curve called an astroid. It


S
is the shape traced out by a point on the edge of a circle that is rolling around
inside of a larger circle, as shown in Figure . .
To nd the slope of the astroid at the point ( , ), we take the deriva ve
implicitly.

20
(8, 8)

20

20

.
Figure .
line.

: An astroid with a tangent

+ y

y =

y = x

x /
y /

y/
y
y = / =
.
x
x

y =

20

Plugging in x = and y = , we get a slope of . The astroid, with its


tangent line at ( , ), is shown in Figure . .

Implicit Dieren a on and the Second Deriva ve


We can use implicit dieren a on to nd higher order deriva ves. In theory,
this is simple: rst nd dy
dx , then take its deriva ve with respect to x. In prac ce,
it is not hard, but it o en requires a bit of algebra. We demonstrate this in an
example.
Finding the second deriva ve
d y
Given x + y = , nd
= y .
dx
Example

Notes:

We found that y =

dy
dx

= x/y in Example

. To nd y ,

Implicit Dieren a on

we apply implicit dieren a on to y .


d ( )
y
dx ( )
d
x
=

(Now use the Quo ent Rule.)


dx
y
y( ) x(y )
=
y

y =

replace y with x/y:


y x(x/y)
y
y + x /y
=
.
y

While this is not a par cularly simple expression, it is usable. We can see that
y > when y < and y < when y > . In Sec on . , we will see how
this relates to the shape of the graph.
y

Logarithmic Dieren a on
Consider the func on y = xx ; it is graphed in Figure . . It is welldened
for x > and we might be interested in nding equa ons of lines tangent and
normal to its graph. How do we take its deriva ve?
The func on is not a power func on: it has a power of x, not a constant.
It is not an exponen al func on: it has a base of x, not a constant.
A dieren a on technique known as logarithmic dieren a on becomes
useful here. The basic principle is this: take the natural log of both sides of an
equa on y = f(x), then use implicit dieren a on to nd y . We demonstrate
this in the following example.
Using Logarithmic Dieren a on
Example
Given y = xx , use logarithmic dieren a on to nd y .
S

Notes:

As suggested above, we start by taking the natural log of

Figure .

: A plot of y = xx .

Chapter

Deriva ves
both sides then applying implicit dieren a on.

y = xx
ln(y) = ln(xx )

(apply logarithm rule)

ln(y) = x ln x
(now use implicit dieren a on)
)
)
d(
d(
ln(y) =
x ln x
dx
dx
y
= ln x + x
y
x
y
= ln x +
y
(
)
y = y ln x +
(subs tute y = xx )
(
)
y = xx ln x + .

Figure . : A graph of y = xx and its tangent line at x = . .

To test our answer, lets use it to nd the equa on of the tangent line at x =
. . The point on the graph our tangent line must pass through is ( . , . . )
( . , .
). Using the equa on for y , we nd the slope as
(
)
( .
) .
.
y = . . ln . + .

Thus the equa on of the tangent line is y = .


. graphs y = xx along with this tangent line.

(x . ) + .

. Figure

Implicit dieren a on proves to be useful as it allows us to nd the instantaneous rates of change of a variety of func ons. In par cular, it extended the
Power Rule to ra onal exponents, which we then extended to all real numbers.
In the next sec on, implicit dieren a on will be used to nd the deriva ves of
inverse func ons, such as y = sin x.

Notes:

Exercises .

Terms and Concepts

. (y + y x) =

. In your own words, explain the dierence between implicit


func ons and explicit func ons.

x +y
=
x+y

. Implicit dieren a on is based on what other dieren aon rule?

sin(x) + y
=
cos(y) + x

. T/F: Implicitdieren a on can be used to nd the derivave of y = x.

. ln(x + y ) = e

. T/F: Implicit dieren a on can be used to nd the derivave of y = x / .

Problems

. f(x) =
. f(x) =

. f(t) =

. g(t) =

. h(x) = x

dy
is the same for each of the following implicitly
. Show that
dx
dened func ons.
(a) xy =

In Exercises
on.

. ln(x + xy + y ) =

, compute the deriva ve of the given func-

(b) x y =
(c) sin(xy) =

x+

x+x

(d) ln(xy) =

In Exercises
, nd the equa on of the tangent line to
the graph of the implicitly dened func on at the indicated
points. As a visual aid, each func on is graphed.

. x

t sin t

+y

(a) At ( , ).

. f(x) = x + x

(b) At ( . , .
) (which does not exactly lie on the
curve, but is very close).

x+
. g(x) =
x
. f(t) =

t(sec t + e )

In Exercises

, nd

. x +y +y=
. x

+y

.5

dy
using implicit dieren a on.
dx
.

. x +y =
(a) At ( , ).

(b) At ( . ,
. ).

x
=
y

(c) At ( , ).
y

y
=
.
x
. xe +

.5
.5

. cos(x) + sin(y) =
.

( . , . 8 )

.5

( 0.6, 0.8)

0.5

. x tan y =
. ( x + y) =

0.5

0.5
0.5

. (x + y ) =

(a) At ( , ).

(b) At ( ,

In Exercises
, an implicitly dened func on is given.
dy
Find
. Note: these are the same problems used in Exerdx
cises through .

).

. x +y +y=

. x
4

( ,

8)

x
=
y

xx
,
x+

x=

. y=

. (x ) + (y ) =
)
(
+
.
,
(a) At

(
)
+
(b) At
.
,
y

, .

x=

. y = xsin(x)+ , x = /

x=

. ,

/x

. y = ( x)x ,

. y = ( + x)

In Exercises
, use logarithmic dieren a on to nd
dy
, then nd the equa on of the tangent line at the indicated
dx
xvalue.

. (x + y + x) = x + y
(a) At ( , ).
)
(
(b) At ,
.

+y

. cos x + sin y =

. y=

x+
,
x+

. y=

(x + )(x + )
, x=
(x + )(x + )

x=

Deriva ves of Inverse Func ons

Deriva ves of Inverse Func ons

Recall that a func on y = f(x) is said to be one to one if it passes the horizontal
line test; that is, for two dierent x values x and x , we do not have f(x ) = f(x ).
In some cases the domain of f must be restricted so that it is one to one. For
instance, consider f(x) = x . Clearly, f( ) = f( ), so f is not one to one on its
regular domain, but by restric ng f to ( , ), f is one to one.
Now recall that one to one func ons have inverses. That is, if f is one to
one, it has an inverse func on, denoted by f , such that if f(a) = b, then
f (b) = a. The domain of f is the range of f, and vice-versa. For ease of
nota on, we set g = f and treat g as a func on of x.
Since f(a) = b implies g(b) = a, when we compose f and g we get a nice
result:
(
)
f g(b) = f(a) = b.
(
)
(
)
In general, f g(x) = x and g f(x) = x. This gives us a convenient way to check
if two func ons are inverses of each other: compose them and if the result is x,
then they are inverses (on the appropriate domains.)
When the point (a, b) lies on the graph of f, the point (b, a) lies on the graph
of g. This leads us to discover that the graph of g is the reec on of f across the
line y = x. In Figure . we see a func on graphed along with its inverse. See
how the point ( , . ) lies on one graph, whereas ( . , ) lies on the other. Because of this rela onship, whatever we know about f can quickly be transferred
into knowledge about g.
For example, consider Figure . where the tangent line to f at the point
(a, b) is drawn. That line has slope f (a). Through reec on across y = x, we
can see that the tangent line to g at the point (b, a) should have slope
This then tells us that g (b) =
Consider:

f (a)

Informa on about f
( . , .

) lies on f

Slope of tangent line to f


at x = . is /
f ( . ) = /

f (a)

y
(1, 1.5)

(1.5, 1)

(0.5, 0.375)

(0.375, 0.5)

Figure . : A func on f along with its inverse f . (Note how it does not ma er
which func on we refer to as f; the other
is f .)

.
y

Informa on about g = f
, . ) lies on g
.
Slope of tangent line to
g at x = .
is /
( .

g ( .

)= /

We have discovered a rela onship between f and g in a mostly graphical


way. We can realize this rela onship analy cally as well. Let y = g(x), where
again g = f . We want to nd y . Since y = g(x), we know that f(y) = x. Using
the Chain Rule and Implicit Dieren a on, take the deriva ve of both sides of

Notes:

1
(a, b)

(b, a)

Figure . : Corresponding tangent lines


drawn to f and f .

Chapter

Deriva ves
this last equality.
)
d(
d( )
f(y) =
x
dx
dx
f (y) y =
y =
y =

f (y)
f (g(x))

This leads us to the following theorem.


Theorem

Deriva ves of Inverse Func ons

Let f be dieren able and one to one on an open interval I, where f (x) =
for all x in I, let J be the range of f on I, let g be the inverse func on of
f, and let f(a) = b for some a in I. Then g is a dieren able func on on
J, and in par cular,
( )
( )
. f (b) = g (b) =
and
. f (x) = g (x) =
f (a)
f (g(x))
The results of Theorem are not trivial; the nota on may seem confusing
at rst. Careful considera on, along with examples, should earn understanding.
In the next example we apply Theorem to the arcsine func on.
Finding the deriva ve of an inverse trigonometric func on
Example
Let y = arcsin x = sin x. Find y using Theorem .
Adop ng our previously dened nota on, let g(x) = arcsin x
S
and f(x) = sin x. Thus f (x) = cos x. Applying the theorem, we have
g (x) =
=

f (g(x))
cos(arcsin x)

This last expression is not immediately illumina ng. Drawing a gure will
help, as shown in Figure . . Recall that the sine func on can be viewed as
taking in an angle and returning a ra o of sides of a right triangle, specically,
the ra o opposite over hypotenuse. This means that the arcsine func on takes
as input a ra o of sides and returns an angle. The equa on y = arcsin x can
be rewri en as y = arcsin(x/ ); that is, consider a right triangle where the

Notes:

Deriva ves of Inverse Func ons

hypotenuse has length and the side opposite


of the angle with measure y has
length x. This means the nal side has length
x , using the Pythagorean
Theorem.
1

Therefore cos(sin x) = cos y =

x / =

x , resul ng in

1 x2

Figure . : A right triangle dened by


y = sin (x/ ) with the length of the
third leg found using the Pythagorean
Theorem.

)
d(
arcsin x = g (x) =
.
dx
x

Remember that the input x of the arcsine func on is a ra o of a side of a right


triangle to its hypotenuse; the absolute value of this ra o will never be greater
than . Therefore the inside of the square root will never be nega ve.
In order to make y = sin x one to one, we restrict its domain to [/ , / ];
on this domain, the range is [ , ]. Therefore the domain of y = arcsin x is
[ , ] and the range is [/ , / ]. When x = , note how the deriva ve of
the arcsine func on is undened; this corresponds to the fact that as x ,
the tangent lines to arcsine approach ver cal lines with undened slopes.
In Figure . we see f(x) = sin x and f
= sin x graphed
on their re/ ) has slope
spec ve domains. The line tangent to sin x at the point (/ ,

cos
/ = / . The slope of the corresponding point on sin x, the point
( / , / ), is

(,

y = sin x

, )

/ )

= ,

verifying yet again that at corresponding points, a func on and its inverse have
reciprocal slopes.
Using similar techniques, we can nd the deriva ves of all the inverse trigonometric func ons. In Figure . we show the restric ons of the domains of the
standard trigonometric func ons that allow them to be inver ble.

Notes:

y = sin x

Figure . : Graphs of sin x and sin x


along with corresponding tangent lines.

Chapter

Deriva ves

Func on

Domain

sin x

[/ , / ]

[ , ]

cos x

[ , ]

[ , ]

tan x
csc x
sec x
cot x

Inverse
Func on

Range

[ , / ) (/ , ]
( , )

tan

(, ] [ , )

(, ] [ , )
(, )

Figure .

cos (x)

(, )

(/ , / )
[/ , ) ( , / ]

sin

(x)
x

csc

sec

(x)

cot

(x)

Domain

Range

[ , ]

[/ , / ]

[ , ]

[ , ]

(, )

(/ , / )

(, ] [ , )

[/ , ) ( , / ]

(, )

( , )

(, ] [ , )

[ , / ) (/ , ]

: Domains and ranges of the trigonometric and inverse trigonometric func ons.

Theorem

Deriva ves of Inverse Trigonometric Func ons

The inverse trigonometric func ons are dieren able on all open sets
contained in their domains (as listed in Figure . ) and their deriva ves
are as follows:
)
d(
sin (x) =
dx
x
(
)
d
sec (x) =
.
dx
|x| x
)
d(

tan (x) =
.
dx
+x
.

)
d(
cos (x) =
dx
x
)
d(
csc (x) =
.
dx
|x| x
)
d(
cot (x) =
.
dx
+x
.

Note how the last three deriva ves are merely the opposites of the rst
three, respec vely. Because of this, the rst three are used almost exclusively
throughout this text.
)
d(
ln x = .
In Sec on . , we stated without proof or explana on that
dx
x
We can jus fy that now using Theorem , as shown in the example.
Example
Use Theorem

Finding the deriva ve of y = ln x


)
d(
ln x .
to compute
dx

View y = ln x as the inverse of y = ex . Therefore, using our


standard nota on, let f(x) = ex and g(x) = ln x. We wish to nd g (x). Theorem
S

Notes:

.
gives:
g (x) =

f (g(x))

eln x

In this chapter we have dened the deriva ve, given rules to facilitate its
computa on, and given the deriva ves of a number of standard func ons. We
restate the most important of these in the following theorem, intended to be a
reference for further work.
Theorem

Glossary of Deriva ves of Elementary Func ons

Let u and v be dieren able func ons, and let a, c and n be real
numbers, a > , n = .
.
.
.
.
.
.
.
.
.
.
.
.

Notes:

( )
cu = cu
(
)
d

dx u v = uv + u v
(
)
d

dx u(v) = u (v)v
( )
d
dx x =
( x)
d
= ex
dx e
(
)
d
dx ln x = x
(
)
d
dx sin x = cos x
(
)
d
dx csc x = csc x cot x
(
)
d
dx tan x = sec x
( )
d
x = x
dx sin
d
dx

d
dx
d
dx

(
(

)
csc x = |x|x
)
tan x =

+x

.
.
.
.
.
.
.
.
.
.
.
.

d
dx

)
u v = u v
( ) u vuv
d u
dx v =
v
( )
d
dx c =
( n)
d
= nxn
dx x
( x)
d
= ln a ax
dx a
(
)
d
dx loga x = ln a x
(
)
d
dx cos x = sin x
(
)
d
dx sec x = sec x tan x
(
)
d
dx cot x = csc x
(
)
d

x = x
dx cos
d
dx

d
dx

)
sec x =

|x| x

)
cot x =

+x

Deriva ves of Inverse Func ons

Exercises .

Terms and Concepts


. T/F: Every func on has an inverse.
. In your own words explain what it means for a func on to
be one to one.
. If ( , ) lies on the graph of y = f(x), what can be said
about the graph of y = f (x)?
. If ( , ) lies on the graph of y = f(x) and f ( ) = , what
can be said about y = f (x)?

In Exercises
on.

, compute the deriva ve of the given func-

. h(t) = sin ( t)
. f(t) = sec ( t)
. g(x) = tan ( x)
. f(x) = x sin x
. g(t) = sin t cos t

Problems

. f(t) = ln tet

In Exercises , verify that the given func ons are inverses.

. h(x) =

sin x
cos x

. f(x) = x + and g(x) = x

. g(x) = tan ( x)

. f(x) = x + x + , x

g(x) = x , x

. f(x) = sec ( /x)

and

. f(x) =

, x = and
x
+ x
g(x) =
, x =
x

. f(x) =

x+
, x =
x

and g(x) = f(x)

. f(x) = sin x, / x /

/ )
Point= (/ ,
( )
Evaluate f
( / )

. f(x) = x x + x
Point= ( , )
( )
Evaluate f ( )
,x

+x
Point= ( , / )
( )
Evaluate f ( / )

. f(x) = e x
Point= ( , )
( )
Evaluate f ( )

(a) By simplifying rst, then taking the deriva ve, and


Verify that the two answers are the same.
. f(x) = sin(sin x)
. f(x) = tan (tan x)

. f(x) = x x + , x
Point= ( , )
( )
Evaluate f ( )

. f(x) =

In Exercises , compute the deriva ve of the given funcon in two ways:


(b) by using the Chain Rule rst then simplifying.

In Exercises , an inver ble func on f(x) is given along


with( a point
) that lies on its graph. Using Theorem , evaluate f (x) at the indicated value.
. f(x) = x +
Point= ( , )
( )
Evaluate f (

. f(x) = sin(sin x)

. f(x) = sin(cos x)
In Exercises
, nd the equa on of the line tangent to
the graph of f at the indicated x value.
. f(x) = sin x

at x =

. f(x) = cos ( x)

at x =

Review
. Find

dy
,
dx

where x y y x = .

. Find the equa on of the line tangent to the graph of x +


y + xy = at the point ( , ).
. Let f(x) = x + x.
f(x + s) f(x)
Evaluate lim
.
s
s

:T

Our study of limits led to con nuous func ons, which is a certain class of funcons that behave in a par cularly nice way. Limits then gave us an even nicer
class of func ons, func ons that are dieren able.
This chapter explores many of the ways we can take advantage of the informa on that con nuous and dieren able func ons provide.

Extreme Values

Given any quan ty described by a func on, we are o en interested in the largest
and/or smallest values that quan ty a ains. For instance, if a func on describes
the speed of an object, it seems reasonable to want to know the fastest/slowest
the object traveled. If a func on describes the value of a stock, we might want
to know how the highest/lowest values the stock a ained over the past year.
We call such values extreme values.
Deni on

(
.2

(a)
y

Extreme Values

Let f be dened on an interval I containing c.


. f(c) is the minimum (also, absolute minimum) of f on I if f(c)
f(x) for all x in I.

(b)

. f(c) is the maximum (also, absolute maximum) of f on I if f(c)


f(x) for all x in I.

The maximum and minimum values are the extreme values, or extrema,
of f on I.
Consider Figure . . The func on displayed in (a) has a maximum, but no
minimum, as the interval over which the func on is dened is open. In (b), the
func on has a minimum, but no maximum; there is a discon nuity in the natural place for the maximum to occur. Finally, the func on shown in (c) has both a
maximum and a minimum; note that the func on is con nuous and the interval
on which it is dened is closed.
It is possible for discon nuous func ons dened on an open interval to have
both a maximum and minimum value, but we have just seen examples where
they did not. On the other hand, con nuous func ons on a closed interval always have a maximum and minimum value.

(c)
Figure . : Graphs of func ons with and
without extreme values.

Note: The extreme values of a func on


are y values, values the func on a ains,
not the input values.

Chapter

The Graphical Behavior of Func ons

Theorem

The Extreme Value Theorem

Let f be a con nuous func on dened on a closed interval I. Then f has


both a maximum and minimum value on I.
This theorem states that f has extreme values, but it does not oer any advice about how/where to nd these values. The process can seem to be fairly
easy, as the next example illustrates. A er the example, we will draw on lessons
learned to form a more general and powerful method for nding extreme values.
y

Approxima ng extreme values


Example
Consider f(x) = x x on I = [ , ], as graphed in Figure . . Approximate
the extreme values of f.

(5, 5)

( , )

( ,

)
( , 7)

.
Figure . : A graph of f(x) = x x as
in Example .

The graph is drawn in such a way to draw a en on to certain


S
points. It certainly seems that the smallest y value is , found when x = .
It also seems that the largest y value is , found at the endpoint of I, x = .
We use the word seems, for by the graph alone we cannot be sure the smallest
value is not less than . Since the problem asks for an approxima on, we
approximate the extreme values to be and .
No ce how the minimum value came at the bo om of a hill, and the maximum value came at an endpoint. Also note that while is not an extreme value,
it would be if we narrowed our interval to [ , ]. The idea that the point ( , )
is the loca on of an extreme value for some interval is important, leading us to
a deni on.
Deni on

Rela ve Minimum and Rela ve Maximum

Let f be dened on an interval I containing c.


. If there is an open interval containing c such that f(c) is the minimum value, then f(c) is a rela ve minimum of f. We also say that
f has a rela ve minimum at (c, f(c)).
Note: The terms local minimum and local
maximum are o en used as synonyms for
rela ve minimum and rela ve maximum.

. If there is an open interval containing c such that f(c) is the maximum value, then f(c) is a rela ve maximum of f. We also say that
f has a rela ve maximum at (c, f(c)).
The rela ve maximum and minimum values comprise the rela ve extrema of f.

Notes:

Extreme Values

We briey prac ce using these deni ons.


Approxima ng rela ve extrema
Example
Consider f(x) = ( x x x + )/ , as shown in Figure . . Approximate
the rela ve extrema of f. At each of these points, evaluate f .
We s ll do not have the tools to exactly nd the rela ve
S
extrema, but the graph does allow us to make reasonable approxima ons. It
seems f has rela ve minima at x = and x = , with values of f( ) = and
f( ) = . . It also seems that f has a rela ve maximum at the point ( , ).

We approximate the rela ve minima to be and . ; we approximate the


rela ve maximum to be .

It is straigh orward to evaluate f (x) = (


and . In each case, f (x) = .

x) at x = ,

y
6

Figure . : A graph of f(x) = ( x x


x + )/ as in Example .

Approxima ng rela ve extrema


Example
Approximate the rela ve extrema of f(x) = (x ) / + , shown in Figure . .
At each of these points, evaluate f .

The gure implies that f does not have any rela ve maxima,
S
but has a rela ve minimum at ( , ). In fact, the graph suggests that not only is
this point a rela ve minimum, y = f( ) = the minimum value of the func on.
We compute f (x) = (x )

. When x = , f is undened.

What can we learn from the previous two examples? We were able to visually approximate rela ve extrema, and at each such point, the deriva ve was
either or it was not dened. This observa on holds for all func ons, leading
to a deni on and a theorem.
Deni on

Cri cal Numbers and Cri cal Points

Let f be dened at c. The value c is a cri cal number (or cri cal value)
of f if f (c) = or f (c) is not dened.
If c is a cri cal number of f, then the point (c, f(c)) is a cri cal point of f.

Notes:

.
Figure . : A graph of f(x) = (x )
as in Example .

Chapter

The Graphical Behavior of Func ons


y

Theorem

Rela ve Extrema and Cri cal Points

Let a func on f have a rela ve extrema at the point (c, f(c)). Then c is a
cri cal number of f.
1

Figure . : A graph of f(x) = x which has


a cri cal value of x = , but no rela ve
extrema.

Be careful to understand that this theorem states All rela ve extrema occur
at cri cal points. It does not say All cri cal numbers produce rela ve extrema.
For instance, consider f(x) = x . Since f (x) = x , it is straigh orward to determine that x = is a cri cal number of f. However, f has no rela ve extrema,
as illustrated in Figure . .
Theorem states that a con nuous func on on a closed interval will have
absolute extrema, that is, both an absolute maximum and an absolute minimum.
These extrema occur either at the endpoints or at cri cal values in the interval.
We combine these concepts to oer a strategy for nding extrema.
Key Idea

Finding Extrema on a Closed Interval

Let f be a con nuous func on dened on a closed interval [a, b]. To nd


the maximum and minimum values of f on [a, b]:
. Evaluate f at the endpoints a and b of the interval.
. Find the cri cal numbers of f in [a, b].
. Evaluate f at each cri cal number.
. The absolute maximum of f is the largest of these values, and the
absolute minimum of f is the least of these values.
y

We prac ce these ideas in the next examples.


Finding extreme values
Example
Find the extreme values of f(x) = x + x
. .
S

.
Figure . : A graph of f(x) = x + x
x on [ , ] as in Example .

f at the endpoints:

x on [ , ], graphed in Figure

We follow the steps outlined in Key Idea . We rst evaluate


f( ) =

and

f( ) =

Next, we nd the cri cal values of f on [ , ]. f (x) = x + x


= (x +
)(x ); therefore the cri cal values of f are x = and x = . Since x =

Notes:

.
does not lie in the interval [ , ], we ignore it. Evalua ng f at the only cri cal
number in our interval gives: f( ) = .
The table in Figure . gives f evaluated at the important x values in [ , ].
We can easily see the maximum and minimum values of f: the maximum value
is and the minimum value is .
Note that all this was done without the aid of a graph; this work followed
an analy c algorithm and did not depend on any visualiza on. Figure . shows
f and we can conrm our answer, but it is important to understand that these
answers can be found without graphical assistance.
We prac ce again.

Figure . : Finding the extreme values of


f in Example .

f( ) = .

Figure . : Finding the extreme values of


f in Example .

We now nd the cri cal numbers of f. We have to dene f in a piecewise


manner; it is
{
(x ) x <
f (x) =
.
x>

Note that while f is dened for all of [ , ], f is not, as the deriva ve of f does
not exist when x = . (From the le , the deriva ve approaches ; from the
right the deriva ve is .) Thus one cri cal number of f is x = .
We now set f (x) = . When x > , f (x) is never . When x < , f (x) is
also never . (We may be tempted to say that f (x) = when x = . However,
this is nonsensical, for we only consider f (x) = (x ) when x < , so we will
ignore a solu on that says x = .)
So we have three important x values to consider: x = , and . Evalua ng f at each gives, respec vely, , and , shown in Figure . . Thus the
absolute minimum of f is ; the absolute maximum of f is . Our answer is conrmed by the graph of f in Figure . .

Notes:

f(x)

Here f is piecewisedened, but we can s ll apply Key Idea


. Evalua ng f at the endpoints gives:
S

and

f(x)

Finding extreme values


Example
Find the maximum and minimum values of f on [ , ], where
{
(x ) x
f(x) =
.
x+
x>

f( ) =

Extreme Values

x
4

Figure . : A graph of f(x) on [ , ] as in


Example .

Chapter

The Graphical Behavior of Func ons


x

f(x)

Finding extreme values


Example
Find the extrema of f(x) = cos(x ) on [ , ].

Figure . : Finding the extrema of


f(x) = cos(x ) in Example .
y

.5
x

.5

We consider one more example.

Figure . : A graph of f(x) = cos(x ) on


[ , ] as in Example .
x

f(x)

Figure . : Finding the extrema of the


halfcircle in Example .
y
1

We again use Key Idea . Evalua ng f at the endpoints of


S
the interval gives: f( ) = f( ) = cos( ) .
. We now nd the cri cal
values of f.
Applying the Chain Rule, we nd f (x) = x sin(x ). Set f (x) = and
solve for x to nd the cri cal values of f.
We have f (x) = when x = and when sin(x ) = . In general, sin t =
when t = . . . , , , , . . . Thus sin(x ) = when x = , , , . . .
(x is
always posi ve so we ignore , etc.) So sin(x ) = when x= , , , . . ..
The only values to fall in the given interval of [ , ] are and , approximately . .
We again construct a table of important values
in Figure . . In this example

we have values to consider: x = , , .


From the table it is clear that the maximum value of f on [ , ] is ; the
minimum value is . The graph in Figure . conrms our results.

Figure . : A graph of f(x) =


on [ , ] as in Example .

Note: We implicitly found the deriva ve


of x + y = , the unit circle, in Exam= x/y. In Example , half
ple as dy
dx
of
the
unit
circle
is given as y = f(x) =

x . We found f (x) = x . Recx

ognize that the denominator of this fracon is y; that is, we again found f (x) =
dy
= x/y.
dx

Finding extreme
Example
values
Find the extreme values of f(x) =
x .
A closed interval is not given, so we nd the extreme values
S
of f on its domain. f is dened whenever x ; thus the domain of f is
[ , ]. Evalua ng f at either endpoint returns .
x
. The cri cal points of f are
Using the Chain Rule, we nd f (x) =
x

found when f (x) = or when f is undened. It is straigh orward to nd that


f (x) = when x = , and f is undened when x = , the endpoints of the
interval. The table of important values is given in Figure . . The maximum
value is , and the minimum value is .
We have seen that con nuous func ons on closed intervals always have a
maximum and minimum value, and we have also developed a technique to nd
these values. In the next sec on, we further our study of the informa on we can
glean from nice func ons with the Mean Value Theorem. On a closed interval,
we can nd the average rate of change of a func on (as we did at the beginning
of Chapter ). We will see that dieren able func ons always have a point at
which their instantaneous rate of change is same as the average rate of change.
This is surprisingly useful, as well see.

Notes:

Exercises .

Terms and Concepts

. f(x) = x

x
y

. Describe what an extreme value of a func on is in your


own words.

(2, 4 2)

. Sketch the graph of a func on f on ( , ) that has both a


maximum and minimum value.
. Describe the dierence between absolute and rela ve
maxima in your own words.
. Sketch the graph of a func on f where f has a rela ve maximum at x = and f ( ) is undened.

(0, 0)

. f(x) = sin x
y
(/ , )

. T/F: If c is a cri cal value of a func on f, then f has either a


rela ve maximum or rela ve minimum at x = c.

Problems

In Exercises , iden fy each of the marked points as being


an absolute maximum or minimum, a rela ve maximum or
minimum, or none of the above.

( / , )

. f(x) = x

B
G

( , )

. f(x) =

y
C

2
B

x
x

x
x>
y

( , )

0.5
E

A
2

In Exercises
the graph.
. f(x) =

, evaluate f (x) at the points indicated in

. f(x) =
x +

(0, 0) 0.5

0.5

0.5

x
x

x
x>
y
1

( , )
0.5

(0, 0) 0.5

0.5

0.5

. f(x) =
y

(x )
x

, +

. f(x) =

[ , ].

. f(x) = ex cos x

on

[ , ].

. f(x) = ex sin x

on

[ , ].

ln x
x

. f(x) = x

In Exercises
, nd the extreme values of the func on
on the given interval.
. f(x) = x + x +
. f(x) = x

. f(x) = sin x

on

. f(x) = x +

on

[ , ].

. f(x) = x

on

. f(x) =

( , )

x
x +

x
x

on

on

[ , ].

x+

on

[/ , / ].
on

[ , ].

[ , ].

[ , ].

on

[ , ].

Review
. Find

dy
,
dx

where x y y x = .

. Find the equa on of the line tangent to the graph of x +


y + xy = at the point ( , ).
. Let f(x) = x + x.
Evaluate lim
s

f(x + s) f(x)
.
s

The Mean Value Theorem

We mo vate this sec on with the following ques on: Suppose you leave your
house and drive to your friends house in a city
miles away, comple ng the
trip in two hours. At any point during the trip do you necessarily have to be going
miles per hour?
In answering this ques on, it is clear that the average speed for the en re
trip is mph (i.e.
miles in hours), but the ques on is whether or not your
instantaneous speed is ever exactly mph. More simply, does your speedometer ever read exactly
mph?. The answer, under some very reasonable assump ons, is yes.
Lets now see why this situa on is in a calculus text by transla ng it into
mathema cal symbols.
First assume that the func on y = f(t) gives the distance (in miles) traveled
from your home at me t (in hours) where t . In par cular, this gives
f( ) = and f( ) =
. The slope of the secant line connec ng the star ng
and ending points ( , f( )) and ( , f( )) is therefore
f
f( ) f( )
=
=
t

mph.

The slope at any point on the graph itself is given by the deriva ve f (t). So,
since the answer to the ques on above is yes, this means that at some me
during the trip, the deriva ve takes on the value of mph. Symbolically,
f (c) =
for some me

c .

f( ) f( )
=

How about more generally? Given any func on y = f(x) and a range a
x b does the value of the deriva ve at some point between a and b have to
match the slope of the secant line connec ng the points (a, f(a)) and (b, f(b))?
Or equivalently, does the equa on f (c) = f(b)f(a)
have to hold for some a <
ba
c < b?
Lets look at two func ons in an example.
Comparing average and instantaneous rates of change
Example
Consider func ons
f (x) =
and f (x) = |x|
x
with a = and b = as shown in Figure . (a) and (b), respec vely. Both
func ons have a value of at a and b. Therefore the slope of the secant line

Notes:

The Mean Value Theorem

Chapter

The Graphical Behavior of Func ons


connec ng the end points is in each case. But if you look at the plots of each,
you can see that there are no points on either graph where the tangent lines
have slope zero. Therefore we have found that there is no c in [ , ] such that

f (c) =

..

(a)
y

The Mean Value Theorem of Dieren a on

Let y = f(x) be con nuous func on on the closed interval [a, b] and
dieren able on the open interval (a, b). There exists a value c, a < c <
b, such that
f(b) f(a)
f (c) =
.
ba

0.5

So what went wrong? It may not be surprising to nd that the discon nuity
of f and the corner of f play a role. If our func ons had been con nuous and
dieren able, would we have been able to nd that special value c? This is our
mo va on for the following theorem.
Theorem

f( ) f( )
= .
( )

That is, there is a value c in (a, b) where the instantaneous rate of change
of f at c is equal to the average rate of change of f on [a, b].

(b)
Figure . : A graph of f (x) = /x and
f (x) = |x| in Example .

Note that the reasons that the func ons in Example


fail are indeed that
f has a discon nuity on the interval [ , ] and f is not dieren able at the
origin.
We will give a proof of the Mean Value Theorem below. To do so, we use a
fact, called Rolles Theorem, stated here.

Theorem

Let f be con nuous on [a, b] and dieren able on (a, b), where f(a) =
f(b). There is some c in (a, b) such that f (c) = .
c

1 b

Figure . : A graph of f(x) = x x +


x + , where f(a) = f(b). Note the existence of c, where a < c < b, where
f (c) = .

Rolles Theorem

Consider Figure . where the graph of a func on f is given, where f(a) =


f(b). It should make intui ve sense that if f is dieren able (and hence, con nuous) that there would be a value c in (a, b) where f (c) = ; that is, there would
be a rela ve maximum or minimum of f in (a, b). Rolles Theorem guarantees at
least one; there may be more.

Notes:

.
Rolles Theorem is really just a special case of the Mean Value Theorem. If
f(a) = f(b), then the average rate of change on (a, b) is , and the theorem
guarantees some c where f (c) = . We will prove Rolles Theorem, then use it
to prove the Mean Value Theorem.
Proof of Rolles Theorem
Let f be dieren able on (a, b) where f(a) = f(b). We consider two cases.
Case : Consider the case when f is constant on [a, b]; that is, f(x) = f(a) = f(b)
for all x in [a, b]. Then f (x) = for all x in [a, b], showing there is at least one
value c in (a, b) where f (c) = .
Case : Now assume that f is not constant on [a, b]. The Extreme Value Theorem
guarantees that f has a maximal and minimal value on [a, b], found either at the
endpoints or at a cri cal value in (a, b). Since f(a) = f(b) and f is not constant, it
is clear that the maximum and minimum cannot both be found at the endpoints.
Assume, without loss of generality, that the maximum of f is not found at the
endpoints. Therefore there is a c in (a, b) such that f(c) is the maximum value
of f. By Theorem , c must be a cri cal number of f; since f is dieren able, we
have that f (c) = , comple ng the proof of the theorem.

We can now prove the Mean Value Theorem.


Proof of the Mean Value Theorem
Dene the func on
g(x) = f(x)

f(b) f(a)
x.
ba

We know g is dieren able on (a, b) and con nuous on [a, b] since f is. We can
show g(a) = g(b) (it is actually easier to show g(b) g(a) = , which suces).
We can then apply Rolles theorem to guarantee the existence of c (a, b) such
that g (c) = . But note that
= g (c) = f (c)
hence
f (c) =
which is what we sought to prove.

f(b) f(a)
;
ba

f(b) f(a)
,
ba

Going back to the very beginning of the sec on, we see that the only assump on we would need about our distance func on f(t) is that it be con nuous and dieren able for t from to hours (both reasonable assump ons). By
the Mean Value Theorem, we are guaranteed a me during the trip where our

Notes:

The Mean Value Theorem

Chapter

The Graphical Behavior of Func ons


instantaneous speed is mph. This fact is used in prac ce. Some law enforcement agencies monitor trac speeds while in aircra . They do not measure
speed with radar, but rather by ming individual cars as they pass over lines
painted on the highway whose distances apart are known. The ocer is able
to measure the average speed of a car between the painted lines; if that average speed is greater than the posted speed limit, the ocer is assured that the
driver exceeded the speed limit at some me.
Note that the Mean Value Theorem is an existence theorem. It states that a
special value c exists, but it does not give any indica on about how to nd it. It
turns out that when we need the Mean Value Theorem, existence is all we need.
Using the Mean Value Theorem
Example
Consider f(x) = x + x + on [ , ]. Find c in [ , ] that sa ses the Mean
Value Theorem.
S

The average rate of change of f on [ , ] is:


f( ) f( )
=
( )

We want to nd c such that f (c) =


equal to and solve for x.

x +

x=

Figure . : Demonstra ng the Mean


Value Theorem in Example .

. We nd f (x) = x + . We set this

x =
.

f (x) =

We have found values c in [ , ] where the instantaneous rate of change


is equal to the average rate of change; the Mean Value Theorem guaranteed at
least one. In Figure . f is graphed with a dashed
line represen ng the averare also given. Note how
age rate of change; the lines tangent to f at x =
these lines are parallel (i.e., have the same slope) as the dashed line.
While the Mean Value Theorem has prac cal use (for instance, the speed
monitoring applica on men oned before), it is mostly used to advance other
theory. We will use it in the next sec on to relate the shape of a graph to its
deriva ve.

Notes:

Exercises .

Terms and Concepts


. Explain in your own words what the Mean Value Theorem
states.
. Explain in your own words what Rolles Theorem states.

Problems
In Exercises , a func on f(x) and interval [a, b] are given.
Check if Rolles Theorem can be applied to f on [a, b]; if so, nd
c in [a, b] such that f (c) = .
. f(x) =

. f(x) = x + x on [ , ].
. f(x) = x + x on [ , ].
. f(x) = x + x on [ , ].
. f(x) = sin x on [/ , / ].
. f(x) = cos x on [ , ].

x x+

. f(x) = x + x on [ , ].
. f(x) = x x + on [ , ].
. f(x) =

. f(x) =

. f(x) =

x
x

x on [ , ].
x on [ , ].
on [ , ].

. f(x) = ln x on [ , ].

on [ , ].

. f(x) = x on [ , ].

. f(x) =

In Exercises
, a func on f(x) and interval [a, b] are
given. Check if the Mean Value Theorem can be applied to f
on [a, b]; if so, nd a value c in [a, b] guaranteed by the Mean
Value Theorem.

on [ , ].

. f(x) = tan x on [/ , / ].
. f(x) = x x + x + on [ , ].
. f(x) = x x + x + on [ , ].
. f(x) = sin x on [ , ].

Review
. Find the extreme values of f(x) = x x + on [ , ].
. Describe the cri cal points of f(x) = cos x.
. Describe the cri cal points of f(x) = tan x.

Chapter

The Graphical Behavior of Func ons

Our study of nice func ons f in this chapter has so far focused on individual
points: points where f is maximal/minimal, points where f (x) = or f does
not exist, and points c where f (c) is the average rate of change of f on some
interval.
In this sec on we begin to study how func ons behave between special
points; we begin studying in more detail the shape of their graphs.
We start with an intui ve concept. Given the graph in Figure . , where
would you say the func on is increasing? Decreasing? Even though we have
not dened these terms mathema cally, one likely answered that f is increasing
when x > and decreasing when x < . We formally dene these terms here.

Increasing and Decreasing Func ons

Figure . : A graph of a func on f used


to illustrate the concepts of increasing
and decreasing.

Deni on

Increasing and Decreasing Func ons

Let f be a func on dened on an interval I.


. f is increasing on I if for every a < b in I, f(a) f(b).
. f is decreasing on I if for every a < b in I, f(a) f(b).
A func on is strictly increasing when a < b in I implies f(a) < f(b), with
a similar deni on holding for strictly decreasing.
Informally, a func on is increasing if as x gets larger (i.e., looking le to right)
f(x) gets larger.
Our interest lies in nding intervals in the domain of f on which f is either
increasing or decreasing. Such informa on should seem useful. For instance, if
f describes the speed of an object, we might want to know when the speed was
increasing or decreasing (i.e., when the object was accelera ng vs. decelerating). If f describes the popula on of a city, we should be interested in when the
popula on is growing or declining.
To nd such intervals, we again consider secant lines. Let f be an increasing,
dieren able func on on an open interval I, such as the one shown in Figure
. , and let a < b be given in I. The secant line on the graph of f from x = a to
x = b is drawn; it has a slope of (f(b) f(a))/(b a). But note:

y
2

(b, f(b))

(a, f(a))

Figure . : Examining the secant line of


an increasing func on.

f(b) f(a)
numerator >

ba
denominator >

slope of the
secant line >

Average rate of
change of f on
[a, b] is > .

We have shown mathema cally what may have already been obvious: when
f is increasing, its secant lines will have a posi ve slope. Now recall the Mean

Notes:

Increasing and Decreasing Func ons

Value Theorem guarantees that there is a number c, where a < c < b, such that
f (c) =

f(b) f(a)
> .
ba

By considering all such secant lines in I, we strongly imply that f (x) on I. A


similar statement can be made for decreasing func ons.
Our above logic can be summarized as If f is increasing, then f is probably
posi ve. Theorem below turns this around by sta ng If f is pos ve, then f is
increasing. This leads us to a method for nding when func ons are increasing
and decreasing.
Theorem

Test For Increasing/Decreasing Func ons

Let f be a con nuous func on on [a, b] and dieren able on (a, b).
. If f (c) >

for all c in (a, b), then f is increasing on [a, b].

. If f (c) <

for all c in (a, b), then f is decreasing on [a, b].

. If f (c) =

for all c in (a, b), then f is constant on [a, b].

Let a and b be in I where f (a) > and f (b) < . It follows from the
Intermediate Value Theorem that there must be some value c between a and b
where f (c) = . This leads us to the following method for nding intervals on
which a func on is increasing or decreasing.
Key Idea

Finding Intervals on Which f is Increasing or Decreasing

Let f be a dieren able func on on an interval I. To nd intervals on


which f is increasing and decreasing:
. Find the cri cal values of f. That is, nd all c in I where f (c) =
or f is not dened.
. Use the cri cal values to divide I into subintervals.
. Pick any point p in each subinterval, and nd the sign of f (p).
(a) If f (p) > , then f is increasing on that subinterval.
(b) If f (p) < , then f is decreasing on that subinterval.

Notes:

Note: Theorem also holds if f (c) =


for a nite number of values of c in I.

Chapter

The Graphical Behavior of Func ons


We demonstrate using this process in the following example.
Finding intervals of increasing/decreasing
Example
Let f(x) = x + x x + . Find intervals on which f is increasing or decreasing.
Using Key Idea , we rst nd the cri cal values of f. We
S
have f (x) = x + x = ( x )(x + ), so f (x) = when x = and
when x = / . f is never undened.
Since an interval was not specied for us to consider, we consider the enre domain of f which is (, ). We thus break the whole real line into
three subintervals based on the two cri cal values we just found: (, ),
( , / ) and ( / , ). This is shown in Figure . .
f > 0 incr

f < 0 decr

f > 0 incr

.
1

Figure .
y

f (x)

5
f(x)
x

.
Figure . : A graph of f(x) in Example
, showing where f is increasing and decreasing.

1/3

: Number line for f in Example

We now pick a value p in each subinterval and nd the sign of f (p). All we
care about is the sign, so we do not actually have to fully compute f (p); pick
nice values that make this simple.
Subinterval , (, ): We (arbitrarily) pick p = . We can compute
f ( ) directly: f ( ) = ( ) + ( ) = > . We conclude that f is
increasing on (, ).
Note we can arrive at the same conclusion without computa on. For instance, we could choose p =
. The rst term in f (
), i.e., (
) is
clearly posi ve and very large. The other terms are small in comparison, so we
know f (
) > . All we need is the sign.
Subinterval , ( , / ): We pick p = since that value seems easy to deal
with. f ( ) = < . We conclude f is decreasing on ( , / ).
Subinterval , ( / , ): Pick an arbitrarily large value for p > / and note
that f (p) = p + p > . We conclude that f is increasing on ( / , ).
We can verify our calcula ons by considering Figure . , where f is graphed.
The graph also presents f ; note how f > when f is increasing and f <
when f is decreasing.
One is jus ed in wondering why so much work is done when the graph
seems to make the intervals very clear. We give three reasons why the above
work is worthwhile.
First, the points at which f switches from increasing to decreasing are not
precisely known given a graph. The graph shows us something signicant hapNotes:

.
pens near x = and x = . , but we cannot determine exactly where from
the graph.
One could argue that just nding cri cal values is important; once we know
the signicant points are x = and x = / , the graph shows the increasing/decreasing traits just ne. That is true. However, the technique prescribed
here helps reinforce the rela onship between increasing/decreasing and the
sign of f . Once mastery of this concept (and several others) is obtained, one
nds that either (a) just the cri cal points are computed and the graph shows
all else that is desired, or (b) a graph is never produced, because determining
increasing/decreasing using f is straigh orward and the graph is unnecessary.
So our second reason why the above work is worthwhile is this: once mastery
of a subject is gained, one has op ons for nding needed informa on. We are
working to develop mastery.
Finally, our third reason: many problems we face in the real world are very
complex. Solu ons are tractable only through the use of computers to do many
calcula ons for us. Computers do not solve problems on their own, however;
they need to be taught (i.e., programmed) to do the right things. It would be
benecial to give a func on to a computer and have it return maximum and
minimum values, intervals on which the func on is increasing and decreasing,
the loca ons of rela ve maxima, etc. The work that we are doing here is easily
programmable. It is hard to teach a computer to look at the graph and see if it
is going up or down. It is easy to teach a computer to determine if a number
is greater than or less than .
In Sec on . we learned the deni on of rela ve maxima and minima and
found that they occur at cri cal points. We are now learning that func ons can
switch from increasing to decreasing (and viceversa) at cri cal points. This new
understanding of increasing and decreasing creates a great method of determining whether a cri cal point corresponds to a maximum, minimum, or neither.
Imagine a func on increasing un l a cri cal point at x = c, a er which it decreases. A quick sketch helps conrm that f(c) must be a rela ve maximum. A
similar statement can be made for rela ve minimums. We formalize this concept in a theorem.
Theorem

First Deriva ve Test

Let f be dieren able on I and let c be a cri cal number in I.


. If the sign of f switches from posi ve to nega ve at c, then f(c) is
a rela ve maximum of f.
. If the sign of f switches from nega ve to posi ve at c, then f(c) is
a rela ve minimum of f.
. If the sign of f does not change at c, then f(c) is not a rela ve
extrema of f.
Notes:

Increasing and Decreasing Func ons

Chapter

The Graphical Behavior of Func ons

Using the First Deriva ve Test


Example
Find the intervals on which f is increasing and decreasing, and use the First
Deriva ve Test to determine the rela ve extrema of f, where
f(x) =

x +
.
x

We start by no ng the domain of f: (, ) ( , ). Key


S
Idea describes how to nd intervals where f is increasing and decreasing when
the domain of f is an interval. Since the domain of f in this example is the union
of two intervals, we apply the techniques of Key Idea to both intervals of the
domain of f.
Since f is not dened at x = , the increasing/decreasing nature of f could
switch at this value. We do not formally consider x = to be a cri cal value of
f, but we will include it in our list of cri cal values that we nd next.
Using the Quo ent Rule, we nd
f (x) =

x x
.
(x )

We need to nd the cri cal values of f; we want to know when f (x) = and
when f is not dened. That la er is straigh orward: when the denominator
of f (x) is , f is undened. That occurs when x = , which weve already
recognized as an important value.
f (x) = when the numerator of f (x) is . That occurs when x x =
(x )(x + ) = ; i.e., when x = , .
We have found that f has two cri cal numbers, x = , , and at x =
something important might also happen. These three numbers divide the real
number line into subintervals:
(, ),

( , ),

( , )

and

( , ).

Pick a number p from each subinterval and test the sign of f at p to determine
whether f is increasing or decreasing on that interval. Again, we do well to avoid
complicated computa ons; no ce that the denominator of f is always posi ve
so we can ignore it during our work.
Interval , (, ): Choosing a very small number (i.e., a nega ve number
with a large magnitude) p returns p p in the numerator of f ; that will
be posi ve. Hence f is increasing on (, ).
Interval , ( , ): Choosing seems simple: f ( ) = < . We conclude
f is decreasing on ( , ).

Notes:

.
Interval , ( , ): Choosing seems simple: f ( ) = < . Again, f is
decreasing.
Interval , ( , ): Choosing an very large number p from this subinterval will
give a posi ve numerator and (of course) a posi ve denominator. So f is increasing on ( , ).
In summary, f is increasing on the set (, )( , ) and is decreasing on
the set ( , ) ( , ). Since at x = , the sign of f switched from posi ve to
nega ve, Theorem states that f( ) is a rela ve maximum of f. At x = , the
sign of f switched from nega ve to posi ve, meaning f( ) is a rela ve minimum.
At x = , f is not dened, so there is no rela ve extrema at x = .
f > 0 incr

rel.
max

f < 0 decr

f < 0 decr

rel.
min

f > 0 incr

.
1

Figure .

: Number line for f in Example

This is summarized in the number line shown in Figure . . Also, Figure .


shows a graph of f, conrming our calcula ons. This gure also shows f , again
demonstra ng that f is increasing when f > and decreasing when f < .
One is o en tempted to think that func ons always alternate increasing,
decreasing, increasing, decreasing,. . . around cri cal values. Our previous example demonstrated that this is not always the case. While x = was not
technically a cri cal value, it was an important value we needed to consider.
We found that f was decreasing on both sides of x = .
We examine one more example.
Using the First Deriva ve Test
Example
Find the intervals on which f(x) = x / x / is increasing and decreasing and
iden fy the rela ve extrema.
We again start with taking deriva ves. Since we know we
S
want to solve f (x) = , we will do some algebra a er taking deriva ves.
f(x) = x x
f (x) =
=
=
=
Notes:

x x
)
(
x x
x (x )

x (x )(x + ).

Increasing and Decreasing Func ons


y
f(x)

f (x)
4

Figure . : A graph of f(x) in Example


, showing where f is increasing and decreasing.

Chapter

The Graphical Behavior of Func ons


This deriva on of f shows that f (x) = when x = and f is not dened when x = . Thus we have cri cal values, breaking the number line into
subintervals as shown in Figure . .
Interval , (, ): We choose p = ; we can easily verify that f ( ) < .
So f is decreasing on (, ).
Interval , ( , ): Choose p = / . Once more we prac ce nding the sign
of f (p) without compu ng an actual value. We have f (p) = ( / )p / (p
)(p + ); nd the sign of each of the three terms.
f (p) =

p (p ) (p + ) .
|{z} | {z } | {z }
<

<

>

We have a nega ve nega ve posi ve giving a posi ve number; f is increasing on ( , ).


Interval , ( , ): We do a similar sign analysis as before, using p in ( , ).
f (p) =

p (p ) (p + ) .
|{z} | {z } | {z }
>

<

>

We have posi ve factors and one nega ve factor; f (p) < and so f is decreasing on ( , ).
Interval , ( , ): Similar work to that done for the other three intervals shows
that f (x) > on ( , ), so f is increasing on this interval.
f < 0 decr

rel.
min

f > 0 incr

rel.
max

f < 0 decr

rel.
min

f > 0 incr

.
y
1

Figure .
f(x)

f (x)

.
Figure . : A graph of f(x) in Example
, showing where f is increasing and decreasing.

: Number line for f in Example

We conclude by sta ng that f is increasing on ( , )( , ) and decreasing


on (, ) ( , ). The sign of f changes from nega ve to posi ve around
x = and x = , meaning by Theorem
that f( ) and f( ) are rela ve
minima of f. As the sign of f changes from posi ve to nega ve at x = , we
have a rela ve maximum at f( ). Figure . shows a graph of f, conrming our
result. We also graph f , highligh ng once more that f is increasing when f >
and is decreasing when f < .
We have seen how the rst deriva ve of a func on helps determine when
the func on is going up or down. In the next sec on, we will see how the
second deriva ve helps determine how the graph of a func on curves.

.
Notes:

Exercises .

Terms and Concepts


. In your own words describe what it means for a func on to
be increasing.

In Exercises

, a func on f(x) is given.

(a) Give the domain of f.


(b) Find the cri cal numbers of f.

. What does a decreasing func on look like?

(c) Create a number line to determine the intervals on


which f is increasing and decreasing.

. Sketch a graph of a func on on [ , ] that is increasing but


not strictly increasing.

(d) Use the First Deriva ve Test to determine whether


each cri cal point is a rela ve maximum, minimum,
or neither.

. Give an example of a func on describing a situa on where


it is bad to be increasing and good to be decreasing.
. A func on f has deriva ve f (x) = (sin x + )ex + , where
f (x) > for all x. Is f increasing, decreasing, or can we not
tell from the given informa on?

. f(x) = x + x +
. f(x) = x + x x +
. f(x) = x x + x

Problems
In Exercises

. f(x) = x + x

. f(x) =
, a func on f(x) is given.

x x+

(a) Compute f (x).

. f(x) =

(b) Graph f and f on the same axes (using technology is


permi ed) and verify Theorem .

x
x

. f(x) =

x
x x

. f(x) =

(x )
x

. f(x) = x +
. f(x) = x x +
. f(x) = cos x
. f(x) = tan x
. f(x) = x x + x
. f(x) = x x + x
. f(x) = x x +
. f(x) =

x +

. f(x) = sin x cos x on (, ).


. f(x) = x x

Review
. Consider f(x) = x x + on [ , ]; nd c guaranteed
by the Mean Value Theorem.
. Consider f(x) = sin x on [/ , / ]; nd c guaranteed by
the Mean Value Theorem.

Chapter

The Graphical Behavior of Func ons


y

Figure . : A func on f with a concave


up graph. No ce how the slopes of the
tangent lines, when looking from le to
right, are increasing.

Concavity and the Second Deriva ve

Our study of nice func ons con nues. The previous sec on showed how the
rst deriva ve of a func on, f , can relay important informa on about f. We
now apply the same technique to f itself, and learn what this tells us about f.
The key to studying f is to consider its deriva ve, namely f , which is the
second deriva ve of f. When f > , f is increasing. When f < , f is
decreasing. f has rela ve maxima and minima where f = or is undened.
This sec on explores how knowing informa on about f gives informa on
about f.

Concavity
We begin with a deni on, then explore its meaning.
Deni on

Note: We o en state that f is concave


up instead of the graph of f is concave
up for simplicity.

Note: A mnemonic for remembering


what concave up/down means is: Concave up is like a cup; concave down is like
a frown. It is admi edly terrible, but it
works.

Figure . : A func on f with a concave


down graph. No ce how the slopes of the
tangent lines, when looking from le to
right, are decreasing.

Concave Up and Concave Down

Let f be dieren able on an interval I. The graph of f is concave up on I


if f is increasing. The graph of f is concave down on I if f is decreasing.
If f is constant then the graph of f is said to have no concavity.
The graph of a func on f is concave up when f is increasing. That means
as one looks at a concave up graph from le to right, the slopes of the tangent
lines will be increasing. Consider Figure . , where a concave up graph is shown
along with some tangent lines. No ce how the tangent line on the le is steep,
downward, corresponding to a small value of f . On the right, the tangent line
is steep, upward, corresponding to a large value of f .
If a func on is decreasing and concave up, then its rate of decrease is slowing; it is leveling o. If the func on is increasing and concave up, then the rate
of increase is increasing. The func on is increasing at a faster and faster rate.
Now consider a func on which is concave down. We essen ally repeat the
above paragraphs with slight varia on.
The graph of a func on f is concave down when f is decreasing. That means
as one looks at a concave down graph from le to right, the slopes of the tangent
lines will be decreasing. Consider Figure . , where a concave down graph is
shown along with some tangent lines. No ce how the tangent line on the le
is steep, upward, corresponding to a large value of f . On the right, the tangent
line is steep, downward, corresponding to a small value of f .
If a func on is increasing and concave down, then its rate of increase is slowing; it is leveling o. If the func on is decreasing and concave down, then the
rate of decrease is decreasing. The func on is decreasing at a faster and faster
rate.

Notes:

Concavity and the Second Deriva ve

Our deni on of concave up and concave down is given in terms of when


the rst deriva ve is increasing or decreasing. We can apply the results of the
previous sec on and to nd intervals on which a graph is concave up or down.
That is, we recognize that f is increasing when f > , etc.
Theorem

f < 0, decreasing
f < 0, c. down

Test for Concavity

Let f be twice dieren able on an interval I. The graph of f is concave up


if f > on I, and is concave down if f < on I.
If knowing where a graph is concave up/down is important, it makes sense
that the places where the graph changes from one to the other is also important.
This leads us to a deni on.
Deni on

f > 0, increasing
f < 0, c. down

Point of Inec on

f < 0, decreasing

f > 0, increasing

f > 0, c. up

f > 0, c. up

Figure . : Demonstra ng the ways


that concavity interacts with increasing/decreasing, along with the rela onships with the rst and second derivaves.

A point of inec on is a point on the graph of f at which the concavity


of f changes.
Figure . shows a graph of a func on with inec on points labeled.
If the concavity of f changes at a point (c, f(c)), then f is changing from
increasing to decreasing (or, decreasing to increasing) at x = c. That means that
the sign of f is changing from posi ve to nega ve (or, nega ve to posi ve) at
x = c. This leads to the following theorem.
Theorem

Points of Inec on

If (c, f(c)) is a point of inec on on the graph of f, then either f =


f is not dened at c.

or

We have iden ed the concepts of concavity and points of inec on. It is


now me to prac ce using these concepts; given a func on, we should be able
to nd its points of inec on and iden fy intervals on which it is concave up or
down. We do so in the following examples.
Finding intervals of concave up/down, inec on points
Example
Let f(x) = x x + . Find the inec on points of f and the intervals on which
it is concave up/down.

Notes:

Note: Geometrically speaking, a func on


is concave up if its graph lies above its tangent lines. A func on is concave down if
its graph lies below its tangent lines.

f >

f >

c. up

c. up
f <
c. down

Figure . : A graph of a func on with


its inec on points marked. The intervals where concave up/down are also indicated.

Chapter

The Graphical Behavior of Func ons

f < 0 c. down

.
0

Figure . : A number line determining


the concavity of f in Example .

f(x)

f (x)

Finding intervals of concave up/down, inec on points


Example
Let f(x) = x/(x ). Find the inec on points of f and the intervals on which
it is concave up/down.
S

plifying, we nd

.
Figure .
ple .

We start by nding f (x) = x and f (x) = x. To nd


the inec on points, we use Theorem
and nd where f (x) = or where

f is undened. We nd f is always dened, and is only when x = . So the


point ( , ) is the only possible point of inec on.
This possible inec on point divides the real line into two intervals, (, )
and ( , ). We use a process similar to the one used in the previous sec on to
determine increasing/decreasing. Pick any c < ; f (c) < so f is concave
down on (, ). Pick any c > ; f (c) > so f is concave up on ( , ). Since
the concavity changes at x = , the point ( , ) is an inec on point.
The number line in Figure . illustrates the process of determining concavity; Figure . shows a graph of f and f , conrming our results. No ce how f
is concave down precisely when f (x) < and concave up when f (x) > .
S

f > 0 c. up

: A graph of f(x) used in Exam-

We need to nd f and f . Using the Quo ent Rule and sim-

f (x) =

( + x )
(x )

and

f (x) =

x(x + )
.
(x )

To nd the possible points of inec on, we seek to nd where f (x) = and


where f is not dened. Solving f (x) = reduces to solving x(x + ) = ;
we nd x = . We nd that f is not dened when x = , for then the
denominator of f is . We also note that f itself is not dened at x = ,
having a domain of (, ) ( , ) ( , ). Since the domain of f is the
union of three intervals, it makes sense that the concavity of f could switch across
intervals. We technically cannot say that f has a point of inec on at x = as
they are not part of the domain, but we must s ll consider these x-values to be
important and will include them in our number line.
The important x-values at which concavity might switch are x = , x =
and x = , which split the number line into four intervals as shown in Figure
. . We determine the concavity on each. Keep in mind that all we are concerned with is the sign of f on the interval.
Interval , (, ): Select a number c in this interval with a large magnitude
(for instance, c =
). The denominator of f (x) will be posi ve. In the
numerator, the (c + ) will be posi ve and the c term will be nega ve. Thus
the numerator is nega ve and f (c) is nega ve. We conclude f is concave down
on (, ).

Notes:

.
Interval , ( , ): For any number c in this interval, the term c in the numerator will be nega ve, the term (c + ) in the numerator will be posi ve, and
the term (c ) in the denominator will be nega ve. Thus f (c) > and f is
concave up on this interval.
Interval , ( , ): Any number c in this interval will be posi ve and small. Thus
the numerator is posi ve while the denominator is nega ve. Thus f (c) <
and f is concave down on this interval.
Interval , ( , ): Choose a large value for c. It is evident that f (c) > , so we
conclude that f is concave up on ( , ).
f < 0 c. down

f > 0 c. up

f < 0 c. down

f > 0 c. up

.
1

Figure .

f(x)

f (x)
x

Figure .
Example

: Number line for f in Example

Concavity and the Second Deriva ve

: A graph of f(x) and f (x) in


.

We conclude that f is concave up on ( , ) ( , ) and concave down on


(, )( , ). There is only one point of inec on, ( , ), as f is not dened
at x = . Our work is conrmed by the graph of f in Figure . . No ce how
f is concave up whenever f is posi ve, and concave down when f is nega ve.
Recall that rela ve maxima and minima of f are found at cri cal points of
f; that is, they are found when f (x) = or when f is undened. Likewise,
the rela ve maxima and minima of f are found when f (x) = or when f is
undened; note that these are the inec on points of f.
What does a rela ve maximum of f mean? The deriva ve measures the
rate of change of f; maximizing f means nding the where f is increasing the
most where f has the steepest tangent line. A similar statement can be made
for minimizing f ; it corresponds to where f has the steepest nega velysloped
tangent line.
We u lize this concept in the next example.
y

Understanding inec on points


Example
The sales of a certain product over a three-year span are modeled by S(t) =
t t + , where t is the me in years, shown in Figure . . Over the rst
two years, sales are decreasing. Find the point at which sales are decreasing at
their greatest rate.
We want to maximize the rate of decrease, which is to say,
we want to nd where S has a minimum. To do this, we nd where S is . We
nd S (t)= t t and S (t) = t . Se ng S (t) = and solving, we
get t =
/ . (we ignore the nega ve value of t since it does not lie in
the domain of our func on S).
This is both the inec on point and the point of maximum decrease. This
S

Notes:

S(t)

Figure . : A graph of S(t) in Example ,


modeling the sale of a product over me.

Chapter

The Graphical Behavior of Func ons

y
S(t)

Not every cri cal point corresponds to a rela ve extrema; f(x) = x has a
cri cal point at ( , ) but no rela ve maximum or minimum. Likewise, just because f (x) = we cannot conclude concavity changes at that point. We were
careful before to use terminology possible point of inec on since we needed
to check to see if the concavity changed. The canonical example of f (x) =
without concavity changing is f(x) = x . At x = , f (x) = but f is always
concave up, as shown in Figure . .

S (t)

.
Figure . : A graph of S(t) in Example
along with S (t).

The Second Deriva ve Test

0.5

.1

0.5

0.5

is the point at which things rst start looking up for the company. A er the
inec on point, it will s ll take some me before sales start to increase, but at
least sales are not decreasing quite as quickly as they had been.
A graph of S(t) and S (t) is given in Figure . . When S (t) < , sales are
decreasing; note how at t . , S (t) is minimized. That is, sales are decreasing at the fastest rate at t . . On the interval of ( . , ), S is decreasing
but concave up, so the decline in sales is leveling o.

Figure . : A graph of f(x) = x . Clearly f


is always concave up, despite the fact that
f (x) = when x = . It this example, the possible point of inec on ( , )
is not a point of inec on.

The rst deriva ve of a func on gave us a test to nd if a cri cal value corresponded to a rela ve maximum, minimum, or neither. The second deriva ve
gives us another way to test if a cri cal point is a local maximum or minimum.
The following theorem ocially states something that is intui ve: if a cri cal
value occurs in a region where a func on f is concave up, then that cri cal value
must correspond to a rela ve minimum of f, etc. See Figure . for a visualizaon of this.
Theorem

The Second Deriva ve Test

Let c be a cri cal value of f where f (c) is dened.


. If f (c) > , then f has a local minimum at (c, f(c)).
. If f (c) < , then f has a local maximum at (c, f(c)).

c. down

rel. max

The Second Deriva ve Test relates to the First Deriva ve Test in the following
way. If f (c) > , then the graph is concave up at a cri cal point c and f itself
is growing. Since f (c) = and f is growing at c, then it must go from nega ve
to posi ve at c. This means the func on goes from decreasing to increasing, indica ng a local minimum at c.

c. up
rel. min

Figure . : Demonstra ng the fact that


rela ve maxima occur when the graph is
concave down and rela ve minima occur
when the graph is concave up.

Notes:

Concavity and the Second Deriva ve


y

Using the Second Deriva ve Test


Example
Let f(x) =
/x + x. Find the cri cal points of f and use the Second Deriva ve
Test to label them as rela ve maxima or minima.
We nd f (x) =
/x + and f (x) =
/x . We set
f (x) = and solve for x to nd the cri cal values (note that f is not dened at
x = , but neither is f so this is not a cri cal value.) We nd the cri cal values
are x = . Evalua ng f at x = gives . > , so there is a local minimum
at x = . Evalua ng f ( ) = . < , determining a rela ve maximum
at x = . These results are conrmed in Figure . .

4
f (

We have been learning how the rst and second deriva ves of a func on
relate informa on about the graph of that func on. We have found intervals of
increasing and decreasing, intervals where the graph is concave up and down,
along with the loca ons of rela ve extrema and inec on points. In Chapter
we saw how limits explained asympto c behavior. In the next sec on we combine all of this informa on to produce accurate sketches of func ons.

f (

) <

Figure . : A graph of f(x) in Example


. The second deriva ve is evaluated
at each cri cal point. When the graph is
concave up, the cri cal point represents
a local minimum; when the graph is concave down, the cri cal point represents a
local maximum.

Notes:

) >

Exercises .

Terms and Concepts


. Sketch a graph of a func on f(x) that is concave up on ( , )
and is concave down on ( , ).
. Sketch a graph of a func on f(x) that is:
(a) Increasing, concave up on ( , ),
(b) increasing, concave down on ( , ),

In Exercises

, a func on f(x) is given.

(a) Find the possible points of inec on of f.


(b) Create a number line to determine the intervals on
which f is concave up or concave down.
. f(x) = x x +
. f(x) = x x +

(c) decreasing, concave down on ( , ) and

. f(x) = x x +

(d) increasing, concave down on ( , ).

. f(x) = x x + x +

. Is is possible for a func on to be increasing and concave


down on ( , ) with a horizontal asymptote of y = ? If
so, give a sketch of such a func on.
. Is is possible for a func on to be increasing and concave up
on ( , ) with a horizontal asymptote of y = ? If so, give
a sketch of such a func on.

. f(x) =

In Exercises

(a) Compute f (x).


(b) Graph f and f on the same axes (using technology is
permi ed) and verify Theorem .
. f(x) = x +
. f(x) = x + x
. f(x) = x + x
. f(x) = x x + x
. f(x) = x + x x +
. f(x) = cos x
. f(x) = sin x
. f(x) = tan x

. f(x) =
. f(x) =

x +
x
x

x+
x+

x +
x
x

. f(x) = sin x + cos x on (, )


. f(x) = x ex
. f(x) = x ln x
. f(x) = ex
In Exercises
, a func on f(x) is given. Find the cri cal
points of f and use the Second Deriva ve Test, when possible, to determine the rela ve extrema. (Note: these are the
same func ons as in Exercises .)
. f(x) = x x +
. f(x) = x x +
. f(x) = x x +
. f(x) = x x + x +
. f(x) =

. f(x) =

. f(x) = x x + x x +

. f(x) =
, a func on f(x) is given.

. f(x) = x + x + x

. f(x) =

Problems

x+

. f(x) = x + x + x
. f(x) = x x + x x +
. f(x) =

x +

x+

. f(x) =

x
x

. f(x) = sin x + cos x on (, )


. f(x) = x ex
. f(x) = x ln x
. f(x) = ex
In Exercises
, a func on f(x) is given. Find the x values where f (x) has a rela ve maximum or minimum. (Note:
these are the same func ons as in Exercises .)
. f(x) = x x +
. f(x) = x x +
. f(x) = x x +
. f(x) = x x + x +

. f(x) =

x+

. f(x) = x + x + x

x+

. f(x) = x x + x x +
. f(x) =
. f(x) =

x +
x
x

. f(x) = sin x + cos x on (, )


. f(x) = x ex
. f(x) = x ln x
. f(x) = ex

Chapter

The Graphical Behavior of Func ons

Curve Sketching

We have been learning how we can understand the behavior of a func on based
on its rst and second deriva ves. While we have been trea ng the proper es
of a func on separately (increasing and decreasing, concave up and concave
down, etc.), we combine them here to produce an accurate graph of the func on
without plo ng lots of extraneous points.
Why bother? Graphing u li es are very accessible, whether on a computer,
a handheld calculator, or a smartphone. These resources are usually very fast
and accurate. We will see that our method is not par cularly fast it will require
me (but it is not hard). So again: why bother?
We are a emp ng to understand the behavior of a func on f based on the
informa on given by its deriva ves. While all of a func ons deriva ves relay
informa on about it, it turns out that most of the behavior we care about is
explained by f and f . Understanding the interac ons between the graph of f
and f and f is important. To gain this understanding, one might argue that all
that is needed is to look at lots of graphs. This is true to a point, but is somewhat
similar to sta ng that one understands how an engine works a er looking only at
pictures. It is true that the basic ideas will be conveyed, but handson access
increases understanding.
The following Key Idea summarizes what we have learned so far that is applicable to sketching graphs of func ons and gives a framework for pu ng that
informa on together. It is followed by several examples.
Key Idea

Curve Sketching

To produce an accurate sketch a given func on f, consider the following


steps.
. Find the domain of f. Generally, we assume that the domain is the
en re real line then nd restric ons, such as where a denominator
is or where nega ves appear under the radical.
. Find the cri cal values of f.
. Find the possible points of inec on of f.
. Find the loca on of any ver cal asymptotes of f (usually done in
conjunc on with item above).
. Consider the limits lim f(x) and lim f(x) to determine the end
x

behavior of the func on.


(con nued)

Notes:

Key Idea

Curve Sketching Con nued

. Create a number line that includes all cri cal points, possible
points of inec on, and loca ons of ver cal asymptotes. For each
interval created, determine whether f is increasing or decreasing,
concave up or down.
. Evaluate f at each cri cal point and possible point of inec on.
Plot these points on a set of axes. Connect these points with curves
exhibi ng the proper concavity. Sketch asymptotes and x and y
intercepts where applicable.
Curve sketching
Example
Use Key Idea to sketch f(x) = x
S

x + x+ .

We follow the steps outlined in the Key Idea.

. The domain of f is the en re real line; there are no values x for which f(x)
is not dened.
. Find the cri cal values of f. We compute f (x) = x x + . Use the
Quadra c Formula to nd the roots of f :

(
)
( ) ( )( )
x .
, .
.
x=
=

( )
. Find the possible points of inec on of f. Compute f (x) =
have
f (x) = x = / .
.

. We

. There are no ver cal asymptotes.


. We determine the end behavior using limits as x approaches innity.
lim f(x) =

lim f(x) = .

We do not have any horizontal asymptotes.

. We place the values x = (


)/ and x =
/ on a number
line, as shown in Figure . . We mark each subinterval as increasing or

Notes:

Curve Sketching

Chapter

The Graphical Behavior of Func ons


decreasing, concave up or down, using the techniques used in Sec ons
. and . .

f > 0 incr

f < 0 decr

f < 0 decr

f > 0 incr

f < 0 c. down

f < 0 c. down

f > 0 c. up

f < 0 c. up

5
1

(10 +
1.

: Number line for f in Example

Curve sketching
x x
.
Sketch f(x) =
x x
S

We again follow the steps outlined in Key Idea .

. In determining the domain, we assume it is all real numbers and looks for
restric ons. We nd that at x = and x = , f(x) is not dened. So the
domain of f is D = {real numbers x | x = , }.

(b)
y

. To nd the cri cal values of f, we rst nd f (x). Using the Quo ent Rule,
we nd
x+
x+
f (x) =
=
.
(x + x )
(x ) (x + )

f (x) = when x = / , and f is undened when x = , . Since f


is undened only when f is, these are not cri cal values. The only cri cal
value is x = / .

. To nd the possible points of inec on, we nd f (x), again employing


the Quo ent Rule:

(c)
Figure .

1.111

Example

. We plot the appropriate points on axes as shown in Figure . (a) and


connect the points with straight lines. In Figure . (b) we adjust these
lines to demonstrate the proper concavity. Our curve crosses the y axis at
y = and crosses the x axis near x = .
. In Figure . (c) we show a
graph of f drawn with a computer program, verifying the accuracy of our
sketch.

(a)

10

Figure .

(10
0.

: Sketching f in Example

f (x) =

x x+
.
(x ) (x + )

We nd that f (x) is never (se ng the numerator equal to and solving


for x, we nd the only roots to this quadra c are imaginary) and f is

Notes:

.
undened when x = , . Thus concavity will possibly only change at
x = and x = .

Curve Sketching

y
5

. The ver cal asymptotes of f are at x = and x = , the places where f


is undened.
. There is a horizontal asymptote of y = , as lim f(x) = and lim f(x) =
x

. We place the values x = / , x = and x = on a number line as


shown in Figure . . We mark in each interval whether f is increasing or
decreasing, concave up or down. We see that f has a rela ve maximum at
x = / ; concavity changes only at the ver cal asymptotes.

f >

incr

c. up

>

f >
f

<

f <

incr
c. down

<

f <

decr

c. down

>

(a)
y

decr
5

c. up

Figure .

: Number line for f in Example

. In Figure . (a), we plot the points from the number line on a set of
axes and connect the points with straight lines to get a general idea of
what the func on looks like (these lines eec vely only convey increasing/decreasing informa on). In Figure . (b), we adjust the graph with
the appropriate concavity. We also show f crossing the x axis at x =
and x = .

..

(b)
y

Figure . (c) shows a computer generated graph of f, which veries the accuracy of our sketch.

Curve sketching
(x )(x + )
.
Sketch f(x) =
x + x+
Example

We again follow Key Idea .

. We assume that the domain of f is all real numbers and consider restricons. The only restric ons come when the denominator is , but this
never occurs. Therefore the domain of f is all real numbers, R.

. We nd the cri cal values of f by se ng f (x) =


nd

Figure .

f (x) =

Notes:

x(x + )
(x + x + )

f (x) =

and solving for x. We


when x = , .

(c)
: Sketching f in Example

Chapter

The Graphical Behavior of Func ons


. We nd the possible points of inec on by solving f (x) =
nd
x +
x
f (x) =
.
(x + x + )

The cubic in the numerator does not factor very nicely. We instead approximate the roots at x = .
,x= .
and x = .
.
5

. There are no ver cal asymptotes.

. We have a horizontal asymptote of y = , as lim f(x) = lim f(x) = .


x

f > 0 incr

f > 0 incr

f < 0 decr

f < 0 decr

f > 0 incr

f > 0 decr

f > 0 c. up

f < 0 c. down

f < 0 c. down

f > 0 c. up

f > 0 c. up

f < 0 c. down

.
(c)
: Sketching f in Example

1. 0

: Number line for f in Example

1.0

. In Figure . (a) we plot the signicant points from the number line as
well as the two roots of f, x = and x = , and connect the points
with straight lines to get a general impression about the graph. In Figure
. (b), we add concavity. Figure . (c) shows a computer generated
graph of f, arming our results.

(b)

Figure .

. We place the cri cal points and possible points on a number line as shown
in Figure . and mark each interval as increasing/decreasing, concave
up/down appropriately.

(a)

Figure .

for x. We

In each of our examples, we found a few, signicant points on the graph of


f that corresponded to changes in increasing/decreasing or concavity. We connected these points with straight lines, then adjusted for concavity, and nished
by showing a very accurate, computer generated graph.
Why are computer graphics so good? It is not because computers are smarter than we are. Rather, it is largely because computers are much faster at compu ng than we are. In general, computers graph func ons much like most students do when rst learning to draw graphs: they plot equally spaced points,
then connect the dots using lines. By using lots of points, the connec ng lines
are short and the graph looks smooth.
This does a ne job of graphing in most cases (in fact, this is the method
used for many graphs in this text). However, in regions where the graph is very
curvy, this can generate no ceable sharp edges on the graph unless a large
number of points are used. High quality computer algebra systems, such as

Notes:

.
Mathema ca, use special algorithms to plot lots of points only where the graph
is curvy.
In Figure . , a graph of y = sin x is given, generated by Mathema ca.
The small points represent each of the places Mathema ca sampled the funcon. No ce how at the bends of sin x, lots of points are used; where sin x
is rela vely straight, fewer points are used. (Many points are also used at the
endpoints to ensure the end behavior is accurate.)
1.0

0.5

-0.5

-1.0

Figure .

: A graph of y = sin x generated by Mathema ca.

How does Mathema ca know where the graph is curvy? Calculus. When
we study curvature in a later chapter, we will see how the rst and second
deriva ves of a func on work together to provide a measurement of curviness. Mathema ca employs algorithms to determine regions of high curvature and plots extra points there.
Again, the goal of this sec on is not How to graph a func on when there
is no computer to help. Rather, the goal is Understand that the shape of the
graph of a func on is largely determined by understanding the behavior of the
func on at a few key places. In Example , we were able to accurately sketch
a complicated graph using only points and knowledge of asymptotes!
There are many applica ons of our understanding of deriva ves beyond curve
sketching. The next chapter explores some of these applica ons, demonstrating just a few kinds of problems that can be solved with a basic knowledge of
dieren a on.

Notes:

Curve Sketching

Exercises .

Terms and Concepts


. Why is sketching curves by hand benecial even though
technology is ubiquitous?
. What does ubiquitous mean?
. T/F: When sketching graphs of func ons, it is useful to nd
the cri cal points.
. T/F: When sketching graphs of func ons, it is useful to nd
the possible points of inec on.
. T/F: When sketching graphs of func ons, it is useful to nd
the horizontal and ver cal asymptotes.

. f(x) = x + x + x +
. f(x) = x x x +
. f(x) = (x ) ln(x )
. f(x) = (x ) ln(x )
. f(x) =

x
x

. f(x) =

x x+
x x+

. f(x) =

x x+
x x+

Problems

. f(x) = x x +

In Exercises , prac ce using Key Idea by applying the


principles to the given func ons with familiar graphs.

. f(x) = x ex

. f(x) = x +
. f(x) = x +
. f(x) = sin x
. f(x) = ex
. f(x) =
. f(x) =

. f(x) = sin x cos x on [, ]


. f(x) = (x )

(x )
x

. f(x) =

In Exercises
, a func on with the parameters a and b
are given. Describe the cri cal points and possible points of
inec on of f in terms of a and b.
. f(x) =

In Exercises , sketch a graph of the given func on using


Key Idea . Show all work; check your answer with technology.
. f(x) = x x + x +
. f(x) = x + x x +

a
x +b

. f(x) = sin(ax + b)
. f(x) = (x a)(x b)
. Given x + y =

, use implicit dieren a on to nd

dy
dx

and ddx y . Use this informa on to jus fy the sketch of the


unit circle.

:A
D 6

In Chapter , we learned how the rst and second deriva ves of a func on inuence its graph. In this chapter we explore other applica ons of the deriva ve.

Newtons Method

Solving equa ons is one of the most important things we do in mathema cs,
yet we are surprisingly limited in what we can solve analy cally. For instance,
equa ons as simple as x + x + = or cos x = x cannot be solved by algebraic
methods in terms of familiar func ons. Fortunately, there are methods that
can give us approximate solu ons to equa ons like these. These methods can
usually give an approxima on correct to as many decimal places as we like. In
Sec on . we learned about the Bisec on Method. This sec on focuses on
another technique (which generally works faster), called Newtons Method.
Newtons Method is built around tangent lines. The main idea is that if x is
suciently close to a root of f(x), then the tangent line to the graph at (x, f(x))
will cross the x-axis at a point closer to the root than x.
We start Newtons Method with an ini al guess about roughly where the
root is. Call this x . (See Figure . (a).) Draw the tangent line to the graph at
(x , f(x )) and see where it meets the x-axis. Call this point x . Then repeat the
process draw the tangent line to the graph at (x , f(x )) and see where it meets
the x-axis. (See Figure . (b).) Call this point x . Repeat the process again to get
x , x , etc. This sequence of points will o en converge rather quickly to a root
of f.
We can use this geometric process to create an algebraic process. Lets look
at how we found x . We started with the tangent line to the graph at (x , f(x )).
The slope of this tangent line is f (x ) and the equa on of the line is

1
0.5

x0

x1

0.5
1
.

(a)
y

.5

.5

(b)
y

y = f (x )(x x ) + f(x ).
This line crosses the x-axis when y = , and the xvalue where it crosses is what
we called x . So let y = and replace x with x , giving the equa on:
= f (x )(x x ) + f(x ).
Now solve for x :
x =x

f(x )
.
f (x )

.5

.5

(c)
Figure . : Demonstra ng the geometric
concept behind Newtons Method.

Chapter

Applica ons of the Deriva ve


Since we repeat the same geometric process to nd x from x , we have

x =x

f(x )
.
f (x )

In general, given an approxima on xn , we can nd the next approxima on, xn+


as follows:
xn+ = xn

f(xn )
.
f (xn )

We summarize this process as follows.

Key Idea

Newtons Method

Let f be a dieren able func on on an interval I with a root in I. To approximate the value of the root, accurate to d decimal places:
. Choose a value x as an ini al approxima on of the root. (This is
o en done by looking at a graph of f.)
Note: Newtons Method is not infallible. The sequence of approximate values
may not converge, or it may converge so
slowly that one is tricked into thinking a
certain approxima on is be er than it actually is. These issues will be discussed at
the end of the sec on.

. Create successive approxima ons itera vely; given an approximaon xn , compute the next approxima on xn+ as
xn+ = xn

f(xn )
.
f (xn )

. Stop the itera ons when successive approxima ons do not dier
in the rst d places a er the decimal point.

Lets prac ce Newtons Method with a concrete example.


Using Newtons Method
Example
Approximate the real root of x x = , accurate to the rst places a er
the decimal, using Newtons Method and an ini al approxima on of x = .
S

Notes:

To begin, we compute f (x) = x x. Then we apply the

Newtons Method

Newtons Method algorithm, outlined in Key Idea .


x =
x =
x =
x =
x =

f( )

=
= ,
f ( )


f( )
=
= .

f ( )

f( .
)
.
.

= .

f ( .
)
.
f( .
)
.

.
f ( .
)
f( .
)
.

.
f ( .
)

,
.

We performed itera ons of Newtons Method to nd a root accurate to the


rst places a er the decimal; our nal approxima on is .
. The exact value
of the root, to six decimal places, is .
; It turns out that our x is accurate
to more than just decimal places.
A graph of f(x) is given in Figure . . We can see from the graph that our
ini al approxima on of x = was not par cularly accurate; a closer guess
would have been x = . . Our choice was based on ease of ini al calcula on,
and shows that Newtons Method can be robust enough that we do not have to
make a very accurate ini al approxima on.
We can automate this process on a calculator that has an Ans key that returns the result of the previous calcula on. Start by pressing 1 and then Enter.
(We have just entered our ini al guess, x = .) Now compute
Ans

f(Ans)
f (Ans)

by entering the following and repeatedly press the Enter key:


Ans-(Ans^3-Ans^2-1)/(3*Ans^2-2*Ans)
Each me we press the Enter key, we are nding the successive approxima ons,
x , x , , and each one is ge ng closer to the root. In fact, once we get past
around x or so, the approxima ons dont appear to be changing. They actually
are changing, but the change is far enough to the right of the decimal point that
it doesnt show up on the calculators display. When this happens, we can be
pre y condent that we have found an accurate approxima on.
Using a calculator in this manner makes the calcula ons simple; many itera ons can be computed very quickly.

Notes:

.
0.5

0.5

1.5

0.5

1
1.5.

Figure . : A graph of f(x) = x x


in Example .

Chapter

Applica ons of the Deriva ve


Using Newtons Method to nd where func ons intersect
Example
Use Newtons Method to approximate a solu on to cos x = x, accurate to
places a er the decimal.

y
1
0.5

0.5

0.5

0.5

Newtons Method provides a method of solving f(x) = ; it


S
is not (directly) a method for solving equa ons like f(x) = g(x). However, this is
not a problem; we can rewrite the la er equa on as f(x) g(x) = and then
use Newtons Method.
So we rewrite cos x = x as cos x x = . Wri en this way, we are nding
a root of f(x) = cos x x. We compute f (x) = sin x . Next we need a
star ng value, x . Consider Figure . , where f(x) = cos x x is graphed. It
seems that x = . is pre y close to the root, so we will use that as our x .
(The gure also shows the graphs of y = cos x and y = x, drawn with dashed
lines. Note how they intersect at the same x value as when f(x) = .)
We now compute x , x , etc. The formula for x is
x = .

Figure . : A graph of f(x) = cos x x


used to nd an ini al approxima on of its
root.

cos( . ) .
sin( . )

Apply Newtons Method again to nd x :


x = .

cos( .
sin( .

) .

We can con nue this way, but it is really best to automate this process. On a calculator with an Ans key, we would start by pressing . , then Enter, inpu ng
our ini al approxima on. We then enter:
Ans - (cos(Ans)-Ans)/(-sin(Ans)-1).
Repeatedly pressing the Enter key gives successive approxima ons. We
quickly nd:
x = .
x = .

Our approxima ons x and x did not dier for at least the rst places a er the
decimal, so we could have stopped. However, using our calculator in the manner described is easy, so nding x was not hard. It is interes ng to see how we
found an approxima on, accurate to as many decimal places as our calculator
displays, in just itera ons.
If you know how to program, you can translate the following pseudocode
into your favorite language to perform the computa on in this problem.

Notes:

.
x = .75
while true
oldx = x
x = x - (cos(x)-x)/(-sin(x)-1)
print x
if abs(x-oldx) < .0000000001
break
This code calculates x , x , etc., storing each result in the variable x. The previous approxima on is stored in the variable oldx. We con nue looping un l
the dierence between two successive approxima ons, abs(x-oldx), is less
than some small tolerance, in this case, .0000000001.

Convergence of Newtons Method

y
0.5

x =

and f ( ) = , we see that x is not well dened.


This problem can also occur if, for instance, it turns out that f (x ) = .
Adjus ng the ini al approxima on x by a very small amount will likely x the
problem.
It is also possible for Newtons Method to not converge while each successive
approxima on is well dened. Consider f(x) = x / , as shown in Figure . . It
is clear that the root is x = , but lets approximate this with x = . . Figure
. (a) shows graphically the calcula on of x ; no ce how it is farther from the
root than x . Figures . (b) and (c) show the calcula on of x and x , which are
even farther away; our successive approxima ons are ge ng worse. (It turns
out that in this par cular example, each successive approxima on is twice as far
from the true answer as the previous approxima on.)
There is no x to this problem; Newtons Method simply will not work and
another method must be used.
While Newtons Method does not always work, it does work most of the
me, and it is generally very fast. Once the approxima ons get close to the root,

0.5

0.5

1.5

0.5

1.5

Figure . : A graph of f(x) = x x ,


showing why an ini al approxima on of
x = with Newtons Method fails.

What should one use for the ini al guess, x ? Generally, the closer to the
actual root the ini al guess is, the be er. However, some ini al guesses should
be avoided. For instance, consider Example
where we sought the root to
f(x) = x x . Choosing x = would have been a par cularly poor choice.
Consider Figure . , where f(x) is graphed along with its tangent line at x = .
Since f ( ) = , the tangent line is horizontal and does not intersect the xaxis.
Graphically, we see that Newtons Method fails.
We can also see analy cally that it fails. Since
f( )

f ( )

Newtons Method

x0

(a)
y

x1 x0

(b)
y

Notes:

(c)

Figure . : Newtons Method fails to nd


a root of f(x) = x / , regardless of the
choice of x .

Chapter

Applica ons of the Deriva ve


Newtons Method can as much as double the number of correct decimal places
with each successive approxima on. A course in Numerical Analysis will introduce the reader to more itera ve root nding methods, as well as give greater
detail about the strengths and weaknesses of Newtons Method.

Notes:

Exercises .

Terms and Concepts


. T/F: Given a func on f(x), Newtons Method produces an
exact solu on to f(x) = .
. T/F: In order to get a solu on to f(x) = accurate to d
places a er the decimal, at least d + itera ons of Newtons Method must be used.

Problems
In Exercises , the roots of f(x) are known or are easily
found. Use itera ons of Newtons Method with the given
ini al approxima on to approximate the root. Compare it to
the known value of the root.
. f(x) = cos x, x = .
. f(x) = sin x, x =

that interval. Use technology to obtain good ini al approxima ons.


. f(x) = x + x x
. f(x) = x + x x x +
. f(x) = x x

x +

on ( , )

. f(x) = x cos x + (x ) sin x on ( , )


In Exercises
, use Newtons Method to approximate
when the given func ons are equal, accurate to places after the decimal. Use technology to obtain good ini al approxima ons.
. f(x) = x , g(x) = cos x
. f(x) = x , g(x) = sin x

. f(x) = x + x , x =

. f(x) = ex , g(x) = cos x

. f(x) = x , x = .

. f(x) = x, g(x) = tan x on [ , ]

. f(x) = ln x, x =

. Why does Newtons Method fail in nding a root of f(x) =


x x + x + when x = ?

In Exercises , use Newtons Method to approximate all


roots of the given func ons accurate to places a er the decimal. If an interval is given, nd only the roots that lie in

. Why does Newtons Method fail in nding a root of f(x) =


x +
x
x +
x+
when x = ?

Chapter

Applica ons of the Deriva ve

Related Rates

When two quan es are related by an equa on, knowing the value of one quanty can determine the value of the other. For instance, the circumference and
radius of a circle are related by C = r; knowing that C = in determines the
radius must be in.
The topic of related rates takes this one step further: knowing the rate
at which one quan ty is changing can determine the rate at which the other
changes.
We demonstrate the concepts of related rates through examples.
Understanding related rates
Example
The radius of a circle is growing at a rate of in/hr. At what rate is the circumference growing?
The circumference and radius of a circle are related by C =
S
r. We are given informa on about how the length of r changes with respect
to me; that is, we are told dr
dt = in/hr. We want to know how the length of C
changes with respect to me, i.e., we want to know dC
dt .
Implicitly dieren ate both sides of C = r with respect to t:
C=
d( )
C =
dt
dC
=
dt

Note: This sec on relies heavily on implicit dieren a on, so referring back to
Sec on . may help.

As we know

dr
dt

r
d( )
r
dt
dr
.
dt

= in/hr, we know
dC
= =
dt

. in/hr.

Consider another, similar example.


Finding related rates
Example
Water streams out of a faucet at a rate of in /s onto a at surface at a constant
rate, forming a circular puddle that is / in deep.
. At what rate is the area of the puddle growing?
. At what rate is the radius of the circle growing?

Notes:

.
S

. We can answer this ques on two ways: using common sense or related
rates. The common sense method states that the volume of the puddle is
growing by in /s, where
volume of puddle = area of circle depth.
Since the depth is constant at / in, the area must be growing by

in /s.

This approach reveals the underlying relatedrates principle. Let V and A


represent the Volume and Area of the puddle. We know V = A . Take
the deriva ve of both sides with respect to t, employing implicit dierena on.
A
( )
d( )
d
V =
A
dt
dt
dA
dV
=
dt
dt
V=

As dV
dt = , we know
growing by in /s.

dA
dt ,

and hence

dA
dt

. Thus the area is

. To start, we need an equa on that relates what we know to the radius.


We just learned something about the surface area of the circular puddle,
and we know A = r . We should be able to learn about the rate at which
the radius is growing with this informa on.
Implicitly derive both sides of A = r with respect to t:
A = r
d( )
d( )
A =
r
dt
dt
dA
dr
= r
dt
dt
Our work above told us that

dA
dt

in /s. Solving for

dr
dt ,

we have

dr
= .
dt
r
Note how our answer is not a number, but rather a func on of r. In other
words, the rate at which the radius is growing depends on how big the

Notes:

Related Rates

Chapter

Applica ons of the Deriva ve


circle already is. If the circle is very large, adding in of water will not
make the circle much bigger at all. If the circle dimesized, adding the
same amount of water will make a radical change in the radius of the circle.
In some ways, our problem was (inten onally) illposed. We need to specify a current radius in order to know a rate of change. When the puddle
has a radius of in, the radius is growing at a rate of
dr
=
dt

in/s.

Studying related rates


Example
Radar guns measure the rate of distance change between the gun and the object
it is measuring. For instance, a reading of mph means the object is moving
away from the gun at a rate of
miles per hour, whereas a measurement of
mph would mean that the object is approaching the gun at a rate of
miles per hour.
If the radar gun is moving (say, a ached to a police car) then radar readouts
are only immediately understandable if the gun and the object are moving along
the same line. If a police ocer is traveling mph and gets a readout of mph,
he knows that the car ahead of him is moving away at a rate of miles an hour,
meaning the car is traveling mph. (This straightline principle is one reason
ocers park on the side of the highway and try to shoot straight back down the
road. It gives the most accurate reading.)
Suppose an ocer is driving due north at
mph and sees a car moving
due east, as shown in Figure . . Using his radar gun, he measures a reading of
mph. By using landmarks, he believes both he and the other car are about
/ mile from the intersec on of their two roads.

If the speed limit on the other road is

B = 1/2

.
A = 1/2

Car

mph, is the other driver speeding?

Ocer

Figure . : A sketch of a police car (at bottom) a emp ng to measure the speed of
a car (at right) in Example
.

Using the diagram in Figure . , lets label what we know


S
about the situa on. As both the police ocer and other driver are / mile
from the intersec on, we have A = / , B = / , and through the Pythagorean
.
.
Theorem, C = /

We know the police ocer is traveling at mph; that is, dA


dt = . The
reason this rate of change is nega ve is that A is ge ng smaller; the distance
between the ocer and the intersec on is shrinking. The radar measurement
is dC
. We want to nd dB
dt =
dt .
We need an equa on that relates B to A and/or C. The Pythagorean Theorem

Notes:

.
is a good choice: A + B = C . Dieren ate both sides with respect to t:
A +B =C
)
d ( )
d (
A +B =
C
dt
dt
dA
dB
dC
A
+ B
= C
dt
dt
dt
We have values for everything except

dB
dt .

Solving for this we have

C dC A dA
dB
dt
= dt

dt
B

mph.

Related Rates

Note: Example
is both interes ng
and imprac cal. It highlights the diculty
in using radar in a nonlinear fashion, and
explains why in real life the police ocer would follow the other driver to determine their speed, and not pull out pencil and paper.
The principles here are important,
though. Many automated vehicles make
judgments about other moving objects
based on perceived distances, radarlike
measurements and the concepts of
related rates.

The other driver appears to be speeding slightly.


Studying related rates
Example
A camera is placed on a tripod
from the side of a road. The camera is to turn
to track a car that is to drive by at
mph for a promo onal video. The videos
planners want to know what kind of motor the tripod should be equipped with
in order to properly track the car as it passes by. Figure . shows the proposed
setup.
How fast must the camera be able to turn to track the car?

100mph

We seek informa on about how fast the camera is to turn;


S
therefore, we need an equa on that will relate an angle to the posi on of the
camera and the speed and posi on of the car.
Figure . suggests we use a trigonometric equa on. Le ng x represent the
distance the car is from the point on the road directly in front of the camera, we
have
x
.
( . )
tan =
mph (as in the last example,
As the car is moving at
mph, we have dx
dt =
since x is ge ng smaller as the car travels, dx
is
nega
ve). We need to convert
dt
the measurements so they use the same units; rewrite mph in terms of /s:
dx
=
dt

m
=
hr

hr

hr
=
s

/s.

Now take the deriva ve of both sides of Equa on ( . ) using implicit dieren -

Notes:

10

Figure . : Tracking a speeding car (at


le ) with a rota ng camera.

Chapter

Applica ons of the Deriva ve


a on:
tan =

)
d(
d ( x )
tan =
dt
dt
dx
d
=
sec
dt
dt
cos dx
d
=
dt
dt

( . )

We want to know the fastest the camera has to turn. Common sense tells us this
is when the car is directly in front of the camera (i.e., when = ). Our mathema cs bears this out. In Equa on ( . ) we see this is when cos is largest; this
is when cos = , or when = .
With dx
.
/s, we have
dt
d
rad
=
dt

/s =

radians/s.

We nd that d
dt is nega ve; this matches our diagram in Figure . for is ge ng
smaller as the car approaches the camera.
What is the prac cal meaning of .
radians/s? Recall that circular
revolu on goes through radians, thus .
rad/s means .
/( )
. revolu ons per second. The nega ve sign indicates the camera is rota ng
in a clockwise fashion.
We introduced the deriva ve as a func on that gives the slopes of tangent
lines of func ons. This chapter emphasizes using the deriva ve in other ways.
Newtons Method uses the deriva ve to approximate roots of func ons; this
sec on stresses the rate of change aspect of the deriva ve to nd a rela onship between the rates of change of two related quan es.
In the next sec on we use Extreme Value concepts to op mize quan es.

Notes:

Exercises .

Terms and Concepts


. T/F: Implicit dieren a on is o en used when solving related rates type problems.
. T/F: A study of related rates is part of the standard police
ocer training.

Problems
. Water ows onto a at surface at a rate of cm /s forming a
circular puddle mm deep. How fast is the radius growing
when the radius is:
(a)
(b)

cm?

cm?
cm?

(c)

(c) Directly overhead?


. An F-

aircra is ying at
mph with an eleva on of
on a straightline path that will take it directly over
an an aircra gun as in Exercise (note the lower elevaon here).
How fast must the gun be able to turn to accurately track
the aircra when the plane is:
(a)
(b)

feet away?
feet away?

. Consider the trac situa on introduced in Example


.
Calculate how fast the other car is traveling in each of the
following situa ons.
(a) The ocer is traveling due north at mph and is
/ mile from the intersec on, while the other car
is mile from the intersec on traveling west and the
radar reading is mph?

(b) The ocer is traveling due north at mph and is


mile from the intersec on, while the other car is
/ mile from the intersec on traveling west and the
radar reading is mph?

. An F- aircra is ying at
mph with an eleva on of
,
on a straightline path that will take it directly over
an an aircra gun.

10,000

How fast must the gun be able to turn to accurately track


the aircra when the plane is:

1 /s

At what rate is the top of the ladder sliding down the side
of the house when the base is:
(a)

cm?

. Consider the trac situa on introduced in Example


.
How fast is the other car traveling if the ocer and the
other car are each / mile from the intersec on, the other
car is traveling due west, the ocer is traveling north at
mph, and the radar reading is mph?

/ mile away?

24

(b)

(b)

. A
. ladder is leaning against a house while the base is
pulled away at a constant rate of /s.

cm?

. A circular balloon is inated with air owing at a rate of


cm /s. How fast is the radius of the balloon increasing
when the radius is:
(a)

mile away?

(c) Directly overhead?

cm?

(c)

(a)

foot from the house?

(b)

feet from the house?

(c)

feet from the house?

(d)

feet from the house?

. A boat is being pulled into a dock at a constant rate of


/min by a winch located
above the deck of the
boat.
.

10

At what rate is the boat approaching the dock when the


boat is:
(a)

feet out?

(b)

feet out?

(c)

foot from the dock?

(d) What happens when the length of rope pulling in the


boat is less than feet long?
. An inverted cylindrical cone,
deep and
across at
the top, is being lled with water at a rate of
/min. At
what rate is the water rising in the tank when the depth of
the water is:
(a)

foot?

(b)

feet?

(c)

feet?

How long will the tank take to ll when star ng at empty?

30

. A rope, a ached to a weight, goes up through a pulley at


the ceiling and back down to a worker. The man holds the
rope at the same height as the connec on point between
rope and weight.

2 /s

Suppose the man stands directly next to the weight (i.e., a


total rope length of
) and begins to walk away at a rate
of /s. How fast is the weight rising when the man has
walked:
(a)

feet?

(b)

feet?

How far must the man walk to raise the weight all the way
to the pulley?
. Consider the situa on described in Exercise . Suppose
the man starts
from the weight and begins to walk
away at a rate of /s.
(a) How long is the rope?

(b) How fast is the weight rising a er the man has walked
feet?
(c) How fast is the weight rising a er the man has walked
feet?
(d) How far must the man walk to raise the weight all the
way to the pulley?
. A hot air balloon li s o from ground rising ver cally. From
feet away, a woman tracks the path of the balloon.
When her sightline with the balloon makes a angle with
the horizontal, she notes the angle is increasing at about

/min.
(a) What is the eleva on of the balloon?
(b) How fast is it rising?
. A company that produces landscaping materials is dumping
sand into a conical pile. The sand is being poured at a rate
of
/sec; the physical proper es of the sand, in conjuncon with gravity, ensure that the cones height is roughly
/ the length of the diameter of the circular base.
How fast is the cone rising when it has a height of feet?

Op miza on

Op miza on

In Sec on . we learned about extreme values the largest and smallest values
a func on a ains on an interval. We mo vated our interest in such values by
discussing how it made sense to want to know the highest/lowest values of a
stock, or the fastest/slowest an object was moving. In this sec on we apply
the concepts of extreme values to solve word problems, i.e., problems stated
in terms of situa ons that require us to create the appropriate mathema cal
framework in which to solve the problem.
We start with a classic example which is followed by a discussion of the topic
of op miza on.

.
.
.
.
.

.
.

One can likely guess the correct answer that is great. We


S
will proceed to show how calculus can provide this answer in a context that
proves this answer is correct.
It helps to make a sketch of the situa on. Our enclosure is sketched twice in
Figure . , either with green grass and nice fence boards or as a simple rectangle.
Either way, drawing a rectangle forces us to realize that we need to know the
dimensions of this rectangle so we can create an area func on a er all, we are
trying to maximize the area.
We let x and y denote the lengths of the sides of the rectangle. Clearly,

Op miza on: perimeter and area


Example
A man has
feet of fencing, a large yard, and a small dog. He wants to create
a rectangular enclosure for his dog with the fencing that provides the maximal
area. What dimensions provide the maximal area?

.
x

Area = xy.
We do not yet know how to handle func ons with variables; we need to
reduce this down to a single variable. We know more about the situa on: the
man has
feet of fencing. By knowing the perimeter of the rectangle must
be
, we can create another equa on:
Perimeter =

= x + y.

We now have equa ons and unknowns. In the la er equa on, we solve
for y:
y=
x.
Now subs tute this expression for y in the area equa on:
Area = A(x) = x(

x).

Note we now have an equa on of one variable; we can truly call the Area a
func on of x.

Notes:

x
Figure . : A sketch of the enclosure in
Example
.

Chapter

Applica ons of the Deriva ve


This func on only makes sense when x , otherwise we get nega ve
values of area. So we nd the extreme values of A(x) on the interval [ , ].
To nd the cri cal points, we take the deriva ve of A(x) and set it equal to
, then solve for x.
A(x) = x(
=
A (x) =

x)

xx
x

We solve x = to nd x = ; this is the only cri cal point. We evaluate


A(x) at the endpoints of our interval and at this cri cal point to nd the extreme
values; in this case, all we care about is the maximum.
Clearly A( ) = and A( ) = , whereas A( ) =
. This is the maximum. Since we earlier found y =
x, we nd that y is also . Thus the
dimensions of the rectangular enclosure with perimeter of
. with maximum area is a square, with sides of length
.
This example is very simplis c and a bit contrived. (A er all, most people
create a design then buy fencing to meet their needs, and not buy fencing and
plan later.) But it models well the necessary process: create equa ons that describe a situa on, reduce an equa on to a single variable, then nd the needed
extreme value.
In real life, problems are much more complex. The equa ons are o en
not reducible to a single variable (hence mul variable calculus is needed) and
the equa ons themselves may be dicult to form. Understanding the principles here will provide a good founda on for the mathema cs you will likely encounter later.
We outline here the basic process of solving these op miza on problems.
Key Idea

Solving Op miza on Problems

. Understand the problem. Clearly iden fy what quan ty is to be


maximized or minimized. Make a sketch if helpful.
. Create equa ons relevant to the context of the problem, using the
informa on given. (One of these should describe the quan ty to
be op mized. Well call this the fundamental equa on.)
. If the fundamental equa on denes the quan ty to be op mized
as a func on of more than one variable, reduce it to a single variable func on using subs tu ons derived from the other equaons.
(con nued). . .

Notes:

Key Idea

Op miza on

Solving Op miza on Problems Con nued

. Iden fy the domain of this func on, keeping in mind the context
of the problem.
. Find the extreme values of this func on on the determined domain.
. Iden fy the values of all relevant quan

es of the problem.

We will use Key Idea in a variety of examples.


Op miza on: perimeter and area
Example
Here is another classic calculus problem: A woman has a
feet of fencing, a
small dog, and a large yard that contains a stream (that is mostly straight). She
wants to create a rectangular enclosure with maximal area that uses the stream
as one side. (Apparently her dog wont swim away.) What dimensions provide
the maximal area?

We again appeal to the perimeter; here the perimeter is


Perimeter =

= x + y.

Note how this is dierent than in our previous example.


. We now reduce the fundamental equa on to a single variable. In the
perimeter equa on, solve for y: y = x/ . We can now write Area as
Area = A(x) = x(

x/ ) =

x x .

Area is now dened as a func on of one variable.

Notes:

.
.

.
.
.

.
.

This is our fundamental equa on. This denes area as a func on of two
variables, so we need another equa on to reduce it to one variable.

Area = xy.

. We want to maximize the area; as in the example before,

. We are maximizing area. A sketch of the region will help; Figure . gives
two sketches of the proposed enclosed area. A key feature of the sketches
is to acknowledge that one side is not fenced.

We will follow the steps outlined by Key Idea .

x
x
y

Figure . : A sketch of the enclosure in


Example
.

Chapter

Applica ons of the Deriva ve


. We want the area to be nonnega ve. Since A(x) = x( x/ ), we want
x and
x/ . The la er inequality implies that x
, so
x
.
. We now nd the extreme values. At the endpoints, the minimum is found,
giving an area of .
Find the cri cal points. We have A (x) =
x; se ng this equal to
and solving for x returns x = . This gives an area of
A(
. We earlier set y =
two sides of length
.

)=

)=

x/ ; thus y = . Thus our rectangle will have


and one side of length , with a total area of

Keep in mind as we do these problems that we are prac cing a process; that
is, we are learning to turn a situa on into a system of equa ons. These equaons allow us to write a certain quan ty as a func on of one variable, which we
then op mize.

1000

.
.

5000

Figure . : Running a power line from


the power sta on to an oshore facility
with minimal cost in Example
.

x +

.
.

Op miza on: minimizing cost


Example
A power line needs to be run from an power sta on located on the beach to an
oshore facility. Figure . shows the distances between the power sta on to
the facility.
It costs $ / . to run a power line along the land, and $
/ . to run a
power line under water. How much of the power line should be run along the
land to minimize the overall cost? What is the minimal cost?
We will follow the strategy of Key Idea implicitly, without
S
specically numbering steps.
There are two immediate solu ons that we could consider, each of which we
will reject through common sense. First, we could minimize the distance by
directly connec ng the two loca ons with a straight line. However, this requires
that all the wire be laid underwater, the most costly op on. Second, we could
minimize the underwater length by running a wire all
. along the beach,
directly across from the oshore facility. This has the undesired eect of having
the longest distance of all, probably ensuring a nonminimal cost.
The op mal solu on likely has the line being run along the ground for a
while, then underwater, as the gure implies. We need to label our unknown
distances the distance run along the ground and the distance run underwater.
Recognizing that the underwater distance can be measured as the hypotenuse
of a right triangle, we choose to label the distances as shown in Figure . .

Figure . : Labeling unknown distances


in Example
.

Notes:

.
By choosing x as we did, we make the expression under the square root simple. We now create the cost func on.
Cost

land cost
land distance
(
x)

+
+
+

water cost
water distance
x +
.

So we have c(x) =
(
x) +
x +
. This func on only
makes sense on the interval [ ,
]. While we are fairly certain the endpoints
will not give a minimal cost, we s ll evaluate c(x) at each to verify.
c( ) =

c(

We now nd the cri cal values of c(x). We compute c (x) as


c (x) =

x
x +

Recognize that this is never undened. Se ng c (x) =


we have:

x
x +

x +
x
x +

=
=
=
x =

(x +

x =

)x =

x=
x=

x =

Evalua ng c(x) at x =
. gives a cost of about $
the power line is laid

. =
.
along land is
.
+

.
water distance is

Notes:

and solving for x,

.
,

. The distance
., and the under-

Op miza on

Chapter

Applica ons of the Deriva ve


In the exercises you will see a variety of situa ons that require you to combine problemsolving skills with calculus. Focus on the process; learn how to
form equa ons from situa ons that can be manipulated into what you need.
Eschew memorizing how to do this kind of problem as opposed to that kind
of problem. Learning a process will benet one far longer than memorizing a
specic technique.
The next sec on introduces our nal applica on of the deriva ve: dierenals. Given y = f(x), they oer a method of approxima ng the change in y a er
x changes by a small amount.

Notes:

Exercises .

Terms and Concepts


. T/F: An op miza on problem is essen ally an extreme
values problem in a story problem se ng.
. T/F: This sec on teaches one to nd the extreme values of
func on that have more than one variable.

What is the maximum volume of a package with a square


cross sec on (w = h) that does not exceed the
standard?
. The strength S of a wooden beam is directly propor onal
to its cross sec onal width w and the square of its height h;
that is, S = kwh for some constant k.
w

Problems
. Find the maximum product of two numbers (not necessarily integers) that have a sum of
.
. Find the minimum sum of two numbers whose product is
.
. Find the maximum sum of two numbers whose product is
.
. Find the maximum sum of two numbers, each of which is
in [ ,
] whose product is
.
. Find the maximal area of a right triangle with hypotenuse
of length .
. A rancher has
feet of fencing in which to construct
adjacent, equally sized rectangular pens. What dimensions
should these pens have to maximize the enclosed area?

. A standard soda can is roughly cylindrical and holds


cm
of liquid. What dimensions should the cylinder be to minimize the material needed to produce the can? Based on
your dimensions, determine whether or not the standard
can is produced to minimize the material costs.
. Find the dimensions of a cylindrical can with a volume of
in that minimizes the surface area.
The # canis a standard sized can used by the restaurant industry that holds about
in with a diameter of
/ in and height of in. Does it seem these dimensions
where chosen with minimiza on in mind?
. The United States Postal Service charges more for boxes
whose combined length and girth exceeds
(the
length of a package is the length of its longest side; the
girth is the perimeter of the cross sec on, i.e., w + h).

Given a circular log with diameter of


inches, what sized
beam can be cut from the log with maximum strength?
. A power line is to be run to an oshore facility in the manner described in Example
. The oshore facility is
miles at sea and miles along the shoreline from the power
plant. It costs $ ,
per mile to lay a power line underground and $ ,
to run the line underwater.
How much of the power line should be run underground to
minimize the overall costs?
. A power line is to be run to an oshore facility in the manner described in Example
. The oshore facility is
miles at sea and miles along the shoreline from the power
plant. It costs $ ,
per mile to lay a power line underground and $ ,
to run the line underwater.
How much of the power line should be run underground to
minimize the overall costs?
. A woman throws a s ck into a lake for her dog to fetch;
the s ck is feet down the shore line and feet into the
water from there. The dog may jump directly into the water and swim, or run along the shore line to get closer to
the s ck before swimming. The dog runs about
/s and
swims about . /s.
How far along the shore should the dog run to minimize
the me it takes to get to the s ck? (Hint: the gure from
Example
can be useful.)
. A woman throws a s ck into a lake for her dog to fetch;
the s ck is feet down the shore line and feet into the
water from there. The dog may jump directly into the water and swim, or run along the shore line to get closer to
the s ck before swimming. The dog runs about
/s and
swims about . /s.
How far along the shore should the dog run to minimize the
me it takes to get to the s ck? (Google calculus dog to learn
more about a dogs ability to minimize mes.)

. What are the dimensions of the rectangle with largest area


that can be drawn inside the unit circle?

Chapter

Applica ons of the Deriva ve

Dieren als

In Sec on . we explored the meaning and use of the deriva ve. This sec on
starts by revisi ng some of those ideas.
Recall that the deriva ve of a func on f can be used to nd the slopes of
lines tangent to the graph of f. At x = c, the tangent line to the graph of f has
equa on
y = f (c)(x c) + f(c).

The tangent line can be used to nd good approxima ons of f(x) for values of x
near c.
For instance, we can approximate sin . using the tangent
line to the graph
of f(x) = sin x at x = / . . Recall that sin(/ ) =
/ .
, and
cos(/ ) = / . Thus the tangent line to f(x) = sin x at x = / is:

(x) =

In Figure . (a), we see a graph of f(x) = sin x graphed along with its tangent line at x = / . The small rectangle shows the region that is displayed in
Figure . (b). In this gure, we see how we are approxima ng sin . with the
tangent line, evaluated at . . Together, the two gures show how close these
values are.
Using this line to approximate sin . , we have:

.5

(x / ) + .

(a)

( . ) =

( . ) sin .

( . / ) + .
( .

)+ .

= .

(We leave it to the reader to see how good of an approxima on this is.)
sin .

We now generalize this concept. Given f(x) and an xvalue c, the tangent
line is (x) = f (c)(x c) + f(c). Clearly, f(c) = (c). Let x be a small number,
represen ng a small change in x value. We assert that:

f(c + x) (c + x),

(b)
Figure . : Graphing f(x) = sin x and its
tangent line at x = / in order to es mate sin . .

since the tangent line to a func on approximates well the values of that func on
near x = c.
As the x value changes from c to c + x, the y value of f changes from f(c)
to f(c + x). We call this change of y value y. That is:
y = f(c + x) f(c).

Notes:

.
Replacing f(c + x) with its tangent line approxima on, we have
y (c + x) f(c)
(
)
= f (c) (c + x) c + f(c) f(c)
= f (c)x

( . )

This nal equa on is important; well come back to it in Key Idea .


We introduce two new variables, dx and dy in the context of a formal denion.
Deni on

Dieren als of x and y.

Let y = f(x) be dieren able. The dieren al of x, denoted dx, is any


nonzero real number (usually taken to be a small number). The dieren al of y, denoted dy, is
dy = f (x)dx.

We can solve for f (x) in the above equa on: f (x) = dy/dx. This states that
the deriva ve of f with respect to x is the dieren al of y divided by the dieren al of x; this is not the alternate nota on for the deriva ve, dy
dx . This la er
nota on was chosen because of the frac onlike quali es of the deriva ve, but
again, it is one symbol and not a frac on.
It is helpful to organize our new concepts and nota ons in one place.
Key Idea

Dieren al Nota on

Let y = f(x) be a dieren able func on.


. x represents a small, nonzero change in x value.
. dx represents a small, nonzero change in x value (i.e., x = dx).
. y is the change in y value as x changes by x; hence
y = f(x + x) f(x).
. dy = f (x)dx which, by Equa on ( . ), is an approxima on of the
change in y value as x changes by x; dy y.

Notes:

Dieren als

Chapter

Applica ons of the Deriva ve


What is the value of dieren als? Like many mathema cal concepts, dieren als provide both prac cal and theore cal benets. We explore both here.
Example
Finding and using dieren als
Consider f(x) = x . Knowing f( ) = , approximate f( . ).
The x value is changing from x = to x = . ; therefore, we
S
see that dx = . . If we know how much the y value changes from f( ) to f( . )
(i.e., if we know y), we will know exactly what f( . ) is (since we already know
f( )). We can approximate y with dy.
y dy

= f ( )dx
= . = . .

We expect the y value to change by about . , so we approximate f( . )


. .
We leave it to the reader to verify this, but the preceding discussion links the
dieren al to the tangent line of f(x) at x = . One can verify that the tangent
line, evaluated at x = . , also gives y = . .
Of course, it is easy to compute the actual answer (by hand or with a calculator): . = . . (Before we get too cynical and say Then why bother?, note
our approxima on is really good!)
So why bother?
In most real life situa ons, we do not know the func on that describes
a par cular behavior. Instead, we can only take measurements of how things
change measurements of the deriva ve.
Imagine water owing down a winding channel. It is easy to measure the
speed and direc on (i.e., the velocity) of water at any loca on. It is very hard
to create a func on that describes the overall ow, hence it is hard to predict
where a oa ng object placed at the beginning of the channel will end up. However, we can approximate the path of an object using dieren als. Over small
intervals, the path taken by a oa ng object is essen ally linear. Dieren als
allow us to approximate the true path by piecing together lots of short, linear
paths. This technique is called Eulers Method, studied in introductory Dierenal Equa ons courses.
We use dieren als once more to approximate the value of a func on. Even
though calculators are very accessible, it is neat to see how these techniques can
some mes be used to easily compute something that looks rather hard.

Notes:

.
Example
Using dieren als to approximate a func on value

Approximate
. .

. , yet we can do be er.Let f(x) = x,


We expect
and let c = . Thus f( ) = . We can compute f (x) = /( x), so f ( ) =
/ .
We approximate the dierence between f( . ) and f( ) using dieren als,
with dx = . :
S

f( . ) f( ) = y dy = f ( ) dx = / / = / = .
The
approximate change in f from x =
. .
.

to x = . is .

, so we approximate

Dieren als are important when we discuss integra on. When we study
that topic, we will use nota on such as

f(x) dx
quite o en. While we dont discuss here what all of that nota on means, note
the existence of the dieren al dx. Proper handling of integrals comes with
proper handling of dieren als.
In light of that, we prac ce nding dieren als in general.
Finding dieren als
Example
In each of the following, nd the dieren al dy.
. y = sin x

. y = ex (x + )

.y=

x + x

. y = sin x:

As f(x) = sin x, f (x) = cos x. Thus


dy = cos(x)dx.

. y = ex (x + ): Let f(x) = ex (x + ). We need f (x), requiring the


Product Rule.
We have f (x) = ex (x + ) + xex , so
)
(
dy = ex (x + ) + xex dx.
Notes:

Dieren als

Chapter

Applica ons of the Deriva ve

. y = x + x :
the Chain Rule.
We have f (x) =

Let f(x) =

x + x ; we need f (x), requiring

(x + x ) ( x + ) =
( x + )dx
dy =
.
x + x

x+
. Thus
x + x

Finding the dieren al dy of y = f(x) is really no harder than nding the


deriva ve of f; we just mul ply f (x) by dx. It is important to remember that we
are not simply adding the symbol dx at the end.
We have seen a prac cal use of dieren als as they oer a good method of
making certain approxima ons. Another use is error propaga on. Suppose a
length is measured to be x, although the actual value is x + x (where we hope
x is small). This measurement of x may be used to compute some other value;
we can think of this as f(x) for some func on f. As the true length is x + x,
one really should have computed f(x + x). The dierence between f(x) and
f(x + x) is the propagated error.
How close are f(x) and f(x + x)? This is a dierence in y values;
f(x + x) f(x) = y dy.
We can approximate the propagated error using dieren als.
Using dieren als to approximate propagated error
Example
A steel ball bearing is to be manufactured with a diameter of cm. The manufacturing process has a tolerance of . mm in the diameter. Given that the
density of steel is about . g/cm , es mate the propagated error in the mass
of the ball bearing.
The mass of a ball bearing is found using the equa on mass
S
= volume density. In this situa on the mass func on is a product of the radius
of the ball bearing, hence it is m = .
r . The dieren al of the mass is
dm =

. r dr.

The radius is to be cm; the manufacturing tolerance in the radius is .


or .
cm. The propagated error is approximately:
m dm

= . ( ) ( .
= .
g

Notes:

mm,

.
Is this error signicant? It certainly depends on the applica on, but we can get
an idea by compu ng the rela ve error. The ra o between amount of error to
the total mass is
dm
.
=
m

.
.
=
.
= .
,
or . %.
We leave it to the reader to conrm this, but if the diameter of the ball was
supposed to be cm, the same manufacturing tolerance would give a propagated error in mass of . g, which corresponds to a percent error of .
%.
While the amount of error is much greater ( . > .
), the percent error
is much lower.
We rst learned of the deriva ve in the context of instantaneous rates of
change and slopes of tangent lines. We furthered our understanding of the
power of the deriva ve by studying how it relates to the graph of a func on
(leading to ideas of increasing/decreasing and concavity). This chapter has put
the deriva ve to yet more uses:
Equa on solving (Newtons Method)
Related Rates (furthering our use of the deriva ve to nd instantaneous
rates of change)
Op miza on (applied extreme values), and
Dieren als (useful for various approxima ons and for something called
integra on).
In the next chapters, we will consider the reverse problem to compu ng
the deriva ve: given a func on f, can we nd a func on whose deriva ve is f?
Be able to do so opens up an incredible world of mathema cs and applica ons.

Notes:

Dieren als

Exercises .

Terms and Concepts

. y=

. T/F: Given a dieren able func on y = f(x), we are generally free to choose a value for dx, which then determines
the value of dy.

. y=

. T/F: The symbols dx and x represent the same concept.

. y = ex sin x

. T/F: The symbols dy and y represent the same concept.


. T/F: Dieren als are important in the study of integra on.
. How are dieren als and tangent lines related?

In Exercises
value by hand.
.

, use dieren als to approximate the given

. y = cos(sin x)
. y=
. y=

x+
x+
x

ln x

. A set of plas c spheres are to be made with a diameter


of cm. If the manufacturing process is accurate to mm,
what is the propagated error in volume of the spheres?
. The distance, in feet, a stone drops in t seconds is given by
d(t) = t . The depth of a hole is to be approximated by
dropping a rock and listening for it to hit the bo om. What
is the propagated error if the me measurement is accurate
to / ths of a second and the measured me is:
(a)
(b)

seconds?
seconds?

. What is the propagated error in the measurement of the


cross sec onal area of a circular log if the diameter is measured at , accurate to / ?

. A wall is to be painted that is high and is measured to


be , long. Find the propagated error in the measurement of the walls surface area if the measurement is accurate to / .

Exercises explore some issues related to surveying in


which distances are approximated using other measured distances and measured angles. (Hint: Convert all angles to radians before compu ng.)

. sin
. cos .
. e

. y = ln( x)

. y = x ln x x

Problems

x
tan x +

In Exercises

, compute the dieren al dy.

. The length l of a long wall is to be approximated. The angle


, as shown in the diagram (not to scale), is measured to be
. , accurate to . Assume that the triangle formed is a
right triangle.

. y=x + x
. y=x x
. y=

. y = ( x + sin x)
. y=x e

l =?

(a) What is the measured length l of the wall?


(b) What is the propagated error?
(c) What is the percent error?

. Answer the ques ons of Exercise , but with a measured

angle of . , accurate to , measured from a point


from the wall.
. The length l of a long wall is to be calculated by measuring
the angle shown in the diagram (not to scale). Assume
the formed triangle is an isosceles triangle. The measured

angle is
, accurate to .

l =?

(a) What is the measured length of the wall?


(b) What is the propagated error?
(c) What is the percent error?
. The length of the walls in Exercises

are essen ally


the same. Which setup gives the most accurate result?
. Consider the setup in Exercises . This me, assume the

angle measurement of
is exact but the measured
from the wall is accurate to . What is the approximate
percent error?

:I
We have spent considerable me considering the deriva ves of a func on and
their applica ons. In the following chapters, we are going to star ng thinking
in the other direc on. That is, given a func on f(x), we are going to consider
func ons F(x) such that F (x) = f(x). There are numerous reasons this will
prove to be useful: these func ons will help us compute areas, volumes, mass,
force, pressure, work, and much more.

An deriva ves and Indenite Integra on

Given a func on y = f(x), a dieren al equa on is one that incorporates y, x,


and the deriva ves of y. For instance, a simple dieren al equa on is:
y = x.
Solving a dieren al equa on amounts to nding a func on y that sa ses
the given equa on. Take a moment and consider that equa on; can you nd a
func on y such that y = x?
Can you nd another?
And yet another?
Hopefully one was able to come up with at least one solu on: y = x . Finding another may have seemed impossible un l one realizes that a func on like
y = x + also has a deriva ve of x. Once that discovery is made, nding yet
another is not dicult; the func on y = x +
,
,
also has a derivave of x. The dieren al equa on y = x has many solu ons. This leads us
to some deni ons.
Deni on

An deriva ves and Indenite Integrals

Let a func on f(x) be given. An an deriva ve of f(x) is a func on F(x)


such that F (x) = f(x).
The set of all an deriva ves of f(x) is the indenite integral of f, denoted
by

f(x) dx.

Make a note about our deni on: we refer to an an deriva ve of f, as opposed to the an deriva ve of f, since there is always an innite number of them.

Chapter

Integra on
We o en use upper-case le ers to denote an deriva ves.
Knowing one an deriva ve of f allows us to nd innitely more, simply by
adding a constant. Not only does this give us more an deriva ves, it gives us all
of them.
Theorem

An deriva ve Forms

Let F(x) and G(x) be an deriva ves of f(x). Then there exists a constant
C such that
G(x) = F(x) + C.
Given a func on f and one of its an deriva ves F, we know all an deriva ves
of f have the form F(x) + C for some constant C. Using Deni on , we can say
that

f(x) dx = F(x) + C.

Lets analyze this indenite integral nota on.


Integra on
symbol

Dieren al
of x

Constant of
integra on

f(x) dx = F(x) + C

Integrand

One
an deriva ve

Figure . : Understanding
the indenite integral nota on.
.

Figure . shows the typical nota on of the indenite integral. The integraon symbol, , is in reality an elongated S, represen ng take the sum. We
will later see how sums and an deriva ves are related.
The func on we want to nd an an deriva ve of is called the integrand. It
contains the dieren al of the variable we are integra ng with respect to. The
symbol and the dieren al dxare not bookends with a func on sandwiched in
between; rather, the symbol means nd all an deriva ves of what follows,
and the func on f(x) and dx are mul plied together; the dx does not just sit
there.
Lets prac ce using this nota on.
Example

Evaluate

Notes:

Evalua ng indenite integrals


sin x dx.

An deriva ves and Indenite Integra on

We are asked to nd all func ons F(x) such that F (x) =


d
sin x. Some thought will lead us to one solu on: F(x) = cos x, because dx
( cos x) =
sin x.
The indenite integral of sin x is thus cos x, plus a constant of integra on.
So:

sin x dx = cos x + C.
S

A commonly asked ques on is What happened to the dx? The unenlightened response is Dont worry about it. It just goes away. A full understanding
includes the following.
This process of an dieren a on is really solving a dieren al ques on. The
integral

sin x dx

presents us with a dieren al, dy = sin x dx. It is asking: What is y? We found


lots of solu ons, all of the form y = cos x + C.
Le ng dy = sin x dx, rewrite

sin x dx as
dy.

This is asking: What func ons have a dieren al of the form dy? The answer
is Func ons of the form y + C, where C is a constant. What is y? We have lots
of choices, all diering by a constant; the simplest choice is y = cos x.
Understanding all of this is more important later as we try to nd an derivaves of more complicated func ons. In this sec on, we will simply explore the
rules of indenite integra on, and one can succeed for now with answering
What happened to the dx? with It went away.
Lets prac ce once more before sta ng integra on rules.
Example
Evaluate

Evalua ng indenite integrals


( x + x + ) dx.

We seek a func on F(x) whose deriva ve is x + x + .


S
When taking deriva ves, we can consider func ons termbyterm, so we can
likely do that here.
What func ons have a deriva ve of x ? Some thought will lead us to a
cubic, specically x + C , where C is a constant.
What func ons have a deriva ve of x? Here the x term is raised to the rst
power, so we likely seek a quadra c. Some thought should lead us to x + C ,
where C is a constant.

Notes:

Chapter

Integra on
Finally, what func ons have a deriva ve of ? Func ons of the form x + C ,
where C is a constant.
Our answer appears to be

( x + x + ) dx = x + C + x + C + x + C .

We do not need three separate constants of integra on; combine them as one
constant, giving the nal answer of

( x + x + ) dx = x + x + x + C.

It is easy to verify our answer; take the deriva ve of x + x + x + C and


see we indeed get x + x + .
This nal step of verifying our answer is important both prac cally and
theore cally. In general, taking deriva ves is easier than nding an deriva ves
so checking our work is easy and vital as we learn.
We also see that taking the deriva ve of our answer returns the func on in
the integrand. Thus we can say that:

d
dx

)
f(x) dx = f(x).

Dieren a on undoes the work done by an dieren a on.


Theorem gave a list of the deriva ves of common func ons we had learned
at that point. We restate part of that list here to stress the rela onship between
deriva ves and an deriva ves. This list will also be useful as a glossary of common an deriva ves as we learn.

Notes:

Theorem

Deriva ves and An deriva ves

Common Dieren a on Rules Common Indenite Integral Rules


(
)

d
. dx
cf(x) = c f (x)
. c f(x) dx = c f(x) dx
)
(
)
(
d
. dx
f(x) g(x) dx =
.
f(x) g(x) =

f(x) dx g(x) dx
f (x) g (x)
( )

d
.
dx = C
C =
. dx
(
)

d
x =
.
dx = dx = x + C
. dx
( n)

d
. dx
x = n xn
. xn dx = n+ xn+ + C (n =
(
)

d
. dx
sin x = cos x
. cos x dx = sin x + C
(
)

d
cos x = sin x
. dx
. sin x dx = cos x + C
(
)

d
. dx
tan x = sec x
. sec x dx = tan x + C
(
)

d
csc x = csc x cot x
. dx
. csc x cot x dx = csc x + C
(
)

d
. dx
sec x = sec x tan x
. sec x tan x dx = sec x + C
(
)

d
cot x = csc x
. dx
. csc x dx = cot x + C
( x)

d
. dx
e = ex
. ex dx = ex + C
( x)

d
. dx
a = ln a ax
. ax dx = ln a ax + C
(
)

d
. dx
ln x = x
. x dx = ln |x| + C

We highlight a few important points from Theorem :

Rule # states c f(x) dx = c f(x) dx. This is the Constant Mul ple
Rule: we can temporarily ignore constants when nding
( ) an deriva ves,
d
x is just as easy to
just as we did when compu ng deriva ves (i.e., dx
( )
d
compute as dx x ). An example:

cos x dx = cos x dx = (sin x + C) = sin x + C.


In the last step we can consider the constant as also being mul plied by

Notes:

An deriva ves and Indenite Integra on

Chapter

Integra on
, but

mes a constant is s ll a constant, so we just write C .

Rule # is the Sum/Dierence Rule: we can split integrals apart when the
integrand contains terms that are added/subtracted, as we did in Example
. So:

( x + x + ) dx =
x dx +
x dx +
dx

=
x dx +
x dx +
dx
=

x + x + x+C

=x + x + x+C

In prac ce we generally do not write out all these steps, but we demonstrate them here for completeness.
Rule # is the Power Rule of indenite integra on. There are two important things to keep in mind:

. No ce the restric on that n = . This is important: x dx =


x + C; rather, see Rule # .
. We are presen ng an dieren a on as the inverse opera on of
dieren a on. Here is a useful quote to remember:
Inverse opera ons do the opposite things in the opposite
order.
When taking a deriva ve using the Power Rule, we rst mul ply by
the power, then second subtract from the power. To nd the anderiva ve, do the opposite things in the opposite order: rst add
one to the power, then second divide by the power.
Note that Rule # incorporates the absolute value of x. The exercises will
work the reader through why this is the case; for now, know the absolute
value is important and cannot be ignored.

Ini al Value Problems


In Sec on . we saw that the deriva ve of a posi on func on gave a velocity
func on, and the deriva ve of a velocity func on describes accelera on. We can
now go the other way: the an deriva ve of an accelera on func on gives a
velocity func on, etc. While there is just one deriva ve of a given func on, there
are innite an deriva ves. Therefore we cannot ask What is the velocity of an
object whose accelera on is
/s ?, since there is more than one answer.
Notes:

.
We can nd the answer if we provide more informa on with the ques on,
as done in the following example. O en the addi onal informa on comes in the
form of an ini al value, a value of the func on that one knows beforehand.
Solving ini al value problems
Example
The accelera on due to gravity of a falling object is
/s . At me t = ,
a falling object had a velocity of
/s. Find the equa on of the objects
velocity.
We want to know a velocity func on, v(t). We know two

things:

The accelera on, i.e., v (t) =

, and

the velocity at a specic me, i.e., v( ) =

Using the rst piece of informa on, we know that v(t) is an an deriva ve of
v (t) = . So we begin by nding the indenite integral of :

( ) dt = t + C = v(t).
Now we use the fact that v( ) =

to nd C:

v(t) =

t+C

v( ) =
( )+C=
C=

Thus v(t) = t + . We can use this equa on to understand the mo on


of the object: when t = , the object had a velocity of v( ) =
/s. Since the
velocity is posi ve, the object was moving upward.
When did the object begin moving down? Immediately a er v(t) = :

t+

t=

s.

Recognize that we are able to determine quite a bit about the path of the object
knowing just its accelera on and its velocity at a single point in me.
Solving ini al value problems
Example
Find f(t), given that f (t) = cos t, f ( ) = and f( ) = .
S

Notes:

We start by nding f (t), which is an an deriva ve of f (t):

f (t) dt = cos t dt = sin t + C = f (t).

An deriva ves and Indenite Integra on

Chapter

Integra on

so:

So f (t) = sin t + C for the correct value of C. We are given that f ( ) = ,


f ( ) =

sin + C =

C= .

Using the ini al value, we have found f (t) = sin t + .


We now nd f(t) by integra ng again.

f(t) = f (t) dt = (sin t + ) dt = cos t + t + C.

We are given that f( ) = , so

cos + ( ) + C =
+C=
C=
Thus f(t) = cos t + t + .
This sec on introduced an deriva ves and the indenite integral. We found
they are needed when nding a func on given informa on about its derivave(s). For instance, we found a posi on func on given a velocity func on.
In the next sec on, we will see how posi on and velocity are unexpectedly
related by the areas of certain regions on a graph of the velocity func on. Then,
in Sec on . , we will see how areas and an deriva ves are closely ed together.

Notes:

Exercises .

Terms and Concepts


. Dene the term an deriva ve in your own words.
. Is it more accurate to refer to the an deriva ve of f(x) or
an an deriva ve of f(x)?
. Use your own words to dene the indenite integral of
f(x).
. Fill in the blanks: Inverse opera ons do the
things in the
order.

( t + ) dt

(t + )(t t) dt

x x dx

e dx

a dx

. What is an ini al value problem?


. The deriva ve of a posi on func on is a
on.

func-

. The an deriva ve of an accelera on func on is a


func on.

Problems
In Exercises

, evaluate the given indenite integral.

e d
t

dt
dt

.
This problem inves gates why Theorem
dx = ln |x| + C.
x
(a) What is the domain of y = ln x?
(
)
d
ln x .
dx

x dx

dt

ds

dx
x

. f (x) =

sec d

. f (x) =

sin d

(sec x tan x + csc x cot x) dx

x dx

x ) dx

(b) Find

(c) What is the domain of y = ln(x)?


(
)
d
ln(x) .
dx

(d) Find

(e) You should nd that /x has two types of an derivaves, depending on whether x > or x < . In

dx that takes
x
these dierent domains into account, and explain
your answer.

one expression, give a formula for

In Exercises

value problem.

, nd f(x) described by the given ini al

. f (x) = sin x and f( ) =


dt
dt

states that

. f (x) = ex and f( ) =
. f (x) = x x and f( ) =
. f (x) = sec x and f(/ ) =
x

and f( ) =
and f ( ) = , f( ) =

. f (x) = x and f ( ) = , f( ) =
. f (x) = ex and f ( ) = , f( ) =
. f () = sin and f () = , f() =

. f (x) =

x +

cos x and f ( ) = , f( ) =

Review
. Use informa on gained from the rst and second derivaves to sketch f(x) =

. f (x) =

and f ( ) = , f( ) =

ex +

. Given y = x ex cos x, nd dy.

The Denite Integral

y ( /s)

We start with an easy problem. An object travels in a straight line at a constant


velocity of
/s for
seconds. How far away from its star ng point is the
object?
We approach this problem with the familiar Distance = Rate Time equaon. In this case, Distance = /s s = feet.
It is interes ng to note that this solu on of feet can be represented graphically. Consider Figure . , where the constant velocity of /s is graphed on the
axes. Shading the area under the line from t = to t =
gives a rectangle
with an area of
square units; when one considers the units of the axes, we
can say this area represents
.
Now consider a slightly harder situa on (and not par cularly realis c): an
object travels in a straight line with a constant velocity of /s for
seconds,
then instantly reverses course at a rate of /s for seconds. (Since the object
is traveling in the opposite direc on when reversing course, we say the velocity
is a constant /s.) How far away from the star ng point is the object what
is its displacement?
Here we use Distance = Rate Time + Rate Time , which is
Distance =

+ ( )

Area below the taxis,

Finding posi on using velocity


Example
The velocity of an object moving straight up/down under the accelera on of
gravity is given as v(t) = t+ , where me t is given in seconds and velocity
is in /s. When t = , the object had a height of
.
. What was the ini al velocity of the object?
. What was the maximum height of the object?
. What was the height of the object at me t = ?
S

Notes:

t (s)

y ( /s)
5

which is easy to calculate as =


feet.
Now consider a more dicult problem.

v( ) =

10

Figure . : The area under a constant


velocity func on corresponds to distance
traveled.

Hence the object is feet from its star ng loca on.


We can again depict this situa on graphically. In Figure . we have the
veloci es graphed as straight lines on [ , ] and [ , ], respec vely. The displacement of the object is
Area above the taxis

The Denite Integral

It is straigh orward to nd the ini al velocity; at me t = ,


=
/s.

t (s)

Figure . : The total displacement is the


area above the taxis minus the area below the taxis.

Chapter

Integra on
To answer ques ons about the height of the object, we need to nd the
objects posi on func on s(t). This is an ini al value problem, which we studied
in the previous sec on. We are told the ini al height is , i.e., s( ) = . We
know s (t) = v(t) = t + . To nd s, we nd the indenite integral of v(t):

v(t) dt = ( t + ) dt = t + t + C = s(t).
Since s( ) = , we conclude that C = and s(t) = t + t.
To nd the maximum height of the object, we need to nd the maximum of
s. Recalling our work nding extreme values, we nd the cri cal points of s by
se ng its deriva ve equal to and solving for t:
s (t) =

t+

t=

= . s.

(No ce how we ended up just nding when the velocity was /s!) The rst
deriva ve test shows this is a maximum, so the maximum height of the object
is found at
s( . ) = ( . ) + ( . ) =
.
The height at me t =

y ( /s)
5

t (s)

5
.

Figure . : A graph of v(t) = t +


; the shaded areas help determine displacement.

is now straigh orward to compute: it is s( ) =

While we have answered all three ques ons, lets look at them again graphically, using the concepts of area that we explored earlier.
Figure . shows a graph of v(t) on axes from t = to t = . It is again
straigh orward to nd v( ). How can we use the graph to nd the maximum
height of the object?
Recall how in our previous work that the displacement of the object (in this
case, its height) was found as the area under the velocity curve, as shaded in the
gure. Moreover, the area between the curve and the taxis that is below the
taxis counted as nega ve area. That is, it represents the object coming back
toward its star ng posi on. So to nd the maximum distance from the star ng
point the maximum height we nd the area under the velocity line that is
above the taxis, i.e., from t = to t = . . This region is a triangle; its area is
Area =

Base Height =

. s

/s =

which matches our previous calcula on of the maximum height.


Finally, we nd the total signed area under the velocity func on from t =
to t = to nd the s( ), the height at t = , which is a displacement, the
distance from the current posi on to the star ng posi on. That is,
Displacement = Area above the taxis Area below taxis.

Notes:

.
The regions are triangles, and we nd
Displacement =

( . s)(

/s) (. s)(

/s) =

This also matches our previous calcula on of the height at t = .


No ce how we answered each ques on in this example in two ways. Our rst
method was to manipulate equa ons using our understanding of an deriva ves
and deriva ves. Our second method was geometric: we answered ques ons
looking at a graph and nding the areas of certain regions of this graph.
The above example does not prove a rela onship between area under a velocity func on and displacement, but it does imply a rela onship exists. Sec on
. will fully establish fact that the area under a velocity func on is displacement.
Given a graph of a func on y = f(x), we will nd that there is great use in
compu ng the area between the curve y = f(x) and the x-axis. Because of this,
we need to dene some terms.
Deni on

The Denite Integral, Total Signed Area

Let y = f(x) be dened on a closed interval [a, b]. The total signed area
from x = a to x = b under f is:
(area under f and above the xaxis on [a, b]) (area above f and under
the xaxis on [a, b]).
The denite integral of f on [a, b] is the total signed area of f on [a, b],
denoted

f(x) dx,

where a and b are the bounds of integra on.

By our deni on, the denite integral gives the signed area under f. We
usually drop the word signed when talking about the denite integral, and
simply say the denite integral gives the area under f or, more commonly,
the area under the curve.
The previous sec on introduced the indenite integral, which related to anderiva ves. We have now dened the denite integral, which relates to areas
under a func on. The two are very much related, as well see when we learn
the Fundamental
Theorem of Calculus in Sec on . . Recall that earlier we said

that the symbol was an elongated S that represented nding a sum. In


the context of the denite integral, this nota on makes a bit more sense, as we
are adding up areas under the func on f.

Notes:

The Denite Integral

Chapter

Integra on
We prac ce using this nota on.

Evalua ng denite integrals


Example
Consider the func on f given in Figure . .
Find:
x

Figure . : A graph of f(x) in Example

f(x) dx

f(x) dx

f(x) dx

f(x) dx
f(x) dx

f(x) dx is the area under f on [ , ]. This is sketched in Figure . .


Again, the region is a triangle, with height
mes that of the height of

the original triangle. Thus the area is


f(x) dx = / = . .

Figure . : A graph of f in Example


.
(Yes, it looks just like the graph of f in Figure . , just with a dierent y-scale.)

f(x) dx is the area under f on the interval [ , ]. This region is a triangle,

so the area is f(x) dx = ( )( ) = . .


f(x) dx represents the area of the triangle found under the xaxis on
[ , ]. The area is ( )( ) = ; since it is found under the xaxis, this is

nega ve area. Therefore f(x) dx = .


f(x) dx is the total signed area under f on [ , ]. This is . +( ) = . .

f(x) dx is the area under f on the interval [ , ]. This describes a line


segment, not a region; it has no width. Therefore the area is .

This example illustrates some of the proper es of the denite integral, given
here.

Notes:

Theorem

Proper es of the Denite Integral

Let f and g be dened on a closed interval I that contains the values a, b


and c, and let k be a constant. The following hold:
a
f(x) dx =
.
a

a
b
a
b
a
b
a

f(x) dx +

f(x) dx =
(

f(x) dx =

a
b

f(x) dx

f(x) dx

)
f(x) g(x) dx =

k f(x) dx = k

b
a

f(x) dx

b
a

g(x) dx

f(x) dx

We give a brief jus ca on of Theorem

here.

. As demonstrated in Example
, there is no area under the curve when
the region has no width; hence this denite integral is .
. This states that total area is the sum of the areas of subregions. It is easily
considered when we let a < b < c. We can break the interval [a, c] into
two subintervals, [a, b] and [b, c]. The total area over [a, c] is the area over
[a, b] plus the area over [b, c].
It is important to note that this s ll holds true even if a < b < c is not
true. We discuss this in the next point.
. This property can be viewed a merely a conven on to make other properes work well. (Later we will see how this property has a jus ca on all its
own, not necessarily in support of other proper es.) Suppose b < a < c.
The discussion from the previous point clearly jus es
c
c
a
f(x) dx.
( . )
f(x) dx =
f(x) dx +
b

However, we s ll claim that, as originally stated,


c
c
b
f(x) dx.
f(x) dx =
f(x) dx +
a

Notes:

( . )

The Denite Integral

Chapter

Integra on
How do Equa ons ( . ) and ( . ) relate? Start with Equa on ( . ):
c
c
a
f(x) dx
f(x) dx =
f(x) dx +
b
a
b
c
a
c
f(x) dx
f(x) dx +
f(x) dx =
b

Property ( ) jus
es changing the sign and switching the bounds of intea
f(x) dx term; when this is done, Equa ons ( . ) and
gra on on the
b

( . ) are equivalent.

The conclusion is this: by adop ng the conven on of Property ( ), Property ( ) holds no ma er the order of a, b and c. Again, in the next sec on
we will see another jus ca on for this property.
, . Each of these may be nonintui ve. Property ( ) states that when one
scales a func on by, for instance, , the area of the enclosed region also
is scaled by a factor of . Both Proper es ( ) and ( ) can be proved using
geometry. The details are not complicated but are not discussed here.
Evalua ng denite integrals using Theorem .
Example
Consider the graph of a func on f(x) shown in Figure . . Answer the following:
c
b
f(x) dx?
f(x) dx or
. Which value is greater:

. Is

f(x) dx greater or less than ?

. Which value is greater:


.
Figure . : A graph of a func on in Example
.

b
a

f(x) dx or

b
c

f(x) dx?

b
f(x) dx has a posi ve value (since the area is above the xaxis) whereas
ac
b
f(x)
dx has a nega ve value. Hence a f(x) dx is bigger.
b
c
. a f(x) dx is the total signed area under f between x = a and x = c. Since
the region below the xaxis looks to be larger than the region above, we
conclude that the denite integral has a value less than .
b
. Note how the second integral has the bounds reversed. Therefore c f(x) dx
represents a posi ve number, greater than the area described by the rst
b
denite integral. Hence c f(x) dx is greater.
.

Notes:

.
The area deni on of the denite integral allows us to use geometry compute the denite integral of some simple func ons.

The Denite Integral


y
(5, 6)

Evalua ng denite integrals using geometry


Example
Evaluate the following denite integrals:

x dx.
( x ) dx
.
.

5
R

R
5

( , 8)

(a)

. It is useful to sketch the func on in the integrand, as shown in Figure


. (a). We see we need to compute the areas of two regions, which we
have labeled R and R . Both are triangles, so the area computa on is
straigh orward:
R :

( )( ) =

R :

y
5

( ) = .

Region R lies under the xaxis, hence it is counted as nega ve area (we
can think of the triangles height as being ), so

( x ) dx = + = .

. 3
(b)

. Recognize that the integrand of this denite integral describes a half circle,
as sketched in Figure . (b), with radius . Thus the area is:

x dx = r = .

Figure . : Agraph of f(x) = x in (a)


x in (b), from Example
and f(x) =
.

Understanding mo on given velocity


Example
Consider the graph of a velocity func on of an object moving in a straight line,
given in Figure . , where the numbers in the given regions gives the area of that
region. Assume that the denite integral of a velocity func on gives displacement. Find the maximum speed of the object and its maximum displacement
from its star ng posi on.

y ( /s)
15
10

38

Since the graph gives velocity, nding the maximum speed


is simple: it looks to be
/s.
At me t = , the displacement is ; the object is at its star ng posi on. At
me t = a, the object has moved backward
feet. Between mes t = a and
S

Notes:

11

t (s)

11

.
Figure . : A graph of a velocity in Example
.

Chapter

Integra on
t = b, the object moves forward
feet, bringing it into a posi on
feet forward of its star ng posi on. From t = b to t = c the object is moving backwards
again, hence its maximum displacement is feet from its star ng posi on.

Figure . : What is the area below y =


x on [ , ]? The region is not a usual geometric shape.

In our examples, we have either found the areas of regions that have nice
geometric shapes (such as rectangles, triangles and circles) or the areas were
given to us. Consider Figure . , where a region below y = x is shaded. What
is its area? The func on y = x is rela vely simple, yet the shape it denes has
an area that is not simple to nd geometrically.
In the next sec on we will explore how to nd the areas of such regions.

Notes:

Exercises .

Terms and Concepts

. What is total signed area?

y = f(x)

. What is displacement?
. What is

sin x dx?

(a)

. Give a single denite integral that has the same value as

( x + ) dx +
( x + ) dx.

(b)
(c)

f(x) dx

(d)

f(x) dx

(e)

x dx
( x ) dx

(f)

( x ) dx

(x ) dx

(d)

(x ) dx

(x ) dx

(e)

f(x) dx

Problems
In Exercises , a graph of a func on f(x) is given. Using
the geometry of the graph, evaluate the denite integrals.

y=x

y
4

y = 2x + 4

.
2

(a)
(b)
(c)

( x + ) dx

(d)

( x + ) dx

(e)

( x + ) dx

(f)

(a)

( x + ) dx

(b)

( x + ) dx
( x +

(c)

) dx

(x ) dx

(f)

(
)
(x ) + dx

f(x) dx

(c)

f(x) dx

f(x) dx

(d)

(x ) dx

x
f(x) =

y = f(x)

(x )

(a)
(b)
(c)

f(x) dx

(d)

f(x) dx

(e)

f(x) dx

(f)

f(x) dx

f(x) dx

(a)

f(x) dx

(b)

f(x) dx

In Exercises
, a graph of a func on f(x) is given; the
numbers inside the shaded regions give the area of that region. Evaluate the denite integrals using this area informaon.

f(x) = x

y
y = f(x)

(a)

59

(b)

(a)
(b)

f(x) dx

(c)

f(x) dx

(d)

f(x) dx

x dx
(x + ) dx

(c)
(d)

(x ) dx
(
)
(x ) + dx

In Exercises , a graph of the velocity func on of an object moving in a straight line is given. Answer the ques ons
based on that graph.

f(x) dx

7/

y ( /s)

.
t (s)

f(x) = sin(x/ )

(a) What is the objects maximum velocity?

(b) What is the objects maximum displacement?

(a)
(b)

f(x) dx

(c)

f(x) dx

(d)

(c) What is the objects total displacement on [ , ]?

f(x) dx

y ( /s)

f(x) dx

y
f(x) = x

t (s)

(a) What is the objects maximum velocity?

(b) What is the objects maximum displacement?


(c) What is the objects total displacement on [ , ]?

(a)

(b)

f(x) dx

(c)

(d)

f(x) dx

f(x) dx

f(x) dx

. An object is thrown straight up with a velocity, in /s, given


by v(t) = t + , where t is in seconds, from a height
of feet.
(a) What is the objects maximum velocity?
(b) What is the objects maximum displacement?
(c) When does the maximum displacement occur?
(d) When will the object reach a height of ? (Hint: nd
when the displacement is
.)

. An object is thrown straight up with a velocity, in /s, given


by v(t) = t + , where t is in seconds, from a height
of feet.

In Exercises
, let

s(t) dt = ,

(a) What is the objects ini al velocity?

(b) When is the objects displacement ?

(c) How long does it take for the object to return to its
ini al height?
(d) When will the object reach a height of
In Exercises

feet?

, let

f(x) dx = ,

f(x) dx = ,

g(x) dx = , and

g(x) dx = .

Use these values to evaluate the given denite integrals.


.

(
)
f(x) + g(x) dx

(
)
f(x) g(x) dx

)
f(x) + g(x) dx

. Find values for a and b such that

(
)
af(x) + bg(x) dx =

s(t) dt = ,

r(t) dt = , and

r(t) dt =

Use these values to evaluate the given denite integrals.


.

(
)
s(t) + r(t) dt

(
)
s(t) r(t) dt

)
s(t) r(t) dt

.
Find values for a and b such that
(
)
ar(t) + bs(t) dt =

Review
In Exercises
.

, evaluate the given indenite integral.

x x + x

)
sin x cos x + sec x dx

t+

t)

dx

dt

)
csc x cot x dx

Chapter

Integra on

.
y

.
Figure . : A graph of f(x) = x x .
What is the area of the shaded region?

RHR

MPR

LHR

other

Figure . : Approxima ng ( xx ) dx
using rectangles. The heights of the
rectangles are determined using dierent
rules.

Riemann Sums

In the previous sec on we dened the denite integral of a func on on [a, b] to


be the signed area between the curve and the xaxis. Some areas were simple
to compute; we ended the sec on with a region whose area was not simple to
compute. In this sec on we develop a technique to nd such areas.
A fundamental calculus technique is to rst answer a given problem with an
approxima on, then rene that approxima on to make it be er, then use limits
in the rening process to nd the exact answer. That is exactly what we will do
here.
Consider the region given in Figure . , which is the area under y = x x
on [ , ]. What is the signed area of this region i.e., what is ( x x ) dx?
We start by approxima ng. We can surround the region with a rectangle
with height and width of and nd the area is approximately
square units.
This is obviously an overapproxima on; we are including area in the rectangle
that is not under the parabola.
We have an approxima on of the area, using one rectangle. How can we
rene our approxima on to make it be er? The key to this sec on is this answer:
use more rectangles.
Lets use rectangles of equal width of . This par ons the interval [ , ]
into subintervals, [ , ], [ , ], [ , ] and [ , ]. On each subinterval we will
draw a rectangle.
There are three common ways to determine the height of these rectangles:
the Le Hand Rule, the Right Hand Rule, and the Midpoint Rule. The Le Hand
Rule says to evaluate the func on at the le hand endpoint of the subinterval
and make the rectangle that height. In Figure . , the rectangle drawn on the
interval [ , ] has height determined by the Le Hand Rule; it has a height of
f( ). (The rectangle is labeled LHR.)
The Right Hand Rule says the opposite: on each subinterval, evaluate the
func on at the right endpoint and make the rectangle that height. In the gure,
the rectangle drawn on [ , ] is drawn using f( ) as its height; this rectangle is
labeled RHR..
The Midpoint Rule says that on each subinterval, evaluate the func on at
the midpoint and make the rectangle that height. The rectangle drawn on [ , ]
was made using the Midpoint Rule, with a height of f( . ). That rectangle is
labeled MPR.
These are the three most common rules for determining the heights of approxima ng rectangles, but one is not forced to use one of these three methods.
The rectangle on [ , ] has a height of approximately f( . ), very close to the
Midpoint Rule. It was chosen so that the area of the rectangle is exactly the area
of the region under f on [ , ]. (Later youll be able to gure how to do this, too.)

The following example will approximate the value of ( x x ) dx using


Notes:

.
these rules.

Riemann Sums

Using the
Example
Le Hand, Right Hand and Midpoint Rules
Approximate the value of ( x x ) dx using the Le Hand Rule, the Right
Hand Rule, and the Midpoint Rule, using equally spaced subintervals.
We break the interval [ , ] into four subintervals as before.
S
In Figure . we see rectangles drawn on f(x) = x x using the Le Hand
Rule. (The areas of the rectangles are given in each gure.)
Note how in the rst subinterval, [ , ], the rectangle has height f( ) = . We
add up the areas of each rectangle (height width) for our Le Hand Rule approxima on:
f( ) + f( ) + f( ) + f( )
+ + +

=
=

Figure . : Approxima ng ( xx ) dx
using the Le Hand Rule in Example
.

Figure . shows rectangles drawn under f using the Right Hand Rule;
note how the [ , ] subinterval has a rectangle of height .
In this example, these rectangle seem to be the mirror image of those found
in Figure . . (This is because of the symmetry of our shaded region.) Our
approxima on gives the same answer as before, though calculated a dierent
way:
f( ) + f( ) + f( ) + f( )

+ + +

=
=

Figure . shows rectangles


drawn under f using the Midpoint Rule.
This gives an approxima on of ( x x ) dx as:
f( . ) + f( . ) + f( . ) + f( . )
.

+ .

+ .

Our three methods provide two approxima ons of

Summa on Nota on

+ .

Figure . : Approxima ng ( xx ) dx
using the Right Hand Rule in Example
.

( x x ) dx:

and

It is hard to tell at this moment which is a be er approxima on:


or ?
We can con nue to rene our approxima on by using more rectangles. The
nota on can become unwieldy, though, as we add up longer and longer lists of
numbers. We introduce summa on nota on to ameliorate this problem.

.7

.7

.7

.7

Figure . : Approxima ng ( xx ) dx
using the Midpoint Rule in Example
.

Notes:

Chapter

Integra on
Suppose we wish to add up a list of numbers a , a , a , , a . Instead of
wri ng
a +a +a +a +a +a +a +a +a ,
we use summa on nota on and write
upper
bound

summand
9

summa on
symbol
(an upper case
sigma)

ai .

i=1

i=index
.
of summa on

Figure .

lower
bound

: Understanding summa on nota on.

The upper case sigma represents the term sum. The index of summa on
in this example is i; any symbol can be used. By conven on, the index takes on
only the integer values between (and including) the lower and upper bounds.
Lets prac ce using this nota on.
Using summa on nota on
Example
Let the numbers {ai } be dened as ai = i for integers i, where i . So
a = , a = , a = , etc. (The output is the posi ve odd integers). Evaluate
the following summa ons:
.

ai

i=

i=

( ai )

(ai )

i=

ai = a + a + a + a + a + a

i=

+ + + + +

. Note the star ng value is dierent than :

i=

Notes:

ai = ( a ) + ( a ) + ( a ) + ( a ) + ( a )
=
=

+
.

.
.

(ai ) = (a ) + (a ) + (a ) + (a )

i=

=
=

It might seem odd to stress a new, concise way of wri ng summa ons only
to write each term out as we add them up. It is. The following theorem gives
some of the proper es of summa ons that allow us to work with them without
wri ng individual terms. Examples will follow.

Theorem

i=

i=m

c = c n, where c is a constant.
(ai bi ) =

c ai = c

ai +

i=m

Proper es of Summa ons

i=m

Example
Revisit Example

i=j+

i=m

ai

bi

i=m

i=

ai

i=m

Evalua ng summa ons using Theorem


and, using Theorem , evaluate

i=

ai =

i=

n(n + )

i =

n(n + )( n + )

i=

i=m

ai =

i=

ai

Notes:

i=

( i ).

i =

n(n + )

Riemann Sums

Chapter

Integra on
S

i=

( i )=
=

i=

i=

( + )

=
=

( )

i=

We obtained the same answer without wri ng out all six terms. When dealing
with small sizes of n, it may be faster to write the terms out by hand. However,
Theorem is incredibly important when dealing with large sums as well soon
see.

Riemann Sums

.
x

x9

Figure . : Dividing [ , ] into


spaced subintervals.

Consider again ( x x ) dx. We will approximate this denite integral


using
equally spaced subintervals and the Right Hand Rule in Example
.
Before doing so, it will pay to do some careful prepara on.
Figure . shows a number line of [ , ] divided into
equally spaced
subintervals. We denote as x ; we have marked the values of x , x , x and
x . We could mark them all, but the gure would get crowded. While it is easy
to gure that x = . , in general, we want a method of determining the value
of xi without consul ng the gure. Consider:

equally

number of
subintervals
between x1 and xi

xi = x1 + (i 1)x
.

star ng
value

So x = x + ( / ) = . .
If we had par oned [ , ] into
terval would have length x = /
x

=x +

subinterval
size

equally spaced subintervals, each subin= . . We could compute x as


( /

)= .

(That was far faster than crea ng a sketch rst.)

Notes:

.
Given any subdivision of [ , ], the rst subinterval is [x , x ]; the second is
[x , x ]; the i th subinterval is [xi , xi+ ].
When using the Le Hand Rule, the height of the i th rectangle will be f(xi ).
When using the Right Hand Rule, the height of the i th rectangle will be f(xi+ ).
(
)
xi + xi+
th
When using the Midpoint Rule, the height of the i rectangle will be f
.

Thus approxima ng ( x x ) dx with equally spaced subintervals can


be expressed as follows, where x = / = / :
Le Hand Rule:

f(xi )x

i=

Right Hand Rule:

f(xi+ )x

i=

Midpoint Rule:

( xi + xi+ )
x
f
i=

We use these formulas in the next two examples. The following example lets
us prac ce using the Right Hand Rule and the summa on formulas introduced
in Theorem .
Approxima ng denite integrals using sums
Example

Approximate ( x x ) dx using the Right Hand Rule and summa on formulas


with and
equally spaced intervals.
Using the formula derived before, using
equally spaced
S
intervals and the Right Hand Rule, we can approximate the denite integral as

f(xi+ )x.

i=

We have x = /

= .

. Since xi =

+ (i )x, we have

(
)
xi+ = + (i + ) x
= ix

Notes:

Riemann Sums

Chapter

Integra on
Using the summa on formulas, consider:

( x x ) dx
=

f(xi+ )x

f(ix)x

i=

i=

(
i=

i=

)
ix (ix) x

( ix i x )

= ( x )

i=

= ( x )
=
=

Figure . : Approxima ng ( xx ) dx
with the Right Hand Rule and
evenly
spaced subintervals.

i x

x
.

( . )

i=

)(

We were able to sum up the areas of


rectangles with very li le computaon. In Figure . the func on and the rectangles are graphed. While some
rectangles overapproximate the area, other underapproximate the area (by
about the same amount). Thus our approximate area of .
is likely a fairly
good approxima on.
No ce Equa on ( . ); by changing the s to ,
s (and appropriately
changing the value of x), we can use that equa on to sum up
rectangles! We do so here, skipping from the original summand to the equivalent of
Equa on ( . ) to save space. Note that x = /
= .
.

( x x ) dx

f(xi+ )x

i=

= ( x )

i=

= ( x )
=
=

Notes:

.
.

i x

i=

x
.

)(
,

.
Using many, many rectangles, we have a likely good approxima on of
x )x. That is,

( x x ) dx .
.

Riemann Sums

( x

Before the above example, we stated what the summa ons for the Le Hand,
Right Hand and Midpoint Rules looked like. Each had the same basic structure,
which was:
. each rectangle has the same width, which we referred to as x, and
. each rectangles height is determined by evalua ng f at a par cular point
in each subinterval. For instance, the Le Hand Rule states that each rectangles height is determined by evalua ng f at the le hand endpoint of
the subinterval the rectangle lives on.
One could par on an interval [a, b] with subintervals that did not have the same
size. We refer to the length of the rst subinterval as x , the length of the second subinterval as x , and so on, giving the length of the i th subinterval as xi .
Also, one could determine each rectangles height by evalua ng f at any point in
the i th subinterval. We refer to the point picked in the rst subinterval as c , the
point picked in the second subinterval as c , and so on, with ci represen ng the
point picked in the i th subinterval. Thus the height of the i th subinterval would
be f(ci ), and the area of the i th rectangle would be f(ci )xi .
Summa ons of rectangles with area f(ci )xi are named a er mathema cian
Georg Friedrich Bernhard Riemann, as given in the following deni on.
Deni on

Riemann Sum

Let f be dened on the closed interval [a, b] and let x be a par


[a, b], with
a = x < x < . . . < xn < xn+ = b.

on of
y

Let xi denote the length of the i th subinterval [xi , xi+ ] and let ci denote
any value in the i th subinterval.
The sum
n

f(ci )xi
i=

is a Riemann sum of f on [a, b].

Figure . shows the approxima ng rectangles of a Riemann sum of ( x


x ) dx. While the rectangles in this example do not approximate well the shaded
area, they demonstrate that the subinterval widths may vary and the heights of
the rectangles can be determined without following a par cular rule.
Notes:

.
Figure . : An example ofa general Riemann sum to approximate ( xx ) dx.

Chapter

Integra on
Usually Riemann sums are calculated using one of the three methods we
have introduced. The uniformity of construc on makes computa ons easier.
Before working another example, lets summarize some of what we have learned
in a convenient way.
Key Idea
Consider

Riemann Sum Concepts

b
a

f(x) dx

f(ci )xi .

i=

. When the n subintervals have equal length, xi = x =


. The i th term of the par
xn+ = b.)

ba
.
n

on is xi = a + (i )x. (This makes

. The Le Hand Rule summa on is:

f(xi )x.

i=

. The Right Hand Rule summa on is:

f(xi+ )x.

i=

(
)
n

xi + xx+
f
. The Midpoint Rule summa on is:
x.
i=

Lets do another example.


Approxima ng denite integrals with sums
Example

Approximate ( x + ) dx using the Midpoint Rule and


equally spaced
intervals.
Following Key Idea , we have

x =

( )

= /

and

xi = ( ) + ( / )(i ) = i/ / .

As we are using the Midpoint Rule, we will also need xi+ and

xi + xi+

xi = i/ / , xi+ = (i + )/ / = i/ . This gives


xi + xi+

Notes:

(i/ / ) + (i/ )

i /

= i/ / .

. Since

Riemann Sums

We now construct the Riemann sum and compute its value using summa on
formulas.

( xi + xi+ )
x
f
( x + ) dx

i=

i=

f(i/ / )x

(
i=

= x

[( )
i=

= x
=
=

(
=

(i/ / ) +

i=

(i)

(
i=

))
7

Note the graph of f(x) = x + in Figure . . The regions whose area is


computed by the denite integral are triangles, meaning we can nd the exact
answer without summa on techniques. We nd that the exact answer is indeed
. . One of the strengths of the Midpoint Rule is that o en each rectangle
includes area that should not be counted, but misses other area that should.
When the par on size is small, these two amounts are about equal and these
errors almost cancel each other out. In this example, since our func on is a
line, these errors are exactly equal and they do cancel each other out, giving us
the exact answer.
Note too that when the func on is nega ve, the rectangles have a nega ve
height. When we compute the area of the rectangle, we use f(ci )x; when f is
nega ve, the area is counted as nega ve.
No ce in the previous example that while we used equally spaced intervals, the number didnt play a big role in the calcula ons un l the very end.
Mathema cians love to abstract ideas; lets approximate the area of another region using n subintervals, where we do not specify a value of n un l the very end.
Approxima ng denite integrals with a formula, using sums
Example

Revisit ( xx ) dx yet again. Approximate this denite integral using the Right
Notes:

Figure . : Approxima ng ( x +
) dx using the Midpoint Rule and
evenly spaced subintervals in Example
.

Chapter

Integra on
Hand Rule with n equally spaced subintervals.
Using Key Idea , we know x = n = /n. We also nd
xi = + x(i ) = (i )/n. The Right Hand Rule uses xi+ , which is
xi+ = i/n.
We construct the Right Hand Rule Riemann sum as follows. Be sure to follow each step carefully. If you get stuck, and do not understand how one line
proceeds to the next, you may skip to the result and consider how this result
is used. You should come back, though, and work through each step for full
understanding.

f(xi+ )x
( x x ) dx
S

i=

( )
n

i
f
x
=
n
i=
[
( ) ]
n

i
i
=

x
n
n
i=
)
)
n (
n (

x
x
=
i
i
n
n
i=
i=
(
) n
) n
(
x
x
i
i
=
n
n
i=
i=
)
(
)
(
n(n + )
x n(n + )( n + )
x

=
n
n
(n + )
(n + )( n + )
=

(now simplify)
n
(n
)
=

recall
x = /n

The result is an amazing, easy to use formula. To approximate the denite


integral with equally spaced subintervals and the Right Hand Rule, set n =
and compute
(
)

( x x ) dx

= . .

Recall how earlier we approximated the denite integral with subintervals;


with n = , the formula gives , our answer as before.
It is now easy to approximate the integral with ,
,
subintervals! Handheld calculators will round o the answer a bit prematurely giving an answer of

Notes:

.
. (The actual answer is

.)

We now take an important leap. Up to this point, our mathema cs has been
limited to geometry and algebra (nding areas and manipula ng expressions).
Now we apply calculus. For any nite n, we know that

( x x ) dx

Both common sense and highlevel mathema cs tell us that as n gets large, the
approxima on gets be er.
In fact, if we take the limit as n , we get the
exact area described by ( x x ) dx. That is,

( x x ) dx = lim

( )

This is a fantas c result. By considering n equallyspaced subintervals, we obtained a formula for an approxima on of the denite integral that involved our
variable n. As n grows large without bound the error shrinks to zero and we
obtain the exact area.
This sec on started with a fundamental calculus technique: make an approxima on, rene the approxima on to make it be er, then use limits in the
rening process to get an exact answer. That is precisely what we just did.
Lets prac ce this again.
Approxima ng denite integrals with a formula, using sums
Example
Find a formula that approximates x dx using the Right Hand Rule and n
equally spaced subintervals, then take the limit as n to nd the exact
area.
)
= /n. We
Following Key Idea , we have x = (
n
have xi = ( ) + (i )x; as the Right Hand Rule uses xi+ , we have xi+ =
( ) + ix.

The Riemann sum corresponding to the Right Hand Rule is (followed by sim-

Notes:

Riemann Sums

Chapter

Integra on
plica ons):

f(xi+ )x
x dx

i=

=
=
=
=

i=
n

i=
n

i=
n

i=

= x

f( + ix)x
( + ix) x
(

(ix) (ix) + ix

i x i x + ix x

i=

= x

i x

n(n + )

i + x

i=

i=

(now distribute x)

(now split up summa on)

i=

n(n + )( n + )

+ x

n(n + )

nx

(use x = /n)

n (n + )

n(n + )( n + )

n(n + )
n

(now do a sizable amount of algebra to simplify)

Figure . : Approxima ng x dx using the Right Hand Rule and


evenly
spaced subintervals.

+
n
n
Once again, we have found a compact formula for approxima ng the denite
integral with n equally spaced subintervals and the Right Hand Rule. Using
subintervals, we have an approxima on of
. (these rectangles are shown
in Figure . ). Using n =
gives an approxima on of
.
.
Now nd the exact answer using a limit:
)
(

+
=
.
x dx = lim
+
n
n
n

Limits of Riemann Sums


We have used limits to evaluate exactly given denite limits. Will this always work? We will show, given notveryrestric ve condi ons, that yes, it will
always work.

Notes:

.
The previous two examples demonstrated how an expression such as
n

f(xi+ )x

i=

can be rewri en as an expression explicitly involving n, such as / ( /n ).


Viewed in this manner, we can think of the summa on as a func on of n.
An n value is given (where n is a posi ve integer), and the sum of areas of n
equally spaced rectangles is returned, using the Le Hand, Right Hand, or Midpoint Rules.
b
Given a denite integral a f(x) dx, let:
SL (n) =

f(xi )x, the sum of equally spaced rectangles formed using

i=

the Le Hand Rule,


SR (n) =

f(xi+ )x, the sum of equally spaced rectangles formed us-

i=

ing the Right Hand Rule, and


SM (n) =

(
)
n

xi + xi+
f
x, the sum of equally spaced rectangles
i=

formed using the Midpoint Rule.

Recall the deni on of a limit as n : lim SL (n) = K if, given any > ,
n
there exists N > such that
|SL (n) K| <

when

n N.

The following theorem states that we can use any of our three rules to nd
b
the exact value of a denite integral a f(x) dx. It also goes two steps further.
The theorem states that the height of each rectangle doesnt have to be determined following a specic rule, but could be f(ci ), where ci is any point in the i th
subinterval, as discussed before Riemann Sums where dened in Deni on .
The theorem goes on to state that the rectangles do not need to be of the
same width. Using the nota on of Deni on , let xi denote the length of
the i th subinterval in a par on of [a, b]. Now let ||x|| represent the length
of the largest subinterval in the par on: that is, ||x|| is the largest of all the
xi s. If ||x|| is small, then [a, b] must be par oned into many subintervals,
since all subintervals must have small lengths. Taking the limit as ||x|| goes
to zero implies that the number n of subintervals in the par on is growing to

Notes:

Riemann Sums

Chapter

Integra on
innity, as the largest subinterval length is becoming arbitrarily small. We then
interpret the expression
n

f(ci )xi
lim
||x||

i=

as the limit of the sum of rectangles, where the width of each rectangle can be
dierent but ge ng small, and the height of each rectangle is not necessarily
determined by a par cular rule. The theorem states that this Riemann Sum
also gives the value of the denite integral of f over [a, b].
Theorem

Denite Integrals and the Limit of Riemann Sums

Let f be con nuous on the closed interval [a, b] and let SL (n), SR (n) and
SM (n) be dened as before. Then:
. lim SL (n) = lim SR (n) = lim SM (n) = lim
n

. lim

lim

f(ci )x =

i=

i=

b
a

f(ci )xi =

f(ci )x,

i=

f(x) dx, and

b
a

f(x) dx.

We summarize what we have learned over the past few sec ons here.
Knowing the area under the curve can be useful. One common example
is: the area under a velocity curve is displacement.
b
We have dened the denite integral, a f(x) dx, to be the signed area
under f on the interval [a, b].
While we can approximate a denite integral many ways, we have focused
on using rectangles whose heights can be determined using: the Le Hand
Rule, the Right Hand Rule and the Midpoint Rule.
Sums of rectangles of this type are called Riemann sums.
The exact value of the denite integral can be computed using the limit of
a Riemann sum. We generally use one of the above methods as it makes
the algebra simpler.

Notes:

.
We rst learned of deriva ves through limits then learned rules that made
the process simpler. We know of a way to evaluate a denite integral using limits;
in the next sec on we will see how the Fundamental Theorem of Calculus makes
the process simpler. The key feature of this theorem is its connec on between
the indenite integral and the denite integral.

Notes:

Riemann Sums

Exercises .

Terms and Concepts

In Exercises
.

. A fundamental calculus technique is to use


ne approxima ons to get an exact answer.

i=

)?

. This sec on approximates denite integrals using what geometric shape?


. T/F: A sum using the Right Hand Rule is an example of a
Riemann Sum.

Problems

sin(i/ )

(i i + i + )

+ + + ... +

( ) cos(i)

+ + +

i=

+
+

i=

e+e e +e

ai +

i=

i=

, write each sum in summa on nota on.

. + + + +

+
+

ai , so

i=k+

ai

ai .

i=

Use this fact, along with other parts of Theorem


uate the summa ons given in Exercises .

In Exercises

ai =

i=k+

i+

i=

states

ai =

i=

i=

i i+

+ + + ... +

( )i i

i=

( i +

i=

i=

( i

Theorem

i=

i=

( i )

( i i)

i=

i=

i=

i=

i=

i=

In Exercises , write out each term of the summa on and


compute the sum.

, evaluate the summa on using Theorem

to re.

. What is the upper bound in the summa on

+
.

i=

, to eval-

In Exercises

, a denite integral

(b) Add to the sketch rectangles using the provided rule.


b
(c) Approximate
f(x) dx by summing the areas of the
a

rectangles.

x dx, with rectangles using the Le Hand Rule.

( x ) dx, with rectangles using the Midpoint Rule.

ln x dx, with rectangles using the Midpoint Rule.

sin x dx, with rectangles using the Right Hand Rule.


x

dx, with rectangles using the Le Hand Rule.

subintervals and the provided rule.


(b) Evaluate the formula using n =

f(x) dx using n
and ,

(c) Find the limit of


the formula, as n , to nd the
exact value of

f(x) dx.

( x ) dx, using the Le Hand Rule.

( x) dx, using the Right Hand Rule.


(x x ) dx, using the Right Hand Rule.

. f(x) =

, do the following.

(a) Find a formula to approximate

( x ) dx, using the Midpoint Rule.

, nd an an deriva ve of the given func-

. f(x) = sec x

f(x) dx is given. As demonstrated in Examples


b

x dx, using the Le Hand Rule.

In Exercises
on.

, a denite integral

x dx, using the Right Hand Rule.

Review

dx, with rectangles using the Right Hand Rule.

In Exercises

and

f(x) dx is given.

(a) Graph f(x) on [a, b].

. g(t) = t t +
. g(t) =

. g(t) = cos t + sin t


. f(x) =

Chapter

Integra on

b
Let f(t) be a con nuous func on dened on [a, b]. The denite integral a f(x) dx
is the area under f on [a, b]. We can turn this concept into a func on by le ng
the upper (or lower) bound vary.
x
Let F(x) = a f(t) dt. It computes the area under f on [a, x] as illustrated
in Figure . . We can study this func on using our knowledge of the denite
a
integral. For instance, F(a) = since a f(t) dt = .

The Fundamental Theorem of Calculus

Figure . : The
x area of the shaded region is F(x) = a f(t) dt.

We can also apply calculus ideas to F(x); in par cular, we can compute its
deriva ve. While this may seem like an innocuous thing to do, it has farreaching
implica ons, as demonstrated by the fact that the result is given as an important
theorem.
Theorem

The Fundamental Theorem of Calculus, Part


x
Let f be con nuous on [a, b] and let F(x) = a f(t) dt. Then F is a dieren able func on on (a, b), and
F (x) = f(x).

Ini ally this seems simple, as demonstrated in the following example.


Example
Let F(x) =
S

x + sin x.

Using the Fundamental Theorem of Calculus, Part


(t + sin t) dt. What is F (x)?

Using the Fundamental Theorem of Calculus, we have F (x) =

This simple example reveals something incredible: F(x) is an an deriva ve


of x + sin x! Therefore, F(x) = x cos x + C for some value of C. (We can
nd C, but generally we do not care. We know that F( ) = , which allows us
to compute C. In this case, C = cos( ) +
.)
We have done more than found a complicated way of compu ng an anderiva ve. Consider a func on f dened on an open interval containing a, b
b
x
and c. Suppose we want to compute a f(t) dt. First, let F(x) = c f(t) dt. Using
Notes:

.
the proper es of the denite integral found in Theorem , we know
b
c
b
f(t) dt
f(t) dt +
f(t) dt =
a

f(t) dt +

= F(a) + F(b)
= F(b) F(a).

f(t) dt

We now see how indenite integrals and denite integrals are related: we can
evaluate a denite integral using an deriva ves! This is the second part of the
Fundamental Theorem of Calculus.
Theorem

The Fundamental Theorem of Calculus, Part

Let f be con nuous on [a, b] and let F be any an deriva ve of f. Then

b
a

f(x) dx = F(b) F(a).

Example
Using the Fundamental Theorem of Calculus,
Part
We spent a great deal of me in the previous sec on studying ( x x ) dx.
Using the Fundamental Theorem of Calculus, evaluate this denite integral.
We need an an deriva ve of f(x) = x x . All an derivaS
ves of f have the form F(x) = x x + C; for simplicity, choose C = .
The Fundamental Theorem of Calculus states

( x x ) dx = F( ) F( ) =

( )

/ .

This is the same answer we obtained using limits in the previous sec on, just
with much less work.
Nota on: A special nota on is o en used in the process of evalua ng denite
integrals using the Fundamental Theorem of Calculus. Instead of explicitly writ b

ing F(b) F(a), the nota on F(x) is used. Thus the solu on to Example
a
would be wri en as:
)
(


) (
)
(
= / .
( x x ) dx =
x x = ( )
Notes:

The Fundamental Theorem of Calculus

Chapter

Integra on
The Constant C: Any an deriva ve F(x) can be chosen when using the Fundamental Theorem of Calculus to evaluate a denite integral, meaning any value
of C can be picked. The constant always cancels out of the expression when
evalua ng F(b) F(a), so it does not ma er what value is picked. This being
the case, we might as well let C = .
Using the Fundamental Theorem of Calculus, Part
Example
Evaluate the following denite integrals.

t
e dt
.
sin x dx
.
x dx
.
.
u du
.

dx

.
.

x dx =

( )

= .


(
)

sin x dx = cos x = cos cos =

= .

(This is interes ng; it says that the area under one hump of a sine curve
is .)



et dt = et = e e = e
.
. .
.

u du =

u du =



dx = x = ( )



u =

= ( )= .

This integral is interes ng; the integrand is a constant func on, hence we
are nding the area of a rectangle with width ( ) = and height .
No ce how the evalua on of the denite integral led to ( ) = .
b
In general, if c is a constant, then a c dx = c(b a).

Understanding Mo on with the Fundamental Theorem of Calculus


We established, star ng with Key Idea , that the deriva ve of a posi on
func on is a velocity func on, and the deriva ve of a velocity func on is an accelera on func on. Now consider denite integrals of velocity and accelera on
b
v(t) dt mean?
func ons. Specically, if v(t) is a velocity func on, what does
a

Notes:

.
The Fundamental Theorem of Calculus states that
b
v(t) dt = V(b) V(a),
a

where V(t) is any an deriva ve of v(t). Since v(t) is a velocity func on, V(t)
must be a posi on func on, and V(b) V(a) measures a change in posi on, or
displacement.
Finding displacement
Example
A ball is thrown straight up with velocity given by v(t) =

v(t) dt.
t is measured in seconds. Find, and interpret,
S

t+

/s, where

Using the Fundamental Theorem of Calculus, we have

( t + ) dt
v(t) dt =
=

t +

= .

Thus if a ball is thrown straight up into the air with velocity v(t) = t + ,
the height of the ball, second later, will be feet above the ini al height. (Note
that the ball has traveled much farther. It has gone up to its peak and is falling
down, but the dierence between its height at t = and t = is .)
Integra ng a rate of change func on gives total change. Velocity is the rate
of posi on change; integra ng velocity gives the total change of posi on, i.e.,
displacement.
Integra ng a speed func on gives a similar, though dierent, result. Speed
is also the rate of posi on change, but does not account for direc on. So integra ng a speed func on gives total change of posi on, without the possibility
of nega ve posi on change. Hence the integral of a speed func on gives distance traveled.
As accelera on is the rate of velocity change, integra ng an accelera on
func on gives total change in velocity. We do not have a simple term for this
analogous to displacement. If a(t) = miles/h and t is measured in hours,
then

a(t) dt =

means the velocity has increased by

Notes:

m/h from t =

to t = .

The Fundamental Theorem of Calculus

Chapter

Integra on

The Fundamental Theorem of Calculus and the Chain Rule


x Part of the Fundamental Theorem of Calculus (FTC) states that given F(x) =
)
d(
f(t) dt, F (x) = f(x). Using other nota on,
F(x) = f(x). While we have
dx
a
just prac ced evalua ng denite integrals, some mes nding an deriva ves is
impossible and we need to rely on other techniques
xto approximate the value
of a denite integral. Func ons wri en as F(x) = a f(t) dt are useful in such
situa ons.
It may be of further use to compose such a func on with another. As an
example, we may compose F(x) with g(x) to get
(
)
F g(x) =

g(x)
a

f(t) dt.

What is the deriva ve of such a func on? The Chain Rule can be employed to
state
))
(
)
(
)
d( (
F g(x) = F g(x) g (x) = f g(x) g (x).
dx
An example will help us understand this.
Example

The FTC, Part , and the Chain Rule


x
ln t dt.
Find the deriva ve of F(x) =

We can view F(x) as being the func on G(x) =


ln t dt
(
)
composed with g(x) = x ; that is, F(x) = G g(x) . The Fundamental Theorem
of Calculus states that G (x) = ln x. The Chain Rule gives us
(
)
F (x) = G g(x) g (x)
= ln(g(x))g (x)
S

= ln(x ) x
= x ln x

Normally, the steps dening G(x) and g(x) are skipped.


Prac ce this once more.
Example

The FTC, Part


, and the Chain Rule

Find the deriva ve of F(x) =

Notes:

cos x

t dt.

.
S

Note that F(x) =

ve of F is straigh orward:

cos x

The Fundamental Theorem of Calculus

t dt. Viewed this way, the derivay

F (x) = sin x cos x.

f(x)

Area Between Curves


g(x)

Consider con nuous func ons f(x) and g(x) dened on [a, b], where f(x)
g(x) for all x in [a, b], as demonstrated in Figure . . What is the area of the
shaded region bounded by the two curves over [a, b]?
The area can be found by recognizing that this area is the area under f
the area under g. Using mathema cal nota on, the area is

f(x) dx

b
a

f(x)

g(x) dx.

g(x)

Proper es of the denite integral allow us to simplify this expression to

Theorem

b
a

(
)
f(x) g(x) dx.

Area Between Curves

Figure . : Finding the area bounded by


two func ons on an interval; it is found
by subtrac ng the area under g from the
area under f.

Let f(x) and g(x) be con nuous func ons dened on [a, b] where f(x)
g(x) for all x in [a, b]. The area of the region bounded by the curves
y = f(x), y = g(x) and the lines x = a and x = b is

b
a

)
f(x) g(x) dx.

y
y=x +x

Finding area between curves


Example
Find the area of the region enclosed by y = x + x and y = x .
It will help to sketch these two func ons, as done in Figure
. . The region whose area we seek is completely bounded by these two
func ons; they seem to intersect at x = and x = . To check, set x +x =

Notes:

.y =

Figure . : Sketching the region enclosed by y = x + x and y = x


in Example
.

Chapter

Integra on
x and solve for x:

x +x

= x

(x + x ) ( x ) =

x x =
(x )(x + ) =

x= , .

.
Figure . : A graph of a func on f to introduce the Mean Value Theorem.

Following Theorem

, the area is

)
x (x + x ) dx =

(x + x + ) dx

(
)

= x + x + x

= (
=

)+ +

(a)
y

The Mean Value Theorem and Average Value


Consider the graph of a func on f in Figure . and the area dened by
f(x) dx. Three rectangles are drawn in Figure . ; in (a), the height of the
rectangle is greater than f on [ , ], hence the area of this rectangle is is greater

than f(x) dx.

In (b), the height of the rectangle is smaller than f on [ , ], hence the area

of this rectangle is less than f(x) dx.


Finally, in (c) the
height of the rectangle is such that the area of the rectangle
is exactly that of
f(x) dx. Since rectangles that are too big, as in (a), and

rectangles that are too li le, as in (b), give areas greater/lesser than f(x) dx,
it makes sense that there is a rectangle, whose top intersects f(x) somewhere
on [ , ], whose area is exactly that of the denite integral.

(b)
y

We state this idea formally in a theorem.


x

(c)

Figure . : Dierently sized rectangles


give upper and lower bounds on

f(x) dx; the last rectangle matches the


area exactly.

Notes:

Theorem

The Fundamental Theorem of Calculus


y

The Mean Value Theorem of Integra on

Let f be con nuous on [a, b]. There exists a value c in [a, b] such that

b
a

sin .69

f(x) dx = f(c)(b a).

This is an existen al statement; c exists, but we do not provide a method


of nding it. Theorem
is directly connected to the Mean Value Theorem of
Dieren a on, given as Theorem ; we leave it to the reader to see how.
We demonstrate the principles involved in this version of the Mean Value
Theorem in the following example.

. . : A graph of y = sin x on
Figure
[ , ] and the rectangle guaranteed by
the Mean Value Theorem.

Using the Mean Value Theorem


Example

Consider
sin x dx. Find a value c guaranteed by the Mean Value Theorem.

We rst need to evaluate


sin x dx. (This was previously
.)



sin x dx = cos x = .

done in Example

y
y = f(x)

Thus we seek a value c in [ , ] such that sin c = .


sin c =

sin c = / c = arcsin( /) .

In Figure . sin x is sketched along with a rectangle with height sin( .


The area of the rectangle is the same as the area under sin x on [ , ].
Let f be a func on on [a, b] with c such that f(c)(ba) =
)
b(
f(x) f(c) dx:
a

)
f(x) f(c) dx =

f(x)

b
a

f(c)

).

f(x) dx. Consider

= f(c)(b a) f(c)(b a)
= .

y = f(x) f(c)

f(c) dx

When f(x) is shi ed by f(c), the amount of area under f above the xaxis on
[a, b] is the same as the amount of area below the xaxis above f; see Figure
. for an illustra on of this. In this sense, we can say that f(c) is the average
value of f on [a, b].

Notes:

f(c)

.
Figure . : On top, a graph of y =
f(x) and the rectangle guaranteed by the
Mean Value Theorem. Below, y = f(x) is
shi ed down by f(c); the resul ng area
under the curve is .

Chapter

Integra on
The value f(c) is the average value in another sense. First, recognize that the
Mean Value Theorem can be rewri en as
b
f(x) dx,
f(c) =
ba a

for some value of c in [a, b]. Next, par on the interval [a, b] into n equally
spaced subintervals, a = x < x < . . . < xn+ = b and choose any ci in
[xi , xi+ ]. The average of the numbers f(c ), f(c ), , f(cn ) is:
n

f(c ) + f(c ) + . . . + f(cn ) =

f(ci ) =

i=

i=
n

f(ci )
f(ci )

i=

=
=

ba
ba

Now take the limit as n :


n

f(ci )x
lim
n b a
i=

f(ci ).

i=

(ba)
(ba) :

Mul ply this last expression by in the form of


n

n
(b a)
n (b a)

i=

f(ci )

ba
n

f(ci )x

(where x = (b a)/n)

i=

ba

b
a

f(x) dx

f(c).

This tells us this: when we evaluate f at n (somewhat) equally spaced points in


[a, b], the average value of these samples is f(c) as n .
This leads us to a deni on.
Deni on

The Average Value of f on [a, b]

Let f be con nuous on [a, b]. The average value of f on [a, b] is f(c),
where c is a value in [a, b] guaranteed by the Mean Value Theorem. I.e.,
Average Value of f on [a, b] =

Notes:

ba

b
a

f(x) dx.

.
An applica on of this deni on is given in the following example.
Finding the average value of a func on
Example
An object moves back and forth along a straight line with a velocity given by
v(t) = (t ) on [ , ], where t is measured in seconds and v(t) is measured
in /s.
What is the average velocity of the object?
By our deni on, the average velocity is:

(t ) dt =

t t+

dt =

)

t t + t =

/s.

We can understand the above example through a simpler situa on. Suppose
you drove
miles in hours. What was your average speed? The answer is
simple: displacement/ me =
miles/ hours = mph.
What was the displacement of the
? We calculate this
object in Example
by integra ng its velocity func on: (t ) dt =
. Its nal posi on was
feet from its ini al posi on a er seconds: its average velocity was
/s.
This sec on has laid the groundwork for a lot of great mathema cs to follow. The most important lesson is this: denite integrals can be evaluated using
an deriva ves. Since the previous sec on established that denite integrals are
the limit of Riemann sums, we can later create Riemann sums to approximate
values other than area under the curve, convert the sums to denite integrals,
then evaluate these using the Fundamental Theorem of Calculus. This will allow
us to compute the work done by a variable force, the volume of certain solids,
the arc length of curves, and more.
The downside is this: generally speaking, compu ng an deriva ves is much
more dicult than compu ng deriva ves. The next chapter is devoted to techniques of nding an deriva ves so that a wide variety of denite integrals can
be evaluated. Before that, the next sec on explores techniques of approximating the value of denite integrals beyond using the Le Hand, Right Hand and
Midpoint Rules.

Notes:

The Fundamental Theorem of Calculus

Exercises .

Terms and Concepts

x dx

x dx

x dx

. How are denite and indenite integrals related?


. What constant of integra on is most commonly used when
evalua ng denite integrals?
. T/F: If f is a con nuous func on, then F(x) =
also a con nuous func on.

x
a

f(t) dt is

. The denite integral can be used to nd the area under a


curve. Give two other uses for denite integrals.

Problems
In Exercises

, evaluate the denite integral.

( x x + ) dx

(x ) dx

(x x ) dx

cos x dx

/
/

sec x dx

dx

x
x

dx

e dx

( cos x sin x) dx

dx

dx

dx
/

csc x cot x dx

. Explain why:

(a)
xn dx = , when n is a posi ve, odd integer, and

( x ) dx

dx

dx

(b)

dx

dx

xn dx =

integer.

In Exercises

Value Theorem.
.

t dt

dt
t

ex dx

x dx

x dx

x dx

x dx

xn dx when n is a posi ve, even

, nd a value c guaranteed by the Mean

In Exercises , nd the average value of the func on on


the given interval.

In Exercises
, an accelera on func on of an object
moving along a straight line is given. Find the change of the
objects velocity over the given me interval.
. a(t) =

. f(x) = sin x on [ , / ]

/s on [ , ]

. a(t) =

. y = sin x on [ , ]

/s on [ , ]

. a(t) = t /s on [ , ]
. y = x on [ , ]

. a(t) = cos t /s on [ , ]

. y = x on [ , ]

In Exercises
, sketch the given func ons and nd the
area of the enclosed region.

. y = x on [ , ]

. y = x, y = x, and x = .

. g(t) = /t on [ , e]

. y = x + , y = x + , x =

In Exercises
, a velocity func on of an object moving
along a straight line is given. Find the displacement of the
object over the given me interval.

. y=x x+ ,y= x .
. y= x + x ,y=x + x+ .
In Exercises

. v(t) =

t+

. v(t) =

t+

. v(t) =

/s on [ , ]
/s on [ ,

. v(t) =

. F(x) =

. F(x) =

. F(x) =

. F(x) =

mph on [ , ]

. v(t) = cos t /s on [ , / ]

t /s on [ ,

and x = .

, nd F (x).

x +x

dt

t dt

x
x
x
ex
ln x

(t + ) dt

sin t dt

Chapter

Integra on

.
y

The Fundamental Theorem of Calculus gives a concrete technique for nding


the exact value of a denite integral. That technique is based on compu ng anderiva ves. Despite the power of this theorem, there are s ll situa ons where
we must approximate the value of the denite integral instead of nding its exact value. The rst situa on we explore is where we cannot compute the anderiva ve of the integrand. The second case is when we actually do not know
the integrand, but only its value when evaluated at certain points.

y=e

.5

An elementary func on is any func on that is a combina on of polynomials, nth roots, ra onal, exponen al, logarithmic and trigonometric func ons. We
can compute the deriva ve of any elementary func on, but there are many elementary func ons of which we cannot compute an an deriva ve. For example,
the following func ons do not have an deriva ves that we can express with elementary func ons:

.5

y
1

Numerical Integra on

y = sin(x3 )

ex ,

0.5

..

0.5

y
1
y=

ex dx,

as pictured in Figure .
15

.
Figure . : Graphically represen ng
three denite integrals that cannot be
evaluated using an deriva ves.

sin x
.
x

sin x
x

10

and

The simplest way to refer to the an deriva ves of ex is to simply write


e
dx.
This sec on outlines three common methods of approxima ng the value of
denite integrals. We describe each as a systema c method of approxima ng
area under a curve. By approxima ng this area accurately, we nd an accurate
approxima on of the corresponding denite integral.
We will apply the methods we learn in this sec on to the following denite
integrals:

0.5

sin(x )

sin(x ) dx,

and

sin(x)
dx,
x

The Le and Right Hand Rule Methods


In Sec on . we addressed the problem of evalua ng denite integrals by
approxima ng the area under the curve using rectangles. We revisit those ideas
here before introducing other methods of approxima ng denite integrals.
We start with a review of nota on. Let f be a con nuous func on on the
b
f(x) dx. We par on [a, b] into n
interval [a, b]. We wish to approximate
a

equally spaced subintervals, each of length x =

Notes:

ba
. The endpoints of these
n

Numerical Integra on

subintervals are labeled as


x = a, x = a + x, x = a + x, . . . , xi = a + (i )x, . . . , xn+ = b.
n

i=

Key Idea

states that to use the Le Hand Rule we use the summa on


n

f(xi+ )x. We review


f(xi )x and to use the Right Hand Rule we use

the use of these rules in the context of examples.


Example
Approximate

i=

Approxima ng denite integrals with rectangles


e

dx using the Le and Right Hand Rules with

y = ex

equally

spaced subintervals.
.

We begin by par oning the interval [ , ] into


S
spaced intervals. We have x = = / = . , so

equally

x = , x = . , x = . , x = . , x = . , and x = .

.8

Using the Le Hand Rule, we have:


y
n

i=

(
)
f(xi )x = f(x ) + f(x ) + f(x ) + f(x ) + f(x ) x

(
)
= f( ) + f( . ) + f( . ) + f( . ) + f( . ) x
(
+ .
+ .
+ .
+ .
)( . )
.
.

Using the Right Hand Rule, we have:


n

i=

.
)

f(xi+ )x = f(x ) + f(x ) + f(x ) + f(x ) + f(x ) x


(
)
= f( . ) + f( . ) + f( . ) + f( . ) + f( ) x
(
.
+ .
+ .
+ .
+ .
)( . )
.
.

Figure . shows the rectangles used in each method to approximate the


denite integral. These graphs show that in this par cular case, the Le Hand
Rule is an over approxima on and the Right Hand Rule is an under approximaon. To get a be er approxima on, we could use more rectangles, as we did in

Notes:

y = ex

Figure .
Example

: Approxima ng
.

.8

ex dx in

Chapter
xi
x
x
x
x
x
x
x
x
x
x
x

Integra on

Exact
/
/
/
/
/
/
/
/
/
/
/

Approx.
.
.
.
.
.
.
.
.
.
.
.

sin(xi )
.
.
.

Sec on . . We could also average the Le and Right Hand Rule results together,
giving
.
+ .
= .
.
The actual answer, accurate to places a er the decimal, is .
our average is a good approxima on.

.
.
.
.
.
.

Example
Approximate

y
y = sin(x3 )

0.5

0.5

Approxima ng denite integrals with rectangles


sin(x ) dx using the Le and Right Hand Rules with

/ (/ )

ba
=
=
.
n

y = sin(x3 )

It is useful to write out the endpoints of the subintervals in a table; in Figure


. , we give the exact values of the endpoints, their decimal approxima ons,
and decimal approxima ons of sin(x ) evaluated at these points.
Once this table is created, it is straigh orward to approximate the denite
integral using the Le and Right Hand Rules. (Note: the table itself is easy to
create, especially with a standard spreadsheet program on a computer. The last
two columns are all that are needed.) The Le Hand Rule sums the rst values
of sin(xi ) and mul plies the sum by x; the Right Hand Rule sums the last
values of sin(xi ) and mul plies by x. Therefore we have:

sin(x ) dx ( . )( .
)= .
.
Le Hand Rule:
Right Hand Rule:

0.5

equally

We begin by nding x:

spaced subintervals.

Figure . : Table of values used to ap


.
proximate sin(x ) dx in Example

, showing

sin(x ) dx ( .

)( .

)= .

Average of the Le and Right Hand Rules: .


.
The actual answer, accurate to places a er the decimal, is .
. Our approxima ons were once again fairly good. The rectangles used in each approxima on are shown in Figure . . It is clear from the graphs that using more
rectangles (and hence, narrower rectangles) should result in a more accurate
approxima on.

The Trapezoidal Rule

0.5

Figure
. :
Approxima ng

sin(x
)
dx
in
Example
.

In Example

we approximated the value of

of equal width. Figure .

Notes:

ex dx with rectangles

shows the rectangles used in the Le and Right Hand

.
Rules. These graphs clearly show that rectangles do not match the shape of the
graph all that well, and that accurate approxima ons will only come by using
lots of rectangles.
Instead of using rectangles to approximate the area, we can instead use
trapezoids. In Figure . , we show the region under f(x) = ex on [ , ] approximated with trapezoids of equal width; the top corners of each trapezoid lies on the graph of f(x). It is clear from this gure that these trapezoids
more accurately approximate
the area under f and hence should give a be er

approxima on of ex dx. (In fact, these trapezoids seem to give a great approxima on of the area!)
The
formula for the area of a trapezoid is given in Figure . . We approximate ex dx with these trapezoids in the following example.
Approxima ng denite integrals
using trapezoids

Example

Use trapezoids of equal width to approximate

dx.

To compute the areas of the trapezoids in Figure . , it


S
will again be useful to create a table of values as shown in Figure . .
The le most trapezoid has legs of length and .
and a height of . .
Thus, by our formula, the area of the le most trapezoid is:
+ .

( . )= .

+ .

( . )= .

y
y = ex

and .

and a height

( . )+

We approximate

+ .

+ .

( . )+
( . )+

ex dx .

Notes:

)+( .

+ .

)+( .

.8

.
Figure .

Area =

a+b
2 h

h
: The area of a trapezoid.

+ .

+ .

( . )+
( . )= .

xi

exi

.
.
.
.

.
.
.
.
.

There are many things to observe in this example. Note how each term in
the nal summa on was mul plied by both / and by x = . . We can factor
these coecients out, leaving a more concise summa on as:
[
( . ) ( + .

Figure . : Approxima ng ex dx using trapezoids of equal widths.

The sum of the areas of all trapezoids is:


+ .

Moving right, the next trapezoid has legs of length .


of . . Thus its area is:
.

Numerical Integra on

+ .

)+( .

+ .

)+( .

+ .

]
) .

Figure .

: A table of values of ex .

Chapter

Integra on
Now no ce that all numbers except for the rst and the last are added twice.
Therefore we can write the summa on even more concisely as
. [

+ ( .

+ .

+ .

+ .

)+ .

This is the heart of the Trapezoidal Rule, wherein a denite integral

b
a

f(x) dx

is approximated by using trapezoids of equal widths to approximate the corresponding area under f. Using n equally spaced subintervals with endpoints x ,
ba
. Thus:
x , . . ., xn+ , we again have x =
n
b
n

f(xi ) + f(xi+ )
f(x) dx
x
a

i=

x (

f(xi ) + f(xi+ )

i=

n
]

x [
=
f(xi ) + f(xn+ ) .
f(x ) +
i=

Using the Trapezoidal Rule

and approximate
sin(x ) dx using the Trapezoidal Rule

Example
Revisit Example
and

equally spaced subintervals.

We refer back to Figure . for the table of values of sin(x ).


S
Recall that x = / .
. Thus we have:

sin(x ) dx

= .

+ ( .

) + ... + .

+ ( .

No ce how quickly the Trapezoidal Rule can be implemented once the table of values is created. This is true for all the methods explored in this sec on;
the real work is crea ng a table of xi and f(xi ) values. Once this is completed, approxima ng the denite integral is not dicult. Again, using technology is wise.
Spreadsheets can make quick work of these computa ons and make using lots
of subintervals easy.
Also no ce the approxima ons the Trapezoidal Rule gives. It is the average
of the approxima ons given by the Le and Right Hand Rules! This eec vely

Notes:

Numerical Integra on

renders the Le and Right Hand Rules obsolete. They are useful when rst learning about denite integrals, but if a real approxima on is needed, one is generally be er o using the Trapezoidal Rule instead of either the Le or Right Hand
Rule.
How can we improve on the Trapezoidal Rule, apart from using more and
more trapezoids? The answer is clear once we look back and consider what we
have really done so far. The Le Hand Rule is not really about using rectangles to
approximate area. Instead, it approximates a func on f with constant func ons
on small subintervals and then computes the denite integral of these constant
func ons. The Trapezoidal Rule is really approxima ng a func on f with a linear
func on on a small subinterval, then computes the denite integral of this linear
func on. In both of these cases the denite integrals are easy to compute in
geometric terms.
So we have a progression: we start by approxima ng f with a constant funcon and then with a linear func on. What is next? A quadra c func on. By
approxima ng the curve of a func on with lots of parabolas, we generally get
an even be er approxima on of the denite integral. We call this process Simpsons Rule, named a er Thomas Simpson (
), even though others had
used this rule as much as
years prior.

Simpsons Rule
Given one point, we can create a constant func on that goes through that
point. Given two points, we can create a linear func on that goes through those
points. Given three points, we can create a quadra c func on that goes through
those three points (given that no two have the same xvalue).
Consider three points (x , y ), (x , y ) and (x , y ) whose xvalues are equally
spaced and x < x < x . Let f be the quadra c func on that goes through these
three points. It is not hard to show that
x
)
x x (
f(x) dx =
( . )
y + y +y .
x

Consider Figure . . A func on f goes through the points shown and the
parabola g that also goes through those points is graphed with a dashed line.
Using our equa on from above, we know exactly that

)
(
+ ( )+ = .
g(x) dx =

Since g is a good approxima on for f on [ , ], we can state that

f(x) dx .

Notes:

Figure . : A graph of a func on f and


a parabola that approximates it well on
[ , ].

Chapter

Integra on

xi

exi

.
.
.
.

No ce how the interval [ , ] was split into two subintervals as we needed


points. Because of this, whenever we use Simpsons Rule, we need to break the
interval into an even number of subintervals.
In general, to approximate

f(x) dx using Simpsons Rule, subdivide [a, b]

into n subintervals, where n is even and each subinterval has width x = (b


a)/n. We approximate f with n/ parabolic curves, using Equa on ( . ) to compute the area under these parabolas. Adding up these areas gives the formula:

(a)

b
a

x [

f(x) dx

]
f(x )+ f(x )+ f(x )+ f(x )+. . .+ f(xn )+ f(xn )+f(xn+ ) .

Note how the coecients of the terms in the summa on have the pa ern , ,
, , , , . . ., , , .
Lets demonstrate Simpsons Rule with a concrete example.

y=e

.5

Example
.

. 5

x
.5

.75

dx using Simpsons Rule and equally spaced subintervals.

We begin by making a table of values as we have in the past,


S
as shown in Figure . (a). Simpsons Rule states that

(b)
Figure . : A table of values to approxi
mate ex dx, along with a graph of the
func on.

xi
.
.
.
.
.
.
.
.
.
.
.

Approximate

Using Simpsons Rule

sin(xi )
.
.
.
.
.
.
.
.
.
.

Figure . : Table of values used to ap


.
proximate sin(x ) dx in Example

ex dx

+ ( .

)+ ( .

)+ ( .

)+ .

= .

Recall in Example
we stated that the correct answer, accurate to places
a er the decimal, was .
. Our approxima on with Simpsons Rule, with
subintervals, is be er than our approxima on with the Trapezoidal Rule using
!
Figure . (b) shows f(x) = ex along with its approxima ng parabolas,
demonstra ng how good our approxima on is. The approxima ng curves are
nearly indis nguishable from the actual func on.
Example
Approximate
vals.

Using Simpsons Rule


sin(x ) dx using Simpsons Rule and

equally spaced inter-

Figure . shows the table of values that we used in the


S
past for this problem, shown here again for convenience. Again, x = (/ +
/ )/ .
.

Notes:

Numerical Integra on
y

Simpsons Rule states that

sin(x ) dx

... + ( .

( .

) + ( .

) + ( .

) + ...
0.5

)+ ( .

) + ( .

= .

Recall that the actual value, accurate to decimal places, is .


. Our apth
proxima on is within one /
of the correct value. The graph in Figure .
shows how closely the parabolas match the shape of the graph.

Summary and Error Analysis


We summarize the key concepts of this sec on thus far in the following Key
Idea.
Key Idea

Let f be a con nuous func on on [a, b], let n be a posi ve integer, and let x =
Set x = a, x = a + x, . . ., xi = a + (i )x, xn+ = b.
b
f(x) dx.
Consider
a
b
[
]
f(x) dx x f(x ) + f(x ) + . . . + f(xn ) .
Le Hand Rule:
a

Right Hand Rule:

Trapezoidal Rule:
Simpsons Rule:

f(x) dx

f(x) dx

x [

x [

Figure
. :
Approxima ng

with
sin(x ) dx in Example
Simpsons Rule and
equally spaced
intervals.

ba
.
n

]
f(x ) + f(x ) + f(x ) + . . . + f(xn ) + f(xn+ ) .

]
f(x ) + f(x ) + f(x ) + . . . + f(xn ) + f(xn+ ) (n even).

. How was the right answer computed?


. If the right answer can be found, what is the point of approxima ng?
. If there is value to approxima ng, how are we supposed to know if the
approxima on is any good?

0.5

[
]
f(x) dx x f(x ) + f(x ) + . . . + f(xn+ ) .

In our examples, we approximated the value of a denite integral using a


given method then compared it to the right answer. This should have raised
several ques ons in the readers mind, such as:

Notes:

Numerical Integra on

y = sin(x3 )

Chapter

Integra on
These are good ques ons, and their answers are educa onal. In the examples, the right answer was never computed. Rather, an approxima on accurate
to a certain number of places a er the decimal was given. In Example
, we
do not know the exact answer, but we know it starts with .
. These more
accurate approxima ons were computed using numerical integra on but with
more precision (i.e., more subintervals and the help of a computer).
Since the exact answer cannot be found, approxima on s ll has its place.
How are we to tell if the approxima on is any good?
Trial and error provides one way. Using technology, make an approximaon with, say, ,
, and
subintervals. This likely will not take much me
at all, and a trend should emerge. If a trend does not emerge, try using yet more
subintervals. Keep in mind that trial and error is never foolproof; you might
stumble upon a problem in which a trend will not emerge.
A second method is to use Error Analysis. While the details are beyond the
scope of this text, there are some formulas that give bounds for how good your
approxima on will be. For instance, the formula might state that the approxima on is within . of the correct answer. If the approxima on is . , then
one knows that the correct answer is between . and . . By using lots of
subintervals, one can get an approxima on as accurate as one likes. Theorem
states what these bounds are.
Theorem

Error Bounds in the Trapezoidal and Simpsons Rules

. Let ET be the error in approxima ng


zoidal Rule.

b
a

f(x) dx using the Trape-

nd
If f has a con nuous
deriva ve on [a, b] and M is any upper

bound of f (x) on [a, b], then

ET

(b a)
M.
n

. Let ES be the error in approxima ng


Rule.

b
a

f(x) dx using Simpsons

If f has a con nuous th deriva ve on [a, b] and M is any upper


bound of f ( ) on [a, b], then
ES

Notes:

(b a)
M.
n

Numerical Integra on

There are some key things to note about this theorem.


. The larger the interval, the larger the error. This should make sense intui vely.
. The error shrinks as more subintervals are used (i.e., as n gets larger).
. The error in Simpsons Rule has a term rela ng to the th deriva ve of f.
Consider a cubic polynomial: its th deriva ve is . Therefore, the error in
approxima ng the denite integral of a cubic polynomial with Simpsons
Rule is Simpsons Rule computes the exact answer!
We revisit Examples
and
orem in the following example.
Example

and compute the error bounds using The-

Compu ng error bounds

Find the error bounds when approxima ng

ex dx using the Trapezoidal

Rule and subintervals, and using Simpsons Rule with subintervals.


S

Trapezoidal Rule with n = :


We start by compu ng the

nd

deriva ve of f(x) = ex :

f (x) = ex ( x ).
Figure . shows a graph of f (x) on [ , ]. It is clear that the largest value of
f , in absolute value, is . Thus we let M = and apply the error formula from
Theorem .
( )
= .
.

Our error es ma on formula states that our approxima on of .


found
in Example
is within .
of the correct answer, hence we know that

ex dx .
= .
+ .
.
.
.
=.

x
.5

y = ex (4x )

ET =

We had earlier computed the exact answer, correct to


.
, arming the validity of Theorem .
Simpsons Rule with n = :
We start by compu ng the

th

deriva ve of f(x) = ex :

f ( ) (x) = ex (

Notes:

decimal places, to be

x +

).

Figure . : Graphing f (x) in Example


to help establish error bounds.

Chapter

Integra on
Figure . shows a graph of f ( ) (x) on [ , ]. It is clear that the largest value of
f ( ) , in absolute value, is . Thus we let M =
and apply the error formula
from Theorem .

y
y = ex (

x 8x +

Es =
x

( )

= .

found
Our error es ma on formula states that our approxima on of .
in Example
is within .
of the correct answer, hence we know that

.
Figure . : Graphing f ( ) (x) in Example
to help establish error bounds.

=.

ex dx .

Once again we arm the validity of Theorem

Time

Speed
(mph)

= .

+ .

At the beginning of this sec on we men oned two main situa ons where
numerical integra on was desirable. We have considered the case where an
an deriva ve of the integrand cannot be computed. We now inves gate the
situa on where the integrand is not known. This is, in fact, the most widely
used applica on of Numerical Integra on methods. Most of the me we observe behavior but do not know the func on that describes it. We instead
collect data about the behavior and make approxima ons based o of this data.
We demonstrate this in an example.
Approxima ng distance traveled
Example
One of the authors drove his daughter home from school while she recorded
their speed every seconds. The data is given in Figure . . Approximate the
distance they traveled.
Recall that by integra ng a speed func on we get distance
S
traveled. We have informa on about v(t); we will use Simpsons Rule to approx b
v(t) dt.
imate
a

The most dicult aspect of this problem is conver ng the given data into the
form we need it to be in. The speed is measured in miles per hour, whereas the
me is measured in second increments.
We need to compute x = (b a)/n. Clearly, n = . What are a and b?
Since we start at me t = , we have that a = . The nal recorded me came
a er periods of seconds, which is minutes or / of an hour. Thus we
have
x =

Figure . : Speed data collected at


second intervals for Example
.

Notes:

/
ba
=
n

.
Thus the distance traveled is approximately:
.
[
]
v(t) dt
f(x ) + f(x ) + f(x ) + + f(xn ) + f(xn+ )
=
.

+
miles.

+ +

We approximate the author drove . miles. (Because we are sure the reader
wants to know, the authors odometer recorded the distance as about .
miles.)
We started this chapter learning about an deriva ves and indenite integrals. We then seemed to change focus by looking at areas between the graph
of a func on and the x-axis. We dened these areas as the denite integral of
the func on, using a nota on very similar to the nota on of the indenite integral. The Fundamental Theorem of Calculus ed these two seemingly separate
concepts together: we can nd areas under a curve, i.e., we can evaluate a definite integral, using an deriva ves.
We ended the chapter by no ng that an deriva ves are some mes more
than dicult to nd: they are impossible. Therefore we developed numerical
techniques that gave us good approxima ons of denite integrals.
We used the denite integral to compute areas, and also to compute displacements and distances traveled. There is far more we can do than that. In
Chapter well see more applica ons of the denite integral. Before that, in
Chapter well learn advanced techniques of integra on, analogous to learning
rules like the Product, Quo ent and Chain Rules of dieren a on.

Notes:

Numerical Integra on

Exercises .

Terms and Concepts


. T/F: Simpsons Rule is a method of approxima ng anderiva ves.

. What are the two basic situa ons where approxima ng the
value of a denite integral is necessary?
. Why are the Le and Right Hand Rules rarely used?

Problems
In Exercises

, a denite integral is given.

(a) Approximate the denite integral with the Trapezoidal


Rule and n = .
(b) Approximate the denite integral with Simpsons Rule
and n = .
(c) Find the exact value of the integral.

dx

(a) the Trapezoidal Rule

x dx

(b) Simpsons Rule


x dx

sin x dx

x dx

(x + x x + ) dx
x dx

cos x dx

( )
cos x dx

ex dx

dx
x

( )
cos x dx
x dx

(a) where the measurements are in cen meters, taken in


cm increments, and

In Exercises
, approximate the denite integral with
the Trapezoidal Rule and Simpsons Rule, with n = .

sin x dx

In Exercises
, a region is given. Find the area of the
region using Simpsons Rule:

x dx

(b) where the measurements are in hundreds of yards,


taken in
yd increments.

x + dx

.1

sin x +

dx

sin x +

.9

In Exercises , nd n such that the error in approximating the given denite integral is less than .
when using:

ln x dx

cos x dx

x sin x dx

.
.

:T
A
The previous chapter introduced the an deriva ve and connected it to signed
areas under a curve through the Fundamental Theorem of Calculus. The next
chapter explores more applica ons of denite integrals than just area. As evalua ng denite integrals will become important, we will want to nd an derivaves of a variety of func ons.
This chapter is devoted to exploring techniques of an dieren a on. While
not every func on has an an deriva ve in terms of elementary func ons (a
concept introduced in the sec on on Numerical Integra on), we can s ll nd
an deriva ves of a wide variety of func ons.

Subs tu on

We mo vate this sec on with an example. Let f(x) = (x + x ) . We can


compute f (x) using the Chain Rule. It is:
f (x) =

(x + x ) ( x + ) = ( x + )(x + x ) .

Now consider this: What is ( x + )(x + x ) dx? We have the answer


in front of us;

( x + )(x + x ) dx = (x + x ) + C.
How would we have evaluated this indenite integral without star ng with f(x)
as we did?
This sec on explores integra on by subs tu on. It allows us to undo the
Chain Rule. Subs tu on allows us to evaluate the above integral without knowing the original func on rst.
The underlying principle is to rewrite a complicated integral of the form
f(x) dx as a notsocomplicated integral h(u) du. Well formally establish
later
how this is done. First, consider again our introductory indenite integral,

( x + )(x + x ) dx. Arguably the most complicated part of the


integrand is (x + x ) . We wish to make this simpler; we do so through a
subs tu on. Let u = x + x . Thus
(x + x ) = u .

We have established u as a func on of x, so now consider the dieren al of u:


du = ( x + )dx.

Chapter

Techniques of An dieren a on
Keep in mind that ( x+ ) and dx are mul plied; the dx is not just si ng there.
Return to the original integral and do some subs tu ons through algebra:

x+

)(x + x ) dx =

( x + )(x + x ) dx
(x + x ) ( x + ) dx
| {z } | {z }
u

du

u du

= u + C (replace u with x
= (x + x ) + C

+ x )

One might well look at this and think I (sort of) followed how that worked,
but I could never come up with that on my own, but the process is learnable.
This sec on contains numerous examples through which the reader will gain
understanding and mathema cal maturity enabling them to regard subs tu on
as a natural tool when evalua ng integrals.
We stated before that integra on by subs tu on undoes the Chain Rule.
Specically, let F(x) and g(x) be dieren able func ons and consider the derivave of their composi on:
))
d( (
F g(x) = F (g(x))g (x).
dx

Thus

F (g(x))g (x) dx = F(g(x)) + C.

Integra on by subs tu on works by recognizing the inside func on g(x) and


replacing it with a variable. By se ng u = g(x), we can rewrite the deriva ve
as
d ( ( ))
F u = F (u)u .
dx
Since du = g (x)dx, we can rewrite the above integral as

F (g(x))g (x) dx =

F (u)du = F(u) + C = F(g(x)) + C.

This concept is important so we restate it in the context of a theorem.

Notes:

Theorem

Integra on by Subs tu on

Let F and g be dieren able func ons, where the range of g is an interval
I contained in the domain of F. Then

F (g(x))g (x) dx = F(g(x)) + C.


If u = g(x), then du = g (x)dx and

F (g(x))g (x) dx = F (u) du = F(u) + C = F(g(x)) + C.


The
point of subs tu on is to make the integra on step easy. Indeed, the
step F (u) du = F(u) + C looks easy, as the an deriva ve of the deriva ve of F
is just F, plus a constant. The work involved is making the proper subs tu on.
There is not a stepbystep process that one can memorize; rather, experience
will be ones guide. To gain experience, we now embark on many examples.
Example

Evaluate

Integra ng by subs tu on
x sin(x + ) dx.

Knowing that subs tu on is related to the Chain Rule, we


S
choose to let u be the inside func on of sin(x + ). (This is not always a good
choice, but it is o en the best place to start.)
Let u = x + , hence du = x dx. The integrand has an x dx term, but
not a x dx term. (Recall that mul plica on is commuta ve, so the x does not
physically have to be next to dx for there to be an x dx term.) We can divide both
sides of the du expression by :
du = x dx

du = x dx.

We can now subs tute.

Notes:

x sin(x + ) dx =

sin(x + ) |{z}
x dx
| {z }
u

sin u du

du

Subs tu on

Chapter

Techniques of An dieren a on
= cos u + C

(now replace u with x + )

= cos(x + ) + C.

Thus x sin(x + ) dx = cos(x + ) + C. We can check our work by evalua ng the deriva ve of the right hand side.
Example
Evaluate

Integra ng by subs tu on
cos( x) dx.

Again let u replace the inside func on. Le ng u = x, we


S
have du = dx. Since our integrand does not have a dx term, we can divide
the previous equa on by to obtain du = dx. We can now subs tute.

cos( x) dx = cos(|{z}
x ) |{z}
dx
u

du

cos u du

sin u + C

sin( x) + C.

We can again check our work through dieren a on.


The previous example exhibited a common, and simple, type of subs tu on.
The inside func on was a linear func on (in this case, y = x). When the
inside func on is linear, the resul ng integra on is very predictable, outlined
here.
Key Idea

Subs tu on With A Linear Func on

Consider F (ax + b) dx, where a = and b are constants. Le ng


u = ax + b gives du = a dx, leading to the result

F (ax + b) dx = F(ax + b) + C.
a

Thus sin( x ) dx = cos( x ) + C. Our next example can use Key


Idea , but we will only employ it a er going through all of the steps.

Notes:

Example
Evaluate

Integra ng by subs tu ng a linear func on


x+

dx.

View this a composi on of func ons f(g(x)), where f(x) =


S
/x and g(x) = x + . Employing our understanding of subs tu on, we let
u = x + , the inside func on. Thus du = dx. The integrand lacks a ;
hence divide the previous equa on by to obtain du/ = dx. We can now
evaluate the integral through subs tu on.

du
dx =
x+
u

du
=
u

=
ln |u| + C
= ln | x + | + C.
Using Key Idea
is faster, recognizing that u is linear and a = . One may
want to con nue wri ng out all the steps un l they are comfortable with this
par cular shortcut.
Not all integrals that benet from subs tu on have a clear inside func on.
Several of the following examples will demonstrate ways in which this occurs.
Example

Evaluate

Integra ng by subs tu on
sin x cos x dx.

There is not a composi on of func on here to exploit; rather,


S
just a product of func ons. Do not be afraid to experiment; when given an integral to evaluate, it is o en benecial to think If I let u be this, then du must be
that and see if this helps simplify the integral at all.
In this example, lets set u = sin x. Then du = cos x dx, which we have as
part of the integrand! The subs tu on becomes very straigh orward:

sin x cos x dx = u du

Notes:

u +C

sin x + C.

Subs tu on

Chapter

Techniques of An dieren a on
One would do well to ask What would happen if we let u = cos x? The result
is just as easy to nd, yet looks very dierent. The challenge to the reader is to
evaluate the integral le ng u = cos x and discover why the answer is the same,
yet looks dierent.
Our examples so far have required basic subs tu on. The next example
demonstrates how subs tu ons can be made that o en strike the new learner
as being nonstandard.
Example
Evaluate

Integra ng by subs tu on

x x + dx.

Recognizing the composi on of func ons, set u = x + .


S
Then du = dx, giving what seems ini ally to be a simple subs tu on. But at this
stage, we have:

x x + dx = x u du.
We cannot evaluate an integral that has both an x and an u in it. We need to
convert the x to an expression involving just u.
Since we set u = x+ , we can also state that u = x. Thus
we can replace
x in the integrand with u . It will also be helpful to rewrite u as u .

x x + dx = (u )u du

(
)
u u du
=
=

u u +C

(x + ) (x + ) + C.

Checking your work is always a good idea. In this par cular case, some algebra
will be needed to make ones answer match the integrand in the original problem.
Example
Evaluate

Integra ng by subs tu on
x ln x

dx.

This is another example where there does not seem to be


S
an obvious composi on of func ons. The line of thinking used in Example
is useful here: choose something for u and consider what this implies du must

Notes:

.
be. If u can be chosen such that du also appears in the integrand, then we have
chosen well.
Choosing u = /x makes du = /x dx; that does not seem helpful. However, se ng u = ln x makes du = /x dx, which is part of the integrand. Thus:

x ln x

dx =

dx
ln x |{z}
x
|{z}
/u

du

du
u
= ln |u| + C
= ln | ln x| + C.
=

The nal answer is interes ng; the natural log of the natural log. Take the derivave to conrm this answer is indeed correct.

Integrals Involving Trigonometric Func ons


Sec on . delves deeper into integrals of a variety of trigonometric funcons; here we use subs tu on to establish a founda on that we will build upon.
The next three examples will help ll in some missing pieces of our an derivave knowledge. We know the an deriva ves of the sine and cosine func ons;
what about the other standard func ons tangent, cotangent, secant and cosecant? We discover these next.
Example

Evaluate

Integra on by subs tu on: an deriva ves of tan x


tan x dx.

The previous paragraph established that we did not know


S
the an deriva ves of tangent, hence we must assume that we have learned
something in this sec on that can help us evaluate this indenite integral.
Rewrite tan x as sin x/ cos x. While the presence of a composi on of funcons may not be immediately obvious, recognize that cos x is inside the /x
func on. Therefore, we see if se ng u = cos x returns usable results. We have

Notes:

Subs tu on

Chapter

Techniques of An dieren a on
that du = sin x dx, hence du = sin x dx. We can integrate:

tan x dx =
=

sin x
dx
cos x
sin
x dx}
| {z
cos
|{z}x du
u

du
=
u
= ln |u| + C

= ln | cos x| + C.

Some texts prefer to bring the inside the logarithm as a power of cos x, as in:
ln | cos x| + C = ln |(cos x) | + C




+C

= ln
cos x
= ln | sec x| + C.

Thus the result they give is


equivalent.
Example

Evaluate

tan x dx = ln | sec x| + C. These two answers are

Integra ng by subs tu on: an deriva ves of sec x


sec x dx.

This example employs a wonderful trick: mul ply the inteS


grand by so that we see how to integrate more clearly. In this case, we write
as
=

sec x + tan x
.
sec x + tan x

This may seem like it came out of le eld, but it works beau fully. Consider:

Notes:

sec x + tan x
dx
sec x + tan x

sec x + sec x tan x


=
dx.
sec x + tan x

sec x dx =

sec x

.
Now let u = sec x + tan x; this means du = (sec x tan x + sec x) dx, which is
our numerator. Thus:

du
=
u
= ln |u| + C
= ln | sec x + tan x| + C.

We can use similar techniques to those used in Examples


and
to nd
an deriva ves of cot x and csc x (which the reader can explore in the exercises.)
We summarize our results here.
Theorem
.
.
.

An deriva ves of Trigonometric Func ons

sin x dx = cos x + C

cos x dx = sin x + C

tan x dx = ln | cos x|+C

csc x dx = ln | csc x + cot x| + C


sec x dx = ln | sec x + tan x| + C
cot x dx = ln | sin x| + C

We explore one more common trigonometric integral.


Example

Evaluate

Integra on by subs tu on: powers of cos x and sin x


cos x dx.

(
)
We have a composi on of func ons as cos x = cos x .
However, se ng u = cos x means du = sin x dx, which we do not have in the
integral. Another technique is needed.
The process well employ is to use a Power Reducing formula for cos x (perhaps consult the back of this text for this formula), which states
S

cos x =

+ cos( x)

The right hand side of this equa on is not dicult to integrate. We have:

+ cos( x)
dx
cos x dx =
)
(
=
+ cos( x) dx.

Notes:

Subs tu on

Chapter

Techniques of An dieren a on
Now use Key Idea

x+

x+

sin( x)
sin( x)

+C

+ C.

Well make signicant use of this powerreducing technique in future sec ons.

Simplifying the Integrand


It is common to be reluctant to manipulate the integrand of an integral; at
rst, our grasp of integra on is tenuous and one may think that working with
the integrand will improperly change the results. Integra on by subs tu on
works using a dierent logic: as long as equality is maintained, the integrand can
be manipulated so that its form is easier to deal with. The next two examples
demonstrate common ways in which using algebra rst makes the integra on
easier to perform.
Example

Evaluate

Integra on by subs tu on: simplifying rst


x + x + x+
dx.
x + x+

One may try to start by se ng u equal to either the numerS


ator or denominator; in each instance, the result is not workable.
When dealing with ra onal func ons (i.e., quo ents made up of polynomial
func ons), it is an almost universal rule that everything works be er when the
degree of the numerator is less than the degree of the denominator. Hence we
use polynomial division.
We skip the specics of the steps, but note that when x + x + is divided
into x + x + x + , it goes in x +
mes with a remainder of x + . Thus
x + x + x+
x + x+

=x+ +

x+
.
x + x+

Integra ng x + is simple. The frac on can be integrated by se ng u = x +


x + , giving du = ( x + ) dx. This is very similar to the numerator. Note that

Notes:

.
du/ = (x + ) dx and then consider the following:
)

(
x + x + x+
x+
x+ +
dx
dx =
x + x+
x + x+

(x + )
= (x + ) dx +
dx
x + x+

du
= x + x+C +
u
=

x + x+C +

x + x+

ln |u| + C

ln |x + x + | + C.

In some ways, we lucked out in that a er dividing, subs tu on was able to be


done. In later sec ons well develop techniques for handling ra onal func ons
where subs tu on is not directly feasible.
Example

Evaluate
S

Rewrite

Integra on by alternate methods


x + x+

dx with, and without, subs tu on.


x

We already know how to integrate this par cular example.


x as x and simplify the frac on:

x + x+
= x + x + x .
x/
We can now integrate using the Power Rule:
(

)
x + x+

dx
=
x
+
x
+
x
dx
x/
=

x + x + x +C

This is a perfectly ne approach. We demonstrate how this can also be solved


using subs tu
on as its implementa on is rather clever.
Let u = x = x ; therefore
du = dx.
x dx = dx
x
x

x + x+

This gives us
dx = (x + x + ) du. What are we to do
x
with the other x terms? Since u = x , u = x, etc. We can then replace x and
du =

Notes:

Subs tu on

Chapter

Techniques of An dieren a on
x with appropriate powers of u. We thus have

x + x+

dx = (x + x + ) du
x

=
(u + u + ) du
=

u + u + u+C

x + x + x + C,

which is obviously the same answer we obtained before. In this situa on, subs tu on is arguably more work than our other method. The fantas c thing is
that it works. It demonstrates how exible integra on is.

Subs tu on and Inverse Trigonometric Func ons


When studying deriva ves of inverse func ons, we learned that
)
d(
tan x =
dx

+x

Applying the Chain Rule to this is not dicult; for instance,


d(
tan
dx

)
x =

We now explore how Subs tu on can be used to undo certain deriva ves that
are the result of the Chain Rule applied to Inverse Trigonometric func ons. We
begin with an example.
Example

Evaluate

Integra ng by subs tu on: inverse trigonometric func ons


+x

dx.

The integrand looks similar to the deriva ve of the arctanS


gent func on. Note:
+x

=
=
=

Notes:

( +
( +
+

(x)

(x) .

.
Thus

+x

dx =

( x ) dx.

This can be integrated using Subs tu on. Set u = x/ , hence du = dx/ or


dx = du. Thus

dx =
( ) dx
+x
+ x

du
=
+u
=
=

tan u + C
(x)
tan
+C

Example
demonstrates a general technique that can be applied to other
integrands that result in inverse trigonometric func ons. The results are summarized here.
Theorem

Integrals Involving Inverse Trigonomentric Func ons

Let a > .

(x)
dx = tan
.
+C
a +x
a
a

(x)

+C
.
dx = sin
a
a x
( )

|x|

dx = sec
+C
.
a
a
x x a
Lets prac ce using Theorem

Integra ng by subs tu on: inverse trigonometric func ons


Example
Evaluate the given indenite integrals.

dx and
dx,
dx.
+x
x
x x

Notes:

Subs tu on

Chapter

Techniques of An dieren a on
S

Theorem

+x

Each can be answered using a straigh orward applica on of

.
dx =

x x

tan
dx =

+ C, as a = .
x + C, as a =

sec

x
.
= sin + C, as a =

Most applica ons of Theorem are not as straigh orward. The next examples show some common integrals that can s ll be approached with this theorem.
Example

Evaluate

Integra ng by subs tu on: comple ng the square


x x+

dx.

Ini ally, this integral seems to have nothing in common with


S
the integrals in Theorem . As it lacks a square root, it almost certainly is not
related to arcsine or arcsecant. It is, however, related to the arctangent func on.
We see this by comple ng the square in the denominator. We give a brief
reminder of the process here.
Start with a quadra c with a leading coecient of . It will have the form of
x +bx+c. Take / of b, square it, and add/subtract it back into the expression.
I.e.,
b
b
+c
x + bx + c = x + bx +
{z
}
|
(x+b/ )

)
(
b
b
+c
= x+

In our example, we take half of and square it, ge ng . We add/subtract it


into the denominator as follows:

x x+

x x+ +
| {z }
(x )

Notes:

(x ) +

.
We can now integrate this using the arctangent rule. Technically, we need to
subs tute rst with u = x , but we can employ Key Idea instead. Thus we
have

x
dx =
dx = tan
+ C.
x x+
(x ) +
Example
Evaluate

Integrals requiring mul ple methods


x

dx.
x

This integral requires two dierent methods to evaluate it.


S
We get to those methods by spli ng up the integral:

x
x

dx =
dx
dx.
x
x
x

The rst integral is handled using a straigh orward applica on of Theorem ;


the second integral is handled by subs tu on, with u =
x . We handle
each separately.
x

dx = sin
+ C.
x

dx: Set u =
x , so du = xdx and xdx = du/ . We
x
have

du/
x

dx =
u
x

du
=
u

= u+C

=
x + C.
Combining these together, we have

x
x

+
dx = sin
x

x + C.

Subs tu on and Denite Integra on

This sec on has focused on evalua ng indenite integrals as we are learning


a new technique for nding an deriva ves. However, much of the me integraon is used in the context of a denite integral. Denite integrals that require
subs tu on can be calculated using the following workow:

Notes:

Subs tu on

Chapter

Techniques of An dieren a on
. Start with a denite integral

b
a

f(x) dx that requires subs tu on.

. Ignore the bounds; use subs tu on to evaluate


deriva ve F(x).

f(x) dx and nd an an-

b

. Evaluate F(x) at the bounds; that is, evaluate F(x) = F(b) F(a).
a

This workow works ne, but subs tu on oers an alterna ve that is powerful
and amazing (and a li le me saving).
At its heart, (using
the nota on of Theorem ) subs tu on converts inte
grals of the form F (g(x))g (x) dx into an integral of the form F (u) du with
the subs tu on of u = g(x). The following theorem states how the bounds of
a denite integral can be changed as the subs tu on is performed.
Theorem

Subs tu on with Denite Integrals

Let F and g be dieren able func ons, where the range of g is an interval
I that is contained in the domain of F. Then
g(b)
b
(
)
F g(x) g (x) dx =
F (u) du.
a

g(a)

In eect, Theorem
states that once you convert to integra ng with respect to u, you do not need to switch back to evalua ng with respect to x. A few
examples will help one understand.
Example

Evaluate

Denite integrals and subs tu on: changing the bounds


cos( x ) dx using Theorem

Observing the composi on of func ons, let u = x ,


S
hence du = dx. As dx does not appear in the integrand, divide the la er
equa on by to get du/ = dx.
By se ng u = x , we are implicitly sta ng that g(x) = x . Theorem
states that the new lower bound is g( ) = ; the new upper bound is

Notes:

.
y

g( ) = . We now evaluate the denite integral:

du
cos u
cos( x ) dx =
=
=

)
sin sin( ) .

No ce how once we converted the integral to be in terms of u, we never went


back to using x.
The graphs in Figure . tell more of the story. In (a) the area dened by the
original integrand is shaded, whereas in (b) the area dened by the new integrand is shaded. In this par cular situa on, the areas look very similar; the new
region is shorter but wider, giving the same area.
Example
Evaluate

y = cos( x )



sin u

Subs tu on

(a)
y

y=

cos(u)

Denite integrals and subs tu on: changing the bounds


sin x cos x dx using Theorem

We saw the corresponding indenite integral in Example


.
S
In that example we set u = sin x but stated that we could have let u = cos x.
For variety, we do the la er here.
Let u = g(x) = cos x, giving du = sin x dx and hence sin x dx = du. The
new upper bound is g(/ ) = ; the new lower bound is g( ) = . Note how
the lower bound is actually larger than the upper bound now. We have

/
u du (switch bounds & change sign)
sin x cos x dx =
=
=

(b)
Figure . : Graphing the areas dened by
the denite integrals of Example
.
y

u du

y = sin x cos x
.5



u = / .

In Figure . we have again graphed the two regions dened by our denite
integrals. Unlike the previous example, they bear no resemblance to each other.
However, Theorem guarantees that they have the same area.

Integra on by subs tu on is a powerful and useful integra on technique.


The next sec on introduces another technique, called Integra on by Parts. As
subs tu on undoes the Chain Rule, integra on by parts undoes the Product
Rule. Together, these two techniques provide a strong founda on on which most
other integra on techniques are based.

.5

(a)
y
y=u

.5

Notes:

.5

(b)

Figure . : Graphing the areas dened by


the denite integrals of Example
.

Exercises .

Terms and Concepts

sec ( x)dx

sec( x)dx

tan (x) sec (x)dx

( )
x cos x dx

tan (x)dx

cot x dx. Do not just refer to Theorem

csc x dx. Do not just refer to Theorem

. Subs tu on undoes what deriva ve rule?


. T/F: One can use algebra to rewrite the integrand of an integral to make it easier to evaluate.

Problems
In Exercises , evaluate the indenite integral to develop
an understanding of Subs tu on.
.

dx

(
)
x x +
dx

x x
dx
x

ln(x)
dx
x

( x ) x x+
(

x+

x + x

dx
.

dx

In Exercises
, use Subs tu on to evaluate the indenite integral involving exponen al func ons.

ex x dx

ex

ex +
dx
ex

e x
dx
x

ex ex
dx
ex

dx

dx

dx
dx

x+

x
dx
x+

x +

dx

+
dx
x

In Exercises
, use Subs tu on to evaluate the indenite integral involving trigonometric func ons.
.

sin (x) cos(x)dx

cos( x)dx

for the answer;

jus fy it through Subs tu on.

x+

for the answer;

jus fy it through Subs tu on.

dx

x+

(x )dx

In Exercises
, use Subs tu on to evaluate the indenite integral involving logarithmic func ons.
ln x
dx
x

)
ln x
dx
x

( )
ln x
dx
x

x ln (x )

dx

In Exercises
, use Subs tu on to evaluate the indenite integral involving ra onal func ons.

)(
)
x + x
x + x +
dx
x
dx
x

x + x+
dx
x

x csc

x +x +x+
dx
x

sin(x)

x
dx
x+

x + x
dx
x

x+

x x+
dx
x+

x + x + x
x + x+

x + x+
dx
x + x + x

x + x x
dx
x x+

x
x +

dx
x x

x
x x+

x +

x
x+

dx

x + x
x x+

dx

x
dx
x +

x x
dx
x + x+

sin(x)
dx
cos (x) +

In Exercises
, use Subs tu on to evaluate the indenite integral involving inverse trigonometric func ons.
.

x +

dx
x x

dx

x
x

dx
dx

x
dx
x

x + x+

.
.

x x+

dx

x + x +
x + x +

In Exercises

dx

x
dx
(x + )

dx
dx

dx
, evaluate the indenite integral.

x +

dx

cos(x)dx

dx
dx

dx

x+
dx
x + x+
( x+ )
dx
x + x+

dx

x +

dx

dx

dx

cos(x)
dx
sin (x) +

cos(x)
dx
sin (x)

x
dx
x x

x
dx
x x+

In Exercises
.

dx

x x dx

, evaluate the denite integral.

sin x cos x dx

x( x ) dx

(x + )ex

+x

+ x+

dx

dx

x x+

dx

dx

Integra on by Parts

Heres a simple integral that we cant yet evaluate:

x cos x dx.

Its a simple ma er to take the deriva ve of the integrand using the Product
Rule, but there is no Product Rule for integrals. However, this sec on introduces
Integra on by Parts, a method of integra on that is based on the Product Rule
for deriva ves. It will enable us to evaluate this integral.
The Product Rule says that if u and v are func ons of x, then (uv) = u v+uv .
For simplicity, weve wri en u for u(x) and v for v(x). Suppose we integrate both
sides with respect to x. This gives

(uv) dx = (u v + uv ) dx.

By the Fundamental Theorem of Calculus, the le side integrates to uv. The right
side can be broken up into two integrals, and we have

uv = u v dx + uv dx.
Solving for the second integral we have

uv dx = uv u v dx.

Using dieren al nota on, we can write du = u (x)dx and dv = v (x)dx and
the expression above can be wri en as follows:

u dv = uv v du.
This is the Integra on by Parts formula. For reference purposes, we state this in
a theorem.
Theorem

Integra on by Parts

Let u and v be dieren able func ons of x on an interval I containing a


and b. Then

u dv = uv v du,
and

Notes:

x=b
x=a

b

u dv = uv
a

x=b
x=a

v du.

Integra on by Parts

Chapter

Techniques of An dieren a on
Lets try an example to understand our new technique.
Example
Evaluate

Integra ng using Integra on by Parts


x cos x dx.

The key to Integra on by Parts is to iden fy part of the inS


tegrand as u and part as dv. Regular prac ce will help one make good idenca ons, and later we will introduce some principles that help. For now, let
u = x and dv = cos x dx.
It is generally useful to make a small table of these values as done below.
Right now we only know u and dv as shown on the le of Figure . ; on the right
we ll in the rest of what we need. If u = x, then du = dx. Since dv = cos x dx,
v is an an deriva ve of cos x. We choose v = sin x.
u=x
du = ?

v=?
dv = cos x dx

u=x
du = dx

v = sin x
dv = cos x dx

Figure . : Se ng up Integra on by Parts.

Now subs tute all of this into the Integra on by Parts formula, giving

x cos x dx = x sin x sin x dx.

We can then integrate sin x to get cos x + C and overall our answer is

x cos x dx = x sin x + cos x + C.

Note how the an deriva ve contains a product, x sin x. This product is what
makes Integra on by Parts necessary.
The example above demonstrates how Integra on by Parts works in general.
We try to iden fy u and dv in the integral we are given, and the key is that we
usually want to choose u and dv so that du is simpler than u and v is hopefully
not too much more complicated than dv. This will mean that the integral on the
right side of the Integra on
by Parts formula, v du will be simpler to integrate
than the original integral u dv.
In the example above, we chose u = x and dv = cos x dx. Then du = dx was
simpler than u and v = sin x is no more complicated than dv. Therefore, instead
of integra ng x cos x dx, we could integrate sin x dx, which we knew how to do.
A useful mnemonic for helping to determine u is LIATE, where
L = Logarithmic, I = Inverse Trig., A = Algebraic (polynomials),
T = Trigonometric, and E = Exponen al.

Notes:

.
If the integrand contains both a logarithmic and an algebraic term, in general
le ng u be the logarithmic term works best, as indicated by L coming before A
in LIATE.
We now consider another example.
Example
Evaluate

Integra ng using Integra on by Parts


x

xe dx.

The integrand contains an Algebraic term (x) and an Exponen al


S
term (ex ). Our mnemonic suggests le ng u be the algebraic term, so we choose
u = x and dv = ex dx. Then du = dx and v = ex as indicated by the tables below.
u=x
du = ?

v=?
dv = ex dx

u=x
du = dx

v = ex
dv = ex dx

Figure . : Se ng up Integra on by Parts.

We see du is simpler than u, while there is no change in going from dv to v.


This is good. The Integra on by Parts formula gives

xex dx = xex ex dx.


The integral on the right is simple; our nal answer is

xex dx = xex ex + C.
Note again how the an deriva ves contain a product term.
Example

Evaluate

Integra ng using Integra on by Parts


x cos x dx.

The mnemonic suggests le ng u = x instead of the trigonoS


metric func on, hence dv = cos x dx. Then du = x dx and v = sin x as shown
below.
u=x
du = ?

v=?
dv = cos x dx

u=x
du = x dx

Figure . : Se ng up Integra on by Parts.

Notes:

v = sin x
dv = cos x dx

Integra on by Parts

Chapter

Techniques of An dieren a on
The Integra on by Parts formula gives

x cos x dx = x sin x

x sin x dx.

At this point, the integral on the right is indeed simpler than the one we started
with, but to evaluate it, we need to do Integra on by Parts again. Here we
choose u = x and dv = sin x and ll in the rest below.
u= x
du = ?

v=?
dv = sin x dx

u= x
du = dx

v = cos x
dv = sin x dx

Figure . : Se ng up Integra on by Parts (again).

(
)

x cos x dx = x sin x x cos x cos x dx .

The integral all the way on the right is now something we can evaluate. It evaluates to sin x. Then going through and simplifying, being careful to keep all
the signs straight, our answer is

x cos x dx = x sin x + x cos x sin x + C.


Example
Evaluate

Integra ng using Integra on by Parts


x

e cos x dx.

This is a classic problem. Our mnemonic suggests le ng u


S
be the trigonometric func on instead of the exponen al. In this par cular example, one can let u be either cos x or ex ; to demonstrate that we do not have
to follow LIATE, we choose u = ex and hence dv = cos x dx. Then du = ex dx
and v = sin x as shown below.
u = ex
du = ?

v=?
dv = cos x dx

u = ex
du = ex dx

v = sin x
dv = cos x dx

Figure . : Se ng up Integra on by Parts.

No ce that du is no simpler than u, going against our general rule (but bear
with us). The Integra on by Parts formula yields

x
x
e cos x dx = e sin x ex sin x dx.

Notes:

.
The integral on the right is not much dierent than the one we started with, so
it seems like we have go en nowhere. Lets keep working and apply Integra on
by Parts to the new integral, using u = ex and dv = sin x dx. This leads us to the
following:
u = ex
du = ?

v=?
dv = sin x dx

u = ex
du = ex dx

v = cos x
dv = sin x dx

Figure . : Se ng up Integra on by Parts (again).

The Integra on by Parts formula then gives:


(
)

ex cos x dx = ex sin x ex cos x ex cos x dx

= ex sin x + ex cos x ex cos x dx.


It
we are back right where we started, as the right hand side contains
seems
ex cos xdx. But this is actually a good thing.
Add ex cos x dx to both sides. This gives

ex cos x dx = ex sin x + ex cos x

Now divide both sides by :

ex cos x dx =

)
ex sin x + ex cos x .

Simplifying a li le and adding the constant of integra on, our answer is thus

ex cos x dx = ex (sin x + cos x) + C.


Example

Evaluate

Integra ng using Integra on by Parts: an deriva ve of ln x


ln x dx.

One may have no ced that we have rules for integra ng the
S
familiar trigonometric func ons and ex , but we have not yet given a rule for
integra ng ln x. That is because ln x cant easily be integrated with any of the
rules we have learned up to this point. But we can nd its an deriva ve by a

Notes:

Integra on by Parts

Chapter

Techniques of An dieren a on
clever applica on of Integra on by Parts. Set u = ln x and dv = dx. This is a
good, sneaky trick to learn as it can help in other situa ons. This determines
du = ( /x) dx and v = x as shown below.
u = ln x
du = ?

v=?
dv = dx

u = ln x
du = /x dx

v=x
dv = dx

Figure . : Se ng up Integra on by Parts.

Pu ng this all together in the Integra on by Parts formula, things work out
very nicely:

ln x dx = x ln x x dx.
x

The new integral simplies to


dx, which is about as simple as things get. Its
integral is x + C and our answer is

ln x dx = x ln x x + C.
Example
Evaluate

Integra ng using Int. by Parts: an deriva ve of arctan x


arctan x dx.

The same sneaky trick we used above works here. Let u =


S
arctan x and dv = dx. Then du = /( + x ) dx and v = x. The Integra on by
Parts formula gives

x
dx.
arctan x dx = x arctan x
+x
The integral on the right can be solved by subs tu on. Taking u =
get du = x dx. The integral then becomes

arctan x dx = x arctan x
du.
u

+ x , we

The integral on the right evaluates to ln |u| + C, which becomes ln( + x ) + C.


Therefore, the answer is

arctan x dx = x arctan x ln( + x ) + C.

Notes:

Subs tu on Before Integra on


When taking deriva ves, it was common to employ mul ple rules (such as
using both the Quo ent and the Chain Rules). It should then come as no surprise
that some integrals are best evaluated by combining integra on techniques. In
par cular, here we illustrate making an unusual subs tu on rst before using
Integra on by Parts.
Example
Evaluate

Integra on by Parts a er subs tu on


cos(ln x) dx.

The integrand contains a composi on of func ons, leading


S
us to think Subs tu on would be benecial. Le ng u = ln x, we have du =
/x dx. This seems problema c, as we do not have a /x in the integrand. But
consider:
du = dx x du = dx.
x
Since u = ln x, we can use inverse func ons and conclude that x = eu . Therefore
we have that
dx = x du
= eu du.
We can thus replace ln x with u and dx with eu du. Thus we rewrite our integral
as

cos(ln x) dx = eu cos u du.


We evaluated this integral in Example
. Using the result there, we have:

cos(ln x) dx = eu cos u du
=
=
=

(
)
eu sin u + cos u + C

(
)
eln x sin(ln x) + cos(ln x) + C
(
)
x sin(ln x) + cos(ln x) + C.

Denite Integrals and Integra on By Parts

So far we have focused only on evalua ng indenite integrals. Of course, we


can use Integra on by Parts to evaluate denite integrals as well, as Theorem

Notes:

Integra on by Parts

Chapter

Techniques of An dieren a on
states. We do so in the next example.
Example
Evaluate

Denite integra on using Integra on by Parts


x ln x dx.

Our mnemonic suggests le ng u = ln x, hence dv = x dx.


S
We then get du = ( /x) dx and v = x / as shown below.
u = ln x
du = ?

v=?

dv = x dx
Figure .

u = ln x
du = /x dx

v=x /
dv = x dx

: Se ng up Integra on by Parts.

The Integra on by Parts formula then gives



x
x
ln x
x ln x dx =

dx


x
=
ln x
dx



x
x

=
ln x
(
)
x
x
=
ln x

(
) (
)
=

ln
ln
x

=
.

ln
.

In general, Integra on by Parts is useful for integra ng certain products of


func ons, like xex dx or x sin x dx. It is also useful for integrals involving
logarithms and inverse trigonometric func ons.
As stated before, integra on is generally more dicult than deriva on. We
are developing tools for handling a large array of integrals, and experience will
tell us when one tool is preferable/necessary over another. For instance, consider the three similarlooking integrals

x
x
and
xex dx.
xe dx,
xe dx

Notes:

.
While the rst is calculated easily with Integra on by Parts, the second is best
approached with Subs tu on. Taking things one step further, the third integral
has no answer in terms of elementary func ons, so none of the methods we
learn in calculus will get us the exact answer.
Integra on by Parts is a very useful method, second only to subs tu on. In
the following sec ons of this chapter, we con nue to learn other integra on
techniques. The next sec on focuses on handling integrals containing trigonometric func ons.

Notes:

Integra on by Parts

Exercises .

Terms and Concepts

sin x dx

. T/F: Integra on by Parts is useful in evalua ng integrands


that contain products of func ons.

x ln x dx

. T/F: Integra on by Parts can be thought of as the opposite


of the Chain Rule.

(x ) ln x dx

x ln(x ) dx

x ln(x ) dx

x ln x dx

. For what is LIATE useful?

Problems
In Exercises

, evaluate the given indenite integral.

x sin x dx

xe

(ln x) dx

x sin x dx

(ln(x + )) dx

x sin x dx

x sec x dx

xex dx

x csc x dx

x e dx

x x dx

x x dx

e sin x dx

sec x tan x dx

e x cos x dx

x sec x tan x dx

e x sin( x) dx

x csc x cot x dx

e x cos( x) dx

sin x cos x dx

sin(ln x) dx

sin x dx

sin( x) dx

tan ( x) dx

ln( x) dx

x tan

xe

dx

dx

x dx

In Exercises , evaluate the indenite integral a er rst


making a subs tu on.

dx

eln x dx

In Exercises , evaluate the denite integral. Note: the


corresponding indenite integrals appear in Exercises .
.

x sin x dx
xex dx

x sin x dx

/
/
/

x sin x dx

ln

x ex dx

xe x dx

xex dx

ex sin x dx

e x cos x dx

Chapter

Techniques of An dieren a on

Trigonometric Integrals

Func ons involving trigonometric func ons are useful as they are good at describing periodic behavior. This sec on describes several techniques for nding
an deriva ves of certain combina ons of trigonometric func ons.

Integrals of the form

sinm x cosn x dx

In learning the technique of Subs tu on, we saw the integral sin x cos x dx
in Example
. The integra on was not dicult, and one could easily evaluate
the indenite integral by le ng u = sin x or by le ng u = cos x. This integral is
easy since the power of both sine and cosine is .

We generalize this integral and consider integrals of the form sinm x cosn x dx,
where m, n are nonnega ve integers. Our strategy for evalua ng these integrals is to use the iden ty cos x + sin x = to convert high powers of one
trigonometric func on into the other, leaving a single sine or cosine term in the
integrand. We summarize the general technique in the following Key Idea.
Key Idea
Consider

Integrals Involving Powers of Sine and Cosine

sinm x cosn x dx, where m, n are nonnega ve integers.

. If m is odd, then m = k + for some integer k. Rewrite


sinm x = sin

k+

x = sin k x sin x = (sin x)k sin x = ( cos x)k sin x.

Then

sinm x cosn x dx =

( cos x)k sin x cosn x dx =

where u = cos x and du = sin x dx.

( u )k un du,

. If n is odd, then using subs tu ons similar to that outlined above we have

sinm x cosn x dx =

where u = sin x and du = cos x dx.

um ( u )k du,

. If both m and n are even, use the powerreducing iden


cos x =

+ cos( x)

and

sin x =

es

cos( x)

to reduce the degree of the integrand. Expand the result and apply the principles
of this Key Idea again.

Notes:

.
We prac ce applying Key Idea
Example
Evaluate
S

Trigonometric Integrals

in the next examples.

Integra ng powers of sine and cosine


sin x cos x dx.
The power of the sine term is odd, so we rewrite sin x as

sin x = sin x sin x = (sin x) sin x = ( cos x) sin x.

Our integral is now ( cos x) cos x sin x dx. Let u = cos x, hence du =

sin x dx. Making the subs tu on and expanding the integrand gives

)
(
)
(
u u +u du.
u +u u du =
( cos ) cos x sin x dx = ( u ) u du =
This nal integral is not dicult to evaluate, giving

(
)
u u + u du = u + u

= cos x +

Example

Evaluate

u +C

cos x

cos x + C.

Integra ng powers of sine and cosine


sin x cos x dx.

The powers of both the sine and cosine terms are odd, thereS
fore we can apply the techniques of Key Idea to either power. We choose to
work with the power of the cosine term since the previous example used the
sine terms power.
We rewrite cos x as
cos x = cos x cos x
= (cos x) cos x
= ( sin x) cos x

= ( sin x + sin x sin x + sin x) cos x.

We rewrite the integral as

(
)
sin x cos x dx = sin x sin x + sin x sin x + sin x cos x dx.

Notes:

Chapter

Techniques of An dieren a on

Now subs tute and integrate, using u = sin x and du = cos x dx.
(
)
sin x sin x + sin x sin x + sin x cos x dx =

(
)
u u + u u + u du
u ( u + u u + u ) du =
=

sin x

.
x

Figure .
Example

cos( x)

f(x) =

sin x +

sin x +

u +

u +C

sin x + . . .

sin x + C.

cos( x)

cos( x)

cos( x)

cos(

x)

which clearly has a dierent form than our answer in Example

f(x)

: A plot of f(x) and g(x) from


and the Technology Note.

Technology Note: The work we are doing here can be a bit tedious, but the
skills developed (problem solving, algebraic manipula on, etc.) are important.
Nowadays problems of this sort are o en solved usinga computer algebra system. The powerful program Mathema ca integrates sin x cos x dx as

g(x)

u +

g(x) =

sin x

sin x +

sin x

sin x +

cos(

x)

cos(

x)

, which is
sin x.

Figure . shows a graph of f and g; they are clearly not equal, but they dier
only by a constant. That is g(x) = f(x) + C for some constant C. So we have
two dierent an deriva ves of the same func on, meaning both answers are
correct.
Example

Evaluate

Integra ng powers of sine and cosine


cos x sin x dx.

The powers of sine and cosine are both even, so we employ


S
the powerreducing formulas and algebra as follows.
) (
)

(
+ cos( x)
cos( x)
cos x sin x dx =
dx

cos( x)
+ cos( x) + cos ( x)

dx
=

(
)
=
+ cos( x) cos ( x) cos ( x) dx

The cos( x) term is easy to integrate, especially with Key Idea . The cos ( x)
term is another trigonometric integral with an even power, requiring the power
reducing formula again. The cos ( x) term is a cosine func on with an odd
power, requiring a subs tu on as done before. We integrate each in turn below.

Notes:

cos( x) dx =

cos ( x) dx =

+ cos( x)

sin( x) + C.
dx =

Finally, we rewrite cos ( x) as


cos ( x) = cos ( x) cos( x) =

(
x+

)
sin( x) + C.

)
sin ( x) cos( x).

Le ng u = sin( x), we have du = cos( x) dx, hence

(
)
sin ( x) cos( x) dx
cos ( x) dx =

=
( u ) du
)
(
=
u u +C
)
(
sin( x) sin ( x) + C
=

Pu ng all the pieces together, we have

(
)
cos x sin x dx =
+ cos( x) cos ( x) cos ( x) dx
[
(
(
)
sin( x)
= x + sin( x) x + sin( x)
[
]
=
x sin( x) + sin ( x) + C

sin ( x)

The process above was a bit long and tedious, but being able to work a problem such as this from start to nish is important.

Integrals of the form sin(mx) sin(nx) dx,

and sin(mx) cos(nx) dx.

cos(mx) cos(nx) dx,

Func ons that contain products of sines and cosines of diering periods are
important in many applica ons including the analysis of sound waves. Integrals
of the form

sin(mx) sin(nx) dx,


cos(mx) cos(nx) dx and
sin(mx) cos(nx) dx

Notes:

)]

+C

Trigonometric Integrals

Chapter

Techniques of An dieren a on
are best approached by rst applying the Product to Sum Formulas found in the
back cover of this text, namely
sin(mx) sin(nx) =
cos(mx) cos(nx) =
sin(mx) cos(nx) =
Example
Evaluate

(
)
(
)]
cos (m n)x cos (m + n)x
[
(
)
(
)]
cos (m n)x + cos (m + n)x
[ (
)
(
)]
sin (m n)x + sin (m + n)x

Integra ng products of sin(mx) and cos(nx)


sin( x) cos( x) dx.

The applica on of the formula and subsequent integra on


S
are straigh orward:

[
]
sin( x) cos( x) dx =
sin( x) + sin( x) dx
= cos( x)

Integrals of the form

cos( x) + C

tanm x secn x dx.

When evalua ng integrals of the form sinm x cosn x dx, the Pythagorean
Theorem allowed us to convert even powers of sine into even powers of cosine,
and viseversa. If, for instance, the power of sine was odd, we pulled out one
sin x and converted the remaining even power of sin x into a func on using powers of cos x, leading to an easy subs tu on.

The same basic strategy applies to integrals of the form tanm x secn x dx,
albeit a bit more nuanced. The following three facts will prove useful:

d
dx (tan x)

= sec x,

d
dx (sec x)

= sec x tan x , and

+ tan x = sec x (the Pythagorean Theorem).

If the integrand can be manipulated to separate a sec x term with the remaining secant power even, or if a sec x tan x term can be separated with the
remaining tan x power even, the Pythagorean Theorem can be employed, leading to a simple subs tu on. This strategy is outlined in the following Key Idea.

Notes:

Key Idea
Consider

Trigonometric Integrals

Integrals Involving Powers of Tangent and Secant

tanm x secn x dx, where m, n are nonnega ve integers.

. If n is even, then n = k for some integer k. Rewrite secn x as


secn x = sec k x = sec

x sec x = ( + tan x)k sec x.

Then

tan x sec x dx =

tan x( + tan x)

sec x dx =

um ( + u )k du,

where u = tan x and du = sec x dx.


. If m is odd, then m = k + for some integer k. Rewrite tanm x secn x as
tanm x secn x = tan

k+

x secn x = tan k x secn x sec x tan x = (sec x )k secn x sec x tan x.

Then

tan x sec x dx =

(sec x ) sec

x sec x tan x dx =

(u )k un du,

where u = sec x and du = sec x tan x dx.


. If n is odd and m is even, then m = k for some integer k. Convert tanm x to (sec x )k . Expand
the new integrand and use Integra on By Parts, with dv = sec x dx.
. If m is even and n = , rewrite tanm x as
tanm x = tanm x tan x = tanm x(sec x ) = tanm sec x tanm x.
So

tan x dx =

tan

sec x dx
{z
}

apply rule #

tanm x dx .
|
{z
}

apply rule # again

The techniques described in items and of Key Idea are rela vely straightforward, but the techniques in items and can be rather tedious. A few examples will help with these methods.

Notes:

Chapter

Techniques of An dieren a on
Example
Evaluate

Integra ng powers of tangent and secant


tan x sec x dx.

Since the power of secant is even, we use rule # from Key


S
Idea and pull out a sec x in the integrand. We convert the remaining powers
of secant into powers of tangent.

tan x sec x dx = tan x sec x sec x dx

(
)
= tan x + tan x sec x dx

Now subs tute, with u = tan x, with du = sec x dx.

)
(
+ u du
= u

We leave the integra on and subsequent subs tu on to the reader. The nal
answer is
=

Example
Evaluate

tan x +

tan x +

tan x + C.

Integra ng powers of tangent and secant


sec x dx.

We apply rule # from Key Idea as the power of secant is


S
odd and the power of tangent is even ( is an even number). We use Integra on
by Parts; the rule suggests le ng dv = sec x dx, meaning that u = sec x.
u = sec x
du = ?

v=?
dv = sec x dx
Figure .

u = sec x

du = sec x tan x dx dv = sec x dx

: Se ng up Integra on by Parts.

Employing Integra on by Parts, we have

sec x dx = sec
|{z}x |sec{zx dx}
u

= sec x tan x

Notes:

dv

v = tan x

sec x tan x dx.

.
This new integral also requires applying rule # of Key Idea :

(
)
= sec x tan x sec x sec x dx

= sec x tan x sec x dx + sec x dx

= sec x tan x sec x dx + ln | sec x + tan x|


In previous applica ons of Integra on by Parts, we have seen where
the original
integral has reappeared in our work. We resolve this by adding sec x dx to
both sides, giving:

sec x dx = sec x tan x + ln | sec x + tan x|

(
)
sec x dx =
sec x tan x + ln | sec x + tan x| + C
We give one more example.
Integra ng powers of tangent and secant

Example

tan x dx.

Evaluate
S

We employ rule # of Key Idea .

tan x dx = tan x tan x dx

(
)
= tan x sec x dx

= tan x sec x dx tan x dx

Integrate the rst integral with subs tu on, u = tan x; integrate the second by
employing rule # again.

= tan x tan x tan x dx

(
)
= tan x tan x sec x dx

= tan x tan x sec x dx + tan x dx

Notes:

Trigonometric Integrals

Chapter

Techniques of An dieren a on
Again, use subs tu on for the rst integral and rule # for the second.

sec x

tan x

tan x +

tan x

tan x + tan x x + C.

dx

These la er examples were admi edly long, with repeated applica ons of
the same rule. Try to not be overwhelmed by the length of the problem, but
rather admire how robust this solu on method is. A trigonometric func on of
a high power can be systema cally reduced to trigonometric func ons of lower
powers un l all an deriva ves can be computed.
The next sec on introduces an integra on technique known as Trigonometric Subs tu on, a clever combina on of Subs tu on and the Pythagorean Theorem.

Notes:

Exercises .

Terms and Concepts


. T/F:

sin x cos x dx cannot be evaluated using the tech-

niques described in this sec on since both powers of sin x


and cos x are even.

. T/F: sin x cos x dx cannot be evaluated using the tech-

niques described in this sec on since both powers of sin x


and cos x are odd.

. T/F: This sec on addresses how to evaluate indenite integrals such as sin x tan x dx.

Problems
In Exercises

, evaluate the indenite integral.

tan x sec x dx

tan x sec x dx

tan x sec x dx

tan x sec x dx

tan x sec x dx

tan x sec x dx

tan x dx

sec x dx

tan x sec x dx

tan x sec x dx

sin x cos x dx

sin x cos x dx

sin x cos x dx

sin x cos x dx

sin x cos x dx

sin x cos x dx

.
.

sin x cos x dx

sin( x) cos( x) dx

sin(x) cos( x) dx

sin( x) sin( x) dx

sin(x) sin( x) dx

cos(x) cos( x) dx

cos

( )
x cos(x) dx

In Exercises , evaluate the denite integral. Note: the


corresponding indenite integrals appear in the previous set.

sin x cos x dx

sin x cos x dx

sin x cos x dx

sin( x) cos( x) dx

cos(x) cos( x) dx

/
/

tan x sec x dx

tan x sec x dx

Chapter

Techniques of An dieren a on

Trigonometric Subs tu on

In Sec on . we dened the denite integral as the signed area under the
curve. In that sec on we had not yet learned the Fundamental Theorem of
Calculus, so we evaluated special denite integrals which described nice, geometric shapes. For instance, we were able to evaluate

( . )
x dx =

as we recognized that f(x) =


x described the upper half of a circle with
radius .
We have since learned a number of integra on techniques, including Subs tu on and Integra on by Parts, yet we are s ll unable to evaluate the above
integral without resor ng to a geometric interpreta on. This sec on introduces
Trigonometric Subs tu on, a method of integra on that lls this gap in our integra on skill. This technique works on the same principle as Subs tu on as found
in Sec on . , though it can feel backward. In Sec on . , we set u = f(x), for
some func on f, and replaced f(x) with u. In this sec on, we will set x = f(),
where f is a trigonometric func on, then replace x with f().
We start by demonstra ng this method in evalua ng the integral in ( . ).
A er the example, we will generalize the method and give more examples.
Example

Using Trigonometric Subs tu on

Evaluate

x dx.

We begin by no ng that sin + cos = , and hence


S
cos = sin . If we let x = sin , then x = sin = cos .
Se ng x = sin gives dx = cos d. We are almost ready to subs tute.
We also wish to change our bounds of integra on. The bound x = corresponds to = / (for when = / , x = sin = ). Likewise, the
bound of x = is replaced by the bound = / . Thus
/

x dx =
sin ( cos ) d

=
=

/
/

cos cos d

/
/

| cos | cos d.

On [/ , / ], cos is always posi ve, so we can drop the absolute value bars,
then employ a powerreducing formula:

Notes:

cos d

/
/

Trigonometric Subs tu on

)
+ cos( ) d

/
)
sin( )

This matches our answer from before.

We now describe in detail Trigonometric


on. This method
Subs tu
excels
when dealing with integrands that contain a x , x a and x + a .
The following Key Idea outlines the procedure for each case, followed by more
examples. Each right triangle acts as a reference to help us understand the rela onships between x and .
Key Idea

Trigonometric Subs tu on

(a) For integrands containing a x :


Let x = a sin ,
Thus = sin

dx = a cos d

(x/a), for / / .

On this interval, cos , so

a x = a cos
(b) For integrands containing
Let x = a tan ,

dx = a sec d

x
.

dx = a sec tan d

Thus = sec (x/a). If x/a , then


if x/a , then / < .

< / ;

We restrict our work to where x a, so x/a , and


< / . On this interval, tan , so

x a = a tan

Notes:

a2 x2
2

On this interval, sec > , so

x + a = a sec

(c) For integrands containing x a :

x +a :

Thus = tan (x/a), for / < < / .

Let x = a sec ,

x2 a2

Chapter

Techniques of An dieren a on
Example
Evaluate

Using Trigonometric Subs tu on

+x

dx.

Using Key
and
set x =
Idea (b), we recognize a =
S
tan
.
This
makes
dx
=
sec

d.
We
will
use
the
fact
that
+x =

+ tan =
sec =
sec . Subs tu ng, we have:

dx =
sec d
+x
+ tan

sec

=
d
sec

= sec d


= ln sec + tan + C.

While the integra on steps are over, we are not yet done. The original problem
was stated in terms of x, whereas our answer is given in terms of . We must
convert back to x.

tan , we
The reference triangle given in Key Idea (b) helps. With x =
have

x +
x
tan =
and sec =
.
This gives



dx = ln sec + tan + C
+x


x +
x

+ + C.
= ln

We can leave this answer as is, or we can use a logarithmic iden ty to simplify
it. Note:




x +

(
)
x


ln
+ + C = ln
x + + x + C








= ln + ln x + + x + C


= ln x + + x + C,
( )
where the ln /
term is absorbed into the constant C. (In Sec on . we
will learn another way of approaching this problem.)

Notes:

.
Example

Evaluate

Using Trigonometric Subs tu on

x dx.

We start by rewri ng the integrand so that it looks like x a


for some value of a:
(
)

x =
x
S

( )

So we have a = / , and following Key Idea (c), we set x = sec , and hence
dx = sec tan d. We now rewrite the integral with these subs tu ons:

( )

dx
x dx =
x
=
=
=
=
=
=

sec

sec tan

(
)
(sec ) sec tan d
(
)
tan sec tan d

tan sec d
(
)
sec sec d

(
)
sec sec d.

We integrated sec in Example


, nding its an deriva ves to be

(
)
sec tan + ln | sec + tan | + C.
sec d =

Thus

x dx =
=
=

Notes:

)
sec sec d

( (

)
)
sec tan + ln | sec + tan | ln | sec + tan | + C

(sec tan ln | sec + tan |) + C.

Trigonometric Subs tu on

Chapter

Techniques of An dieren a on
We are not yet done. Our original integral is given in terms of x, whereas our
nal answer, as given, is in terms of . We need to rewrite our answer in terms
of x. With a = / , and x = sec , the reference triangle in Key Idea (c)
shows that
/

tan = x / ( / ) =
x /
and sec = x.

Thus
(


)
sec tan ln sec + tan + C =


)
x / ln x +
x / +C
(


)
x x / ln x +
x / + C.
x

The nal answer is given in the last line above, repeated here:

(


)
x dx =
x x / ln x +
x / + C.
Example
Evaluate
S

and hence

Using Trigonometric Subs tu on


x
dx.
x
We use Key Idea (a) with a = , x = sin , dx = cos
x = cos . This gives

x
cos
dx =
( cos ) d
x
sin

= cot d

= (csc ) d
= cot + C.

We need to rewrite our answer in terms


triangle found
of x. Using the reference
x /x and = sin (x/ ). Thus
in Key Idea (a), we have cot =


(x)
x
x
dx =
sin
+ C.
x
x
Trigonometric
Substu on can be

applied in many situa ons, even those not


of the form a x , x a or x + a . In the following example, we apply it to an integral we already know how to handle.

Notes:

.
Example
Evaluate

Using Trigonometric Subs tu on


x +

dx.

We know the answer already as tan x+C. We apply Trigonometric Subs tu on here to show that we get the same answer without inherently relying on knowledge of the deriva ve of the arctangent func on.
Using Key Idea (b), let x = tan , dx = sec d and note that x + =
tan + = sec . Thus

dx =
sec d
x +
sec

=
d
S

= + C.

Since x = tan , = tan x, and we conclude that

x +

dx = tan x + C.

The next example is similar to the previous one in that it does not involve a
squareroot. It shows how several techniques and iden es can be combined
to obtain a solu on.
Example

Evaluate

Using Trigonometric Subs tu on


(x + x +

dx.

We start by comple ng the square, then make the subs tuS


on u = x + , followed by the trigonometric subs tu on of u = tan :

(x + x +

dx =

(x + ) +

) dx =

(u + )

du.

Now make the subs tu on u = tan , du = sec d:


=
=
=

Notes:

(tan + )
(sec )
cos d.

sec d

sec d

Trigonometric Subs tu on

Chapter

Techniques of An dieren a on
Applying a power reducing formula, we have
=

cos( )

sin( ) + C. ( . )

We need to return to the variable x. As u = tan , = tan u. Using the


iden ty sin( ) = sin cos and using the reference triangle found in Key
Idea (b), we have
sin( ) =

u
u +

u +

u
.
u +

Finally, we return to x with the subs tu on u = x + . We start with the expression in Equa on ( . ):
+

sin( ) + C =
=

tan u +

u
u +

tan (x + ) +

+C
x+
(x + x +

Sta ng our nal result in one line,

x+
dx = tan (x + ) +
(x + x + )
(x + x +

+ C.

+ C.

Our last example returns us to denite integrals, as seen in our rst example.
Given a denite integral that can be evaluated using Trigonometric Subs tu on,
we could rst evaluate the corresponding indenite integral (by changing from
an integral in terms of x to one in terms of , then conver ng back to x) and then
evaluate using the original bounds. It is much more straigh orward, though, to
change the bounds as we subs tute.
Example

Evaluate

Denite integra on and Trigonometric Subs tu on


x

dx.
x +

S
Using Key Idea (b), we set x = tan , dx = sec d,
= sec . As we subs tute, we can also change the
and note that x +
bounds of integra on.
The lower bound of the original integral is x = . As x = tan , we solve for
and nd = tan (x/ ). Thus the new lower bound is = tan ( ) = . The

Notes:

.
original upper bound is x = , thus the new upper bound is = tan ( / ) =
/ .
Thus we have

/
x
tan

dx =
sec d
sec
x +
/
tan sec d.
=
We encountered this indenite integral in Example
where we found

(
)
sec tan ln | sec + tan | .
tan sec d =

So

/
)
sec tan ln | sec + tan |

(
)
ln(
+ )
(

tan sec d =
=
.

The following equali es are very useful when evalua ng integrals using Trigonometric Subs tu on.
Key Idea

Useful Equali es with Trigonometric Subs tu on

. sin( ) = sin cos


. cos( ) = cos sin = cos = sin

(

)
sec tan + ln sec + tan + C
.
sec d =

cos d =

)
+ cos( ) d =

)
+ sin cos + C.

The next sec on introduces Par al Frac on Decomposi on, which is an algebraic technique that turns complicated frac ons into sums of simpler fracons, making integra on easier.

Notes:

Trigonometric Subs tu on

Exercises .

Terms and Concepts

In Exercises
, evaluate the indenite integrals. Some
may be evaluated without Trigonometric Subs tu on.

. Trigonometric Subs tu on works on the same principles as


Integra on by Subs tu on, though it can feel
.
. If one uses
Trigonometric Subs tu on on an integrand containing
x , then one should set x =
.

(x + )

x
(x + )

(x + x +

x ( x )

. Consider the Pythagorean Iden ty sin + cos = .


(a) What iden ty is obtained when both sides are divided by cos ?
(b) Use the new iden ty to simplify tan + .
. Why does Key Idea
and not |a cos |?

(a) state that

a x = a cos ,

Problems
In Exercises , apply Trigonometric Subs tu on to evaluate the indenite integrals.

x + dx

x + dx

x dx

x dx

x dx

dx

x + dx

x dx
x +
x
x

dx

x
x

dx
dx
dx

x dx
dx

x
x

dx

)
/

dx
dx

x
dx
x
x
x +

dx

In Exercises
, evaluate the denite integrals by making the proper trigonometric subs tu on and changing the
bounds of integra on. (Note: each of the corresponding
indenite integrals has appeared previously in this Exercise
set.)
.

x dx

x
x

dx

dx

(x + )

dx

x + dx

x dx

dx

x dx

x dx

Par al Frac on Decomposi on

In this sec on we inves gate the an deriva ves of ra onal func ons. Recall that
p(x)
ra onal func ons are func ons of the form f(x) = q(x)
, where p(x) and q(x) are
polynomials and q(x) = . Such func ons arise in many contexts, one of which
is the solving of certain fundamental dieren al equa ons.
We begin with an examplethat demonstrates the mo va on behind this
sec on. Consider the integral
dx. We do not have a simple formula
x
for this (if the denominator were x + , we would recognize the an deriva ve
as being the arctangent func on). It can be solved using Trigonometric Subs tu on, but note how the integral is easy to evaluate once we realize:

/
x

/
.
x+

Thus

dx =
=

/
x

dx

ln |x |

/
x+

dx

ln |x + | + C.

This sec on teaches how to decompose

into

/
x

/
.
x+

We start with a ra onal func on f(x) = p(x)


q(x) , where p and q do not have any
common factors and the degree of p is less than the degree of q. It can be shown
that any polynomial, and hence q, can be factored into a product of linear and
irreducible quadra c terms. The following Key Idea states how to decompose a
ra onal func on into a sum of ra onal func ons whose denominators are all of
lower degree than q.

Notes:

Par al Frac on Decomposi on

Chapter

Techniques of An dieren a on

Key Idea

Par al Frac on Decomposi on

p(x)
be a ra onal func on, where the degree of p is less than the
q(x)
degree of q.

Let

. Linear Terms: Let (x a) divide q(x), where (x a)n is the highest


p(x)
power of (x a) that divides q(x). Then the decomposi on of q(x)
will contain the sum
A
A
An
+
+ +
.
(x a) (x a)
(x a)n
. Quadra c Terms: Let x + bx + c divide q(x), where (x + bx + c)n
is the highest power of x + bx + c that divides q(x). Then the
decomposi on of p(x)
q(x) will contain the sum
B x+C
Bn x + Cn
B x+C
+
+ +
.
x + bx + c (x + bx + c)
(x + bx + c)n
To nd the coecients Ai , Bi and Ci :
. Mul ply all frac ons by q(x), clearing the denominators. Collect
like terms.
. Equate the resul ng coecients of the powers of x and solve the
resul ng system of linear equa ons.

The following examples will demonstrate how to put this Key Idea into pracce. Example
stresses the decomposi on aspect of the Key Idea.
Example

Decomposing into par al frac ons

Decompose f(x) =

(x + )(x ) (x + x + )(x + x + )
for the resul ng coecients.

without solving

The denominator is already factored, as both x + x + and


S
x + x + cannot be factored further. We need to decompose f(x) properly.
Since (x + ) is a linear term that divides the denominator, there will be a
A
x+

Notes:

.
term in the decomposi on.
As (x ) divides the denominator, we will have the following terms in the
decomposi on:
B
D
C
and
.
,
x
(x )
(x )
The x + x + term in the denominator results in a
Finally, the (x + x + ) term results in the terms
Gx + H
x +x+

and

Ex + F
x +x+

term.

Ix + J
.
(x + x + )

All together, we have


(x + )(x ) (x + x + )(x + x + )

A
B
C
D
+
+
+
+
x+
x
(x )
(x )
Ex + F
Gx + H
Ix + J
+
+
x +x+
x +x+
(x + x + )
=

Solving for the coecients A, B . . . J would be a bit tedious but not hard.
Example

Decomposing into par al frac ons

Perform the par al frac on decomposi on of

The denominator factors into two linear terms: x


S
(x )(x + ). Thus
A
B
=
+
.
x
x
x+

To solve for A and B, rst mul ply through by x

= (x )(x + ):

A(x )(x + ) B(x )(x + )


+
x
x+
= A(x + ) + B(x )
= Ax + A + Bx B
=

Now collect like terms.


= (A + B)x + (A B).
The next step is key. Note the equality we have:
= (A + B)x + (A B).

Notes:

Par al Frac on Decomposi on

Chapter

Techniques of An dieren a on
For claritys sake, rewrite the le hand side as
x+

= (A + B)x + (A B).

On the le , the coecient of the x term is ; on the right, it is (A + B). Since


both sides are equal, we must have that = A + B.
Likewise, on the le , we have a constant term of ; on the right, the constant
term is (A B). Therefore we have = A B.
We have two linear equa ons with two unknowns. This one is easy to solve
by hand, leading to
A+B=
A= /

.
AB=
B= /

Thus

x
Example

/
x

/
.
x+

Integra ng using par al frac ons

Use par al frac on decomposi on to integrate

Note: Equa on . oers a direct route to


nding the values of A, B and C. Since the
equa on holds for all values of x, it holds
in par cular when x = . However, when
x = , the right hand side simplies to
A( + ) = A. Since the le hand side
is s ll , we have = A. Hence A = / .
Likewise, the equality holds when x =
; this leads to the equa on = C.
Thus C = / .
Knowing A and C, we can nd the value of
B by choosing yet another value of x, such
as x = , and solving for B.

Idea

(x )(x + )

dx.

We decompose the integrand as follows, as described by Key


(x )(x + )

A
x

B
x+

C
.
(x + )

To solve for A, B and C, we mul ply both sides by (x )(x + ) and collect like
terms:
= A(x + ) + B(x )(x + ) + C(x )

= Ax + Ax + A + Bx + Bx B + Cx C

= (A + B)x + ( A + B + C)x + ( A B C)
We have
x + x+

= (A + B)x + ( A + B + C)x + ( A B C)

leading to the equa ons


A+B= ,

A+B+C=

and

A BC= .

These three equa ons of three unknowns lead to a unique solu on:
A= / ,

Notes:

B= /

and

C= / .

( . )

.
Thus

(x )(x + )

dx =

/
x

dx +

/
dx +
x+

Par al Frac on Decomposi on

/
dx.
(x + )

Each can be integrated with a simple subs tu on with u = x or u = x+


(or by directly applying Key Idea
as the denominators are linear func ons).
The end result is

dx = ln |x | ln |x + | +
+ C.
(x )(x + )
(x + )
Example

Integra ng using par al frac ons

x
dx.
(x )(x + )

Use par al frac on decomposi on to integrate

Key Idea
presumes that the degree of the numerator is
S
less than the degree of the denominator. Since this is not the case here, we
begin by using polynomial division to reduce the degree of the numerator. We
omit the steps, but encourage the reader to verify that
x+
x
=x+ +
.
(x )(x + )
(x )(x + )
Using Key Idea

, we can rewrite the new ra onal func on as:


A
x+
=
(x )(x + )
x

B
x+

for appropriate values of A and B. Clearing denominators, we have


x+

= A(x + ) + B(x )

= (A + B)x + ( A B).

This implies that:


=A+B
= A B.
Solving this system of linear equa ons gives
/ =A
/ = B.

Notes:

Note: The values of A and B can be quickly


found using the technique described in
the margin of Example
.

Chapter

Techniques of An dieren a on
We can now integrate.
)

(
x
/
/
x+ +
dx
dx =
+
(x )(x + )
x
x+
x
+ x+
ln |x | +
ln |x + | + C.
=
Integra ng using par al frac
ons

Example

Use par al frac on decomposi on to evaluate

x + x+
(x + )(x + x +

dx.

The degree of the numerator is less than the degree of the


S
denominator so we begin by applying Key Idea . We have:
x + x+
(x + )(x + x +

A
x+

Bx + C
x + x+

Now clear the denominators.


x +

x+

= A(x + x +

) + (Bx + C)(x + )

= (A + B)x + ( A + B + C)x + (

A + C).

This implies that:


=A+B
= A+B+C
= A + C.
Solving this system of linear equa ons gives the nice result of A = , B =
C = . Thus
)

(
x + x+
x
dx.
dx =
+
(x + )(x + x + )
x+
x + x+

and

The rst term of this new integrand is easy to evaluate; it leads to a ln |x+ |
term. The second term is not hard, but takes several steps and uses subs tu on
techniques.
x
The integrand
has a quadra c in the denominator and a linear
x + x+
term in the numerator. This leads us to try subs tu on. Let u = x + x + , so
du = ( x + ) dx. The numerator is x , not x + , but we can get a x +

Notes:

.
term in the numerator by adding in the form of .
x
x + x+

x +
x + x+
x+
=

x + x+
x + x+
=

We can now integrate the rst term with subs tu on, leading to a ln |x + x+ |
term. The nal term can be integrated using arctangent. First, complete the
square in the denominator:
x + x+

(x + ) +

An an deriva ve of the la er term can be found using Theorem


tu on:
(
)

x+

dx = tan
+ C.
x + x+

and subs -

Lets start at the beginning and put all of the steps together.

x + x+
(x + )(x + x +

)
x
dx
x+
x + x+

x+
=
dx +
dx
dx
x+
x + x+
x + x+
(
)
x+

+ C.
= ln |x + | + ln |x + x + | tan

dx =

As with many other problems in calculus, it is important to remember that one


is not expected to see the nal answer immediately a er seeing the problem.
Rather, given the ini al problem, we break it down into smaller problems that
are easier to solve. The nal answer is a combina on of the answers of the
smaller problems.
Par al Frac on Decomposi on is an important tool when dealing with ra onal func ons. Note that at its heart, it is a technique of algebra, not calculus,
as we are rewri ng a frac on in a new form. Regardless, it is very useful in the
realm of calculus as it lets us evaluate a certain set of complicated integrals.
The next sec on introduces new func ons, called the Hyperbolic Func ons.
They will allow us to make subs tu ons similar to those found when studying
Trigonometric Subs tu on, allowing us to approach even more integra on problems.

Notes:

Par al Frac on Decomposi on

Exercises .

Terms and Concepts

x +x+
x +x

x
x x

without solving for the coecients, as


x x
done in Example
.

x x+
x x+

x
without solving for the coecients, as
. Decompose
x
done in Example
.

x + x + x

x +x+
x + x+

x + x+
dx
(x + )( x + x )

x + x
(x )(x + x +

x +x+
dx
(x + )(x + )

x x
(x )(x + x +

x x+
(x )(x x +

x + x+
(x + )(x + x +

. Fill in the blank: Par al Frac on Decomposi on is a method


of rewri ng
func ons.
. T/F: It is some mes necessary to use polynomial division
before using Par al Frac on Decomposi on.
. Decompose

x
x
done in Example

without solving for the coecients, as

. Decompose

x+
without solving for the coecients, as
x + x
done in Example
.

. Decompose

Problems
In Exercises

, evaluate the indenite integral.

x+
x + x

x
dx
x +x

x x+
dx
(x )(x + )( x)

x x
dx
( x + )( x )( x )

dx

dx

In Exercises

x+
dx
(x + )
x
(x + )

dx

x + x+
x(x + )

dx

dx

dx

dx

dx

dx

x+
dx
(x + )(x + )

x+
dx
( x + )(x + )

(x

dx

dx

dx

dx

, evaluate the denite integral.

x + x
dx
)(x + x + )

x
dx
(x + )(x + x + )

Hyperbolic Func ons

Hyperbolic Func ons

y
x2 + y 2 = 1

The hyperbolic func ons are a set of func ons that have many applica ons to
mathema cs, physics, and engineering. Among many other applica ons, they
are used to describe the forma on of satellite rings around planets, to describe
the shape of a rope hanging from two points, and have applica on to the theory
of special rela vity. This sec on denes the hyperbolic func ons and describes
many of their proper es, especially their usefulness to calculus.
These func ons are some mes referred to as the hyperbolic trigonometric
func ons as there are many, many connec ons between them and the standard trigonometric func ons. Figure . demonstrates one such connec on.
Just as cosine and sine are used to dene points on the circle dened by x +y =
, the func ons hyperbolic cosine and hyperbolic sine are used to dene points
on the hyperbola x y = .
We begin with their deni on.
Deni on
. cosh x =

y
2

x y =1
2

(cosh ,sinh )

. sech x =

. sinh x =

ex ex

. csch x =

. tanh x =

sinh x
cosh x

. coth x =

cosh x
sinh x
cosh x
sinh x

These hyperbolic func ons are graphed in Figure . . In the graphs of cosh x
and sinh x, graphs of ex / and ex / are included with dashed lines. As x gets
large, cosh x and sinh x each act like ex / ; when x is a large nega ve number,
cosh x acts like ex / whereas sinh x acts like ex / .
No ce the domains of tanh x and sech x are (, ), whereas both coth x
and csch x have ver cal asymptotes at x = . Also note the ranges of these
func ons, especially tanh x: as x , both sinh x and cosh x approach ex / ,
hence tanh x approaches .
The following example explores some of the proper es of these func ons
that bear remarkable resemblance to the proper es of their trigonometric counterparts.

Notes:

(cos ,sin )

Hyperbolic Func ons


ex + ex

.
Figure . : Using trigonometric funcons to dene points on a circle and hyperbolic func ons to dene points on a
hyperbola. The area of the shaded regions are included in them.

Pronuncia on Note:
cosh rhymes with gosh,
sinh rhymes with pinch, and
tanh rhymes with ranch.

Chapter

Techniques of An dieren a on

5
x

f(x) = cosh x

f(x) = coth x

f(x) = tanh x

.
Figure .

Use Deni on

Notes:

: Graphs of the hyperbolic func ons.

to rewrite the following expressions.

. tanh x + sech x

Exploring proper es of hyperbolic


func ons
.

. cosh x sinh x
.

f(x) = csch x

f(x) = sech x
x

Example
.

f(x) = sinh x

cosh x sinh x

)
cosh x
(
)
d
sinh x
. dx
(
)
d
. dx
tanh x
.

d
dx

cosh x sinh x =

=
=

ex + ex

e x + ex ex + e

ex ex
x

e x ex ex + e

= .

So cosh x sinh x = .
sinh x
+
cosh x cosh x
sinh x +
Now use iden ty from # .
=
cosh x
cosh x
=
= .
cosh x

tanh x + sech x =

So tanh x + sech x = .

cosh x sinh x =

=
=

ex + ex

e x e

e x e

Thus cosh x sinh x = sinh( x).


(
)
)
d(
d ex + ex
cosh x =
dx
dx
ex ex
=

= sinh x.
So
.

(
d

cosh x = sinh x.

(
d

)
sinh x = cosh x.

dx

(
)
)
d(
d ex ex
sinh x =
dx
dx
ex + ex
=
= cosh x.

So

Notes:

dx

)(

ex ex

= sinh( x).

Hyperbolic Func ons

Chapter

Techniques of An dieren a on
(
)
)
d(
d sinh x
tanh x =
dx
dx cosh x
cosh x cosh x sinh x sinh x
=
cosh x

cosh x
= sech x.

So

d
dx

)
tanh x = sech x.

The following Key Idea summarizes many of the important iden es rela ng
to hyperbolic func ons. Each can be veried by referring back to Deni on .
Key Idea
Basic Iden

Useful Hyperbolic Func on Proper es


es

. cosh x sinh x =
. tanh x + sech x =
. coth x csch x =
. cosh x = cosh x + sinh x
. sinh x = sinh x cosh x
. cosh x =
. sinh x =

cosh x +
cosh x

Integrals
Deriva ves

(
)
d
. dx
cosh x = sinh x
.
cosh x dx = sinh x + C
(
)
d
. dx sinh x = cosh x

(
)
.
sinh x dx = cosh x + C
d
tanh x = sech x
. dx

(
)
d
sech x = sech x tanh x
. dx
.
tanh x dx = ln(cosh x) + C
(
)
d

. dx csch x = csch x coth x


.
coth x dx = ln | sinh x | + C
(
)
d
. dx
coth x = csch x

We prac ce using Key Idea

Deriva ves and integrals of hyperbolic func ons


Example
Evaluate the following deriva ves and integrals.
)
d(
cosh x
dx

.
sech ( t ) dt

Notes:

ln

cosh x dx

.
S

. Using the Chain Rule directly, we have

d
dx

)
cosh x = sinh x.

Just to demonstrate that it works, lets also use the Basic Iden ty found in
Key Idea : cosh x = cosh x + sinh x.
)
)
d(
d(
cosh x =
cosh x + sinh x = cosh x sinh x + sinh x cosh x
dx
dx
= cosh x sinh x.
Using another Basic Iden ty, we can see that cosh x sinh x =
We get the same answer either way.

sinh x.

. We employ subs tu on, with u = t and du = dt. Applying Key


Ideas and we have:

sech ( t ) dt = tanh( t ) + C.
.

ln

ln

cosh x dx = sinh x = sinh(ln ) sinh = sinh(ln ).

We can simplify this last expression as sinh x is based on exponen als:


sinh(ln ) =

eln e ln

Inverse Hyperbolic Func ons


Just as the inverse trigonometric func ons are useful in certain integra ons,
the inverse hyperbolic func ons are useful with others. Figure . shows the
restric ons on the domains to make each func on one-to-one and the resul ng
domains and ranges of their inverse func ons. Their graphs are shown in Figure
. .
Because the hyperbolic func ons are dened in terms of exponen al funcons, their inverses can be expressed in terms of logarithms as(shown
in Key)Idea
. It is o en more convenient to refer to sinh x than to ln x + x + , especially when one is working on theory and does not need to compute actual
values. On the other hand, when computa ons are needed, technology is o en
helpful but many hand-held calculators lack a convenient sinh x bu on. (Often it can be accessed under a menu system, but not conveniently.) In such a
situa on, the logarithmic representa on is useful. The reader is not encouraged
to memorize these, but rather know they exist and know how to use them when
needed.

Notes:

Hyperbolic Func ons

Chapter

Techniques of An dieren a on
Func on
cosh x
sinh x
tanh x
sech x
csch x
coth x

Domain
[ , )
(, )
(, )
[ , )
(, ) ( , )
(, ) ( , )
Figure .

Func on

Range
[ , )
(, )
( , )
( , ]
(, ) ( , )
(, ) ( , )

cosh
sinh
tanh
sech
csch
coth

x
x
x
x
x
x

Domain

Range

[ , )
(, )
( , )
( , ]
(, ) ( , )
(, ) ( , )

[ , )
(, )
(, )
[ , )
(, ) ( , )
(, ) ( , )

: Domains and ranges of the hyperbolic and inverse hyperbolic func ons.

y
y = cosh x

10

10
y = sinh x
5

5
y = cosh1 x

10

y = sinh1 x

10

10

10

2
y = coth1 x

y = sech x

y = csch x

y = tanh1 x

.
.

.
Figure .

Key Idea

: Graphs of the hyperbolic func ons and their inverses.

Logarithmic deni ons of Inverse Hyperbolic Func ons

(
)
. sinh x = ln x + x +

(
)
. cosh x = ln x + x ; x
. tanh
. sech

x=

ln

x = ln

+x
x

; |x| <

Notes:

. coth

<x

. csch

x=

ln

x = ln

x+
x

; |x| >

+x
|x|

; x =

.
The following Key Ideas give the deriva ves and integrals rela ng to the inverse hyperbolic func ons. In Key Idea , both the inverse hyperbolic and logarithmic func on representa ons of the an deriva ve are given, based on Key
Idea . Again, these la er func ons are o en more useful than the former.
Note how inverse hyperbolic func ons can be used to solve integrals we used
Trigonometric Subs tu on to solve in Sec on . .
Key Idea
.
.
.

Deriva ves Involving Inverse Hyperbolic Func ons

)
d(
cosh x =
; x>
dx
x
)
d(
sinh x =
dx
x +
)
d(
tanh x =
dx

Key Idea

.
.

.
.

; |x| <

d(
sech x =
; <x<
dx
x
x
)
d(

; x =
csch x =
dx
|x|
+x
)
d(
coth x =
dx

Integrals Involving Inverse Hyperbolic Func ons


x a
x +a

a x

dx =

cosh

dx =

sinh

dx

x a x

dx =

dx =
x x +a

(x)

a
(x)
a

tanh

a coth

sech
a

csch
a

+ C;

<a<x

+ C; a >
(x)
a

(x)
a

(x)

+C

x <a

+C

a <x

+ C; < x < a
a
x

+ C; x = , a >
a



= ln x + x a + C



= ln x + x + a + C
=



a + x
+C

ln
a x

)
x

+C
a
a+ a x




x
+C

= ln
a
a+ a +x
=

ln

We prac ce using the deriva ve and integral formulas in the following example.

Notes:

; |x| >

Hyperbolic Func ons

Chapter

Techniques of An dieren a on
Deriva ves and integrals involving inverse hyperbolic funcExample
ons
Evaluate the following.
[
(
)]
d
x

.
cosh
dx

dx
.
x

dx

x +

. Applying Key Idea with the Chain Rule gives:


[
(
)]
x
d

cosh
= (
)
dx
x

. Mul plying the numerator and denominator by ( ) gives:


dx =
x

dx. The second integral can be solved with a direct applica on


x
of item # from Key Idea , with a = . Thus

dx =
dx
x
x

x <
tanh (x) + C
=

coth (x) + C
<x


x +
+C
= ln
x


x
+ C.

= ln
( . )
x+
We should note that this exact problem was solved at the beginning of
Sec on . . In that example the answer was given as ln |x | ln |x+
| + C. Note that this is equivalent to the answer given in Equa on . , as
ln(a/b) = ln a ln b.

. This requires a subs tu on, then item # of Key Idea


Let u = x, hence du = dx. We have

dx =
x +
u +

Notes:

du.

can be applied.

.
Note a =

, hence a =

. Now apply the integral rule.

=
=

sinh

+C


x +
ln x +



+ C.

This sec on covers a lot of ground. New func ons were introduced, along
with some of their fundamental iden es, their deriva ves and an deriva ves,
their inverses, and the deriva ves and an deriva ves of these inverses. Four
Key Ideas were presented, each including quite a bit of informa on.
Do not view this sec on as containing a source of informa on to be memorized, but rather as a reference for future problem solving. Key Idea contains
perhaps the most useful informa on. Know the integra on forms it helps evaluate and understand how to use the inverse hyperbolic answer and the logarithmic answer.
The next sec on takes a brief break from demonstra ng new integra on
techniques. It instead demonstrates a technique of evalua ng limits that return indeterminate forms. This technique will be useful in Sec on . , where
limits will arise in the evalua on of certain denite integrals.

Notes:

Hyperbolic Func ons

Exercises .

Terms and Concepts


. In Key Idea

, the equa on

. f(x) = tanh (cos x)


tanh x dx = ln(cosh x)+C is

given. Why is ln | cosh x| not used i.e., why are absolute


values not necessary?
. The hyperbolic func ons are used to dene points on the
right hand por on of the hyperbola x y = , as shown
in Figure . . How can we use the hyperbolic func ons to
dene points on the le hand por on of the hyperbola?

. coth x csch x =
. cosh x = cosh x + sinh x

cosh x

d
[coth x] = csch x
dx

.
tanh x dx = ln(cosh x) + C

. f(x) = cosh x at x = ln

. f(x) = cosh x at x =
In Exercises

coth x dx = ln | sinh x| + C

In Exercises

, nd the deriva ve of the given func on.

tanh( x) dx

cosh( x ) dx

sinh x cosh x dx

x cosh x dx

x sinh x dx

. f(x) = ln(sinh x)

x +x

ex
ex+

sinh x dx

tanh x dx

. f(x) = sinh x cosh x


. f(x) = x sinh x cosh x
. f(x) = sech (x )
. f(x) = sinh ( x)
. f(x) = cosh
. f(x) = tanh

( x)
(x + )

x
x
x

dx
dx

x
dx
+x

dx
dx
dx

, evaluate the given indenite integral.

. f(x) = cosh x
. f(x) = tanh(x )

cosh x +

d
[sech x] = sech x tanh x
dx

. f(x) = sinh x at x =

. f(x) = sinh x at x =

In Exercises , verify the given iden ty using Deni on


, as done in Example
.

. sinh x =

In Exercises
, nd the equa on of the line tangent to
the func on at the given x-value.

. f(x) = sech x at x = ln

Problems

. cosh x =

. f(x) = cosh (sec x)

sech x dx

In Exercises
.

(Hint: mu ply by

sinh x dx

cosh x
;
cosh x

set u = sinh x.)

, evaluate the given denite integral.

ln
ln

cosh x dx

tanh x dx

Chapter

Techniques of An dieren a on

LHpitals Rule

While this chapter is devoted to learning techniques of integra on, this sec on
is not about integra on. Rather, it is concerned with a technique of evalua ng
certain limits that will be useful in the following sec on, where integra on is
once more discussed.
Our treatment of limits exposed us to / , an indeterminate form. If lim f(x) =
xc

and lim g(x) = , we do not conclude that lim f(x)/g(x) is / ; rather, we use
xc

xc

/ as nota on to describe the fact that both the numerator and denominator
approach . The expression / has no numeric value; other work must be done
to evaluate the limit.
Other indeterminate forms exist; they are: /, , , , and
. Just as / does not mean divide by , the expression / does
not mean divide innity by innity. Instead, it means a quan ty is growing
without bound and is being divided by another quan ty that is growing without
bound. We cannot determine from such a statement what value, if any, results
in the limit. Likewise, does not mean mul ply zero by innity. Instead,
it means one quan ty is shrinking to zero, and is being mul plied by a quan ty
that is growing without bound. We cannot determine from such a descrip on
what the result of such a limit will be.
This sec on introduces lHpitals Rule, a method of resolving limits that produce the indeterminate forms / and /. Well also show how algebraic
manipula on can be used to convert other indeterminate expressions into one
of these two forms so that our new rule can be applied.
Theorem
Let lim f(x) =
xc

LHpitals Rule, Part


and lim g(x) = , where f and g are dieren able funcxc

ons on an open interval I containing c, and g (x) =


bly at c. Then
f (x)
f(x)
= lim .
lim
xc g (x)
xc g(x)

on I except possi-

We demonstrate the use of lHpitals Rule in the following examples; we


will o en use LHR as an abbrevia on of lHpitals Rule.

Notes:

.
Example
Using lHpitals Rule
Evaluate the following limits, using lHpitals Rule as needed.
sin x
x

x+
. lim
x
x
. lim

. lim

. lim
x

x
cos x

x +x
x x+

. We proved this limit is in Example using the Squeeze Theorem. Here


we use lHpitals Rule to show its power.
lim

.
.

lim

sin x
x

x+
x

by LHR

lim

=
by LHR

cos x

lim

= .

(x + )

= .

by LHR
x
x
= lim
.
x
x sin x
cos x
This la er limit also evaluates to the / indeterminate form. To evaluate
it, we apply lHpitals Rule again.

lim

lim

Thus lim
x

x
sin x

by LHR

cos x

= .

x
= .
cos x

. We already know how to evaluate this limit; rst factor the numerator and
denominator. We then have:
lim

x +x
x x+

= lim
x

(x )(x + )
x+
= lim
(x )(x ) x x

= .

We now show how to solve this using lHpitals Rule.


lim

x +x
x x+

by LHR

lim

x+
x

= .

Note that at each step where lHpitals Rule was applied, it was needed: the
ini al limit returned the indeterminate form of / . If the ini al limit returns,
for example, / , then lHpitals Rule does not apply.

Notes:

LHpitals Rule

Chapter

Techniques of An dieren a on
The following theorem extends our ini al version of lHpitals Rule in two
ways. It allows the technique to be applied to the indeterminate form /
and to limits where x approaches .
Theorem

LHpitals Rule, Part

. Let lim f(x) = and lim g(x) = , where f and g are dierxa
xa
en able on an open interval I containing a. Then
f(x)
f (x)
= lim .
xa g(x)
xa g (x)
lim

. Let f and g be dieren able func ons on the open interval (a, )
for some value a, where g (x) = on (a, ) and lim f(x)/g(x)
x

returns either / or /. Then

f(x)
f (x)
= lim .
x g(x)
x g (x)
lim

A similar statement can be made for limits where x approaches


.
Using lHpitals Rule with limits involving
Example
Evaluate the following limits.
. lim

ex
.
x x

x
x+
x + x

. lim

. We can evaluate this limit already using Theorem ; the answer is / .


We apply lHpitals Rule to demonstrate its applicability.
lim

x
x+
x + x

by LHR

lim

x
x+

by LHR

lim

ex by LHR
ex
ex by LHR
ex by LHR
=
lim
=
lim
=
lim
= .
x x
x
x x
x x
Recall that this means that the limit does not exist; as x approaches ,
the expression ex /x grows without bound. We can infer from this that
ex grows faster than x ; as x gets large, ex is far larger than x . (This

. lim

Notes:

.
has important implica ons in compu ng when considering eciency of
algorithms.)

Indeterminate Forms and


LHpitals Rule can only be applied to ra os of func ons. When faced with
an indeterminate form such as or , we can some mes apply algebra
to rewrite the limit so that lHpitals Rule can be applied. We demonstrate the
general idea in the next example.
Applying lHpitals Rule to other indeterminate forms
Example
Evaluate the following limits.
. lim x e

/x

. lim x e

/x

. lim ln(x + ) ln x
x

. lim x ex
x

. As x

,x

and e

. Thus we have the indeterminate form


e /x
. We rewrite the expression x e /x as
; now, as x + , we get
/x
the indeterminate form / to which lHpitals Rule can be applied.
+

lim x e

/x

= lim
x

/x

by LHR

/x

/x

lim

( /x )e
/x

/x

= lim e
x

/x

= .

Interpreta on: e /x grows faster than x shrinks to zero, meaning their


product grows without bound.
. As x , x and e /x e . The the limit evaluates to
which is not an indeterminate form. We conclude then that
lim x e

/x

= .

. This limit ini ally evaluates to the indeterminate form . By applying


a logarithmic rule, we can rewrite the limit as
)
(
x+
.
lim ln(x + ) ln x = lim ln
x
x
x
As x , the argument of the ln term approaches /, to which we
can apply lHpitals Rule.
lim

Notes:

x+
x

by LHR

= .

LHpitals Rule

Chapter

Techniques of An dieren a on
Since x implies
x

x+
x

, it follows that

implies

ln

x+
x

Thus
lim ln(x + ) ln x = lim ln

ln = .

x+
x

= .

Interpreta on: since this limit evaluates to , it means that for large x,
there is essen ally no dierence between ln(x + ) and ln x; their dierence is essen ally .
. The limit lim x ex ini ally returns the indeterminate form . We
x
(
)
ex
x
can rewrite the expression by factoring out x ; x e = x

.
x
We need to evaluate how ex /x behaves as x :
ex
x x
lim

by LHR

lim

ex
x

by LHR

lim

ex

= .

Thus limx x ( ex /x ) evaluates to (), which is not an indeterminate form; rather, () evaluates to . We conclude that
lim x ex = .
x

Interpreta on: as x gets large, the dierence between x and ex grows


very large.

Indeterminate Forms

and

When faced with an indeterminate form that involves a power, it o en helps


to employ the natural logarithmic func on. The following Key Idea expresses the
concept, which is followed by an example that demonstrates its use.
Key Idea

Evalua ng Limits Involving Indeterminate Forms


, and

(
)
If lim ln f(x) = L,
xc

Notes:

then lim f(x) = lim eln(f(x)) = e L .


xc

xc

.
Using lHpitals Rule with indeterminate forms involving
Example
exponents
Evaluate the following limits.
)x
(
. lim xx .
. lim
+
x
x
x +
S

. This equivalent to a special limit given in Theorem ; these limits have


important applica ons within mathema cs and nance. Note that the
exponent approaches while the base approaches , leading to the inx
determinate form . Let f(x) = ( +
( /x)
) ; the problem asks to evaluate
lim f(x). Lets rst evaluate lim ln f(x) .
x

(
)
lim ln f(x) = lim ln

)x

= lim x ln
+
x
x
(
)
ln + x
= lim
x
/x

This produces the indeterminate form / , so we apply lHpitals Rule.


= lim

= lim

( /x )

+ /x

( /x )
+ /x

= .
(
)
Thus lim ln f(x) = . We return to the original limit and apply Key Idea
x
.
lim

)x

= lim f(x) = lim eln(f(x)) = e = e.


x

. This limit leads to the indeterminate form

Notes:

. Let f(x) = xx and consider

LHpitals Rule

Chapter

Techniques of An dieren a on
(
)
rst lim ln f(x) .

(
)
lim ln f(x) = lim ln (xx )

= lim x ln x
x

= lim
x

f(x) = xx

ln x
.
/x

This produces the indeterminate form / so we apply lHpitals Rule.

/x
/x
= lim x
= lim

Figure . : A graph of f(x) = xx supporting the fact that as x + , f(x) .

= .
(
)
Thus lim ln f(x) = . We return to the original limit and apply Key Idea
.

lim xx = lim f(x) = lim eln(f(x)) = e = .

This result is supported by the graph of f(x) = xx given in Figure .

Our brief revisit of limits will be rewarded in the next sec on where we consider improper integra on. So far, we have only
considered denite integrals
where the bounds are nite numbers, such as

f(x) dx. Improper integra on

considers integrals where one, or both, of the bounds are innity. Such integrals have many uses and applica ons, in addi on to genera ng ideas that are
enlightening.

Notes:

Exercises .

Terms and Concepts

. lim

. List the dierent indeterminate forms described in this secon.


. T/F: lHpitals Rule provides a faster method of compu ng
deriva ves.
[
]
f (x)
d f(x)
. T/F: lHpitals Rule states that
= .
dx g(x)
g (x)
. Explain what the indeterminate form

f(x)
when
g(x)
; lHpitals Rule is applied when taking
.

taking
certain

. lim

. Create (but do not evaluate!) a limit that returns .

x
ex

ex
. lim
x
x
ex

. lim

means.

. Fill in the blanks: The Quo ent Rule is applied to

x
ex

ex

. lim

. lim
x

x x + x+
x x + x

. lim

x + x + x
x + x + x+

. lim

ln x
x

. lim

ln(x )
x

x +x
x

. lim

. lim

x +x
x x+

. lim x ln x

. lim

sin x
x

. lim

. Create a func on f(x) such that lim f(x) returns .


x

Problems
In Exercises
. lim
x

lim

x/

. lim
x

sin x cos x
cos( x)

sin( x)
x

sin( x)
. lim
x x +
. lim
x

. lim
x

sin( x)
sin( x)
sin(ax)
sin(bx)

. lim

x +

. lim

x +

. lim

x +

, evaluate the given limit.

ex
x

)
ln x
x

x +

x +

x ln x

. lim xe

/x

x +

. lim x x
x

. lim

x ln x

lim xex

. lim

x +

/x

. lim ( + x)

/x

x +

. lim ( x)x
x +

e x
x

. lim ( /x)x

x sin x
x x

. lim (sin x)x

x +

x +

Hint: use the Squeeze Theorem.

. lim ( x)

x +

. lim (x)

/x

x +

. lim (ln x)

. lim

x +

. lim ( + x)

. lim ( + x )

/x

lim tan x cos x

x/

. lim x tan( /x)


/x

ln x

x +

lim tan x sin( x)

x/

. lim

. lim ( /x)

. lim

(ln x)
x

. lim

x +x
ln x

x
x

Improper Integra on

Improper Integra on

We begin this sec on by considering the following denite integrals:

dx .
,
+x

dx .
,

+x
,
dx .
.

+x
No ce how the integrand is /( + x ) in each integral (which is sketched in
Figure . ). As the upper bound gets larger, one would expect the area under
the curve would also grow. While the denite integrals do increase in value as
the upper bound grows, they are not increasing by much. In fact, consider:

+x

b

dx = tan x = tan b tan

= tan b.

As b , tan b / . Therefore it seems that as the upper bound b grows,


b
dx approaches / .
. This
the value of the denite integral
+x
should strike the reader as being a bit amazing: even though the curve extends
to innity, it has a nite amount of area underneath it.
b
f(x) dx, we made two s pula ons:
When we dened the denite integral
a

. The interval over which we integrated, [a, b], was a nite interval, and

. The func on f(x) was con nuous on [a, b] (ensuring that the range of f
was nite).
In this sec on we consider integrals where one or both of the above condions do not hold. Such integrals are called improper integrals.

Notes:

y
1

0.5

.
Figure .

: Graphing f(x) =

10

+x

Chapter

Techniques of An dieren a on

Improper Integrals with Innite Bounds


Deni on

Improper Integrals with Innite Bounds; Converge,


Diverge

. Let f be a con nuous func on on [a, ). Dene



b
f(x) dx to be lim
f(x) dx.
b

. Let f be a con nuous func on on (, b]. Dene


b
b
f(x) dx to be
lim
f(x) dx.
a

. Let f be a con nuous func on on (, ). Let c be any real number; dene


b

c
f(x) dx.
f(x) dx to be
lim
f(x) dx + lim
a

An improper integral is said to converge if its corresponding limit exists;


otherwise, it diverges. The improper integral in part converges if and
only if both of its limits exist.

Evalua ng improper integrals


Example
Evaluate the following improper integrals.

f(x) =

.5

.
x

Figure . : A graph of f(x) =


ample
.

in Ex-

dx

dx

ex dx

+x

dx

dx = lim

b x

= lim
b b
= .

dx = lim

A graph of the area dened by this integral is given in Figure .


Notes:

dx = lim

dx
x
b

= lim ln |x|
b

y
1

= .

0.5

dx diverges.
x
Compare the graphs in Figures . and . ; no ce how the graph of
f(x) = /x is no ceably larger. This dierence is enough to cause the
improper integral to diverge.

x
ex dx
e dx = lim
.
The limit does not exist, hence the improper integral

Figure .
ple
.

10

: A graph of f(x) =

in Exam-



= lim ex
a

1
x

f(x) =

= lim ln(b)

Improper Integra on

= lim e ea
a

= .
A graph of the area dened by this integral is given in Figure .

f(x) = e

. We will need to break this into two improper integrals and choose a value
of c as in part of Deni on . Any value of c is ne; we choose c = .

+x

dx = lim

dx + lim
dx
b
+x
+x
b



= lim tan x + lim tan x
a
b
a
(
)
(
= lim tan tan a + lim tan b tan
a
b
) (
(
)

+
.
=

. 10

Figure .
ple
.

: A graph of f(x) = ex in Exam-

)
y

Each limit exists, hence the original integral converges and has value:
f(x) =

= .
A graph of the area dened by this integral is given in Figure .

.
.

Notes:

+x

Figure . : A graph of f(x) =


ample
.

+x

in Ex-

Chapter

Techniques of An dieren a on
The previous sec on introduced lHpitals Rule, a method of evalua ng limits that return indeterminate forms. It is not uncommon for the limits resul ng
from improper integrals to need this rule as demonstrated next.

.4
f(x) =

ln x
x

Improper integra
on and lHpitals Rule
ln x
Evaluate the improper integral
dx.
x
Example

Figure . : A graph of f(x) =


ample
.

ln x
x

This integral will require the use of Integra on by Parts. Let


S
u = ln x and dv = /x dx. Then

in Ex-

ln x
dx = lim
b
x

ln x
dx
x
(
)
b
ln x b
= lim
dx
+
b
x
x
) b
(

ln x

= lim
b
x
x
)
(
ln b
( ln ) .
= lim
b
b
b

The /b and ln terms go to , leaving lim


lim

ln b
with lHpitals Rule. We have:
b
ln b
b b
lim

by LHR

lim

ln b
+ . We need to evaluate
b

/b

= .
Thus the improper integral evaluates as:

ln x
dx = .
x

Improper Integrals with Innite Range


We have just considered denite integrals where the interval of integra on
was innite. We now consider another type of improper integra on, where the
range of the integrand is innite.

Notes:

Deni on

Improper Integra on

Improper Integra on with Innite Range

Let f(x) be a con nuous func on on [a, b] except at c, a c b, where


x = c is a ver cal asymptote of f. Dene

b
a

f(x) dx = lim

tc

t
a

f(x) dx + lim

tc+

b
t

f(x) dx.

Improper integra on of func ons with innite range


Example
Evaluate the following improper integrals:

dx.
dx
.
.
x
x

Note: In Deni on , c can be one of the


endpoints (a or b). In that case, there is
only one limit to consider as part of the
deni on.

y
10

. A graph of f(x) = / x is given in Figure . . No ce that f has a ver cal


asymptote at x = ; in some sense, we are trying to compute the area of
a region that has no top. Could this have a nite value?

dx = lim
a +
x
= lim
a

= lim
a

= .

dx
x

x
a
(
)
a
a

1
f(x) =
x

Figure . : A graph of f(x) =


ample
.

Notes:

f(x) =

= ( )
= !

in Ex-



dx =
x
x

It turns out that the region does have a nite area even though it has no
upper bound (strange things can occur in mathema cs when considering
the innite).
. The func on f(x) = /x has a ver cal asymptote at x = , as shown
in Figure . , so this integral is an improper integral. Lets eschew using
limits for a moment and proceed without recognizing the improper nature
of the integral. This leads to:

0.5

x
.5

.5

Figure . : A graph of f(x) =


ample
.

in Ex-

Chapter

Techniques of An dieren a on
Clearly the area in ques on is above the x-axis, yet the area is supposedly
nega ve! Why does our answer not match our intui on? To answer this,
evaluate the integral using Deni on .
t

dx = lim
dx + lim
dx
t x
t + t x
x
t



= lim + lim
x
x t
t
t +
= lim + lim +
t
t
t +
t
(
)
(
)

+ + .

Neither limit converges hence the original improper integral diverges. The
nonsensical answer we obtained by ignoring the improper nature of the
integral is just that: nonsensical.

Understanding Convergence and Divergence


O en mes we are interested in knowing simply whether or not an improper
integral converges, and not necessarily the value of a convergent integral. We
provide here several tools that help determine the convergence or divergence
of improper integrals without integra ng.
Our rst tool is to understand the behavior of func ons of the form

Determine the values of p for which


f(x) =

1
xq

p<1<q

Improper integra on of /xp

Example
y

xp

xp

dx converges.

We begin by integra ng and then evalua ng the limit.


b

dx = lim
dx
b
xp
xp
b
= lim
xp dx
(assume p = )
b

1
f(x) = p
x
1

= lim

= lim

Figure . : Plo ng func ons of the form


/x p in Example
.

p +
p

b

xp+

(
b

)
.

When does this limit converge i.e., when is this limit not ? This limit converges precisely when the power of b is less than : when p < < p.

Notes:

.
Our analysis shows that if p > , then

Improper Integra on

dx converges. When p <


xp
the improper integral diverges; we showed in Example
that when p = the
integral also diverges.
Figure . graphs y = /x with a dashed line, along with graphs of y = /xp ,
p < , and y = /xq , q > . Somehow the dashed line forms a dividing line
between convergence and divergence.
The result of Example
provides an important tool in determining the
convergence of other integrals. A similar result is proved in the exercises about
improper integrals of the form
following Key Idea.

Key Idea

dx. These results are summarized in the

xp

Convergence of Improper Integrals

. The improper integral

. The improper integral

xp
xp

xp

dx converges when p >

dx converges when p <

dx and

xp

and diverges when p .


and diverges when p .

A basic technique in determining convergence of improper integrals is to


compare an integrand whose convergence is unknown to an integrand whose
convergence is known. We o en use integrands of the form /x p to compare
to as their convergence on certain intervals is known. This is described in the
following theorem.
Theorem

Direct Comparison Test for Improper Integrals

Let f and g be con nuous on [a, ) where f(x) g(x) for all x in
[a, ).


f(x) dx converges.
g(x) dx converges, then
. If
a

. If

Notes:

f(x) dx diverges, then

g(x) dx diverges.

dx.

Note: We used the upper and lower


bound of in Key Idea
for convenience. It can be replaced by any a where
a> .

Chapter

Techniques of An dieren a on
Example
Determining convergence of improper integrals
Determine
the
convergence
of the
following improper integrals.

f(x) =

ex dx

x x

dx

f(x) = e

Figure . : Graphs of f(x) = ex and


f(x) = /x in Example
.

. The func on f(x) = ex does not have an an deriva ve expressible in


terms of elementary func ons, so we cannot integrate directly. It is comparable to g(x) = /x , and as demonstrated in Figure . , ex < /x
on [ , ). We know from Key Idea

ex dx also converges.

. Note that for large values of x,


Idea

f(x) =

and the subsequent note that

dx diverges, so we seek to

It is easy to see that when x > , we have x =


reciprocals reverses the inequality, giving

x x

dx converges, hence

= . We know from Key


x
x x
x

compare the original integrand to /x.


y

that

x >

x x. Taking

.4

f(x) =

x
x

Using Theorem

<

x x

, we conclude that since

diverges as well. Figure .

illustrates this.

dx diverges,

x x

Figure . : Graphs of f(x) = / x x


and f(x) = /x in Example
.

Being able to compare unknown integrals to known integrals is very useful in determining convergence. However, some of our examples were a li le
, but what if the
<
x
x x
x were replaced
with a + x + ? That is, what can we say about the con

too nice. For instance, it was convenient that


vergence of

x + x+

dx? We have

>

x + x+

, so we cannot

use Theorem .
In cases like this (and many more) it is useful to employ the following theorem.

Notes:

dx

Theorem

Improper Integra on

Limit Comparison Test for Improper Integrals

Let f and g be con nuous func ons on [a, ) where f(x) > and g(x) >
for all x. If
f(x)
lim
= L,
< L < ,
x g(x)
then

f(x) dx

and

either both converge or both diverge.

g(x) dx

Determining
of improper integrals
convergence

dx.
Determine the convergence of
x + x+
Example

As x gets large, the quadra c inside the square root func on


to with
will begin to behave much like y = x. So we compare
S

x + x+

the Limit Comparison Test:


lim

/ x + x+
/x

= lim
x

x
.
x + x+

The immediate evalua on of this limit returns /, an indeterminate form.


Using lHpitals Rule seems appropriate, but in this situa on, it does not lead
to useful results. (We encourage the reader to employ lHpitals Rule at least
once to verify this.)
The trouble is the square root func on. To get rid of it, we employ the following fact: If lim f(x) = L, then lim f(x) = L . (This is true when either c or L
xc

lim

x + x+

x
.
x + x+

f(x) =

dx diverges, by the Limit Comparison Test we know that

dx

x + x+

also diverges. Figure . graphs f(x) = / x + x + and f(x) = /x, illustra ng that as x gets large, the func ons become indis nguishable.

Notes:

x + x+5

This converges to , meaning the original limit also converged to . As x gets


looks very much like . Since we know that
very large, the func on

f(x) =

xc

is .) So we consider now the limit

.
Figure .
and f(x) =

: Graphing f(x) =
x

in Example

x + x+

Chapter

Techniques of An dieren a on
Both the Direct and Limit Comparison Tests were given in terms of integrals
over an innite interval. There are versions that apply to improper integrals with
an innite range, but as they are a bit wordy and a li le more dicult to employ,
they are omi ed from this text.
This chapter has explored many integra on techniques. We learned Subs tu on, which undoes the Chain Rule of dieren a on, as well as Integra on
by Parts, which undoes the Product Rule. We learned specialized techniques
for handling trigonometric func ons and introduced the hyperbolic func ons,
which are closely related to the trigonometric func ons. All techniques eecvely have this goal in common: rewrite the integrand in a new way so that the
integra on step is easier to see and implement.
As stated before, integra on is, in general, hard. It is easy to write a func on
whose an deriva ve is impossible to write in terms of elementary func ons,
and even when a func on does have an an deriva ve expressible by elementary
func ons, it may be really hard to discover what it is. The powerful computer
algebra system Mathema ca has approximately ,
pages of code dedicated
to integra on.
Do not let this diculty discourage you. There is great value in learning integra on techniques, as they allow one to manipulate an integral in ways that
can illuminate a concept for greater understanding. There is also great value
in understanding the need for good numerical techniques: the Trapezoidal and
Simpsons Rules are just the beginning of powerful techniques for approximating the value of integra on.
The next chapter stresses the uses of integra on. We generally do not nd
an deriva ves for an deriva ves sake, but rather because they provide the solu on to some type of problem. The following chapter introduces us to a number
of dierent problems whose solu on is provided by integra on.

Notes:

Exercises .

Terms and Concepts


. The denite integral was dened with what two s pulaons?
. If lim

said to

f(x) dx exists, then the integral


.

f(x) dx = , and

know that
g(x) dx

. If

. For what values of p will

. For what values of p will

xp
xp

xp

dx converge?
dx converge?

dx

x +

x
x +
x
x +

(x )
(x )

dx

dx

dx

sec x dx

dx

dx

|x|

dx

xex dx

xex dx

xex dx

ex + ex

ln x
dx
x

ln x
dx
x

dx

dx

dx

dx

dx

dx

( )x

x dx

, evaluate the given improper integral.

dx

dx converge?

Problems

f(x) dx is

g(x) f(x) for all x, then we

. For what values of p will

In Exercises

dx

x ln x dx

ln x
dx
x

ln x dx

ex sin x dx

ex cos x dx

In Exercises
, use the Direct Comparison Test or the
Limit Comparison Test to determine whether the given definite integral converges or diverges. Clearly state what test
is being used and what func on the integrand is being compared to.
.

x + x
x x

dx

x+
x x +x+

ex ln x dx

dx

dx

ex

+ x+

dx

x
dx
ex

x + sin x

dx

x
dx
x + cos x
x + ex

ex x

dx
dx

:A

We begin this chapter with a reminder of a few key concepts from Chapter .
Let f be a con nuous func on on [a, b] which is par oned into n equally spaced
subintervals as
a < x < x < < xn < xn+ = b.

Let x = (b a)/n denote the length of the subintervals, and let ci be any
x-value in the i th subinterval. Deni on states that the sum
n

f(ci )x

i=

is a Riemann Sum. Riemann Sums are o en used to approximate some quanty (area, volume, work, pressure, etc.). The approxima on becomes exact by
taking the limit
n

f(ci )x.
lim
n

Theorem

i=

connects limits of Riemann Sums to denite integrals:


lim

f(ci )x =

i=

b
a

f(x) dx.

Finally, the Fundamental Theorem of Calculus states how denite integrals can
be evaluated using an deriva ves.
This chapter employs the following technique to a variety of applica ons.
Suppose the value Q of a quan ty is to be calculated. We rst approximate the
value of Q using a Riemann Sum, then nd the exact value via a denite integral.
We spell out this technique in the following Key Idea.
Key Idea

Applica on of Denite Integrals Strategy

Let a quan ty be given whose value Q is to be computed.


. Divide the quan ty into n smaller subquan

es of value Qi .

. Iden fy a variable x and func on f(x) such that each subquan ty


can be approximated with the product f(ci )x, where x represents a small change in x. Thus Qi f(ci )x. A sample approxima on f(ci )x of Qi is called a dieren al element.
. Recognize that Q =
Sum.

i=

Qi

f(ci )x, which is a Riemann

i=

. Taking the appropriate limit gives Q =

b
a

f(x) dx

This Key Idea will make more sense a er we have had a chance to use it
several mes. We begin with Area Between Curves, which we addressed briey
in Sec on . . .

Chapter

Applica ons of Integra on

.
y

f(x)

g(x)

(a)
y

f(x)

g(x)

Area Between Curves

We are o en interested in knowing the area of a region. Forget momentarily


that we addressed this already in Sec on . . and approach it instead using
the technique described in Key Idea .
Let Q be the area of a region bounded by con nuous func ons f and g. If we
break the region into many subregions, we have an obvious equa on:
Total Area = sum of the areas of the subregions.
The issue to address next is how to systema cally break a region into subregions. A graph will help. Consider Figure . (a) where a region between two
curves is shaded. While there are many ways to break this into subregions, one
par cularly ecient way is to slice it ver cally, as shown in Figure . (b), into
n equally spaced slices.
We now approximate the area of a slice. Again, we have many op ons, but
using a rectangle seems simplest. Picking any x-value ci in the i th slice, we set
the height of the rectangle to be f(ci ) g(ci ), the dierence of the corresponding y-values. The width of the rectangle is a small dierence in x-values, which
we represent with x. Figure . (c) shows sample points ci chosen in each
subinterval and appropriate rectangles drawn. (Each of these rectangles represents
a dieren
al element.) Each slice has an area approximately equal to
(
)
f(ci ) g(ci ) x; hence, the total area is approximately the Riemann Sum
Q=

(b)

(
i=

)
f(ci ) g(ci ) x.

Taking the limit as n gives the exact area as


f(x)

Theorem

)
f(x) g(x) dx.

Area Between Curves (restatement of Theorem

Let f(x) and g(x) be con nuous func ons dened on [a, b] where f(x)
g(x) for all x in [a, b]. The area of the region bounded by the curves
y = f(x), y = g(x) and the lines x = a and x = b is

g(x)

b(

(c)
Figure . : Subdividing a region into vercal slices and approxima ng the areas
with rectangles.

b
a

)
f(x) g(x) dx.

Finding area enclosed by curves


Example
Find the area of the region bounded by f(x) = sin x + , g(x) =
x = and x = , as shown in Figure . .
S

Notes:

cos( x) ,

The graph veries that the upper boundary of the region is

.
given by f and the lower bound is given by g. Therefore the area of the region is
the value of the integral
(

))
(
(
)
cos( x)
dx
sin x +
f(x) g(x) dx =
= cos x
=



sin( x) + x

f(x)

. units .

Area Between Curves

Finding total area enclosed by curves


Example
Find the total area of the region enclosed by the func ons f(x) = x + and
g(x) = x x + x as shown in Figure . .

4
g(x)

Figure . : Graphing an enclosed region


in Example
.

A quick calcula on shows


that f = g at x = , and . One
(
)
f(x) g(x) dx, but this ignores
can proceed thoughtlessly by compu ng
S

the fact that on [ , ], g(x) > f(x). (In fact, the thoughtless integra on returns
/ , hardly the expected value of an area.) Thus we compute the total area by
breaking the interval [ , ] into two subintervals, [ , ] and [ , ] and using the
proper integrand in each.

(
)
(
)
f(x) g(x) dx
g(x) f(x) dx +
Total Area =
=

= /
=

x x +

+ /
= .

dx +

x + x

x+

dx

units .

.
.

Figure . : Graphing a region enclosed by


two func ons in Example
.

The previous example makes note that we are expec ng area to be posi ve.
When rst learning about the denite integral, we interpreted it as signed area
under the curve, allowing for nega ve area. That doesnt apply here; area is
to be posi ve.
The previous example also demonstrates that we o en have to break a given
region into subregions before applying Theorem . The following example
shows another situa on where this is applicable, along with an alternate view
of applying the Theorem.

y
y=

x+

y = (x ) +

Finding area: integra ng with respect to


Example
y
Find the area of the region enclosed by the func ons y = x + , y = (x
) + and y = , as shown in Figure . .
.

Notes:

Figure . : Graphing a region for Example


.

Chapter

Applica ons of Integra on

y
x = (y )

x=

We give two approaches to this problem. In the rst apS


proach, we no ce that the regionstop is dened by two dierent curves.
On [ , ], the top func on is y = x + ; on [ , ], the top func on is y =
(x ) + . Thus we compute the area as the sum of two integrals:
(
(
)
)
(
)
)
(
(x ) +
dx
x+
dx +
Total Area =

y+

= / + /

= / .

Figure . : The region used in Example


with boundaries relabeled as funcons of y.

The second approach is clever and very useful in certain situa ons. We are
used to viewing curves as func ons of x; we input an x-value and a y-value is returned. Some curves can also be described as func ons of y: input a y-value and
an x-value is returned. We can rewrite the equa ons describing the boundary
by solving for x:

x = (y )
y= x+

y = (x ) +
x=
y+ .
Figure . shows the region with the boundaries relabeled. A dieren al
element, a horizontal rectangle, is also pictured. The width of the rectangle is
a small change in y: y. The height of the rectangle is a dierence in x-values.
The top x-value is the largest value, i.e., the rightmost. The bo om x-value
is the smaller, i.e., the le most. Therefore the height of the rectangle is
)
(
y + (y ) .

The area is found by integra ng the above func on with respect to y with
the appropriate bounds. We determine these by considering the y-values the
region occupies. It is bounded below by y = , and bounded above by y = .
That is, both the top and bo om func ons exist on the y interval [ , ]. Thus

(
)
Total Area =
y + (y ) dy
=

( y)

= / .

+ y (y )

This calculusbased technique of nding area can be useful even with shapes
that we normally think of as easy. Example
computes the area of a triangle. While the formula base height is well known, in arbitrary triangles
it can be nontrivial to compute the height. Calculus makes the problem simple.

Notes:

.
Example
Finding the area of a triangle
Compute the area of the regions bounded by the lines
y = x + , y = x + and y = x + , as shown in Figure . .
Recognize that there are two top func ons to this region,
causing us to use two denite integrals.

(
(
)
)
(x + ) ( x + ) dx +
Total Area =
( x + ) ( x + ) dx

Area Between Curves

y=x+

y= x+7

= / + /
= / .

We can also approach this by conver ng each func on into a func on of y. This
also requires integrals, so there isnt really any advantage to doing so. We do
it here for demonstra on purposes.
The top func on is always x = y while there are two bo om funcons. Being mindful of the proper integra on bounds, we have

( y
)
)
( y
( y) dy +
(y ) dy
Total Area =

y= x+

Figure . : Graphing a triangular region in


Example
.

= / + /
= / .

Of course, the nal answer is the same. (It is interes ng to note that the area of
all subregions used is / . This is coincidental.)

The measurements of length can be viewed as measuring


S
top
minus
bo
om
of two func ons. The exact answer is found by integra ng

(
)
f(x) g(x) dx, but of course we dont know the func ons f and g. Our
discrete measurements instead allow us to approximate.

Notes:

(a)

Numerically approxima ng area


Example
To approximate the area of a lake, shown in Figure . (a), the length of the
lake is measured at
-foot increments as shown in Figure . (b), where the
lengths are given in hundreds of feet. Approximate the area of the lake.

While we have focused on producing exact answers, we are also able to make
approxima ons using the principle of Theorem . The integrand in the theorem is a distance (top minus bo om); integra ng this distance func on gives
an area. By taking discrete measurements of distance, we can approximate an
area using numerical integra on techniques developed in Sec on . . The following example demonstrates this.

(b)
Figure . : (a) A sketch of a lake, and (b)
the lake with length measurements.

Chapter

Applica ons of Integra on


We have the following data points:
( , ), ( , .

), ( , .

We also have that x =


Area
=

(
.

+ .
units .

ba
n

), ( , .

), ( , .

), (

, .

), (

, ).

= , so Simpsons Rule gives

+ .

+ .

+ .

+ .

Since the measurements are in hundreds of feet, units = (


) =
,
, giving a total area of
,
. (Since we are approxima ng, wed
likely say the area was about
,
, which is a li le more than acres.)
In the next sec on we apply our applica onsofintegra on techniques to
nding the volumes of certain solids.

Notes:

Exercises .

Terms and Concepts

y
y = sin x + 1

. T/F: The area between curves is always posi ve.


.

1
y = sin x

. T/F: Calculus can be used to nd the area of basic geometric


shapes.

. In your own words, describe how to nd the total area enclosed by y = f(x) and y = g(x).
.

y
y = sec2 x

Problems
In Exercises
given graph.

, nd the area of the shaded region in the

/2

1
y = sin(4x)

y=

.
.

y
x+

/8

/4

y = sin x

.
.
y=

cos x +

.
/

x
/

y = cos x

.
y

y= x + x+

y=

.
.
x

y=

y=x +x

x
.

y
y=2

In Exercises
ons f and g.

. f(x) = x + x , g(x) = x + x

y=1

/2

, nd the total area enclosed by the func-

. f(x) = x x + , g(x) = x +
x

. f(x) = sin x, g(x) = x/


. f(x) = x x + x , g(x) = x + x

. f(x) = x, g(x) =

y=x

. f(x) = x + x + x + , g(x) = x + x +
. The func ons f(x) = cos( x) and g(x) = sin x intersect
innitely many mes, forming an innite number of repeated, enclosed regions. Find the areas of these regions.
In Exercises
, nd the area of the enclosed region in
two ways:
. by trea ng the boundaries as func ons of x, and
. by trea ng the boundaries as func ons of y.
y
y=

(x ) +

.5

y=

.
In Exercises
three points.
. ( , ),

y=x +

( , )

( , ), and

( , )

. ( , ),

( , ), and

( , )

. ( , ),

( , ), and

( , )

. Use the Trapezoidal Rule to approximate the area of the


pictured lake whose lengths, in hundreds of feet, are measured in
-foot increments.

y
y=

, nd the area triangle formed by the given

y=

x /

.5

( , ), and

. ( , ),

.5

y= x+

.9

7.

y= x

.5

y
4

y=x+2
y = x2

. Use Simpsons Rule to approximate the area of the pictured


lake whose lengths, in hundreds of feet, are measured in
-foot increments.

.2

1
1
x=
2

1 2
2y

.9

.
x

x = 12 y + 1

Volume by Cross-Sec onal Area; Disk and Washer Methods

Volume by Cross-Sec onal Area; Disk and Washer


Methods

The volume of a general right cylinder, as shown in Figure . , is


Area of the base height.
We can use this fact as the building block in nding volumes of a variety of
shapes.
Given an arbitrary solid, we can approximate its volume by cu ng it into n
thin slices. When the slices are thin, each slice can be approximated well by a
general right cylinder. Thus the volume of each slice is approximately its crosssec onal area thickness. (These slices are the dieren al elements.)
By orien ng a solid along the x-axis, we can let A(xi ) represent the crosssec onal area of the i th slice, and let xi represent the thickness of this slice (the
thickness is a small change in x). The total volume of the solid is approximately:
Volume
=

n [
]

Area thickness

Figure . : The volume of a general right


cylinder

i=

A(xi )xi .

i=

Recognize that this is a Riemann Sum. By taking a limit (as the thickness of
the slices goes to ) we can nd the volume exactly.
Theorem

Volume By Cross-Sec onal Area

The volume V of a solid, oriented along the x-axis with cross-sec onal
area A(x) from x = a to x = b, is
V=

b
a

A(x) dx.

Finding the volume of a solid


Example
Find the volume of a pyramid with a square base of side length
of in.

in and a height

There are many ways to orient the pyramid along the xS


axis; Figure . gives one such way, with the pointed top of the pyramid at the
origin and the x-axis going through the center of the base.
Each cross sec on of the pyramid is a square; this is a sample dieren al
element. To determine its area A(x), we need to determine the side lengths of

Notes:

Figure . : Orien ng a pyramid along the


x-axis in Example
.

Chapter

Applica ons of Integra on


the square.
When x = , the square has side length ; when x = , the square has side
length . Since the edges of the pyramid are lines, it is easy to gure that each
cross-sec onal square has side length x, giving A(x) = ( x) = x .
If one were to cut a slice out of the pyramid at x = , as shown in Figure
. , one would have a shape with square bo om and top with sloped sides. If
the slice were thin, both the bo om and top squares would have sides lengths
of about , and thus the crosssec onal area of the bo om and top would be
about in . Le ng xi represent the thickness of the slice, the volume of this
slice would then be about xi in .
Cu ng the pyramid into n slices divides the total volume into n equally
spaced smaller pieces, each with volume ( xi ) x, where xi is the approximate
loca on of the slice along the x-axis and x represents the thickness of each
slice. One can approximate total volume of the pyramid by summing up the
volumes of these slices:

Figure . : Cu ng a slice in they pyramid in Example


at x = .

Approximate volume =

( xi ) x.

i=

Taking the limit as n gives the actual volume of the pyramid; recoginizing
this sum as a Riemann Sum allows us to nd the exact answer using a denite
integral, matching the denite integral given by Theorem .
We have
V = lim

=
=
=

( xi ) x

i=

x dx


x

in

in .

We can check our work by consul ng the general equa on for the volume of a
pyramid (see the back cover under Volume of A General Cone):
area of base height.
Certainly, using this formula from geometry is faster than our new method, but
the calculusbased method can be applied to much more than just cones.
An important special case of Theorem is when the solid is a solid of revolu on, that is, when the solid is formed by rota ng a shape around an axis.
Start with a func on y = f(x) from x = a to x = b. Revolving this curve
about a horizontal axis creates a three-dimensional solid whose cross sec ons

Notes:

Volume by Cross-Sec onal Area; Disk and Washer Methods

are disks (thin circles). Let R(x) represent the radius of the cross-sec onal disk
at x; the area of this disk is R(x) . Applying Theorem gives the Disk Method.
Key Idea

The Disk Method

Let a solid be formed by revolving the curve y = f(x) from x = a to x = b


around a horizontal axis, and let R(x) be the radius of the cross-sec onal
disk at x. The volume of the solid is
b
R(x) dx.
V=
a

Finding volume using the Disk Method


Example
Find the volume of the solid formed by revolving the curve y = /x, from x =
to x = , around the x-axis.
A sketch can help us understand this problem. In Figure
S
. (a) the curve y = /x is sketched along with the dieren al element a
disk at x with radius R(x) = /x. In Figure . (b) the whole solid is pictured,
along with the dieren al element.
The volume of the dieren al element shown in part (a) of the gure is approximately R(xi ) x, where R(xi ) is the radius of the disk shown and x is
the thickness of that slice. The radius R(xi ) is the distance from the x-axis to the
curve, hence R(xi ) = /xi .
Slicing the solid into n equallyspaced slices, we can approximate the total
volume by adding up the approximate volume of each slice:
( )
n

x.

Approximate volume =
xi
i=

(a)

Taking the limit of the above sum as n gives the actual volume; recognizing this sum as a Riemann sum allows us to evaluate the limit with a denite
integral, which matches the formula given in Key Idea :
( )
n

V = lim
n
xi
i=
( )
=
dx
x

dx
=
x

Notes:

(b)
Figure .
.

: Sketching a solid in Example

Chapter

Applica ons of Integra on


[
]

=
x
[
]
= ( )
=

units .

While Key Idea


is given in terms of func ons of x, the principle involved
can be applied to func ons of y when the axis of rota on is ver cal, not horizontal. We demonstrate this in the next example.
(a)

(b)
Figure .
.

: Sketching a solid in Example

(a)

Finding volume using the Disk Method


Example
Find the volume of the solid formed by revolving the curve y = /x, from x =
to x = , about the y-axis.
Since the axis of rota on is ver cal, we need to convert the
S
func on into a func on of y and convert the x-bounds to y-bounds. Since y =
/x denes the curve, we rewrite it as x = /y. The bound x = corresponds to
the y-bound y = , and the bound x = corresponds to the y-bound y = / .
Thus we are rota ng the curve x = /y, from y = / to y = about the
y-axis to form a solid. The curve and sample dieren al element are sketched
in Figure . (a), with a full sketch of the solid in Figure . (b). We integrate
to nd the volume:

dy
V=
/ y

=
y /
= units .
We can also compute the volume of solids of revolu on that have a hole in
the center. The general principle is simple: compute the volume of the solid
irrespec ve of the hole, then subtract the volume of the hole. If the outside
radius of the solid is R(x) and the inside radius (dening the hole) is r(x), then
the volume is
b
b
b
(
)
R(x) r(x) dx.
r(x) dx =
R(x) dx
V=
a

One can generate a solid of revolu on with a hole in the middle by revolving
a region about an axis. Consider Figure . (a), where a region is sketched along

Notes:

(b)
Figure . : Establishing the Washer
Method; see also Figure . .

Volume by Cross-Sec onal Area; Disk and Washer Methods

with a dashed, horizontal axis of rota on. By rota ng the region about the axis,
a solid is formed as sketched in Figure . (b). The outside of the solid has radius
R(x), whereas the inside has radius r(x). Each cross sec on of this solid will be a
washer (a disk with a hole in the center) as sketched in Figure . (c). This leads
us to the Washer Method.
Key Idea

The Washer Method

Let a region bounded by y = f(x), y = g(x), x = a and x = b be rotated about a horizontal axis that does not intersect the region, forming
a solid. Each cross sec on at x will be a washer with outside radius R(x)
and inside radius r(x). The volume of the solid is
V=

b
a

R(x) r(x)

dx.

Figure . : Establishing the Washer


Method; see also Figure . .

Even though we introduced it rst, the Disk Method is just a special case of
the Washer Method with an inside radius of r(x) = .
Finding volume with the Washer Method
Example
Find the volume of the solid formed by rota ng the region bounded by y =
x x + and y = x about the x-axis.
A sketch of the region will help, as given in Figure . (a).
S
Rota ng about the x-axis will produce cross sec ons in the shape of washers,
as shown in Figure . (b); the complete solid is shown in part (c). The outside
radius of this washer is R(x) = x + ; the inside radius is r(x) = x x + . As
the region is bounded from x = to x = , we integrate as follows to compute
the volume.
(
)
( x ) (x x + ) dx
V=
=

x + x x + x

dx

[
]

= x +x x + x x

(a)

units .

(b)

When rota ng about a ver cal axis, the outside and inside radius func ons
must be func ons of y.

Notes:

(c)
Figure . : Sketching the dieren al element and solid in Example
.

Chapter

Applica ons of Integra on

Finding volume with the Washer Method


Example
Find the volume of the solid formed by rota ng the triangular region with verces at ( , ), ( , ) and ( , ) about the y-axis.

(a)

The triangular region is sketched in Figure . (a); the difS


feren al element is sketched in (b) and the full solid is drawn in (c). They help us
establish the outside and inside radii. Since the axis of rota on is ver cal, each
radius is a func on of y.
The outside radius R(y) is formed by the line connec ng ( , ) and ( , ); it
is a constant func on, as regardless of the y-value the distance from the line to
the axis of rota on is . Thus R(y) = .
The inside radius is formed by the line connec ng ( , ) and ( , ). The equaon of this line is y = x , but we need to refer to it as a func on of y. Solving
for x gives r(y) = (y + ).
We integrate over the y-bounds of y = to y = . Thus the volume is
(
) )
(
dy
(y + )
V=
=

[
=

(b)
Figure .
ple
.

(c)
: Sketching the solid in Exam-

y y+
y y +
.

units .

dy

]

y

This sec on introduced a new applica on of the denite integral. Our default view of the denite integral is that it gives the area under the curve. However, we can establish denite integrals that represent other quan es; in this
sec on, we computed volume.
The ul mate goal of this sec on is not to compute volumes of solids. That
can be useful, but what is more useful is the understanding of this basic principle
of integral calculus, outlined in Key Idea : to nd the exact value of some
quan ty,
we start with an approxima on (in this sec on, slice the solid and approximate the volume of each slice),
then make the approxima on be er by rening our original approximaon (i.e., use more slices),
then use limits to establish a denite integral which gives the exact value.
We prac ce this principle in the next sec on where we nd volumes by slicing solids in a dierent way.

Notes:

Exercises .

Terms and Concepts

y
1

. T/F: A solid of revolu on is formed by revolving a shape


around an axis.
. In your own words, explain how the Disk and Washer Methods are related.
. Explain the how the units of volume are found in the integral
of Theorem : if A(x) has units of in , how does

A(x) dx have units of in ?

Problems
In Exercises , a region of the Cartesian plane is shaded.
Use the Disk/Washer Method to nd the volume of the solid
of revolu on formed by revolving the region about the xaxis.

y=

x
y=x

0.5

0.5

In Exercises , a region of the Cartesian plane is shaded.


Use the Disk/Washer Method to nd the volume of the solid
of revolu on formed by revolving the region about the yaxis.
y
y=

y
y=

y = 5x

y = 5x

x
.5

.5

.5

.5

y
1

.
.

y = cos x

0.5

.
.

y = cos x

0.5

0.5

1.5

0.5

1.5

(Hint: Integra on By Parts will be necessary, twice. First let


u = arccos x, then let u = arccos x.)

In Exercises , a solid is described. Orient the solid along


the x-axis such that a cross-sec onal area func on A(x) can
be obtained, then apply Theorem
to nd the volume of
the solid.

1
y=

x
y=x

0.5

. A right circular cone with height of

0.5

In Exercises
, a region of the Cartesian plane is described. Use the Disk/Washer Method to nd the volume of
the solid of revolu on formed by rota ng the region about
each of the given axes.
. Region bounded by: y =
Rotate about:

x, y =

and x = .

(a) the x-axis

(c) the y-axis

(b) y =

(d) x =

. Region bounded by: y =


Rotate about:
(a) the x-axis
(b) y =

and base radius of .

. A skew right circular cone with height of


of . (Hint: all cross-sec ons are circles.)

and base radius

x and y = .
. A right triangular cone with height of and whose base is
a right, isosceles triangle with side length .

(c) y =

(d) x =

. The triangle with ver ces ( , ), ( , ) and ( , ).


Rotate about:
(a) the x-axis

(c) the y-axis

(b) y =

(d) x =

. A solid with length with a rectangular base and triangular top, wherein one end is a square with side length and
the other end is a triangle with base and height of .

. Region bounded by y = x x + and y = x .


Rotate about:
(a) the x-axis

(c) y =

(b) y =

. Region bounded by y = / x + , x = , x =
the x-axis.
Rotate about:
(a) the x-axis
(b) y =

(c) y =

. Region bounded by y = x, y = x and x = .


Rotate about:
(a) the x-axis

(c) the y-axis

(b) y =

(d) x =

and

The Shell Method

The Shell Method

O en a given problem can be solved in more than one way. A par cular method
may be chosen out of convenience, personal preference, or perhaps necessity.
Ul mately, it is good to have op ons.
The previous sec on introduced the Disk and Washer Methods, which computed the volume of solids of revolu on by integra ng the crosssec onal area
of the solid. This sec on develops another method of compu ng volume, the
Shell Method. Instead of slicing the solid perpendicular to the axis of rota on
crea ng cross-sec ons, we now slice it parallel to the axis of rota on, crea ng
shells.
Consider Figure . , where the region shown in (a) is rotated around the
y-axis forming the solid shown in (b). A small slice of the region is drawn in (a),
parallel to the axis of rota on. When the region is rotated, this thin slice forms
a cylindrical shell, as pictured in part (c) of the gure. The previous sec on
approximated a solid with lots of thin disks (or washers); we now approximate
a solid with many thin cylindrical shells.
To compute the volume of one shell, rst consider the paper label on a soup
can with radius r and height h. What is the area of this label? A simple way of
determining this is to cut the label and lay it out at, forming a rectangle with
height h and length r. Thus the area is A = rh; see Figure . (a).
Do a similar process with a cylindrical shell, with height h, thickness x, and
approximate radius r. Cu ng the shell and laying it at forms a rectangular solid
with length r, height h and depth x. Thus the volume is V rhx; see
Figure . (b). (We say approximately since our radius was an approximaon.)
By breaking the solid into n cylindrical shells, we can approximate the volume
of the solid as
n

ri hi xi ,
V=

(a)

(b)

i=

where ri , hi and xi are the radius, height and thickness of the i th shell, respecvely.
This is a Riemann Sum. Taking a limit as the thickness of the shells approaches leads to a denite integral.

(c)
Figure .
Method.

Notes:

Introducing the Shell

Applica ons of Integra on


x

2r

V 2rhx

(a)

A = 2rh

cut here

cut here

2r

..

Chapter

(b)
Figure .

: Determining the volume of a thin cylindrical shell.

Key Idea

The Shell Method

Let a solid be formed by revolving a region R, bounded by x = a and


x = b, around a ver cal axis. Let r(x) represent the distance from the axis
of rota on to x (i.e., the radius of a sample shell) and let h(x) represent
the height of the solid at x (i.e., the height of the shell). The volume of
the solid is

V=

r(x)h(x) dx.

Special Cases:
. When the region R is bounded above by y = f(x) and below by y = g(x),
then h(x) = f(x) g(x).

y
1

h(x)

. When the axis of rota on is the y-axis (i.e., x = ) then r(x) = x.

1
y=
1 + x2

Figure .
.

{z

r(x)

Lets prac ce using the Shell Method.

}x

Finding volume using the Shell Method


Example
Find the volume of the solid formed by rota ng the region bounded by y = ,
y = /( + x ), x = and x = about the y-axis.
1

: Graphing a region in Example

This is the region used to introduce the Shell Method in FigS


ure . , but is sketched again in Figure . for closer reference. A line is drawn
in the region parallel to the axis of rota on represen ng a shell that will be

Notes:

The Shell Method

carved out as the region is rotated about the y-axis. (This is the dieren al element.)
The distance this line is from the axis of rota on determines r(x); as the
distance from x to the y-axis is x, we have r(x) = x. The height of this line
determines h(x); the top of the line is at y = /( + x ), whereas the bo om
of the line is at y = . Thus h(x) = /( + x ) = /( + x ). The region is
bounded from x = to x = , so the volume is
V=

x
dx.
+x

This requires subs tu on. Let u = + x , so du = x dx. We also change the


bounds: u( ) = and u( ) = . Thus we have:

u


= ln u

du

= ln .

x+

y=

units .

Note: in order to nd this volume using the Disk Method, two integrals would
be needed to account for the regions above and below y = / .

h(x).

{z

r(x)

(a)

With the Shell Method, nothing special needs to be accounted for to compute the volume of a solid that has a hole in the middle, as demonstrated next.
Finding volume using the Shell Method
Example
Find the volume of the solid formed by rota ng the triangular region determined
by the points ( , ), ( , ) and ( , ) about the line x = .
The region is sketched in Figure . (a) along with the difS
feren al element, a line within the region parallel to the axis of rota on. In part
(b) of the gure, we see the shell traced out by the dieren al element, and in
part (c) the whole solid is shown.

(b)

The height of the dieren al element is the distance from y = to y = x +


, the line that connects the points ( , ) and ( , ). Thus h(x) = x+ = x.
The radius of the shell formed by the dieren al element is the distance from
x to x = ; that is, it is r(x) = x. The x-bounds of the region are x = to

Notes:

(c)
Figure .
.

: Graphing a region in Example

Chapter

Applica ons of Integra on


x = , giving
V=
=

( x)( x) dx
(

x x ) dx
)

x x

units .

Finding volume using the Shell Method


Example
Find the volume of the solid formed by rota ng the region given in Example
about the x-axis.

y
x=

{z

h(y)

r(y)
x

(a)

The region is sketched in Figure . (a) with a sample difS


feren al element. In part (b) of the gure the shell formed by the dieren al
element is drawn, and the solid is sketched in (c). (Note that the triangular region looks short and wide here, whereas in the previous example the same
region looked tall and narrow. This is because the bounds on the graphs are
dierent.)
The height of the dieren al element is an x-distance, between x = y
and x = . Thus h(y) = ( y ) = y+ . The radius is the distance from
y to the x-axis, so r(y) = y. The y bounds of the region are y = and y = ,
leading to the integral
V=

: Graphing a region in Example

)]
[ (
dy
y y+

]
y + y dy
=
]
[

= y + y
]
[

=
=

(c)

(b)

Figure .
.

When revolving a region around a horizontal axis, we must consider the radius and height func ons in terms of y, not x.

Notes:

units .

The Shell Method

At the beginning of this sec on it was stated that it is good to have op ons.
The next example nds the volume of a solid rather easily with the Shell Method,
but using the Washer Method would be quite a chore.
Example
Finding volume using the Shell Method
Find the volume of the solid formed by revolving the region bounded by y = sin x
and the x-axis from x = to x = about the y-axis.

y
1

The region and a dieren al element, the shell formed by


this dieren al element, and the resul ng solid are given in Figure . . The
radius of a sample shell is r(x) = x; the height of a sample shell is h(x) = sin x,
each from x = to x = . Thus the volume of the solid is
S

V=

x sin x dx.

r(x)

z }| {

h(x)

(a)

This requires Integra on By Parts. Set u = x and dv = sin x dx; we leave it to


the reader to ll in the rest. We have:

[

= x cos x +
]
[

= + sin x
[
]
= +
=

cos x dx

units .

(b)

Note that in order to use the Washer Method, we would need to solve y =
sin x for x, requiring the use of the arcsine func on. We leave it to the reader
to verify that the outside radius func on is R(y) = arcsin y and the inside
radius func on is r(y) = arcsin y. Thus the volume can be computed as

[
]
( arcsin y) (arcsin y) dy.

This integral isnt terrible given that the arcsin y terms cancel, but it is more
onerous than the integral created by the Shell Method.
We end this sec on with a table summarizing the usage of the Washer and
Shell Methods.

Notes:

(c)
Figure .
.

: Graphing a region in Example

Chapter

Applica ons of Integra on

Key Idea

Summary of the Washer and Shell Methods

Let a region R be given with x-bounds x = a and x = b and y-bounds


y = c and y = d.
Washer Method
Horizontal
Axis

Ver cal
Axis

b
a
d

R(x) r(x)

dx

R(y) r(y)

dy

(
(

Shell Method
d
r(y)h(y) dy

b
a

r(x)h(x) dx

As in the previous sec on, the real goal of this sec on is not to be able to
compute volumes of certain solids. Rather, it is to be able to solve a problem
by rst approxima ng, then using limits to rene the approxima on to give the
exact value. In this sec on, we approximate the volume of a solid by cu ng it
into thin cylindrical shells. By summing up the volumes of each shell, we get an
approxima on of the volume. By taking a limit as the number of equally spaced
shells goes to innity, our summa on can be evaluated as a denite integral,
giving the exact value.
We use this same principle again in the next sec on, where we nd the
length of curves in the plane.

Notes:

Exercises .

Terms and Concepts

y
1

. T/F: A solid of revolu on is formed by revolving a shape


around an axis.

y = cos x

0.5

. T/F: The Shell Method can only be used when the Washer
Method fails.

0.5

. T/F: The Shell Method works by integra ng crosssec onal


areas of a solid.

1.5

y
1

. T/F: When nding the volume of a solid of revolu on that


was revolved around a ver cal axis, the Shell Method integrates with respect to x.

y=

y=x

0.5

Problems
In Exercises , a region of the Cartesian plane is shaded.
Use the Shell Method to nd the volume of the solid of revolu on formed by revolving the region about the y-axis.

0.5

In Exercises , a region of the Cartesian plane is shaded.


Use the Shell Method to nd the volume of the solid of revolu on formed by revolving the region about the x-axis.

y
y=

y=

y
y = 5x
y = 5x

x
.5

.5

x
.5

.5

. Region bounded by: y =

Rotate about:

y = cos x

0.5

x and y = .

(a) x =

(c) the x-axis

(b) x =

(d) y =

. The triangle with ver ces ( , ), ( , ) and ( , ).

0.5

1.5

Rotate about:
(a) the y-axis

(c) the x-axis

(b) x =

(d) y =

1
y=

. Region bounded by y = x x + and y = x .

Rotate about:

y=x

0.5

(a) the y-axis


(b) x =

0.5

In Exercises
, a region of the Cartesian plane is described. Use the Shell Method to nd the volume of the solid
of revolu on formed by rota ng the region about each of the
given axes.
. Region bounded by: y =
Rotate about:

x, y =

and x = .

(c) x =

. Region bounded by y = / x + , x =
y-axes.

and the x and

Rotate about:
(a) the y-axis

(b) x =

. Region bounded by y = x, y = x and x = .


Rotate about:

(a) the y-axis

(c) the x-axis

(a) the y-axis

(c) the x-axis

(b) x =

(d) y =

(b) x =

(d) y =

Arc Length and Surface Area

Arc Length and Surface Area

In previous sec ons we have used integra on to answer the following ques ons:
. Given a region, what is its area?
. Given a solid, what is its volume?
In this sec on, we address a related ques on: Given a curve, what is its
length? This is o en referred to as arc length.
Consider the graph of y = sin x on [ , ] given in Figure . (a). How long is
this curve? That is, if we were to use a piece of string to exactly match the shape
of this curve, how long would the string be?
As we have done in the past, we start by approxima ng; later, we will rene
our answer using limits to get an exact solu on.
The length of straightline segments is easy to compute using the Distance
Formula. We can approximate the length of the given curve by approxima ng
the curve with straight lines and measuring their lengths.
In Figure . (b), the curve y = sin x has been approximated with line
segments (the interval [ , ] has been divided into equallylengthed subintervals). It is clear that these four line segments approximate y = sin x very well
on the rst and last subinterval, though not so well in the middle. Regardless,
the sum of the lengths of the line segments is . , so we approximate the arc
length of y = sin x on [ , ] to be . .
In general, we can approximate the arc length of y = f(x) on [a, b] in the
following manner. Let a = x < x < . . . < xn < xn+ = b be a par on
of [a, b] into n subintervals. Let xi represent the length of the i th subinterval
[xi , xi+ ].
Figure . zooms in on the i th subinterval where y = f(x) is approximated
by a straight line segment. The dashed lines show that we can view this line
segment as they hypotenuse of a right triangle whose sides have length xi
and yi . Using the Pythagorean Theorem, the length of this line segment is

xi + yi . Summing over all subintervals gives an arc length approxima on


L

i=

xi + yi .

As shown here, this is not a Riemann Sum. While we could conclude that
taking a limit as the subinterval length goes to zero gives the exact arc length,
we would not be able to compute the answer with a denite integral. We need
rst to do a li le algebra.

Notes:

(a)
y

(b)
Figure . : Graphing y = sin x on [ , ]
and approxima ng the curve with line
segments.

y
yi+1
yi
yi
xi

xi

xi+1

Figure . : Zooming in on the i th subinterval [xi , xi+ ] of a par on of [a, b].

Chapter

Applica ons of Integra on


In the above expression factor out a xi term:

(
)
n
n

yi
+
xi + yi =
xi
.
xi
i=
i=
Now pull the xi term out of the square root:
=

i=

yi
xi .
xi

This is nearly a Riemann Sum. Consider the yi /xi term. The expression
yi /xi measures the change in y/change in x, that is, the rise over run of
f on the i th subinterval. The Mean Value Theorem of Dieren a on (Theorem
) states that there is a ci in the i th subinterval where f (ci ) = yi /xi . Thus
we can rewrite our above expression as:
=

+ f (ci ) xi .

i=

This is a Riemann Sum. As long as f is con nuous, we can invoke Theorem


and conclude
=

Key Idea

b
a

+ f (x) dx.

Arc Length

Let f be dieren able on an open interval containing [a, b], where f is


also con nuous on [a, b]. Then the arc length of f from x = a to x = b is
L=

b
a

+ f (x) dx.

As the integrand contains a square root, it is o en dicult to use the formula in Key Idea to nd the length exactly. When exact answers are dicult
to come by, we resort to using numerical methods of approxima ng denite integrals. The following examples will demonstrate this.

Notes:

.
Example
Finding arc length
Find the arc length of f(x) = x / from x =

We begin by nding f (x) = x


nd the arc length L as

L=
=
=

+ x

A graph of f is given in Figure .

. Using the formula, we

4
2

dx

dx

)
+ x
/

Figure .
Example

units.

Finding arc length

Find the arc length of f(x) =

x ln x from x =

to x = .

This func on was chosen specically because the resul ng


S
integral can be evaluated exactly. We begin by nding f (x) = x/ /x. The
arc length is

L=
=
=
=

Notes:

+ x dx

Example

to x = .

(
x

)
+

dx

dx

dx

Arc Length and Surface Area

dx

: A graph of f(x) = x
.

4
/

from

Chapter

Applica ons of Integra on

x
(
)
x

=
+ ln x

=

.5

.
Figure . : A graph of f(x) = x ln x
from Example
.

+ ln .

A graph of f is given in Figure .


problem is in bold.

dx

units.

; the por on of the curve measured in this

The previous examples found the arc length exactly through careful choice
of the func ons. In general, exact answers are much more dicult to come by
and numerical approxima ons are necessary.
Approxima ng arc length numerically
Example
Find the length of the sine curve from x = to x = .
This is somewhat of a mathema cal curiosity; in Example
we found the area under one hump of the sine curve is square units;
now we are measuring its arc length.
The setup is straigh orward: f(x) = sin x and f (x) = cos x. Thus

+ cos x dx.
L=
S

x
/
/
/

+ cos x

Figure . : A table of values of y =

+ cos x to evaluate a denite integral in Example


.

This integral cannot be evaluated in terms of elementary func ons sowe will ap+ cos x
proximate it with Simpsons Method with n = . Figure . gives
evaluated at evenly spaced points in [ , ]. Simpsons Rule then states that

)
(
+ cos x dx
+
/ + ( )+
/ +

= .
.
Using a computer with n =
the approxima on is L .
ma on with n = is quite good.

Notes:

; our approxi-

Arc Length and Surface Area

Surface Area of Solids of Revolu on


We have already seen how a curve y = f(x) on [a, b] can be revolved around
an axis to form a solid. Instead of compu ng its volume, we now consider its
surface area.
We begin as we have in the previous sec ons: we par on the interval [a, b]
with n subintervals, where the i th subinterval is [xi , xi+ ]. On each subinterval,
we can approximate the curve y = f(x) with a straight line that connects f(xi )
and f(xi+ ) as shown in Figure . (a). Revolving this line segment about the xaxis creates part of a cone (called a frustum of a cone) as shown in Figure . (b).
The surface area of a frustum of a cone is

length average of the two radii R and r.


The length is given by L; we use the material just covered by arc length to
state that

+ f (ci )xi
L

.
a

xi

xi+1

(a)

for some ci in the i th subinterval. The radii are just the func on evaluated at the
endpoints of the interval. That is,
R = f(xi+ )

and

r = f(xi ).

Thus the surface area of this sample frustum of the cone is approximately

f(xi ) + f(xi+ )

+ f (ci ) xi .

Since f is a con nuous func on, the Intermediate Value Theorem states there
f(xi ) + f(xi+ )
is some di in [xi , xi+ ] such that f(di ) =
; we can use this to rewrite
the above equa on as

+ f (ci ) xi .
f(di )

Summing over all the subintervals we get the total surface area to be approximately
n

f(di )
Surface Area
+ f (ci ) xi ,
i=

which is a Riemann Sum. Taking the limit as the subinterval lengths go to zero
gives us the exact surface area, given in the following Key Idea.

Notes:

(b)
Figure . : Establishing the formula for
surface area.

Chapter

Applica ons of Integra on

Key Idea

Surface Area of a Solid of Revolu on

Let f be dieren able on an open interval containing [a, b] where f is


also con nuous on [a, b].
. The surface area of the solid formed by revolving the graph of y =
f(x), where f(x) , about the x-axis is
Surface Area =

b
a

f(x)

+ f (x) dx.

. The surface area of the solid formed by revolving the graph of y =


f(x) about the y-axis, where a, b , is
Surface Area =

b
a

+ f (x) dx.

(When revolving y = f(x) about the y-axis, the radii of the resul ng frustum
are xi and xi+ ; their average value is simply the midpoint of the interval. In the
limit, this midpoint is just x. This gives the second part of Key Idea .)
Example
Finding surface area of a solid of revolu on
Find the surface area of the solid formed by revolving y = sin x on [ , ] around
the x-axis, as shown in Figure . .
The setup is rela vely straigh orward. Using Key Idea
S
we have the surface area SA is:

+ cos x dx
sin x
SA =

)
(


=
+ cos x
sinh (cos x) + cos x
(
)
=
+ sinh

Figure . : Revolving y = sin x on [ , ]


about the x-axis.

units .

The integra on step above is nontrivial, u lizing an integra on method called


Trigonometric Subs tu on.
It is interes ng to see that the surface area of a solid, whose shape is dened
by a trigonometric func on, involves both a square root and an inverse hyperbolic trigonometric func on.

Notes:

Arc Length and Surface Area

Finding surface area of a solid of revolu on


Example
Find the surface area of the solid formed by revolving the curve y = x on [ , ]
about:
. the x-axis
. the y-axis.
S

. The integral is straigh orward to setup:

x
SA =
+ ( x) dx.

Like the integral in Example


, this requires Trigonometric Subs tu on.
)

(

=
( x + x)
+ x sinh ( x)
)
(
sinh
=
.

units .

(a)

The solid formed by revolving y = x around the x-axis is graphed in Figure


. (a).
. Since we are revolving around the y-axis, the radius of the solid is not
f(x) but rather x. Thus the integral to compute the surface area is:

+ ( x) dx.
x
SA =
This integral can be solved using subs tu on. Set u =
bounds are u = to u = . We then have

=
u du
=
=

u
(

units .

+ x ; the new

The solid formed by revolving y = x about the y-axis is graphed in Figure


. (b).
Our nal example is a famous mathema cal paradox.
Notes:

(b)
Figure .
.

: The solids used in Example

Chapter

Applica ons of Integra on


Example
The surface area and volume of Gabriels Horn
Consider the solid formed by revolving y = /x about the x-axis on [ , ). Find
the volume and surface area of this solid. (This shape, as graphed in Figure . ,
is known as Gabriels Horn since it looks like a very long horn that only a supernatural person, such as an angel, could play.)
S

We have:

To compute the volume it is natural to use the Disk Method.


V=

= lim
b

= lim
b

Figure .

= lim

: A graph of Gabriels Horn.

= units .

dx
b

dx
) b



)
b

Gabriels Horn has a nite volume of cubic units. Since we have already seen
that regions with innite length can have a nite area, this is not too dicult to
accept.
We now consider its surface area. The integral is straigh orward to setup:

+ /x dx.
SA =
x

Integra ng this expression is not


trivial. We can, however, compare it to other
improper integrals. Since <
+ /x on [ , ), we can state that


+ /x dx.

dx <
x
x

By Key Idea , the improper integral on the le diverges. Since the integral
on the right is larger, we conclude it also diverges, meaning Gabriels Horn has
innite surface area.
Hence the paradox: we can ll Gabriels Horn with a nite amount of paint,
but since it has innite surface area, we can never paint it.
Somehow this paradox is striking when we think about it in terms of volume and area. However, we have seen a similar paradox before, as referenced
above. We know that the area under the curve y = /x on [ , ) is nite, yet
the shape has an innite perimeter. Strange things can occur when we deal with
the innite.
A standard equa on from physics is Work = force distance, when the
force applied is constant. In the next sec on we learn how to compute work
when the force applied is variable.
Notes:

Exercises .

Terms and Concepts


. T/F: The integral formula for compu ng Arc Length was
found by rst approxima ng arc length with straight line
segments.
. T/F: The integral formula for compu ng Arc Length includes
a squareroot, meaning the integra on is probably easy.

x on [ , ]. (Note: this describes the top


. f(x) =
half of a circle with radius .)

. f(x) =
x / on [ , ]. (Note: this describes the top
half of an ellipse with a major axis of length and a minor
axis of length .)
. f(x) =

on [ , ].

. f(x) = sec x on [/ , / ].

Problems
In Exercises
given interval.

, nd the arc length of the func on on the

In Exercises
, use Simpsons Rule, with n = , to approximate the arc length of the func on on the given interval.
Note: these are the same problems as in Exercises .

. f(x) = x on [ , ].

. f(x) = x on [ , ].

. f(x) = x on [ , ].

. f(x) =
. f(x) =

x on [ , ].

. f(x) =

x +

. f(x) = x

on [ , ].

on [ , ].

x on [ , ].

)
( x
e + ex on [ , ln ].

. f(x) =

x +

on [. , ].

)
. f(x) = ln sin x on [/ , / ].
(

(
)
. f(x) = ln cos x on [ , / ].

In Exercises
, set up the integral to compute the arc
length of the func on on the given interval. Do not evaluate
the integral.
. f(x) = x on [ , ].
. f(x) = x on [ , ].
. f(x) =

x on [ , ]. (Note: f (x) is not dened at x = .)

. f(x) = ln x on [ , e].

. f(x) = cosh x on [ ln , ln ].
. f(x) =

. f(x) =

x on [ , ].

. f(x) = ln x on [ , e].

. f(x) =
x on [ , ]. (Note: f (x) is not dened at
the endpoints.)

x / on [ , ]. (Note: f (x) is not dened


. f(x) =
at the endpoints.)
. f(x) =

on [ , ].

. f(x) = sec x on [/ , / ].
In Exercises
, nd the surface area of the described
solid of revolu on.
. The solid formed by revolving y = x on [ , ] about the
x-axis.
. The solid formed by revolving y = x on [ , ] about the
y-axis.
. The solid formed by revolving y = x on [ , ] about the
x-axis.
. The solid formed by revolving y =
x-axis.

. The sphere formed by revolving y =


about the x-axis.

x on [ , ] about the

x on [ , ]

Chapter

Applica ons of Integra on

Note: Mass and weight are closely related, yet dierent, concepts. The mass
m of an object is a quan ta ve measure
of that objects resistance to accelera on.
The weight w of an object is a measurement of the force applied to the object by
the accelera on of gravity g.
Since the two measurements are propor onal, w = m g, they are o en
used interchangeably in everyday conversa on. When compu ng work, one must
be careful to note which is being referred
to. When mass is given, it must be mul plied by the accelera on of gravity to reference the related force.

Work

Work is the scien c term used to describe the ac on of a force which moves
an object. When a constant force F is applied to move an object a distance d,
the amount of work performed is W = F d.
The SI unit of force is the Newton, (kgm/s ), and the SI unit of distance is
a meter (m). The fundamental unit of work is one Newtonmeter, or a joule
(J). That is, applying a force of one Newton for one meter performs one joule
of work. In Imperial units (as used in the United States), force is measured in
pounds (lb) and distance is measured in feet ( ), hence work is measured in
lb.
When force is constant, the measurement of work is straigh orward. For
instance, li ing a
lb object
performs
=
lb of work.
What if the force applied is variable? For instance, imagine a climber pulling
a
rope up a ver cal face. The rope becomes lighter as more is pulled in,
requiring less force and hence the climber performs less work.
In general, let F(x) be a force func on on an interval [a, b]. We want to measure the amount of work done applying the force F from x = a to x = b. We can
approximate the amount of work being done by par oning [a, b] into subintervals a = x < x < < xn+ = b and assuming that F is constant on each
subinterval. Let ci be a value in the i th subinterval [xi , xi+ ]. Then the work done
on this interval is approximately Wi F(ci ) (xi+ xi ) = F(ci )xi , a constant
force the distance over which it is applied. The total work is
W=

i=

Wi

F(ci )xi .

i=

This, of course, is a Riemann sum. Taking a limit as the subinterval lengths go


to zero give an exact value of work which can be evaluated through a denite
integral.
Key Idea

Work

Let F(x) be a con nuous func on on [a, b] describing the amount of force
being applied to an object in the direc on of travel from distance x = a
to distance x = b. The total work W done on [a, b] is
W=

Notes:

b
a

F(x) dx.

.
Example
Compu ng work performed: applying variable force
A m climbing rope is hanging over the side of a tall cli. How much work
is performed in pulling the rope up to the top, where the rope has a mass of
g/m?
We need to create a force func on F(x) on the interval [ , ].
S
To do so, we must rst decide what x is measuring: it is the length of the rope
s ll hanging or is it the amount of rope pulled in? As long as we are consistent,
either approach is ne. We adopt for this example the conven on that x is the
amount of rope pulled in. This seems to match intui on be er; pulling up the
rst meters of rope involves x = to x =
instead of x =
to x = .
As x is the amount of rope pulled in, the amount of rope s ll hanging is x.
This length of rope has a mass of
g/m, or .
kg/m. The the mass of the
rope s ll hanging is .
( x) kg; mul plying this mass by the accelera on
of gravity, . m/s , gives our variable force func on
F(x) = ( . )( .

)(

x) = .

x).

Thus the total work performed in pulling up the rope is


W=

x) dx = ,

J.

By comparison, consider the work done in li ing the en re rope meters.


The rope weights .
. = .
N, so the work applying this force
for meters is .
= ,
. J. This is exactly twice the work calculated before (and we leave it to the reader to understand why.)
Compu ng work performed: applying variable force
Example
Consider again pulling a m rope up a cli face, where the rope has a mass of
g/m. At what point is exactly half the work performed?
From Example
we know the total work performed is
S
,
. J. We want to nd a height h such that the work in pulling the rope
from a height of x = to a height of x = h is
. , half the total work. Thus
we want to solve the equa on

for h.

Notes:

x) dx =

Work

Chapter

Applica ons of Integra on

Note: In Example
, we nd that half of
the work performed in pulling up a
m
rope is done in the last
. m. Why is it
= . ?
not coincidental that /

(
.

x .

.
.

h +

x) dx =

h .
h

) h
=

h =

= .

Apply the Quadra c Formula.


h=

and

As the rope is only m long, the only sensible answer is h = . . Thus about
half the work is done pulling up the rst . m the other half of the work is done
pulling up the remaining . m.
Compu ng work performed: applying variable force
Example
A box of
lb of sand is being pulled up at a uniform rate a distance of
over minute. The sand is leaking from the box at a rate of lb/s. The box itself
weighs lb and is pulled by a rope weighing . lb/ .
. How much work is done li ing just the rope?
. How much work is done li ing just the box and sand?
. What is the total amount of work performed?
S

. We start by forming the force func on Fr (x) for the rope (where the subscript denotes we are considering the rope). As in the previous example,
let x denote the amount of rope, in feet, pulled in. (This is the same as
saying x denotes the height of the box.) The weight of the rope with x
feet pulled in is Fr (x) = . ( x) =
. x. (Note that we do not
have to include the accelera on of gravity here, for the weight of the rope
per foot is given, not its mass per meter as before.) The work performed
li ing the rope is
Wr =

Notes:

. x) dx =

lb.

.
. The sand is leaving the box at a rate of lb/s. As the ver cal trip is to take
one minute, we know that lb will have le when the box reaches its nal
height of
. Again le ng x represent the height of the box, we have
two points on the line that describes the weight of the sand: when x = ,
the sand weight is
lb, producing the point ( ,
); when x = , the
sand in the box weighs
lb, producing the point ( , ). The slope of
this line is = . , giving the equa on of the weight of the sand
at height x as w(x) = . x +
. The box itself weighs a constant lb,
so the total force func on is Fb (x) = . x +
. Integra ng from x =
to x =
gives the work performed in li ing box and sand:
Wb =

( . x +

) dx =

. The total work is the sum of Wr and Wb :


can also arrive at this via integra on:
W=
=
=
=

lb.
+

lb. We

(Fr (x) + Fb (x)) dx


(

. x . x+

( . x +

) dx

) dx

lb.

Hookes Law and Springs


Hookes Law states that the force required to compress or stretch a spring x
units from its natural length is propor onal to x; that is, this force is F(x) = kx
for some constant k. For example, if a force of N stretches a given spring
cm, then a force of N will stretch the spring
cm. Conver ng the distances to meters, we have that stretching this spring . m requires a force
of F( . ) = k( . ) = N, hence k = / . =
N/m.
Compu ng work performed: stretching a spring
Example
A force of
lb stretches a spring from a natural length of inches to a length
of
inches. How much work was performed in stretching the spring to this
length?
In many ways, we are not at all concerned with the actual
S
length of the spring, only with the amount of its change. Hence, we do not care

Notes:

Work

Chapter

Applica ons of Integra on


that
lb of force stretches the spring to a length of
inches, but rather that
a force of
lb stretches the spring by in. This is illustrated in Figure . ;
we only measure the change in the springs length, not the overall length of the
spring.
F

Figure . : Illustra ng the important aspects of stretching a spring in compu ng work


in Example
.

Conver ng the units of length to feet, we have


F( /

)= /

k=

lb.

Thus k =
lb/ and F(x) = x.
We compute the total work performed by integra ng F(x) from x =
x= / :
/
x dx
W=
=
=

to

/ .

lb.

Pumping Fluids
Fluid
Concrete
Fuel Oil
Gasoline
Iodine
Methanol
Mercury
Milk
Water
Figure .

lb/

kg/m

.
.

.
.
.

. .
.

: Weight and Mass densi es

Another useful example of the applica on of integra on to compute work


comes in the pumping of uids, o en illustrated in the context of emptying a
storage tank by pumping the uid out the top. This situa on is dierent than
our previous examples for the forces involved are constant. A er all, the force
required to move one cubic foot of water (about . lb) is the same regardless
of its loca on in the tank. What is variable is the distance that cubic foot of
water has to travel; water closer to the top travels less distance than water at
the bo om, producing less work.
We demonstrate how to compute the total work done in pumping a uid out
of the top of a tank in the next two examples.

Notes:

Work

Example
Compu ng work performed: pumping uids
A cylindrical storage tank with a radius of
and a height of
is lled with
water, which weighs approximately . lb/ . Compute the amount of work
performed by pumping the water up to a point feet above the top of the tank.

yi (

th

yi ).

To approximate the total work performed in pumping out all the water from the
tank, we sum all the work Wi performed in pumping the water from each of the
n subintervals of [ , ]:
W

Wi =

yi ).

yi (

i=

i=

This is a Riemann sum. Taking the limit as the subinterval length goes to gives

( y) dy
W=
=(
=
.

Notes:

y / y

lb

lb.

10

30

Consider the work Wi of pumping only the water residing in the i subinterval,
illustrated in Figure . . The force required to move this water is equal to its
weight which we calculate as volume density. The volume of water in this
subinterval is Vi =
yi ; its density is . lb/ . Thus the required force is
yi lb.
We approximate the distance the force is applied by using any y-value contained in the i th subinterval; for simplicity, we arbitrarily use yi for now (it will
not ma er later on). The water will be pumped to a point feet above the top
of the tank, that is, to the height of y =
. Thus the distance the water at
height yi travels is yi .
In all, the approximate work Wi peformed in moving the water in the i th
subinterval to a point feet above the tank is
Wi

35

. .

= y < y < < yn+ =

35 yi

We will refer o en to Figure . which illustrates the salient


S
aspects of this problem.
We start as we o en do: we par on an interval into subintervals. We orient
our tank ver cally since this makes intui ve sense with the base of the tank at
y = . Hence the top of the water is at y = , meaning we are interested in
subdividing the y-interval [ , ] into n subintervals as

yi+1 }
yi

yi

Figure . : Illustra ng a water tank in


order to compute the work required to
empty it in Example
.

Chapter

Applica ons of Integra on

y
35
10
35 y

30

V(y) = 100dy

Figure . : A simplied illustra on for


compu ng work.

Compu ng work performed: pumping uids


Example
A conical water tank has its top at ground level and its base feet below ground.
The radius of the cone at ground level is
. It is lled with water weighing .
lb/ and is to be emp ed by pumping the water to a spigot feet above ground
level. Find the total amount of work performed in emptying the tank.
The conical tank is sketched in Figure . . We can orient
S
the tank in a variety of ways; we could let y = represent the base of the tank
and y = represent the top of the tank, but we choose to keep the conven on
of the wording given in the problem and let y = represent ground level and
hence y = represents the bo om of the tank. The actual height of the
water does not ma er; rather, we are concerned with the distance the water
travels.
The gure also sketches a dieren al element, a crosssec onal circle. The
radius of this circle is variable, depending on y. When y = , the circle has
radius ; when y = , the circle has radius . These two points, ( , ) and
( , ), allow us to nd the equa on of the line that gives the radius of the cross
sec onal circle, which is r(y) = / y + . Hence the volume of water at this
height is V(y) = ( / y + ) dy, where dy represents a very small height of
the dieren al element. The force required to move the water at height y is
F(y) = . V(y).
The distance the water at height y travels is given by h(y) = y. Thus the
total work done in pumping the water from the tank is

. ( / y + ) ( y) dy
W=

We can streamline the above process a bit as we may now recognize what
the important features of the problem are. Figure . shows the tank from
Example
without the i th subinterval iden ed. Instead, we just draw one
dieren al element. This helps establish the height a small amount of water
must travel along with the force required to move it (where the force is volume
density).
We demonstrate the concepts again in the next examples.

V(y) =

( 5y

+ ) dy

Figure . : A graph of the conical water


tank in Example
.

Notes:

y y+
lb.

dy

.
Example
Compu ng work performed: pumping uids
A rectangular swimming pool is
wide and has a
shallow end and a
deep end. It is to have its water pumped out to a point
above the current
top of the water. The crosssec onal dimensions of the water in the pool are
given in Figure . ; note that the dimensions are for the water, not the pool
itself. Compute the amount of work performed in draining the pool.
For the purposes of this problem we choose to set y =
to represent the bo om of the pool, meaning the top of the water is at y = .
Figure . shows the pool oriented with this y-axis, along with dieren al
elements as the pool must be split into two dierent regions.
The top region lies in the y-interval of [ , ], where the length of the dierenal element is
as shown. As the pool is
wide, this dieren al element
represents a this slice of water with volume V(y) =

dy. The water is


to be pumped to a height of y = , so the height func on is h(y) = y. The
work done in pumping this top region of water is

( y) dy =
,
lb.
Wt = .

The total work in empy ng the pool is


W = Wb + Wt =

lb.

No ce how the emptying of the bo om of the pool performs almost as much


work as emptying the top. The top por on travels a shorter distance but has
more water. In the end, this extra water produces more work.
The next sec on introduces one nal applica on of the denite integral, the
calcula on of uid force on a plate.

Notes:

6 .

The bo om region lies in the y-interval of [ , ]; we need to compute the


length of the dieren al element in this interval.
One end of the dieren al element is at x = and the other is along the line
segment joining the points ( , ) and ( , ). The equa on of this line is y =
/ (x ); as we will be integra ng with respect to y, we rewrite this equa on
as x = / y + . So the length of the dieren al element is a dierence of
x-values: x = and x = / y + , giving a length of x = / y + .
Again, as the pool is
wide, this dieren al element represents a thin
slice of water with volume V(y) =
( / y + ) dy; the height func on is
the same as before at h(y) = y. The work performed in emptying this part
of the pool is

( / y + )( y) dy =
,
lb.
Wb = .

Work

Figure . : The crosssec on of a swimming pool lled with water in Example


.

y
8
6
3
y
0

(15, 3)

.
0

(10, 0)
10

15

Figure . : Orien ng the pool and showing dieren al elements for Example
.

Exercises .

Terms and Concepts


. What are the typical units of work?
. If a man has a mass of
kg on Earth, will his mass on the
moon be bigger, smaller, or the same?
. If a woman weighs
lb on Earth, will her weight on the
moon be bigger, smaller, or the same?

Problems
. A
rope, weighing . lb/ , hangs over the edge of a
tall building.
(a) How much work is done pulling the en re rope to the
top of the building?
(b) How much rope is pulled in when half of the total
work is done?
. A
m rope, with a mass density of . kg/m, hangs over
the edge of a tall building.
(a) How much work is done pulling the en re rope to the
top of the building?
(b) How much work is done pulling in the rst

m?

. A rope of length hangs over the edge of tall cli. (Assume the cli is taller than the length of the rope.) The
rope has a weight density of d lb/ .
(a) How much work is done pulling the en re rope to the
top of the cli?
(b) What percentage of the total work is done pulling in
the rst half of the rope?
(c) How much rope is pulled in when half of the total
work is done?
. A m rope with mass density of . kg/m hangs over the
edge of a
m building. How much work is done pulling
the rope to the top?
. A crane li s a ,
lb load ver cally
weighing . lb/ .

with a cable

(a) How much work is done li ing the cable alone?


(b) How much work is done li ing the load alone?
(c) Could one conclude that the work done li ing the cable is negligible compared to the work done li ing the
load?
. A
lb bag of sand is li ed uniformly
in one minute.
Sand leaks from the bag at a rate of / lb/s. What is the
total work done in li ing the bag?

. A box weighing lb li s lb of sand ver cally


. A crack
in the box allows the sand to leak out such that lb of sand
is in the box at the end of the trip. Assume the sand leaked
out at a uniform rate. What is the total work done in li ing
the box and sand?
. A force of
lb compresses a spring in. How much work
is performed in compressing the spring?
. A force of N stretches a spring cm. How much work is
performed in stretching the spring?
. A force of lb compresses a spring from a natural length of
in to in. How much work is performed in compressing
the spring?
. A force of
lb stretches a spring from a natural length of
in to in. How much work is performed in stretching the
spring?
. A force of N stretches a spring from a natural length of
cm to
cm. How much work is performed in stretching
the spring from a length of cm to cm?
. A force of f N stretches a spring d m from its natural length.
How much work is performed in stretching the spring?
. A
lb weight is a ached to a spring. The weight rests on
the spring, compressing the spring from a natural length of
to in.
How much work is done in li ing the box .
(i.e, the
spring will be stretched
beyond its natural length)?
. A
lb weight is a ached to a spring. The weight rests on
the spring, compressing the spring from a natural length of
to in.
How much work is done in li ing the box in (i.e, bringing
the spring back to its natural length)?
. A m tall cylindrical tank with radius of m is lled with
m of gasoline, with a mass density of
. kg/m . Compute the total work performed in pumping all the gasoline
to the top of the tank.
. A
cylindrical tank with a radius of
is lled with water, which has a weight density of . lb/ . The water is
to be pumped to a point
above the top of the tank.
(a) How much work is performed in pumping all the water from the tank?
(b) How much work is performed in pumping
ter from the tank?

of wa-

(c) At what point is / of the total work done?


. A gasoline tanker is lled with gasoline with a weight density of . lb/ . The dispensing valve at the base is
jammed shut, forcing the operator to empty the tank via

pumping the gas to a point


above the top of the tank.
Assume the tank is a perfect cylinder,
long with a diameter of . . How much work is performed in pumping
all the gasoline from the tank?

mensions given below, and is lled with water with a weight


density of . lb/ . Find the work performed in pumping
all water to a point
above the top of the tank.

. A fuel oil storage tank is


deep with trapezoidal sides,
at the top and
at the bo om, and is
wide (see
diagram below). Given that fuel oil weighs . lb/ , nd
the work performed in pumping all the oil from the tank to
a point
above the top of the tank.
. A water tank has the shape of an inverted pyramid, with dimensions given below, and is lled with water with a mass
density of
kg/m . Find the work performed in pumping all water to a point m above the top of the tank.
m
m
m

. A conical water tank is m deep with a top radius of m.


(This is similar to Example
.) The tank is lled with pure
water, with a mass density of
kg/m .
(a) Find the work performed in pumping all the water to
the top of the tank.
(b) Find the work performed in pumping the top . m
of water to the top of the tank.

. A water tank has the shape of an truncated, inverted pyramid, with dimensions given below, and is lled with water with a mass density of
kg/m . Find the work performed in pumping all water to a point m above the top
of the tank.
m
m
m

(c) Find the work performed in pumping the top half of


the water, by volume, to the top of the tank.
. A water tank has the shape of a truncated cone, with di-

m
m

Chapter

Applica ons of Integra on

Fluid Forces

In the unfortunate situa on of a car driving into a body of water, the convenonal wisdom is that the water pressure on the doors will quickly be so great
that they will be eec vely unopenable. (Survival techniques suggest immediately opening the door, rolling down or breaking the window, or wai ng un l
the water lls up the interior at which point the pressure is equalized and the
door will open. See Mythbusters episode # to watch Adam Savage test these
op ons.)
How can this be true? How much force does it take to open the door of
a submerged car? In this sec on we will nd the answer to this ques on by
examining the forces exerted by uids.
We start with pressure, which is related to force by the following equa ons:
Pressure =

Force = Pressure Area.

In the context of uids, we have the following deni on.

Deni on

Fluid Pressure

Let w be the weightdensity of a uid. The pressure p exerted on an


object at depth d in the uid is p = w d.

.
Figure .
.

Force
Area

: A cylindrical tank in Example

We use this deni on to nd the force exerted on a horizontal sheet by considering the sheets area.
Example

Compu ng uid force

. A cylindrical storage tank has a radius of


and holds
of a uid with
a weightdensity of lb/ . (See Figure . .) What is the force exerted
on the base of the cylinder by the uid?

. A rectangular tank whose base is a


square has a circular hatch at the
bo om with a radius of
. The tank holds
of a uid with a weight
density of
lb/ . (See Figure . .) What is the force exerted on the
hatch by the uid?

Figure .
ple
.

S
5

: A rectangular tank in Exam-

. Using Deni on , we calculate that the pressure exerted on the cylinders base is w d =
lb/
=
lb/ . The area of the base is

Notes:

Fluid Forces

. So the force exerted by the uid is


F=

lb.

Note that we eec vely just computed the weight of the uid in the tank.
. The dimensions of the tank in this problem are irrelevant. All we are concerned with are the dimensions of the hatch and the depth of the uid.
Since the dimensions of the hatch are the same as the base of the tank
in the previous part of this example, as is the depth, we see that the uid
force is the same. That is, F =
lb.
A key concept to understand here is that we are eec vely measuring the
weight of a
column of water above the hatch. The size of the tank
holding the uid does not ma er.
The previous example demonstrates that compu ng the force exerted on a
horizontally oriented plate is rela vely easy to compute. What about a ver cally
oriented plate? For instance, suppose we have a circular porthole located on the
side of a submarine. How do we compute the uid force exerted on it?
Pascals Principle states that the pressure exerted by a uid at a depth is
equal in all direc ons. Thus the pressure on any por on of a plate that is
below the surface of water is the same no ma er how the plate is oriented.
(Thus a hollow cube submerged at a great depth will not simply be crushed
from above, but the sides will also crumple in. The uid will exert force on all
sides of the cube.)
So consider a ver cally oriented plate as shown in Figure . submerged in
a uid with weightdensity w. What is the total uid force exerted on this plate?
We nd this force by rst approxima ng the force on small horizontal strips.
Let the top of the plate be at depth b and let the bo om be at depth a. (For
now we assume that surface of the uid is at depth , so if the bo om of the
plate is
under the surface, we have a = . We will come back to this later.)
We par on the interval [a, b] into n subintervals

di

a = y < y < < yn+ = b,


with the i th subinterval having length yi . The force Fi exerted on the plate in
the i th subinterval is Fi = Pressure Area.
The pressure is depth w. We approximate the depth of this thin strip by
choosing any value di in [yi , yi+ ]; the depth is approximately di . (Our convenon has di being a nega ve number, so di is posi ve.) For convenience, we let
di be an endpoint of the subinterval; we let di = yi .
The area of the thin strip is approximately length width. The width is yi .
The length is a func on of some y-value ci in the i th subinterval. We state the

Notes:

yi

(ci )

.
Figure . : A thin, ver cally oriented
plate submerged in a uid with weight
density w.

Chapter

Applica ons of Integra on


length is (ci ). Thus
Fi = Pressure Area

= yi w (ci ) yi .

To approximate the total force, we add up the approximate forces on each of


the n thin strips:
F=

i=

Fi

i=

w yi (ci ) yi .

This is, of course, another Riemann Sum. We can nd the exact force by taking
a limit as the subinterval lengths go to ; we evaluate this limit with a denite
integral.
Key Idea

Fluid Force on a Ver cally Oriented Plate

Let a ver cally oriented plate be submerged in a uid with weight


density w where the top of the plate is at y = b and the bo om is at
y = a. Let (y) be the length of the plate at y.
. If y = corresponds to the surface of the uid, then the force
exerted on the plate by the uid is
F=

b
a

w (y) (y) dy.

. In general, let d(y) represent the distance between the surface of


the uid and the plate at y. Then the force exerted on the plate by
the uid is

F=

.
Figure . : A thin plate in the shape of
an isosceles triangle in Example
.

w d(y) (y) dy.

Finding uid force


Example
Consider a thin plate in the shape of an isosceles triangle as shown in Figure .
submerged in water with a weightdensity of . lb/ . If the bo om of the
plate is
below the surface of the water, what is the total uid force exerted
on this plate?
We approach this problem in two dierent ways to illustrate
S
the dierent ways Key Idea can be implemented. First we will let y = represent the surface of the water, then we will consider an alternate conven on.

Notes:

.
. We let y = represent the surface of the water; therefore the bo om of
the plate is at y = . We center the triangle on the y-axis as shown in
Figure . . The depth of the plate at y is y as indicated by the Key Idea.
We now consider the length of the plate at y.

The total uid force is then:



F=

y
water line
x

( , 6)

( , 6)
y
8

. (y)(y +

) dy

( ,

. lb.

The car door, as a rectangle, is drawn in Figure . . Its


and its height is .
. We adopt the conven on that the

6
y

Notes:

water line

d(y) =

Finding uid force


Example
Find the total uid force on a car door submerged up to the bo om of its window
in water, where the car door is a rectangle long and high (based on the
dimensions of a
Fiat Grande Punto.)
S

. lb.

The correct answer is, of course, independent of the placement of the plate in
the coordinate plane as long as we are consistent.

Figure . : Sketching the triangular


plate in Example
with the conven on
that the water level is at y = .

As the surface of the water is


above the base of the plate, we have
that the surface of the water is at y = . Thus the depth func on is the
distance between y =
and y; d(y) =
y. We compute the total
uid force as:

. ( y)(y) dy
F=

length is

. Some mes it seems easier to orient the thin plate nearer the origin. For
instance, consider the conven on that the bo om of the triangular plate
is at ( , ), as shown in Figure . . The equa ons of the le and right
hand sides are easy to nd. They are y = x and y = x, respec vely,
which we rewrite as x = / y and x = / y. Thus the length func on
is (y) = / y ( / y) = y.

d(y) = y

We need to nd equa ons of the le and right edges of the plate. The
right hand side is a line that connects the points ( , ) and ( , ):
that line has equa on x = / (y + ). (Find the equa on in the familiar
y = mx+b format and solve for x.) Likewise, the le hand side is described
by the line x = / (y + ). The total length is the distance between
these two lines: (y) = / (y + ) ( / (y + )) = y + .

Fluid Forces

( , 4)

( , 4)
y

Figure . : Sketching the triangular


plate in Example
with the conven on
that the base of the triangle is at ( , ).

Chapter

Applica ons of Integra on


top of the door is at the surface of the water, both of which are at y = . Using
the weightdensity of water of . lb/ , we have the total force as
y

(0, 0)

(3.3, 0)

F=

y
(0, 2.25)

(3.3, 2.25)

y dy

Figure . : Sketching a submerged car


door in Example
.

/ dy

. (y)
.

. lb.

Most adults would nd it very dicult to apply over


lb of force to a car
door while seated inside, making the door eec vely impossible to open. This
is counterintui ve as most assume that the door would be rela vely easy to
open. The truth is that it is not, hence the survival ps men oned at the beginning of this sec on.
y

water line

5
d(y) = 5 y
y

not to scale

Figure . : Measuring the uid force on


an underwater porthole in Example
.

Finding uid force


Example
An underwater observa on tower is being built with circular viewing portholes
enabling visitors to see underwater life. Each ver cally oriented porthole is to
have a
diameter whose center is to be located
underwater. Find the
total uid force exerted on each porthole. Also, compute the uid force on a
horizontally oriented porthole that is under
of water.
We place the center of the porthole at the origin, meaning
S
the surface of the water is at y = and the depth func on will be d(y) = y;
see Figure .
The equa on of
a circle with a radius of . is x + y = . ; solving for
. y , where the posi ve square root corresponds to
x we have x =
the right side of the circle and the nega ve square root corresponds
to the le
side of the circle. Thus the length func on at depth y is (y) =
. y .
Integra ng on [ . , . ] we have:
F=
=
=

.
.

.
.

.
.

Notes:

y)

y dy

y y

dy

.
.

dy

(
y
.

dy.

.
The second integral above can be evaluated using Subs tu on. Let u = . y
with du = y dy. The new bounds are: u( . ) = and u( . ) = ; the new
integral will integrate from u = to u = , hence the integral is .
The rst integral above nds the area of half a circle of radius . , thus the
rst integral evaluates to
. / = ,
. Thus the total uid force
on a ver cally oriented porthole is ,
lb.
Finding the force on a horizontally oriented porthole is more straigh orward:
F = Pressure Area =

lb.

That these two forces are equal is not coincidental; it turns out that the uid
force applied to a ver cally oriented circle whose center is at depth d is the
same as force applied to a horizontally oriented circle at depth d.
We end this chapter with a reminder of the true skills meant to be developed
here. We are not truly concerned with an ability to nd uid forces or the volumes of solids of revolu on. Work done by a variable force is important, though
measuring the work done in pulling a rope up a cli is probably not.
What we are actually concerned with is the ability to solve certain problems
by rst approxima ng the solu on, then rening the approxima on, then recognizing if/when this rening process results in a denite integral through a limit.
Knowing the formulas found inside the special boxes within this chapter is benecial as it helps solve problems found in the exercises, and other mathema cal
skills are strengthened by properly applying these formulas. However, more importantly, understand how each of these formulas was constructed. Each is the
result of a summa on of approxima ons; each summa on was a Riemann sum,
allowing us to take a limit and nd the exact answer through a denite integral.
The next chapter addresses an en rely dierent topic: sequences and series.
In short, a sequence is a list of numbers, where a series is the summa on of a list
of numbers. These seeminglysimple ideas lead to very powerful mathema cs.

Notes:

Fluid Forces

Exercises .

Terms and Concepts


. State in your own words Pascals Principle.

. State in your own words how pressure is dierent from


force.

Problems
In Exercises , nd the uid force exerted on the given
plate, submerged in water with a weight density of .
lb/ .
.

.
.

.
.

y=

In Exercises
, the side of a container is pictured. Find
the uid force exerted on this plate when the container is full
of:

. water, with a weight density of

. lb/

. concrete, with a weight density of

lb/

.
y=

, and
.

y=

x
x

y=

. How deep must the center of a ver cally oriented circular


plate with a radius of
be submerged in water, with a
weight density of . lb/ , for the uid force on the plate
to reach ,
lb?
.

y=x

. How deep must the center of a ver cally oriented square


plate with a side length of
be submerged in water, with
a weight density of . lb/ , for the uid force on the
plate to reach ,
lb?

:S

This chapter introduces sequences and series, important mathema cal construc ons that are useful when solving a large variety of mathema cal problems. The content of this chapter is considerably dierent from the content of
the chapters before it. While the material we learn here denitely falls under
the scope of calculus, we will make very li le use of deriva ves or integrals.
Limits are extremely important, though, especially limits that involve innity.
One of the problems addressed by this chapter is this: suppose we know
informa on about a func on and its deriva ves at a point, such as f( ) = ,
f ( ) = , f ( ) = , f ( ) = , and so on. What can I say about f(x) itself?
Is there any reasonable approxima on of the value of f( )? The topic of Taylor
Series addresses this problem, and allows us to make excellent approxima ons
of func ons when limited knowledge of the func on is available.

Sequences

We commonly refer to a set of events that occur one a er the other as a sequence of events. In mathema cs, we use the word sequence to refer to an
ordered set of numbers, i.e., a set of numbers that occur one a er the other.
For instance, the numbers , , , , , form a sequence. The order is important; the rst number is , the second is , etc. It seems natural to seek a formula
that describes a given sequence, and o en this can be done. For instance, the
sequence above could be described by the func on a(n) = n, for the values of
n = , , . . . To nd the th term in the sequence, we would compute a( ).
This leads us to the following, formal deni on of a sequence.
Deni on

Sequence

A sequence is a func on a(n) whose domain is N. The range of a


sequence is the set of all dis nct values of a(n).
The terms of a sequence are the values a( ), a( ), , which are usually
denoted with subscripts as a , a , .
A sequence a(n) is o en denoted as {an }.

Nota on: We use N to describe the set of


natural numbers, that is, the integers , ,
,

Factorial: The expression ! refers to the


number = .
In general, n! = n(n )(n ) ,
where n is a natural number.
We dene ! = . While this does not
immediately make sense, it makes many
mathema cal formulas work properly.

Chapter

Sequences and Series


Example
Lis ng terms of a sequence
List the rst four terms of the following sequences.

. {an } =

n!

. {an } = { +( )n }

. {an } =

( )n(n+
n

)/

S
an =

= ;
a =
= ;
a =
= ;
a =
=
!
!
!
!
We can plot the terms of a sequence with a sca er plot. The x-axis is
used for the values of n, and the values of the terms are plo ed on the
y-axis. To visualize this sequence, see Figure . (a).

. a =

n!

(a)
y

. a =

a =

+ ( ) = ;

+ ( ) = ;

a = +( ) = ;
a = +( ) = . Note that the range of this
sequence is nite, consis ng of only the values and . This sequence is
plo ed in Figure . (b).

an =

. a =

( )

( )/

a =

( )

( )/

a =

( )

( )/

+ ( )

(b)
y

a =

= ;

a =

=
=

( )

( )

( )/

( )/

We gave one extra term to begin to show the pa ern of signs is , , +,


+, , , . . ., due to the fact that the exponent of is a special quadra c.
This sequence is plo ed in Figure . (c).

/
/
n

an =

( )n(n+
n

)/

(c)

Figure . : Plo ng sequences in Example


.

Determining a formula for a sequence


Example
Find the nth term of the following sequences, i.e., nd a func on that describes
each of the given sequences.
.

, , ,

, ...

, ,

, , , ,

, ,

Notes:

, ,

,
,
, ...

, ...
, ...

.
We should rst note that there is never exactly one func on that
describes a nite set of numbers as a sequence. There are many sequences
that start with , then , as our rst example does. We are looking for a simple
formula that describes the terms given, knowing there is possibly more than one
answer.
S

. Note how each term is more than the previous one. This implies a linear
func on would be appropriate: a(n) = an = n + b for some appropriate
value of b. As we want a = , we set b = . Thus an = n .
. First no ce how the sign changes from term to term. This is most commonly accomplished by mul plying the terms by either ( )n or ( )n+ .
Using ( )n mul plies the odd terms by ( ); using ( )n+ mul plies
the even terms by ( ). As this sequence has nega ve even terms, we
will mul ply by ( )n+ .
A er this, we might feel a bit stuck as to how to proceed. At this point,
we are just looking for a pa ern of some sort: what do the numbers , ,
, , etc., have in common? There are many correct answers, but the
one that well use here is that each is one more than a perfect square.
That is, =
+ , =
+ ,
=
+ , etc. Thus our formula is
an = ( )n+ (n + ).
. One who is familiar with the factorial func on will readily recognize these
numbers. They are !, !, !, !, etc. Since our sequences start with n = ,
we cannot write an = n!, for this misses the ! term. Instead, we shi by
, and write an = (n )!.
. This one may appear dicult, especially as the rst two terms are the
same, but a li le sleuthing will help. No ce how the terms in the numerator are always mul ples of , and the terms in the denominator are
always powers of . Does something as simple as an = nn work?
When n = , we see that we indeed get / as desired. When n = ,
we get / = / . Further checking shows that this formula indeed
matches the other terms of the sequence.
A common mathema cal endeavor is to create a new mathema cal object
(for instance, a sequence) and then apply previously known mathema cs to the
new object. We do so here. The fundamental concept of calculus is the limit, so
we will inves gate what it means to nd the limit of a sequence.

Notes:

Sequences

Chapter

Sequences and Series

Deni on

Limit of a Sequence, Convergent, Divergent

Let {an } be a sequence and let L be a real number. Given any > , if
an m can be found such that |an L| < for all n > m, then we say the
limit of {an }, as n approaches innity, is L, denoted
lim an = L.

If lim an exists, we say the sequence converges; otherwise, the sen


quence diverges.
This deni on states, informally, that if the limit of a sequence is L, then if
you go far enough out along the sequence, all subsequent terms will be really
close to L. Of course, the terms far enough and really close are subjec ve
terms, but hopefully the intent is clear.
This deni on is reminiscent of the proofs of Chapter . In that chapter
we developed other tools to evaluate limits apart from the formal deni on; we
do so here as well.
Theorem

Limit of a Sequence

Let {an } be a sequence and let f(x) be a func on whose domain contains
the posi ve real numbers where f(n) = an for all n in N.
If lim f(x) = L, then lim an = L.
x

Theorem allows us, in certain cases, to apply the tools developed in Chapter to limits of sequences. Note two things not stated by the theorem:
. If lim f(x) does not exist, we cannot conclude that lim an does not exist.
x

It may, or may not, exist. For instance, we can dene a sequence {an } =
{cos( n)}. Let f(x) = cos( x). Since the cosine func on oscillates
over the real numbers, the limit lim f(x) does not exist.
x

However, for every posi ve integer n, cos( n) = , so lim an = .


n

. If we cannot nd a func on f(x) whose domain contains the posi ve real


numbers where f(n) = an for all n in N, we cannot conclude lim an does
n
not exist. It may, or may not, exist.

Notes:

.
Example
Determining convergence/divergence of a sequence
Determine the convergence or divergence of the following sequences.
{
}
{
}
n n+
( )n
. {an } =
. {an } = {cos n}
. {an } =
n
n

Sequences

x x+
= . (We could
x x
have also directly applied lHpitals Rule.) Thus the sequence {an } converges, and its limit is . A sca er plot of every values of an is given in
Figure . (a). The values of an vary widely near n = , ranging from
about to
, but as n grows, the values approach .

. Using Theorem

, we can state that lim

an =

value in [ , ] innitely many mes). Thus we cannot apply Theorem

an = cos n

The fact that the cosine func on oscillates strongly hints that cos n, when
n is restricted to N, will also oscillate. Figure . (b), where the sequence is
plo ed, shows that this is true. Because only discrete values of cosine are
plo ed, it does not bear strong resemblance to the familiar cosine wave.

.
(b)

. We cannot actually apply Theorem here, as the func on f(x) = ( ) /x


is not well dened. (What does ( ) mean? In actuality, there is an answer, but it involves complex analysis, beyond the scope of this text.) So
for now we say that we cannot determine the limit. (But we will be able
to very soon.) By looking at the plot in Figure . (c), we would like to
conclude that the sequence converges to . That is true, but at this point
we are unable to decisively say so.
It seems that {( )n /n} converges to but we lack the formal tool to prove
it. The following theorem gives us that tool.

.
y

.5

5
an =

Determining the convergence/divergence of a sequence


Example
Determine the convergence or divergence of the following sequences.

Notes:

( )n
n

(c)

Let {an } be a sequence. If lim |an | = , then lim an =


n

.5

Absolute Value Theorem

We conclude that lim an does not exist.

Theorem

x x+
x

(a)

. The limit lim cos x does not exist, as cos x oscillates (and takes on every
x

Figure . : Sca er plots of the sequences


in Example
.

Chapter

Sequences and Series


. {an } =

( )n
n

. {an } =

( )n (n + )
n

. This appeared in Example


the limit of {|an |}:

. We want to apply Theorem

, so consider



( )n

lim |an | = lim
n
n
n
= lim

= .
Since this limit is , we can apply Theorem

and state that lim an = .


n

. Because of the alterna ng nature of this sequence (i.e., every other term
( )x (x + )
is mul plied by ), we cannot simply look at the limit lim
.
x
x
We can try to apply the techniques of Theorem :


( )n (n + )


lim |an | = lim

n
n
n
n+
= lim
n
n
= .

y
an =

( )n (n + )
n

We have concluded that when we ignore the alterna ng sign, the sequence
approaches . This means we cannot apply Theorem ; it states the the
limit must be in order to conclude anything.

Since we know that the signs of the terms alternate and we know that
the limit of |an | is , we know that as n approaches innity, the terms
will alternate between values close to and , meaning the sequence
diverges. A plot of this sequence is given in Figure . .

Figure . : A plot of a sequence in Example


, part .

We con nue our study of the limits of sequences by considering some of the
proper es of these limits.

Notes:

Theorem

Proper es of the Limits of Sequences

Let {an } and {bn } be sequences such that lim an = L, lim bn = K, and
n
n
let c be a real number.
. lim (an bn ) = L K

. lim (an /bn ) = L/K, K =

. lim (an bn ) = L K

. lim c an = c L

Applying proper es of limits of sequences


Example
Let the following sequences, and their limits, be given:
{
}
n+
{an } =
, and lim an = ;
n
n
{(
)n }
{bn } =
+
, and lim bn = e; and
n
n
{
}
{cn } = n sin( /n) , and lim cn = .
n

Evaluate the following limits.


. lim (an + bn )
n

. lim (bn cn )
n

We will use Theorem

. lim (
n

an )

to answer each of these.

. Since lim an = and lim bn = e, we conclude that lim (an + bn ) =


n
n
n
+ e = e. So even though we are adding something to each term of the
sequence bn , we are adding something so small that the nal limit is the
same as before.
. Since lim bn = e and lim cn = , we conclude that lim (bn cn ) =
n
n
n
e = e.
. Since lim an = , we have lim
an =
= . It does not
n
n
ma er that we mul ply each term by
; the sequence s ll approaches
. (It just takes longer to get close to .)
There is more to learn about sequences than just their limits. We will also
study their range and the rela onships terms have with the terms that follow.
We start with some deni ons describing proper es of the range.

Notes:

Sequences

Chapter

Sequences and Series

Deni on

Bounded and Unbounded Sequences

A sequence {an } is said to be bounded if there exists real numbers m


and M such that m < an < M for all n in N.
A sequence {an } is said to be unbounded if it is not bounded.
A sequence {an } is said to be bounded above if there exists an M such
that an < M for all n in N; it is bounded below if there exists an m such
that m < an for all n in N.

It follows from this deni on that an unbounded sequence may be bounded


above or bounded below; a sequence that is both bounded above and below is
simply a bounded sequence.

an =

Determining boundedness of sequences


Example
Determine the boundedness of the following sequences.

. {an } =

/
/

{ }
n

. {an } = { n }

(a)

. The terms of this sequence are always posi ve but are decreasing, so we
have < an < for all n. Thus this sequence is bounded. Figure . (a)
illustrates this.

an =

. The terms of this sequence obviously grow without bound. However, it is


also true that these terms are all posi ve, meaning < an . Thus we can
say the sequence is unbounded, but also bounded below. Figure . (b)
illustrates this.

.
4

(b)
Figure . : A plot of {an } = { /n} and
{an } = { n } from Example
.

The previous example produces some interes ng concepts. First, we can


recognize that the sequence { /n} converges to . This says, informally, that
most of the terms of the sequence are really close to . This implies that
the sequence is bounded, using the following logic. First, most terms are near
, so we could nd some sort of bound on these terms (using Deni on , the
bound is ). That leaves a few terms that are not near (i.e., a nite number
of terms). A nite list of numbers is always bounded.
This logic implies that if a sequence converges, it must be bounded. This is
indeed true, as stated by the following theorem.
Notes:

Theorem

Sequences

Note: Keep in mind what Theorem


does not say. It does not say that
bounded sequences must converge, nor
does it say that if a sequence does not
converge, it is not bounded.

Convergent Sequences are Bounded

Let {an } be a convergent sequence. Then {an } is bounded.


{
n}
In Example
we saw the sequence {bn } = ( + /n) , where it was
stated that lim bn = e. (Note that this is simply resta ng part of Theorem .)
n
Even though it may be dicult to intui vely grasp the behavior of this sequence,
we know immediately that it is bounded.
Another interes ng concept to come out of Example
again involves the
sequence { /n}. We stated, without proof, that the terms of the sequence were
decreasing. That is, that an+ < an for all n. (This is easy to show. Clearly
n < n + . Taking reciprocals ips the inequality: /n > /(n + ). This is the
same as an > an+ .) Sequences that either steadily increase or decrease are
important, so we give this property a name.
Deni on

Monotonic Sequences

. A sequence {an } is monotonically increasing if an an+ for all


n, i.e.,
a a a an an+
. A sequence {an } is monotonically decreasing if an an+ for all
n, i.e.,
a a a an an+
. A sequence is monotonic if it is monotonically increasing or monotonically decreasing.
Determining monotonicity
Example
Determine the monotonicity of the following sequences.
. {an } =

n+
n

. {an } =

n +
n+

}
}

n
n n+
{ }
n
. {an } =
n!
. {an } =

In each of the following, we will examine an+ an . If an+


S
an > , we conclude that an < an+ and hence the sequence is increasing. If

Notes:

Note: It is some mes useful to call


a monotonically increasing sequence
strictly increasing if an < an+ for all
n; i.e, we remove the possibility that
subsequent terms are equal.
A similar statement holds for strictly decreasing.

Chapter

Sequences and Series


an+ an < , we conclude that an > an+ and the sequence is decreasing. Of
course, a sequence need not be monotonic and perhaps neither of the above
will apply.
We also give a sca er plot of each sequence. These are useful as they suggest a pa ern of monotonicity, but analy c work should be done to conrm a
graphical trend.

y
n+
n

an =

n+
n+

n+
n
(n + )(n) (n + )
=
(n + )n

=
n(n + )
<
for all n.

an+ an =

(a)

Since an+ an < for all n, we conclude that the sequence is decreasing.

(n + ) +
n +

n+
n+
(
)
(n + ) + (n + ) (n + )(n + )
=
(n + )(n + )
n + n+
=
(n + )(n + )
>
for all n.

an+ an =

an =

n +
n+

(b)
y
5
an =

n 9
n n+ 6

Since an+ an >

(c)
Figure . : Plots of sequences in Example
.

for all n, we conclude the sequence is increasing.

. We can clearly see in Figure . (c), where the sequence is plo ed, that
it is not monotonic. However, it does seem that a er the rst terms
it is decreasing. To understand why, perform the same analysis as done
before:
n
(n + )

an+ an =
(n + ) (n + ) +
n n+
n + n
n
=

n n+
n n+
(n + n )(n n + ) (n )(n n + )
=
(n n + )(n n + )
n + n
.
=
(n n + )(n n + )
We want to know when this is greater than, or less than, . The denominator is always posi ve, therefore we are only concerned with the numerator. Using the quadra c formula, we can determine that n + n

Notes:

Sequences

= when n . , . . So for n < . , the sequence is decreasing. Since we are only dealing with the natural numbers, this means that
a >a .
Between . and . , i.e., for n = , and , we have that an+ >
an and the sequence is increasing. (That is, when n = , and , the
numerator n + n + from the frac on above is > .)

When n > . , i.e, for n , we have that n +


hence an+ an < , so the sequence is decreasing.

n+

< ,

In short, the sequence is simply not monotonic. However, it is useful to


note that for n , the sequence is monotonically decreasing.
. Again, the plot in Figure . shows that the sequence is not monotonic,
but it suggests that it is monotonically decreasing a er the rst term. We
perform the usual analysis to conrm this.
n
(n + )

(n + )!
n!
(n + ) n (n + )
=
(n + )!
n + n +
=
(n + )!

an+ an =

When n = , the above expression is > ; for n , the above expression is < . Thus this sequence is not monotonic, but it is monotonically
decreasing a er the rst term.
Knowing that a sequence is monotonic can be useful. In par cular, if we
know that a sequence is bounded and monotonic, we can conclude it converges!
Consider, for example, a sequence that is monotonically decreasing and is bounded
below. We know the sequence is always ge ng smaller, but that there is a
bound to how small it can become. This is enough to prove that the sequence
will converge, as stated in the following theorem.
Theorem

Bounded Monotonic Sequences are Convergent

. Let {an } be a bounded, monotonic sequence. Then {an } converges; i.e., lim an exists.
n

. Let {an } be a monotonically increasing sequence that is bounded


above. Then {an } converges.
. Let {an } be a monotonically decreasing sequence that is bounded
below. Then {an } converges.
Notes:

.5
an =

n
n!

.5

Figure . : A plot of {an } = {n /n!} in


Example
.

Chapter

Sequences and Series


Consider once again the sequence {an } = { /n}. It is easy to show it is
monotonically decreasing and that it is always posi ve (i.e., bounded below by
). Therefore we can conclude by Theorem that the sequence converges. We
already knew this by other means, but in the following sec on this theorem will
become very useful.
Sequences are a great source of mathema cal inquiry. The On-Line Encyclopedia of Integer Sequences (http://oeis.org) contains thousands of sequences and their formulae. (As of this wri ng, there are
,
sequences
in the database.) Perusing this database quickly demonstrates that a single sequence can represent several dierent real life phenomena.
Interes ng as this is, our interest actually lies elsewhere. We are more interested in the sum of a sequence. That is, given a sequence {an }, we are very
interested in a + a + a + . Of course, one might immediately counter with
Doesnt this just add up to innity? Many mes, yes, but there are many important cases where the answer is no. This is the topic of series, which we begin
to inves gate in the next sec on.

Notes:

Exercises .

Terms and Concepts

. {an } =

. Use your own words to dene a sequence.


. The domain of a sequence is the

numbers.

. Use your own words to describe the range of a sequence.


. Describe what it means for a sequence to be bounded.

Problems
In Exercises
quence.

, give the rst ve terms of the given sen

. {an } =

. {bn } =

{(

. {cn } =

nn+
n+

. {dn } =

(n + )!

In Exercises
quence.
.
.

, ,

, ,

)n }

)n

)n )}

, determine the nth term of the given se-

, ...

, , ...

, ,

, ...
, ...

In Exercises
, use the following informa on to determine the limit of the given sequences.
{ n
}

{an } =
;
lim an =
n
n

{bn } =

{(

)n }

{cn } = {sin( /n)};


. {an } =

. {an } =

n
n+

n n+
n +

. {an } =

. {an } =

n
n

. {an } =

. {an } =

( )n

n
n

)n }

. {an } =

{(

. {an } =

. {an } =

. {an } =

. n
n

. {an } =

n
n+

. {an } =

( )n

,n

}
}

}
}
n
n

In Exercises
, determine whether the sequence is
bounded, bounded above, bounded below, or none of the
above.

. {an } = {tan n}

)n }

( )n+
n

lim cn =

n
n +

. {an } =

. {an } = {sin n}

sin( /n)

In Exercises
, determine whether the sequence converges or diverges. If convergent, give the limit of the sequence.

lim bn = e

. {an } = { bn an }
{

) n}

. {an } = {ln(n)}

((

{(

. {an } =

( )n

. {an } =

n
n

. {an } = {n cos n}

n
n
}

. {an } = {

n!}

. {an } =

In Exercises
, determine whether the sequence is
monotonically increasing or decreasing. If it is not, determine
if there is an m such that it is monotonic for all n m.
. {an } =

n
n+

. {an } =

n n+
n

. {an } =

( )n

n
n

. Prove Theorem ; that is, use the deni on of the limit of a


sequence to show that if lim |an | = , then lim an = .
n

. Let {an } and {bn } be sequences such that lim an = L and


n

lim bn = K.

(a) Show that if an < bn for all n, then L K.

(b) Give an example where L = K.

. Prove the Squeeze Theorem for sequences: Let {an } and


{bn } be such that lim an = L and lim bn = L, and let
n

{cn } be such that an cn bn for all n. Then lim cn = L


n

Innite Series

Given the sequence {an } = { / n } = / , / , / , . . ., consider the following sums:


a
a +a
a +a +a
a +a +a +a

=
=
=
=

/
/ + /
/ + / + /
/ + / + / + /

=
=
=
=

/
/
/
/

In general, we can show that


a + a + a + + an =

Let Sn be the sum of the rst n terms of the sequence { / n }. From the above,
we see that S = / , S = / , etc. Our formula at the end shows that Sn =
/ n.
)
(
/ n = . This limit
Now consider the following limit: lim Sn = lim
n
n
can be interpreted as saying something amazing: the sum of all the terms of the
sequence { / n } is .
This example illustrates some interes ng concepts that we explore in this
sec on. We begin this explora on with some deni ons.
Innite Series, nth Par al Sums, Convergence, Divergence

Deni on

Let {an } be a sequence.


. The sum

an is an innite series (or, simply series).

n=

. Let Sn =

i=

ai ; the sequence {Sn } is the sequence of nth par al sums of {an }.

. If the sequence {Sn } converges to L, we say the series


and we write

an = L.

n=

n=

. If the sequence {Sn } diverges, the series

Notes:

n=

an diverges.

an converges to L,

Innite Series

Chapter

Sequences and Series

Using our new terminology, we can state that the series


and

converges,

n=

= .

n=

We will explore a variety of series in this sec on. We start with two series
that diverge, showing how we might discern divergence.
Showing series diverge

Example

. Let {an } = {n }. Show

an diverges.

n=

. Let {bn } = {( )n+ }. Show

bn diverges.

n=

. Consider Sn , the nth par al sum.


Sn = a + a + a + + an
=

By Theorem

+ n .

, this is
=

n(n + )( n + )

Since lim Sn = , we conclude that the series


n

instruc ve to write

n=

n diverges. It is

n=

n = for this tells us how the series diverges: it

grows without bound.

A sca er plot of the sequences {an } and {Sn } is given in Figure . (a).
The terms of {an } are growing, so the terms of the par al sums {Sn } are
growing even faster, illustra ng that the series diverges.
. The sequence {bn } starts with , , , , . . .. Consider some of the

Notes:

Innite Series

par al sums Sn of {bn }:


S =
S =

S =
S =
{

n is odd
. As {Sn } osciln is even
lates, repea ng , , , , . . ., we conclude that lim Sn does not exist,

This pa ern repeats; we nd that Sn =

hence

( )

n+

diverges.

n=

A sca er plot of the sequence {bn } and the par al sums {Sn } is given in
Figure . (b). When n is odd, bn = Sn so the marks for bn are drawn
oversized to show they coincide.
While it is important to recognize when a series diverges, we are generally
more interested in the series that converge. In this sec on we will demonstrate
a few general techniques for determining convergence; later sec ons will delve
deeper into this topic.

. an

y
1
0.5

One important type of series is a geometric series.


Geometric Series

A geometric series is a series of the form

n=

rn =

+ r + r + r + + rn +

Note that the index starts at n = , not n = .

We started this sec on with a geometric series, although we dropped the


rst term of . One reason geometric series are important is that they have nice
convergence proper es.

Notes:

Sn

(a)

Geometric Series

Deni on

10

0.5
1
.

. bn

Sn

(b)
Figure . : Sca er plots rela ng to Example
.

Chapter

Sequences and Series

Theorem

Convergence of Geometric Series

Consider the geometric series

rn .

n=

r n+
.
r

. The nth par al sum is: Sn =

. The series converges if, and only if, |r| < . When |r| < ,

rn =

n=

According to Theorem

n=

converges as r = / , and

, the series
( )

n=

= . This concurs with our intro /


ductory example; while there we got a sum of , we skipped the rst term of .
n=

Exploring geometric series


Example
Check the convergence of the following series. If the series converges, nd its
sum.
.

( )n

n=

)n
(

n=

n=

. an

. Since r = / < , this series converges. By Theorem


( )n

=
= .
/
n=

, we have that

However, note the subscript of the summa on in the given series: we are
to start with n = . Therefore we subtract o the rst two terms, giving:
( )n

= = .

Sn

Figure . : Sca er plots rela ng to the series in Example


.

n=

This is illustrated in Figure . .

Notes:

.
. Since |r| = / < , this series converges, and by Theorem
)n
(

= .
=
( / )
n=

The par al sums of this series are plo ed in Figure . (a). Note how the
par al sums are not purely increasing as some of the terms of the sequence {( / )n } are nega ve.

. Since r > , the series diverges. (This makes common sense; we expect
the sum
+ + + + +
+

Innite Series

8
.

. an

Sn

(a)

to diverge.) This is illustrated in Figure . (b).

pSeries
,

Another important type of series is the p-series.


Deni on

pSeries, General pSeries

. A pseries is a series of the form

n=

np

where p > .

. an

. A general pseries is a series of the form

n=

(an + b)p

where p >

and a, b are real numbers.

Sn

(b)
Figure . : Sca er plots rela ng to the series in Example
.

Like geometric series, one of the nice things about pseries is that they have
easy to determine convergence proper es.
Theorem

Convergence of General pSeries

A general pseries

n=

Notes:

(an + b)p

will converge if, and only if, p > .

Note: Theorem assumes that an+b =


for all n. If an + b = for some n, then
of course the series does not converge regardless of p as not all of the terms of the
sequence are dened.

Chapter

Sequences and Series

Determining convergence of series


Example
Determine the convergence of the following series.

n=

n=

n=

( )n
n
n=

n=

n=

( n )
n

. This is a pseries with p = . By Theorem

, this series diverges.

This series is a famous series, called the Harmonic Series, so named because of its rela onship to harmonics in the study of music and sound.
. This is a pseries with p = . By Theorem , it converges. Note that
the theorem does not give a formula by which we can determine what
the series converges to; we just know it converges. A famous, unexpected
result is that this series converges to / .
. This is a pseries with p = / ; the theorem states that it diverges.
. This is not a pseries; the deni on does not allow for alterna ng signs.
Therefore we cannot apply Theorem . (Another famous result states
that this series, the Alterna ng Harmonic Series, converges to ln .)
. This is a general pseries with p = , therefore it converges.
. This is not a pseries, but a geometric series with r = / . It converges.
Later sec ons will provide tests by which we can determine whether or not
a given series converges. This, in general, is much easier than determining what
a given series converges to. There are many cases, though, where the sum can
be determined.
Telescoping series
)
(

Evaluate the sum


.

n n+
n=
Example

Notes:

.
It will help to write down some of the rst few par al sums

of this series.
S =
S =
S =
S =

(
(
(

)
)
)

+
+
+

(
(
(

)
)
)

+
+

(
(

Note how most of the terms in each par al sum are canceled out! In general,
. The sequence {Sn } converges, as lim Sn =
we see that Sn =
n
n+
)
)
(
(

= , and so we conclude that


= . Par

lim
n
n+
n n+
n=
al sums of the series are plo ed in Figure . .
The series in Example
is an example of a telescoping series. Informally, a
telescoping series is one in which the par al sums reduce to just a nite number
of terms. The par al sum Sn did not contain n terms, but rather just two: and
/(n + ).
When possible, seek a way to write an explicit formula for the nth par al sum
Sn . This makes evalua ng the limit lim Sn much more approachable. We do so
n
in the next example.
Evalua ng series
Example
Evaluate each of the following innite series.
.

n=

n + n

n=

ln

n+
n

. We can decompose the frac on /(n + n) as


n + n

n+

(See Sec on . , Par al Frac on Decomposi on, to recall how this is done,
if necessary.)

Notes:

Innite Series

. an

Sn

Figure . : Sca er plots rela ng to the


series of Example
.

Chapter

Sequences and Series


Expressing the terms of {Sn } is now more instruc ve:
S =
S
S
S
S

(
=

(
=

(
=

(
=

)
)
)
)

+
+
+
+

(
(
(
(

)
)
)
)

+
+
+

(
(
(

)
)
)

+
+

(
(

)
)

We again have a telescoping series. In each par al sum, most of the terms

cancel and we obtain the formula Sn =

. Taking
n+
n+
limits allows us to determine the convergence of the series:
(
)

+
= , so
lim Sn = lim

= .
n
n
n+
n+
n + n
n=

. an

This is illustrated in Figure .

(a).

S = ln ( )
S = ln ( ) + ln

S = ln ( ) + ln

S = ln ( ) + ln

( )
( )

( )

+ ln
+ ln

( )
( )

+ ln

( )

At rst, this does not seem helpful, but recall the logarithmic iden ty:
ln x + ln y = ln(xy). Applying this to S gives:
( )
( )
( )
(
)
S = ln ( )+ln
+ln
+ln
= ln
= ln ( ) .

. an

. We begin by wri ng the rst few par al sums of the series:

Sn

(a)

Sn

}
ln(n + ) . This sequence does not con(
)

n+
ln
verge, as lim Sn = . Therefore
= ; the series din
n
n=
verges. Note in Figure . (b) how the sequence of par al sums grows
We can conclude that {Sn } =

(b)
Figure . : Sca er plots rela ng to the
series in Example
.

Notes:

.
slowly; a er
terms, it is not yet over . Graphically we may be fooled
into thinking the series converges, but our analysis above shows that it
does not.
We are learning about a new mathema cal object, the series. As done before, we apply old mathema cs to this new topic.
Theorem

Let

Proper es of Innite Series


an = L,

bn = K, and let c be a constant.

n=

n=

. Constant Mul ple Rule:

n=

. Sum/Dierence Rule:

(
n=

c an = c

n=

an = c L.

)
bn = L K.
an
an bn =
n=

n=

Before using this theorem, we provide a few famous series.


Key Idea
.

n!

n=

n=

Notes:

= e.

(Note that the index starts with n = .)


.

( )n+

=
.
n
n=

( )n
= .
n
+
n=

n=

Important Series

diverges.

(This is called the Harmonic Series.)

( )n+
= ln . (This is called the Alterna ng Harmonic Series.)
n
n=

Innite Series

Chapter

Sequences and Series

Evalua ng series
Example
Evaluate the given series.
(
)

( )n+ n n
.
n
n=
y

n=

n!

. We start by using algebra to break the series apart:

(
)
)
(

( )n+ n n
( )n+ n
( )n+ n
=

n
n
n
n=
n=

. an

Sn

( )n+
( )n+

n
n
n=
n=

= ln( )

(a)
y

This is illustrated in Figure .

(a).

. This looks very similar to the series that involves e in Key Idea . Note,
however, that the series given in this example starts with n = and not
n = . The rst term of the series in the Key Idea is / ! = , so we will
subtract this from our result below:

,5

,
5

.
5

. an

n=

Sn

n!

=
=

n=

n!

(e )

This is illustrated in Figure . (b). The graph shows how this par cular
series converges very rapidly.

(b)
Figure . : Sca er plots rela ng to the
series in Example
.

. The denominators in each term are perfect squares; we are adding

n
(note we start with n = , not n = ). This series will converge. Using the
n=

Notes:

.
formula from Key Idea

, we have the following:

n=

n=

n=

n
n
)

n=

=
=

n=

n=

n=

n=

n=

n
n
n
n

It may take a while before one is comfortable with this statement, whose
truth lies at the heart of the study of innite series: it is possible that the sum of
an innite list of nonzero numbers is nite. We have seen this repeatedly in this
sec on, yet it s ll may take some ge ng used to.
As one contemplates the behavior of series, a few facts become clear.
. In order to add an innite list of nonzero numbers and get a nite result,
most of those numbers must be very near .
. If a series diverges, it means that the sum of an innite list of numbers is
not nite (it may approach or it may oscillate), and:
(a) The series will s ll diverge if the rst term is removed.
(b) The series will s ll diverge if the rst
(c) The series will s ll diverge if the rst ,

terms are removed.


,

terms are removed.

(d) The series will s ll diverge if any nite number of terms from anywhere in the series are removed.
These concepts are very important and lie at the heart of the next two theorems.

Notes:

Innite Series

Chapter

Sequences and Series

nth Term Test for Convergence/Divergence

Theorem

Consider the series

an .

n=

. If

an converges, then lim an = .


n

n=

. If lim an = , then
n

an diverges.

n=

Note that the two statements in Theorem


are really the same. In order
to converge, the limit of the terms of the sequence must approach ; if they do
not, the series will not converge.
Looking back, we can apply this theorem to the series in Example
. In
that example, the nth terms of both sequences do not converge to , therefore
we can quickly conclude that each series diverges.

an
Important! This theorem does not state that if lim an = then
n

n=

converges. The standard example of this is the Harmonic Series, as given in Key
Idea . The Harmonic Sequence, { /n}, converges to ; the Harmonic Series,

/n, diverges.
n=

Theorem

Innite Nature of Series

The convergence or divergence remains unchanged by the addi on or


subtrac on of any nite number of terms. That is:
. A divergent series will remain divergent with the addi on or subtrac on of any nite number of terms.
. A convergent series will remain convergent with the addi on or
subtrac on of any nite number of terms. (Of course, the sum will
likely change.)

Consider once more the Harmonic Series

n=

Notes:

which diverges; that is, the

.
sequence of par al sums {Sn } grows (very, very slowly) without bound. One
might think that by removing the large terms of the sequence that perhaps
the series will converge. This is simply not the case. For instance, the sum of the
rst million terms of the Harmonic Series is about . . Removing the rst
million terms from the Harmonic Series changes the nth par al sums, eec vely
subtrac ng . from the sum. However, a sequence that is growing without
bound will s ll grow without bound when . is subtracted from it.
The equa ons below illustrate this. The rst line shows the innite sum of
the Harmonic Series split into the sum of the rst million terms plus the sum
of everything else. The next equa on shows us subtrac ng these rst million terms from both sides. The nal equa on employs a bit of psuedomath:
subtrac ng . from innity s ll leaves one with innity.

n=

n=

,
,

n=

,
,

=
n=

n=

n=

This sec on introduced us to series and dened a few special types of series
whose convergence proper es are well known: we know when a p-series or
a geometric series converges or diverges. Most series that we encounter are
not one of these types, but we are s ll interested in knowing whether or not
they converge. The next three sec ons introduce tests that help us determine
whether or not a given series converges.

Notes:

Innite Series

Exercises .

Terms and Concepts

In Exercises
diverges.

. Use your own words to describe how sequences and series


are related.

n=

. Use your own words to dene a par al sum.


. Given a series

n=

an , describe the two sequences related

n=

to the series that are important.

. T/F: If {an } is convergent, then

an is also convergent.

In Exercises

, a series

an is given.

(a) Give the rst par al sums of the series.


(b) Give a graph of the rst
same axes.

n=

n=

( )n
n

terms of an and Sn on the


.

n=

n=

)n

, state whether the given series converges

n
n

n=

n!

n=

n=

n=

n=

n=

n=

cos(n)

+
n+

n=

n=

n
+n

In Exercises
or diverges.

n=

n=

n=

n=

Problems

)n
)n

n=

n!

n=

to show the given series

n!

n=

n=

, use Theorem

n
n(n + )

n=

. Use your own words to explain what a geometric series is.

n=

n!

( n+ )

n=

In Exercises

n=

, a series is given.

(a) Find a formula for Sn , the nth par al sum of the series.
(b) Determine whether the series converges or diverges.
If it converges, state what it converges to.
.

.
.

n(n + )

n(n + )

n=

n=

n=

+
)

+
+

sin

)n

n
n+

n=

>

n=

(Compare each nth par al sum.)

(b) Show why


< +
n
n
n=
n=

(c) Explain why (a) and (b) demonstrate that the series
of odd terms is convergent, if, and only if, the series
of even terms is also convergent. (That is, show both
converge or both diverge.)

( n )( n + )
ln

(a) Show why

n=

The goal is to show that each of the series on the right diverge.

n=

. Break the Harmonic Series into the sum of the odd and even
terms:

=
+
.
n
n
n
n=
n=
n=

n=

n=

( )n n

n=

n+
n (n + )

n=

n=

(d) Explain why knowing the Harmonic Series is divergent determines that the even and odd series are also
divergent.
. Show the series

n=

n
diverges.
( n )( n + )

Chapter

Sequences and Series

Integral and Comparison Tests

Knowing whether or not a series converges is very important, especially when


we discuss Power Series in Sec on . . Theorems
and
give criteria for
when Geometric and p-series converge, and Theorem
gives a quick test to
determine if a series diverges. There are many important series whose convergence cannot be determined by these theorems, though, so we introduce a set
of tests that allow us to handle a broad range of series. We start with the Integral Test.

Integral Test

Note: Theorem
does not state that
the integral and the summa on have the
same value.

We stated in Sec on . that a sequence {an } is a func on a(n) whose domain is N, the set of natural numbers. If we can extend a(n) to R, the real numbers, and it is both posi ve and decreasing on [ , ), then the convergence of

a(x) dx.
an is the same as
n=

Theorem

Integral Test

Let a sequence {an } be dened by an = a(n), where a(n) is con nuous,

an converges, if, and only if,


posi ve and decreasing on [ , ). Then
n=

a(x) dx converges.

y = a(x)

(a)
y

We can demonstrate the truth of the Integral Test with two simple graphs.
In Figure . (a), the height of each rectangle is a(n) = an for n = , , . . .,
and clearly the rectangles enclose more area than the area under y = a(x).
Therefore we can conclude that

y = a(x)

a(x) dx <

an .

( . )

n=

In Figure . (b), we draw rectangles under y = a(x) with the Right-Hand rule,
star ng with n = . This me, the area of the rectangles is less than the area

a(x) dx. Note how this summa on starts


an <
under y = a(x), so
n=

with n = ; adding a to both sides lets us rewrite the summa on star ng with

(b)
Figure . : Illustra ng the truth of the
Integral Test.

Notes:

.
n= :

an < a +

n=

a(x) dx.

Integral and Comparison Tests

( . )

Combining Equa ons ( . ) and ( . ), we have

an < a +

n=

a(x) dx < a +

an .

( . )

n=

From Equa on ( . ) we can make the following two statements:


. If

an diverges, so does

n=

. If

an converges, so does

n=

a(x) dx

(because

an < a +

n=

a(x) dx

(because

a(x) dx <

a(x) dx)

an .)

n=

Therefore the series and integral either both converge or both diverge. Theorem
allows us to extend this theorem to series where a(n) is posi ve and
decreasing on [b, ) for some b > .
Using the Integral Test

ln n
. (The terms of the sequence {an } =
Determine the convergence of
n
n=

Example

th

{ln n/n } and the n par al sums are given in Figure .

Notes:

ln x
dx = lim
b
x

.4

ln x
dx
x
b

= lim ln x +
b
x

.6

.)

Figure . implies that a(n) = (ln n)/n is posi ve and deS


creasing on [ , ). We can determine this analy cally, too. We know a(n) is
posi ve as both ln n and n are posi ve on [ , ). To determine that a(n) is
decreasing, consider a (n) = ( ln n)/n , which is nega ve for n . Since
a (n) is nega ve, a(n) is decreasing.

ln x
Applying the Integral Test, we test the convergence of
dx. Integratx
ing this improper integral requires the use of Integra on by Parts, with u = ln x
and dv = /x dx.

y
.8

dx

. an

Sn

Figure . : Plo ng the sequence and


series in Example
.

Chapter

Sequences and Series


b

= lim ln x
b
x
x
ln b
. Apply LHpitals Rule:
= lim
b
b
b
= .
Since

ln n
ln x
dx converges, so does
.
x
n
n=

Theorem

was given without jus ca on, sta ng that the general p-series

converges if, and only if, p > . In the following example, we


(an + b)p
prove this to be true by applying the Integral Test.
n=

Using the Integral Test to establish Theorem .

Use the Integral Test to prove that


converges if, and only if, p > .
(an + b)p
n=
Example

Consider the integral

(ax + b)p

dx = lim

= lim

= lim

(ax + b)p

(ax + b)p

a( p)
a( p)

dx

(ax + b)
(

dx; assuming p = ,

(ac + b)

(a + b)

)
.

This limit converges if, and only if, p > . It is easy to show that the integral also
diverges in the case of p = . (This result is similar to the work preceding Key
Idea .)

converges if, and only if, p > .


Therefore
(an + b)p
n=
We consider two more convergence tests in this sec on, both comparison
tests. That is, we determine the convergence of one series by comparing it to
another series with known convergence.

Notes:

Integral and Comparison Tests

Direct Comparison Test


Theorem

Direct Comparison Test

Let {an } and {bn } be posi ve sequences where an bn for all n N,


for some N .
. If

bn converges, then

n=

an converges.

n=

n=

. If

an diverges, then

bn diverges.

n=

Applying the Direct Comparison Test

Determine the convergence of


.
n+n
n=
Example

This series is neither a geometric or p-series, but seems reS


lated. We predict it will converge, so we look for a series with larger terms that
converges. (Note too that the Integral Test seems dicult to apply here.)

for all n . The series


is a
Since n < n + n , n > n
n
+n
n=

convergent geometric series; by Theorem ,


converges.
n+n
n=
Example

Applying the Direct Comparison Test

Determine the convergence of


.
n ln n
n=

diverges, and it seems


n
that the given series is closely related to it, hence we predict it will diverge.
S

We know the Harmonic Series

n=

Since n n ln n for all n ,

n ln n

for all n .

The Harmonic Series diverges, so we conclude that


well.

Notes:

n=

n ln n

diverges as

Note: A sequence {an } is a posi ve


sequence if an > for all n.
Because of Theorem , any theorem that
relies on a posi ve sequence s ll holds
true when an > for all but a nite number of values of n.

Chapter

Sequences and Series


The concept of direct comparison is powerful and o en rela vely easy to
apply. Prac ce helps one develop the necessary intui on to quickly pick a proper
series with which to compare. However, it is easy to construct a series for which
it is dicult to apply the Direct Comparison Test.

Consider
. It is very similar to the divergent series given in Exn + ln n
n=

for large n. Hown


n + ln n
ever, the inequality that we naturally want to use goes the wrong way: since
ample

. We suspect that it also diverges, as

n n + ln n for all n ,
for all n . The given series has terms
n
n + ln n
less than the terms of a divergent series, and we cannot conclude anything from
this.
Fortunately, we can apply another test to the given series to determine its
convergence.

Limit Comparison Test

Theorem

Limit Comparison Test

Let {an } and {bn } be posi ve sequences.

an
an and
= L, where L is a posi ve real number, then
n bn
n=

bn either both converge or both diverge.

. If lim

n=

an
an .
bn converges, then so does
= , then if
n bn
n=
n=

. If lim

an
an .
bn diverges, then so does
= , then if
n bn
n=
n=

. If lim

Theorem is most useful when the convergence of the series from {bn } is
known and we are trying to determine the convergence of the series from {an }.
We use the Limit Comparison Test in the next example to examine the series

which mo vated this new test.


n + ln n
n=

Notes:

.
Example

Applying the Limit Comparison Test

Determine the convergence of


using the Limit Comparison Test.
n + ln n
n=
We compare the terms of

Harmonic Sequence

n=

n=

n + ln n

to the terms of the

/(n + ln n)
n
= lim
n n + ln n
/n
=
(a er applying LHpitals Rule).

lim

Since the Harmonic Series diverges, we conclude that

n=

well.

n + ln n

diverges as

Applying the Limit Comparison Test

Determine the convergence of


nn
n=
Example

This series is similar to the one in Example


, but now we
S
are considering n n instead of n + n . This dierence makes applying
the Direct Comparison Test dicult.

:
Instead, we use the Limit Comparison Test and compare with the series
n
n=

lim

/(

n )
n

= lim

We know
as well.

n=

n
(a er applying LHpitals Rule twice).

is a convergent geometric series, hence

n=

converges

As men oned before, prac ce helps one develop the intui on to quickly
choose a series with which to compare. A general rule of thumb is to pick a
series based on the dominant term in the expression of {an }. It is also helpful
to note that factorials dominate exponen als, which dominate algebraic funcons (e.g., polynomials), which dominate logarithms. In the previous example,

Notes:

Integral and Comparison Tests

Chapter

Sequences and Series

the dominant term of

was

. It is
n
n
n=
hard to apply the Limit Comparison Test to series containing factorials, though,
as we have not learned how to apply LHpitals Rule to n!.
n

, so we compared the series to

Applying the Limit Comparison Test

x+
.
Determine the convergence of
x

x+
n=
Example

We navely a empt to apply the rule of thumb given above


S
and note that the dominant term in the expression of the series is /x . Knowing

converges, we a empt to apply the Limit Comparison Test:


that
n
n=

( x + )/(x x + )
x ( x+ )
lim
= lim
n
n x x +
/x
= (Apply LHpitals Rule).
Theorem

part ( ) only applies when

bn diverges; in our case, it con-

n=

verges. Ul mately, our test has not revealed anything about the convergence
of our series.
The problem is that we chose a poor series with which to compare. Since
the numerator and denominator of the terms of the series are both algebraic
func ons, we should have compared our series to the dominant term of the
numerator divided by the dominant term of the denominator.
The dominant term of the numerator is x / and the dominant term of the
denominator is x . Thus we should compare the terms of the given series to
x / /x = /x / :

x / ( x+ )
( x + )/(x x + )
=
lim
lim
n x x +
n
/x /
(Apply LHpitals Rule).

x+
converges, we conclude that
x x+
n=
=

Since the p-series


verges as well.

n=

con-

We men oned earlier that the Integral Test did not work well with series
containing factorial terms. The next sec on introduces the Ra o Test, which
does handle such series well. We also introduce the Root Test, which is good for
series where each term is raised to a power.

Notes:

Exercises .

Terms and Concepts

. In order to apply the Integral Test to a sequence {an }, the


func on a(n) = an must be
,
and
.

n=

. T/F: The Integral Test can be used to determine the sum of


a convergent series.
. What test(s) in this sec on do not work well with factorials?
. Suppose

an is convergent, and there are sequences

n=

In Exercises , use the Integral Test to determine the convergence of the given series.
.

n=

n=

n=

n +

n(ln n)

n
n

n
n

n ln n

In Exercises
, use the Limit Comparison Test to determine the convergence of the given series; state what series is
used for comparison.
.

n + n

n=

n +n+
n
n=

n=

ln n
n
n=

n=

n=

n=

In Exercises , use the Direct Comparison Test to determine the convergence of the given series; state what series is
used for comparison.
.

n
n=

n
n +

n=

n ln n

n=

n=

n=

n! + n

n=

n=

Problems

n=

+n n

ln n
n
n=

n=

n=

{bn } and {cn } such that bn an cn for all n. What

cn ?
bn and
can be said about the series

n+

n=

n=

ln n
n

n=

n=

n n+

n=

n +n

n
n + n+
sin

/n

n+
n

n=

n+
n
+
n=

n=

n+

In Exercises , determine the convergence of the given


series. State the test used; more than one test may be appropriate.
.

n=

cos( /n)

n
n=

. Given that

an converges, state which of the following

n=

(b)

an
n
n=

an an+

(c)

(an )

n=

ln n
n!
n=

n=

n=

n=

n!

(a)

( n+ )

n=

n
n+
n=

series converges, may converge, or does not converge.

n
n=

+n

(d)

nan

n=

(e)

n=

an

Ra o and Root Tests

Ra o and Root Tests

The nth Term Test of Theorem

states that in order for a series

an to con-

n=

verge, lim an = . That is, the terms of {an } must get very small. Not only must
n

the terms approach , they must approach fast enough: while lim
the Harmonic Series

n=

/n = ,

diverges as the terms of { /n} do not approach

fast enough.
The comparison tests of the previous sec on determine convergence by comparing terms of a series to terms of another series whose convergence is known.
This sec on introduces the Ra o and Root Tests, which determine convergence
by analyzing the terms of a series to see if they approach fast enough.

Ra o Test

Theorem

Ra o Test

Let {an } be a posi ve sequence where lim

. If L < , then

an+
= L.
an

an converges.

n=

. If L >

or L = , then

Note: Theorem allows us to apply the


Ra o Test to series where {an } is posi ve
for all but a nite number of terms.

an diverges.

n=

. If L = , the Ra o Test is inconclusive.


an+
= L < , then for large n,
n an
each term of {an } is signicantly smaller than its previous term which is enough
to ensure convergence.
The principle of the Ra o Test is this: if lim

Applying the Ra o Test


Example
Use the Ra o Test to determine the convergence of the following series:
.

n=

Notes:

n!

n=

n=

n +

Chapter

Sequences and Series


S

n=

n!

lim

n+

n+

/(n + )!
= lim
n /n!
n
= lim

n (n

n!
+ )!

n+

= .
Since the limit is

< , by the Ra o Test

n=

n=

n!

converges.

lim

n+

n+

/(n + )
= lim
n /n
n
= lim

n (n

n
+ )

n
(n + )

= .
Since the limit is

> , by the Ra o Test

n=

n=

n +

diverges.

lim

(
)
/ (n + ) +
n +
= lim
n
/(n + )
(n + ) +
= .

Since the limit is , the Ra o Test is inconclusive. We can easily show this
series converges using the Direct or Limit Comparison Tests, with each

comparing to the series


.
n
n=

Notes:

.
The Ra o Test is not eec ve when the terms of a series only contain algebraic func ons (e.g., polynomials). It is most eec ve when the terms contain some factorials or exponen als. The previous example also reinforces our
developing intui on: factorials dominate exponen als, which dominate algebraic func ons, which dominate logarithmic func ons. In Part of the example,
the factorial in the denominator dominated the exponen al in the numerator,
causing the series to converge. In Part , the exponen al in the numerator dominated the algebraic func on in the denominator, causing the series to diverge.
While we have used factorials in previous sec ons, we have not explored
them closely and one is likely to not yet have a strong intui ve sense for how
they behave. The following example gives more prac ce with factorials.
Applying the Ra o Test

n!n!
Determine the convergence of
.
(
n)!
n=
Example

Before we begin, be sure to note the dierence between


S
( n)! and n!. When n = , the former is ! = . . . = ,
,
whereas the la er is ( ) = .
Applying the Ra o Test:

(
)
(n + )!(n + )!/ (n + ) !
(n + )!(n + )!( n)!
= lim
lim
n
n
n!n!/( n)!
n!n!( n + )!
No ng that ( n + )! = ( n + ) ( n + ) ( n)!, we have
= lim

(n + )(n + )
( n + )( n + )

= / .

Since the limit is / < , by the Ra o Test we conclude

Root Test

n!n!
converges.
(
n)!
n=

The nal test we introduce is the Root Test, which works par cularly well on
series where each term is raised to a power, and does not work well with terms
containing factorials.

Notes:

Ra o and Root Tests

Chapter

Sequences and Series

Theorem

Root Test

Let {an } be a posi ve sequence, and let lim (an )

/n

. If L < , then

Note: Theorem allows us to apply the


Root Test to series where {an } is posi ve
for all but a nite number of terms.

= L.

an converges.

n=

. If L >

or L = , then

an diverges.

n=

. If L = , the Root Test is inconclusive.


Example
Applying the Root Test
Determine the convergence of the following series using the Root Test:
.

)n
(

n+
n
n=

n=

n
(ln n)n

n=

)n ) /n
n+
n+
= lim
= .
n
n n
n
Since the limit is less than , we conclude the series converges. Note: it is
dicult to apply the Ra o Test to this series.
( /n )
(
) /n
n
n
. lim
= lim
.
n (ln n)n
n
ln n
As n grows, the numerator approaches (apply LHpitals Rule) and the
denominator grows to innity. Thus
( /n )
n
= .
lim
n
ln n
Since the limit is less than , we conclude the series converges.
( n ) /n
. lim
= lim (
) = .
n n
n n /n
Since this is greater than , we conclude the series diverges.
. lim

((

Each of the tests we have encountered so far has required that we analyze
series from posi ve sequences. The next sec on relaxes this restric on by considering alterna ng series, where the underlying sequence has terms that alternate between being posi ve and nega ve.

Notes:

Exercises .

Terms and Concepts


. The Ra o Test is not eec ve when the terms of a sequence
only contain
func ons.

In Exercises , determine the convergence of the given


series using the Root Test. If the Root Test is inconclusive,
state so and determine convergence with another test.
.

. The Ra o Test is most eec ve when the terms of a sequence contains


and/or
func ons.
. What three convergence tests do not work well with terms
containing factorials?
. The Root Test works par cularly well on series where each
term is
to a
.

n
.
n!
n=
n

.
n
.

n=

n=

n=

n=

n+
n

)n
(

n n
n +n
n=
(

ln n

n=

+n
+n

n +

n=

n=

n+

(
(

)n

)n

n
)n
ln n

In Exercises , determine the convergence of the given


series. State the test used; more than one test may be appropriate.

n=

n=

n=

n=

n
n

n=

n! n
.
( n)!
n=

nn

n=

n=

n=

n=

Problems
In Exercises , determine the convergence of the given
series using the Ra o Test. If the Ra o Test is inconclusive,
state so and determine convergence with another test.

)n

)n
(

. n n
.
n +n+
n=

n+
n
+
n=

( )n

n!

n n
n!
n=

n=

n
n

( n)

n
+n

nn

n=

n + n
n + n n+

n=

n
n + n+

n!n!n!
( n)!
n=

n=

ln n

)n
(

n+
n=

n+

n=

n=

n
)n
ln n
n

n+

Alterna ng Series and Absolute Convergence

Alterna ng Series and Absolute Convergence

All of the series convergence tests we have used require that the underlying
sequence {an } be a posi ve sequence. (We can relax this with Theorem and
state that there must be an N > such that an > for all n > N; that is, {an } is
posi ve for all but a nite number of values of n.)
In this sec on we explore series whose summa on includes nega ve terms.
We start with a very specic form of series, where the terms of the summa on
alternate between being posi ve and nega ve.
Deni on

Alterna ng Series

Let {an } be a posi ve sequence. An alterna ng series is a series of either


the form

( )n+ an .
( )n an
or
n=

n=

Recall the terms of Harmonic Series come from the Harmonic Sequence {an } =
{ /n}. An important alterna ng series is the Alterna ng Harmonic Series:

( )n+

n=

Geometric Series can also be alterna ng series when r < . For instance, if
r = / , the geometric series is
)n
(

n=

Theorem
states that geometric series converge when |r| <

rn =
. When r = / as above, we nd
the sum:
r
n=
)n
(

n=

( / )

and gives

A powerful convergence theorem exists for other alterna ng series that meet
a few condi ons.

Notes:

Chapter

Sequences and Series

Theorem

Alterna ng Series Test

Let {an } be a posi ve, decreasing sequence where lim an = . Then


n

and

converge.

( )n+ an

n=

n=

The basic idea behind Theorem


is illustrated in Figure .
decreasing sequence {an } is shown along with the par al sums

L
.

( )n an

. an

Sn =

Sn

Figure . : Illustra ng convergence with


the Alterna ng Series Test.

i=

. A posi ve,

( )i+ ai = a a + a a + + ( )n+ an .

Because {an } is decreasing, the amount by which Sn bounces up/down decreases.


Moreover, the odd terms of Sn form a decreasing, bounded sequence, while the
even terms of Sn form an increasing, bounded sequence. Since bounded, monotonic sequences converge (see Theorem ) and the terms of {an } approach ,
one can show the odd and even terms of Sn converge to the same common limit
L, the sum of the series.
Applying the Alterna ng Series Test
Example
Determine if the Alterna ng Series Test applies to each of the following series.
.

n=

( )n+

n=

( )n

ln n
n

n=

( )n+

| sin n|
n

. This is the Alterna ng Harmonic Series as seen previously. The underlying


sequence is {an } = { /n}, which is posi ve, decreasing, and approaches
as n . Therefore we can apply the Alterna ng Series Test and
conclude this series converges.
While the test does not state what the series converges to, we will see

( )n+ = ln .
later that
n
n=

. The underlying sequence is {an } = {ln n/n}. This is posi ve and approaches as n (use LHpitals Rule). However, the sequence is not
decreasing for all n. It is straigh orward to compute a = , a .
,

Notes:

.
a .
, and a .
rst terms.

Alterna ng Series and Absolute Convergence

: the sequence is increasing for at least the

We do not immediately conclude that we cannot apply the Alterna ng


Series Test. Rather, consider the longterm behavior of {an }. Trea ng
an = a(n) as a con nuous func on of n dened on [ , ), we can take
its deriva ve:
ln n
a (n) =
.
n
The deriva ve is nega ve for all n (actually, for all n > e), meaning a(n) = an is decreasing on [ , ). We can apply the Alterna ng
Series Test to the series when we start with n = and conclude that

ln n
( )n
converges; adding the terms with n = and n = do not
n
n=
change the convergence (i.e., we apply Theorem ).
The important lesson here is that as before, if a series fails to meet the
criteria of the Alterna ng Series Test on only a nite number of terms, we
can s ll apply the test.
. The underlying sequence is {an } = | sin n|/n. This sequence is posi ve
and approaches as n . However, it is not a decreasing sequence;
the value of | sin n| oscillates between and as n . We cannot
remove a nite number of terms to make {an } decreasing, therefore we
cannot apply the Alterna ng Series Test.
Keep in mind that this does not mean we conclude the series diverges;
in fact, it does converge. We are just unable to conclude this based on
Theorem .
Key Idea

n=

gives the sum of some important series. Two of these are


=

and

( )n+

=
.
n
n=

These two series converge to their sums at dierent rates. To be accurate to


two places a er the decimal, we need
terms of the rst series though only
of the second. To get places of accuracy, we need
terms of the rst
series though only of the second. Why is it that the second series converges
so much faster than the rst?
While there are many factors involved when studying rates of convergence,
the alterna ng structure of an alterna ng series gives us a powerful tool when
approxima ng the sum of a convergent series.

Notes:

Chapter

Sequences and Series

Theorem

The Alterna ng Series Approxima on Theorem

Let {an } be a sequence that sa ses the hypotheses of the Alterna ng


Series Test, and let Sn and L be the nth par al sums and sum, respec vely,

( )n+ an . Then
( )n an or
of either
n=

n=

. |Sn L| < an+ , and

. L is between Sn and Sn+ .


Part of Theorem
states that the nth par al sum of a convergent alterna ng series will be within an+ of its total sum. Consider the alterna ng se

( )n+
ries we looked at before the statement of the theorem,
. Since
n
n=
a = /
.
, we know that S is within .
of the total sum.
Moreover, Part of the theorem states that since S .
and S
.
, we know the sum L lies between .
and .
. One use of this is
the knowledge that S is accurate to two places a er the decimal.
Some alterna ng series converge slowly. In Example
we determined the

ln
n
( )n+
series
converged. With n =
, we nd ln n/n .
,
n
n=
meaning that S
.
is accurate to one, maybe two, places a er the
decimal. Since S
.
, we know the sum L is .
L .
.
Approxima ng the sum of convergent alterna ng series
Example
Approximate the sum of the following series, accurate to within .
.
.

n=

( )n+

n=

( )n+

ln n
.
n

. Using Theorem

, we want to nd n where /n < .


.
n
n

n
n .

Notes:

Alterna ng Series and Absolute Convergence

Let L be the sum of this series. By Part of the theorem, |S L| < a =


/
. We can compute S = .
, which our theorem states is
within .
of the total sum.
We can use Part of the theorem to obtain an even more accurate result.
As we know the th term of the series is /
, we can easily compute
S = .
. Part of the theorem states that L is between S and S ,
so .
<L< .
.
. We want to nd n where ln(n)/n < .
. We start by solving ln(n)/n =
.
for n. This cannot be solved algebraically, so we will use Newtons
Method to approximate a solu on.
Let f(x) = ln(x)/x .
; we want to know where f(x) = . We make a
guess that x must be large, so our ini al guess will be x =
. Recall
how Newtons Method works: given an approximate solu on xn , our next
approxima on xn+ is given by
xn+ = xn
We nd f (x) =

f(xn )
.
f (xn )

)
ln(x) /x . This gives

x =

ln(
(

)/
ln(

.
)
) /

Using a computer, we nd that Newtons Method seems to converge to a


solu on x =
. a er itera ons. Taking the next integer higher, we
have n =
, where ln(
)/
= .
< .
.
Again using a computer, we nd S
= .
. Part of the theorem states that this is within .
of the actual sum L. Already knowing
th
the ,
term, we can compute S
= .
, meaning .
<
L< .
.
No ce how the rst series converged quite quickly, where we needed only
terms to reach the desired accuracy, whereas the second series took over ,
terms.
One of the famous results of mathema cs is that the Harmonic Series,
diverges, yet the Alterna ng Harmonic Series,

n=

Notes:

n=

( )n+

, converges. The

Chapter

Sequences and Series


no on that alterna ng the signs of the terms in a series can make a series converge leads us to the following deni ons.
Deni on

Note: In Deni on

an is not nec-

. A series

n=

essarily an alterna ng series; it just may


have some nega ve terms.

Absolute and Condi onal Convergence

an converges absolutely if

n=

n=

. A series

n=

an converges condi onally if

|an | converges.

an converges but

n=

n=

|an | diverges.

Thus we say the Alterna ng Harmonic Series converges condi onally.


Determining absolute and condi onal convergence.
Example
Determine if the following series converge absolutely, condi onally, or diverge.
.

n=

n+
( )
n + n+
n

( )

nn

n=

+ n+
n

n=

( )n

n
n

. We can show the series



( )n

n=


n+
n+
=

n + n+
n + n+
n=

diverges using the Limit Comparison Test, comparing with /n.

n+
( )n
The series
converges using the Alterna ng Series
n
+
n+
n=
Test; we conclude it converges condi onally.

. We can show the series



( )n n + n +

n
n=

converges using the Ra o Test.

Notes:


n + n+
=

n
n=

Therefore we conclude

( )n

n + n+
n

n=

Alterna ng Series and Absolute Convergence

converges absolutely.

. The series


( )n n

n
n=


n
=

n

n=

diverges using the nth Term Test, so it does not converge absolutely.

n
fails the condi ons of the Alterna ng Series
n

n=
Test as ( n )/( n ) does not approach as n . We can state
further that this series diverges; as n , the series eec vely adds and
subtracts / over and over. This causes the sequence of par al sums to
oscillate and not converge.
The series

( )n

Therefore the series

n
n

( )n

n=

diverges.

Knowing that a series converges absolutely allows us to make two important statements, given in the following theorem. The rst is that absolute convergence is stronger than regular convergence. That is, just because
converges, we cannot conclude that

n=

converges absolutely tells us that

n=

an

n=

|an | will converge, but knowing a series

an will converge.

One reason this is important is that our convergence tests all require that the
underlying sequence of terms be posi ve. By taking the absolute value of the
terms of a series where not all terms are posi ve, we are o en able to apply an
appropriate test and determine absolute convergence. This, in turn, determines
that the series we are given also converges.
The second statement relates to rearrangements of series. When dealing
with a nite set of numbers, the sum of the numbers does not depend on the
order which they are added. (So + + = + + .) One may be surprised to
nd out that when dealing with an innite set of numbers, the same statement
does not always hold true: some innite lists of numbers may be rearranged in
dierent orders to achieve dierent sums. The theorem states that the terms of
an absolutely convergent series can be rearranged in any way without aec ng
the sum.

Notes:

Chapter

Sequences and Series

Theorem
Let

Absolute Convergence Theorem

an be a series that converges absolutely.

n=

an converges.

n=

. Let {bn } be any rearrangement of the sequence {an }. Then

bn =

an .

n=

n=

In Example
, we determined the series in part converges absolutely.
Theorem tells us the series converges (which we could also determine using
the Alterna ng Series Test).
The theorem states that rearranging the terms of an absolutely convergent
series does not aect its sum. This implies that perhaps the sum of a condi onally convergent series can change based on the arrangement of terms. Indeed,
it can. The Riemann Rearrangement Theorem (named a er Bernhard Riemann)
states that any condi onally convergent series can have its terms rearranged so
that the sum is any desired value, including !

As an example, consider the Alterna ng Harmonic Series once more. We


have stated that

( )n+

n=

(see Key Idea

or Example

= ln ,

).

Consider the rearrangement where every posi ve term is followed by two


nega ve terms:

(Convince yourself that these are exactly the same numbers as appear in the
Alterna ng Harmonic Series, just in a dierent order.) Now group some terms

Notes:

.
and simplify:
(
)

+ =

+ =
)
+ =

Alterna ng Series and Absolute Convergence

ln .

By rearranging the terms of the series, we have arrived at a dierent sum!


(One could try to argue that the Alterna ng Harmonic Series does not actually
converge to ln , because rearranging the terms of the series shouldnt change
the sum. However, the Alterna ng Series Test proves this series converges to
L, for some number L, and if the rearrangement does not change the sum, then
L = L/ , implying L = . But the Alterna ng Series Approxima on Theorem
quickly shows that L > . The only conclusion is that the rearrangement did
change the sum.) This is an incredible result.
We end here our study of tests to determine convergence. The back cover
of this text contains a table summarizing the tests that one may nd useful.
While series are worthy of study in and of themselves, our ul mate goal
within calculus is the study of Power Series, which we will consider in the next
sec on. We will use power series to create func ons where the output is the
result of an innite summa on.

Notes:

Exercises .

Terms and Concepts


. Why is

sin n not an alterna ng series?


.
n

( ) an converges when {an } is

and lim an =

. Give an example of a series where

n=

an converges but

|an | does not.

In Exercises

, an alterna ng series

(a) Determine if the series converges or diverges.


(b) Determine if

n=

(c) If

|an | converges or diverges.

n=

( )n+
n
n=

n=

( )n

n=

n+
n

n=

(e)n

( )n+

n+
n n+

( )n

n
ln n

( )n n
n!
n=

( )n

( )n

n
n=

)n
n!

Let Sn be the nth par al sum of a series. In Exercises , a


convergent alterna ng series is given and a value of n. Compute Sn and Sn+ and use these values to nd bounds on the
sum of the series.

( )n
.
ln n +
n=

)n

n=

n=

( )n+

n!
( )n

an converges, determine if the convergence is

condi onal or absolute.

n=

n=

an is given.

n=i

(
)

sin (n + / )
n ln n
n=

n=

Problems

( )
cos n

n=

n=

. The sum of a
convergent series can be changed by
rearranging the order of its terms.

n=

n=

( )n+
+ + + + ( n )

n=

n=

. A series

( )n
,
ln(n + )
n=

( )n+
,
n
n=

( )n
,
n!
n=
(

n=

)n

n=

n=

n=

n=

In Exercises
, a convergent alterna ng series is given
along with its sum and a value of . Use Theorem
to nd
n such that the nth par al sum of the series is within of the
sum of the series.
.

( )n+

=
,
n
n=

= .

( )n
= ,
n!
e
n=

( )n

= ,
n
+
n=

= .

= .

( )n
= cos ,
( n)!
n=

Chapter

Sequences and Series

Power Series

So far, our study of series has examined the ques on of Is the sum of these
innite terms nite?, i.e., Does the series converge? We now approach series
from a dierent perspec ve: as a func on. Given a value of x, we evaluate f(x)
by nding the sum of a par cular series that depends on x (assuming the series
converges). We start this new approach to series with a deni on.
Deni on

Power Series

Let {an } be a sequence, let x be a variable, and let c be a real number.


. The power series in x is the series

an xn = a + a x + a x + a x + . . .

n=

. The power series in x centered at c is the series

n=

an (x c)n = a + a (x c) + a (x c) + a (x c) + . . .

Examples of power series


Example
Write out the rst ve terms of the following power series:
.

n=

xn

( )n+

n=

(x + )n
n

n=

( )n+

(x ) n
.
( n)!

. One of the conven ons we adopt is that x = regardless of the value of


x. Therefore

xn = + x + x + x + x + . . .
n=

This is a geometric series in x.

. This series is centered at c = . Note how this series starts with n = .


We could rewrite this series star ng at n = with the understanding that

Notes:

.
a = , and hence the rst term is .

( )n+

n=

(x + )n
(x + ) (x + ) (x + ) (x + )
= (x+ )
+

+
...
n

. This series is centered at c = . Recall that ! = .

( )n+

n=

(x )
( n)!

= +

(x )

(x )

(x ) (x )

...
!
!

We introduced power series as a type of func on, where a value of x is given


and the sum of a series is returned. Of course, not every series converges. For

xn as a geometric
instance, in part of Example
, we recognized the series
n=

series in x. Theorem

states that this series converges only when |x| < .

This raises the ques on: For what values of x will a given power series converge?, which leads us to a theorem and deni on.
Theorem

Convergence of Power Series

Let a power series


true:

n=

an (x c)n be given. Then one of the following is

. The series converges only at x = c.


. There is an R > such that the series converges for all x in
(c R, c + R) and diverges for all x < c R and x > c + R.
. The series converges for all x.

The value of R is important when understanding a power series, hence it is


given a name in the following deni on. Also, note that part of Theorem
makes a statement about the interval (c R, c + R), but the not the endpoints
of that interval. A series may/may not converge at these endpoints.

Notes:

Power Series

Chapter

Sequences and Series

Deni on

Radius and Interval of Convergence

. The number R given in Theorem is the radius of convergence of


a given series. When a series converges for only x = c, we say the
radius of convergence is , i.e., R = . When a series converges
for all x, we say the series has an innite radius of convergence,
i.e., R = .
. The interval of convergence is the set of all values of x for which
the series converges.

To nd the values of x for which a given series converges, we will use the convergence tests we studied previously (especially the Ra o Test). However, the
tests all required that the terms of a series be posi ve. The following theorem
gives us a workaround to this problem.

Theorem
Convergence
The series

The Radius of Convergence of a Series and Absolute

n=

an (x c)n and

convergence R.



an (x c)n have the same radius of
n=

Theorem allows us to nd the radius of convergence R of a series by applying the Ra o Test (or any applicable test) to the absolute value of the terms
of the series. We prac ce this in the following example.
Determining the radius and interval of convergence.
Example
Find the radius and interval of convergence for each of the following series:

x
n=

Notes:

n!

n=

( )n+

xn
n

n=

(x )n

n=

n!xn

.
n

x
:
. We apply the Ra o Test to the series
n!
n=

n+

n+
x /(n + )!
x


= lim n
lim
xn /n!
n
n
x

x
= lim
n n +
=

for all x.


n!

(n + )!



The Ra o Test shows us that regardless of the choice of x, the series converges. Therefore the radius of convergence is R = , and the interval of
convergence is (, ).

n

n

x

n+ x
:

. We apply the Ra o Test to the series
n
=
( )
n
n=
n=

n+
n+
x /(n + )
x
n


lim
= lim n
xn /n
n
n
x
n+
n
= lim |x|
n
n+
= |x|.

The Ra o Test states a series converges if the limit of |an+ /an | = L < .
We found the limit above to be |x|; therefore, the power series converges
when |x| < , or when x is in ( , ). Thus the radius of convergence is
R= .
To determine the interval of convergence, we need to check the endpoints
of ( , ). When x = , we have the opposite of the Harmonic Series:

( )n+

n=

( )n
=
n
n
n=
= .

The series diverges when x = .

( )n
, which is the Alterna ng
n
n=
Harmonic Series, which converges. Therefore the interval of convergence
is ( , ].

When x = , we have the series

Notes:

( )n+

Power Series

Chapter

Sequences and Series


n
(x )n :
. We apply the Ra o Test to the series
n=

lim

n+

n+

(x )


n (x )n

n+

= lim n
n

= lim (x
n


(x )n+
(x )n

) .



(x ) < =
According
to
the
Ra
o
Test,
the
series
converges
when


x < / . The series is centered at , and x must be within / of
in order for the series to converge. Therefore the radius of convergence
is R = / , and we know that the series converges absolutely for all x in
( / , + / ) = ( . , . ).

We check for convergence at the endpoints to nd the interval of convergence. When x = . , we have:

n=

( . )n =
=

( / )n

n=

( )n ,

n=

which diverges. A similar process shows that the series also diverges at
x = . . Therefore the interval of convergence is ( . , . ).
. We apply the Ra o Test to

n
n!x :
n=



(n + )!xn+


(n + )x


lim
=
lim
n!xn
n
n

= for all x, except x = .

The Ra o Test shows that the series diverges for all x except x = . Therefore the radius of convergence is R = .
We can use a power series to dene a func on:
f(x) =

a n xn

n=

where the domain of f is a subset of the interval of convergence of the power


series. One can apply calculus techniques to such func ons; in par cular, we
can nd deriva ves and an deriva ves.

Notes:

Theorem

Deriva ves and Indenite Integrals of Power Series


Func ons

Let f(x) =

n=

an (x c)n be a func on dened by a power series, with

radius of convergence R.
. f(x) is con nuous and dieren able on (c R, c + R).
. f (x) =

n=

an n (x c)n , with radius of convergence R.

f(x) dx = C +

an

n=

A few notes about Theorem

(x c)n+
, with radius of convergence R.
n+

. The theorem states that dieren a on and integra on do not change the
radius of convergence. It does not state anything about the interval of
convergence. They are not always the same.
. No ce how the summa on for f (x) starts with n = . This is because the
constant term a of f(x) goes to .
. Dieren a on and integra on are simply calculated termbyterm using
the Power Rules.
Example
Let f(x) =

n=

Deriva ves and indenite integrals of power series

x . Find f (x) and F(x) =


f(x) dx, along with their respec ve

intervals of convergence.
S

ing Theorem
. f (x) =

n=

In Example

We nd the deriva ve and indenite integral of f(x), follow-

nxn =

+ x + x + x + .

, we recognized that

n=

xn is a geometric series in x. We

know that such a geometric series converges when |x| < ; that is, the
interval of convergence is ( , ).

Notes:

Power Series

Chapter

Sequences and Series


To determine the interval of convergence of f (x), we consider the endpoints of ( , ):
f ( ) =
f ( ) =

+ + ,
+ + + + ,

which diverges.
which diverges.

Therefore, the interval of convergence of f (x) is ( , ).

xn+
x
x
. F(x) = f(x) dx = C +
=C+x+
+
+
n+
n=

To nd the interval of convergence of F(x), we again consider the endpoints of ( , ):


F( ) = C + / / + / +

The value of C is irrelevant; no ce that the rest of the series is an Alterna ng Series that whose terms converge to . By the Alterna ng Series
Test, this series converges. (In fact, we can recognize that the terms of the
series a er C are the opposite of the Alterna ng Harmonic Series. We can
thus say that F( ) = C ln .)
F( ) = C + + / + / + / +

No ce that this summa on is C + the Harmonic Series, which diverges.


Since F converges for x = and diverges for x = , the interval of
convergence of F(x) is [ , ).
The previous example showed how to take the deriva ve and indenite integral of a power series without mo va on for why we care about such operaons. We may care for the sheer mathema cal enjoyment that we can, which
is mo va on enough for many. However, we would be remiss to not recognize
that we can learn a great deal from taking deriva ves and indenite integrals.
Recall that f(x) =
Theorem

xn in Example

is a geometric series. According to

n=

, this series converges to /( x) when |x| < . Thus we can say


f(x) =

n=

xn =

on

( , ).

Integra ng the power series, (as done in Example


F(x) = C +

n=

Notes:

xn+
,
n+

,) we nd
( . )

.
while integra ng the func on f(x) = /( x) gives
F(x) = ln | x| + C .

( . )

Equa ng Equa ons ( . ) and ( . ), we have


F(x) = C +

xn+
n+
n=

= ln | x| + C .

Le ng x = , we have F( ) = C = C . This implies that we can drop the


constants and conclude

xn+
n+
n=

= ln | x|.

We established in Example
that the series on the le converges at x = ;
subs tu ng x = on both sides of the above equality gives

+ = ln .

On the le we have the opposite of the Alterna ng Harmonic Series; on the


right, we have ln . We conclude that

+ = ln .

Important: We stated in Key Idea


(in Sec on . ) that the Alterna ng Harmonic Series converges to ln , and referred to this fact again in Example
of
Sec on . . However, we never gave an argument for why this was the case.
The work above nally shows how we conclude that the Alterna ng Harmonic
Series converges to ln .
We use this type of analysis in the next example.
Example
Let f(x) =
ior of f(x).

x
n=

n!

Analyzing power series func ons

. Find f (x) and f(x) dx, and use these to analyze the behav-

We start by making two notes: rst, in Example


, we
S
found the interval of convergence of this power series is (, ). Second,
we will nd it useful later to have a few terms of the series wri en out:
n

x
n=

Notes:

n!

+x+

( . )

Power Series

Chapter

Sequences and Series


We now nd the deriva ve:
f (x) =
=

xn
n
n!
n=

n=

xn
=
(n )!

+x+

x
!

+ .

Since the series starts at n = and each term refers to (n ), we can re-index
the series star ng with n = :
=

x
n=

n!

= f(x).

We found the deriva ve of f(x) is f(x). The only func ons for which this is true
are of the form y = cex for some constant c. As f( ) = (see Equa on ( . )), c
must be . Therefore we conclude that
f(x) =

x
n=

for all x.

We can also nd f(x) dx:

f(x) dx = C +

n!

= ex

n=

=C+

n=

xn+
n!(n + )
xn+
(n + )!

We write out a few terms of this last series:


C+

n=

xn+
x
x
x
=C+x+
+
+
+
(n + )!

The integral of f(x) diers from f(x) only by a constant, again indica ng that
f(x) = ex .
Example
and the work following Example
established rela onships
between a power series func on and regular func ons that we have dealt
with in the past. In general, given a power series func on, it is dicult (if not

Notes:

.
impossible) to express the func on in terms of elementary func ons. We chose
examples where things worked out nicely.
In this sec ons last example, we show how to solve a simple dieren al
equa on with a power series.
Solving a dieren al equa on with a power series.
Example
Give the rst terms of the power series solu on to y = y, where y( ) = .
The dieren al equa on y = y describes a func on y =
f(x) where the deriva ve of y is twice y and y( ) = . This is a rather simple
dieren al equa on; with a bit of thought one should realize that if y = Ce x ,
then y = Ce x , and hence y = y. By le ng C = we sa sfy the ini al
condi on of y( ) = .
Lets ignore the fact that we already know the solu on and nd a power
series func on that sa ses the equa on. The solu on we seek will have the
form

an xn = a + a x + a x + a x +
f(x) =
S

n=

for unknown coecients an . We can nd f (x) using Theorem


f (x) =

n=

an n xn = a + a x + a x + a x .

Since f (x) = f(x), we have

a + a x + a x + a x =

a + a x + a x + a x +

= a + a x + a x + a x +

The coecients of like powers of x must be equal, so we nd that


a = a ,

a = a ,

a = a ,

a = a ,

etc.

The ini al condi on y( ) = f( ) = indicates that a = ; with this, we can


nd the values of the other coecients:
a =

and a = a a = ;

a = and a = a a = / = ;
a = and a = a a = /( ) = / ;
a = / and a = a a = /( ) = / .

Thus the rst


y = y is

terms of the power series solu on to the dieren al equa on


f(x) =

Notes:

+ x + x + x + x +

Power Series

Chapter

Sequences and Series


In Sec on . , as we study Taylor Series, we will learn how to recognize this series as describing y = e x .
Our last example illustrates that it can be dicult to recognize an elementary
func on by its power series expansion. It is far easier to start with a known funcon, expressed in terms of elementary func ons, and represent it as a power
series func on. One may wonder why we would bother doing so, as the la er
func on probably seems more complicated. In the next two sec ons, we show
both how to do this and why such a process can be benecial.

Notes:

Exercises .

Terms and Concepts

. We adopt the convenc on that x =


the value of x.

. If the radius of convergence of


dius of convergence of

, regardless of
.

n=

. If the radius of convergence of


dius of convergence of

an xn is , what is the ra-

an xn is , what is the ra-

n=

In Exercises , write out the sum of the rst terms of the


given power series.

.
n

n!

n=

n=

xn

xn

In Exercises

.
n

n!

nxn

( )n (x )n
n
n=

(x + )n
n!
n=

(x )n

( )n n!(x

)n

xn
n
n=

(x + )n
n
n=

n!

( x )n

x+

)n
, a func on f(x) =

an xn is given.

n=

(a) Give a power series for f (x) and its interval of convergence.

(b) Give a power series for f(x) dx and its interval of convergence.

( )n+ n
x
n!
n=

n=

xn

In Exercises

(b) Find the interval of convergence.

n=

, a power series is given.

(a) Find the radius of convergence.

n=

( )n
.
x
( n)!
n=

nxn

n=

n n

(x )n

( )n xn

n=

n=

n=

n=

Problems

n=

n=

( )n an xn ?

n=

n=

n an x

( )n (x )n
n=

n=

. What is the dierence between the radius of convergence


and the interval of convergence?

xn

nxn

n=

xn
n
n=
( )

x n
n=

( x)n

n=

( )n x
.
( n)!
n=

. y = y,

y( ) =

. y = y,

y( ) =

( )n xn
n!
n=

In Exercises , give the rst terms of the series that is


a solu on to the given dieren al equa on.

. y = y , y( ) =
. y = y + , y( ) =
. y = y,

y( ) = , y ( ) =

. y = y, y( ) = , y ( ) =

Taylor Polynomials

Taylor Polynomials

(
)
Consider a func on y = f(x) and a point c, f(c) . The deriva ve, f (c), gives
the instantaneous
rate of change of f at x = c. Of all lines that pass through the
(
)
point c, f(c) , the line that best approximates f at this point is the tangent line;
that is, the line whose slope (rate of change) is f (c).
In Figure . , we see a func on y = f(x) graphed. The table below the
graph shows that f( ) = and f ( ) = ; therefore, the tangent line to f at
x = is p (x) = (x ) + = x + . The tangent line is also given in the gure.
Note that near x = , p (x) f(x); that is, the tangent line approximates f
well.
One shortcoming of this approxima on is that the tangent line only matches
the slope of f; it does not, for instance, match the concavity of f. We can nd a
polynomial, p (x), that does match the concavity without much diculty, though.
The table in Figure . gives the following informa on:
f( ) =

f ( ) =

f ( ) = .

5
y = f(x)

f( ) =
f ( ) =
f ( ) =

f ( ) =
f ( )( ) =
f ( )( ) =

Figure . : Plo ng y = f(x) and a table


of deriva ves of f evaluated at .

Therefore, we want our polynomial p (x) to have these same proper es. That
is, we need
p ( )=
p ( ) =
p ( ) = .
This is simply an ini alvalue problem. We can solve this using the techniques rst described in Sec on . . To keep p (x) as simple as possible, well
assume that not only p ( ) = , but that p (x) = . That is, the second derivave of p is constant.
If p (x) = , then p (x) = x + C for some constant C. Since we have
determined that p ( ) = , we nd that C = and so p (x) = x + . Finally,
we can compute p (x) = x + x + C. Using our ini al values, we know p ( ) =
so C = . We conclude that p (x) = x + x + . This func on is plo ed with f in
Figure . .
We can repeat this approxima on process by crea ng polynomials of higher
degree that match more of the deriva ves of f at x = . In general, a polynomial
of degree n can be created to match the rst n deriva ves of f. Figure . also
shows p (x) = x / x / +x +x+ , whose rst four deriva ves at match
( )
those of f. (Using the table in Figure . , start with p (x) = and solve
the related ini alvalue problem.)
As we use more and more deriva ves, our polynomial approxima on to f
gets be er and be er. In this example, the interval on which the approxima on
is good gets bigger and bigger. Figure . shows p (x); we can visually arm
that this polynomial approximates f very well on [ , ]. (The polynomial p (x)
is not par cularly nice. It is
x

5
y = p2 (x)

y = p4 (x)

Figure .

: Plo ng f, p and p .

+x +x+ .)

y = p (x)

.
.
Notes:

4
2
y = p1 (x)

Figure .

: Plo ng f and p .

Chapter

Sequences and Series

The polynomials we have created are examples of Taylor polynomials, named


a er the Bri sh mathema cian Brook Taylor who made important discoveries
about such func ons. While we created the above Taylor polynomials by solving
ini alvalue problems, it can be shown that Taylor polynomials follow a general
pa ern that make their forma on much more direct. This is described in the
following deni on.
Deni on

Taylor Polynomial, Maclaurin Polynomial

Let f be a func on whose rst n deriva ves exist at x = c.


. The Taylor polynomial of degree n of f at x = c is
pn (x) = f(c)+f (c)(xc)+

f (c)
f (n) (c)
f (c)
(xc) +
(xc) + +
(xc)n .
!
!
n!

. A special case of the Taylor polynomial is the Maclaurin polynomial, where c =


. That is, the Maclaurin polynomial of degree n of f is
pn (x) = f( ) + f ( )x +

f ( )
f ( )
f (n) ( ) n
x +
x + +
x .
!
!
n!

We will prac ce crea ng Taylor and Maclaurin polynomials in the following


examples.
f(x) = ex
f (x) = ex
f (x) = ex
..
.
f (n) (x) = ex

Example

f( ) =
f ( ) =
f ( ) =
..
.
f (n) ( ) =

Finding and using Maclaurin polynomials

. Find the nth Maclaurin polynomial for f(x) = ex .


. Use p (x) to approximate the value of e.
x

Figure . : The deriva ves of f(x) = e


evaluated at x = .

. We start with crea ng a table of the deriva ves of ex evaluated at x = .


In this par cular case, this is rela vely simple, as shown in Figure . . By
the deni on of the Maclaurin series, we have

Notes:

Taylor Polynomials

f ( )
f ( )
f n( ) n
pn (x) = f( ) + f ( )x +
x +
x + +
x
!
!
n!
+x+ x + x +

x + +

n!

xn .
5

. Using our answer from part , we have


p =

+x+ x + x +

x +

x .

y = p5 (x)

To approximate the value of e, note that e = e = f( ) p ( ). It is very


straigh orward to evaluate p ( ):
p ( )=

+ +

A plot of f(x) = ex and p (x) is given in Figure .

Figure . : A plot of f(x) = ex and its


degree Maclaurin polynomial p (x).

th

Finding and using Taylor polynomials

Example

. Find the nth Taylor polynomial of y = ln x at x = .


. Use p (x) to approximate the value of ln . .
. Use p (x) to approximate the value of ln .
S

. We begin by crea ng a table of deriva ves of ln x evaluated at x = .


While this is not as straigh orward as it was in the previous example, a
pa ern does emerge, as shown in Figure . .
Using Deni on , we have
f (c)
f n (c)
f (c)
(x c) +
(x c) + +
(x c)n
!
!
n!
( )n+
(x ) + (x ) (x ) + +
(x )n .
n

pn (x) = f(c) + f (c)(x c) +


=

+ (x )

Note how the coecients of the (x ) terms turn out to be nice.

. We can compute p (x) using our work above:

p (x) = (x ) (x ) + (x ) (x ) + (x ) (x ) .

Notes:

f(x) = ln x
f (x) = /x
f (x) = /x
f (x) = /x
f ( ) (x) = /x
..
.
f (n) (x) =
( )n+ (n )!
xn

f( ) =
f ( ) =
f ( ) =
f ( ) =
f ( )( ) =
..
.
f (n) ( ) =
( )n+ (n )!

Figure . : Deriva ves of ln x evaluated


at x = .

Chapter

Sequences and Series


Since p (x) approximates ln x well near x = , we approximate ln .
p ( . ):

y
y = ln x

p ( . ) = ( . ) ( . ) + ( . ) ( . ) +

+ ( . ) ( . )

y = p6 (x)

..

This is a good approxima on as a calculator shows that ln . .


Figure . plots y = ln x with y = p (x). We can see that ln .
p ( . ).

th

Figure . : A plot of y = ln x and its


degree Taylor polynomial at x = .

.
.

. We approximate ln with p ( ):
p ( ) = ( ) ( ) + ( ) ( ) +
+ ( ) ( )
y

=
y = ln x

This approxima on is not terribly impressive: a hand held calculator shows


that ln .
. The graph in Figure . shows that p (x) provides
less accurate approxima ons of ln x as x gets close to or .

y = p (x)

Surprisingly enough, even the th degree Taylor polynomial fails to approximate ln x for x > , as shown in Figure . . Well soon discuss why
this is.

=
x

Figure . : A plot of y = ln x and its


degree Taylor polynomial at x = .

th

Taylor polynomials are used to approximate func ons f(x) in mainly two situa ons:
. When f(x) is known, but perhaps hard to compute directly. For instance,
we can dene y = cos x as either the ra o of sides of a right triangle
(adjacent over hypotenuse) or with the unit circle. However, neither of
these provides a convenient way of compu ng cos . A Taylor polynomial
of suciently high degree can provide a reasonable method of compu ng
such values using only opera ons usually hardwired into a computer (+,
, and ).

Notes:

Taylor Polynomials

. When f(x) is not known, but informa on about its deriva ves is known.
This occurs more o en than one might think, especially in the study of
dieren al equa ons.
In both situa ons, a cri cal piece of informa on to have is How good is my
approxima on? If we use a Taylor polynomial to compute cos , how do we
know how accurate the approxima on is?
We had the same problem when studying Numerical Integra on. Theorem
provided bounds on the error when using, say, Simpsons Rule to approximate
a denite integral. These bounds allowed us to determine that, for instance,
using subintervals provided an approxima on within . of the exact value.
The following theorem gives similar bounds for Taylor (and hence Maclaurin)
polynomials.
Theorem

Note: Even though Taylor polynomials


could be used in calculators and computers to calculate values of trigonometric func ons, in prac ce they generally
arent. Other more ecient and accurate
methods have been developed, such as
the CORDIC algorithm.

Taylors Theorem

. Let f be a func on whose n + th deriva ve exists on an interval I and let c be in I.


Then, for each x in I, there exists zx between x and c such that
f(x) = f(c) + f (c)(x c) +

f (n) (c)
f (c)
(x c) + +
(x c)n + Rn (x),
!
n!

f (n+ ) (zx )
(x c)(n+ ) .
(n + )!


max f (n+ ) (z)


(x c)(n+ )
. Rn (x)
(n + )!
where Rn (x) =

The rst part of Taylors Theorem states that f(x) = pn (x) + Rn (x), where
pn (x) is the nth order Taylor polynomial and Rn (x) is the remainder, or error, in
the Taylor approxima on. The second part gives bounds on how big that error
can be. If the (n + )th deriva ve is large, the error may be large; if x is far from
c, the error may also be large. However, the (n + )! term in the denominator
tends to ensure that the error gets smaller as n increases.
The following example computes error es mates for the approxima ons of
ln . and ln made in Example
.
Finding error bounds of a Taylor polynomial
Example
Use Theorem
to nd error bounds when approxima ng ln . and ln with
p (x), the Taylor polynomial of degree of f(x) = ln x at x = , as calculated in
Example
.

Notes:

Chapter

Sequences and Series


S

. We start with the approxima on of ln . with p ( . ). The theorem references an open interval I that contains both x and c. The smaller the
interval we use the be er; it will give us a more accurate (and smaller!)
approxima on of the error. We let I = ( . , . ), as this interval contains
both c = and x = . .


The theorem references max f (n+ ) (z) . In our situa on, this is asking
How big can the th deriva ve of y = ln x be on the interval ( . , . )?
The seventh deriva ve is y = !/x . The largest value it a ains on I is
about
. Thus we can bound the error as:
()





R ( . ) max f (z) ( . )
!

We computed p ( . ) = .
; using a calculator, we nd ln .
.
, so the actual error is about .
, which is less than our
bound of .
. This arms Taylors Theorem; the theorem states that
our approxima on would be within about thousandths of the actual
value, whereas the approxima on was actually closer.
. We again nd an interval I that contains both c = and x = ; we choose
I = ( . , . ). The maximum value of the seventh deriva ve of f on this
interval is again about
(as the largest values come near x = . ).
Thus
()





R ( ) max f (z) ( )
!

This bound is not as nearly as good as before. Using the degree Taylor
polynomial at x = will bring us within . of the correct answer. As
p ( ) .
, our error es mate guarantees that the actual value of
ln is somewhere between .
and .
. These bounds are not
par cularly useful.
In reality, our approxima on was only o by about . . However, we
are approxima ng ostensibly because we do not know the real answer. In
order to be assured that we have a good approxima on, we would have
to resort to using a polynomial of higher degree.

Notes:

Taylor Polynomials

We prac ce again. This me, we use Taylors theorem to nd n that guarantees our approxima on is within a certain amount.
Finding suciently accurate Taylor polynomials
Example
Find n such that the nth Taylor polynomial of f(x) = cos x at x = approximates
cos to within .
of the actual answer. What is pn ( )?
Following Taylors theorem, we need bounds on the size of
S
the deriva ves of f(x) = cos x. In the case of this trigonometric func on, this is
easy. All deriva ves of cosine are sin x or cos x. In all cases, these func ons
are never greater than in absolute value. We want the error to be less than
.
. To nd the appropriate n, consider the following inequali es:



max f (n+ ) (z)
( )(n+ ) .
(n + )!
(n + )!

(n+ )

We nd an n that sa ses this last inequality with trialanderror. When n = ,


we have
.

; when n = , we have

( + )!
( + )!
. Thus we want to approximate cos with p ( ).

<

We now set out to compute p (x). We again need a table of the deriva ves
of f(x) = cos x evaluated at x = . A table of these values is given in Figure . .
No ce how the deriva ves, evaluated at x = , follow a certain pa ern. All the
odd powers of x in the Taylor polynomial will disappear as their coecient is .
While our error bounds state that we need p (x), our work shows that this will
be the same as p (x).
Since we are forming our polynomial at x = , we are crea ng a Maclaurin
polynomial, and:
p (x) = f( ) + f ( )x +
=

x +

f(x) = cos x
f (x) = sin x
f (x) = cos x
f (x) = sin x
f ( ) (x) = cos x
f ( ) (x) = sin x
f ( ) (x) = cos x
f ( ) (x) = sin x
f ( ) (x) = cos x
f ( ) (x) = sin x

x +

f( ) =
f ( ) =
f ( ) =
f ( ) =
f ( )( ) =
f ( )( ) =
f ( )( ) =
f ( )( ) =
f ( )( ) =
f ( )( ) =

Figure . : A table of the deriva ves of


f(x) = cos x evaluated at x = .

f ( )
f( )
f ( )
x +
x + +
x
!
!
!

y = p8 (x)

x
x

We nally approximate cos :


cos p ( ) =

Our error bound guarantee that this approxima on is within .


of the correct
answer. Technology shows us that our approxima on is actually within about
.
of the correct answer.
Figure . shows a graph of y = p (x) and y = cos x. Note how well the
two func ons agree on about (, ).
Notes:

f(x) = cos x

Figure . : A graph of f(x) = cos x and


its degree Maclaurin polynomial.

Chapter

Sequences and Series


Example

f(x) =

f (x) =

f (x) =
x/
f (x) =

x/

f ( ) (x) =
x

Finding and using Taylor polynomials

. Find the degree Taylor polynomial, p (x), for f(x) = x at x = .

.
. Use p (x) to approximate

. Find bounds on the error when approxima ng


with p ( ).

f( ) =

f ( ) =

f ( ) =

f ( )=

f ( )( ) =

. We begin by evalua ng the deriva ves of f at x = . This is done in Figure


. . These values allow us to form the Taylor polynomial p (x):
p (x) = + (x )+

Figure . : A table of the deriva ves of


f(x) = x evaluated at x = .

. As p (x)

/
!

/
!

x near x = , we approximate

(x ) +

/
!

with p ( ) = .

(x ) .
.

. To nd a bound on the error, we need an open interval that contains x =


and x = . We set I = ( . , . ). The largest value the h deriva ve of
f(x) = x takes on this interval is near x = . , at about .
. Thus



R ( ) .

.
.

(x ) +



( ) .

This shows our approxima on is accurate to at least the rst places a er


the decimal. (It turns out that our approxima on
is actually accurate to
places a er the decimal.) A graph of f(x) = x and p (x) is given in
Figure . . Note how the two func ons are nearly indis nguishable on
( , ).

y= x
y = p (x)
x

Figure . : A graph of f(x) = x and its


degree Taylor polynomial at x = .

Our nal example gives a brief introduc on to using Taylor polynomials to


solve dieren al equa ons.
Approxima ng an unknown func on
Example
A func on y = f(x) is unknown save for the following two facts.
. y( ) = f( ) = , and
. y = y
(This second fact says that amazingly, the deriva ve of the func on is actually
the func on squared!)
Find the degree Maclaurin polynomial p (x) of y = f(x).

Notes:

Taylor Polynomials

One might ini ally think that not enough informa on is given
S
to nd p (x). However, note how the second fact above actually lets us know
what y ( ) is:
y = y y ( ) = y ( ).

Since y( ) = , we conclude that y ( ) = .


Now we nd informa on about y . Star ng with y = y , take deriva ves of
both sides, with respect to x. That means we must use implicit dieren a on.
y = y
d( )
d ( )
y =
y
dx
dx
y = y y .
Now evaluate both sides at x = :
y ( ) = y( ) y ( )
y ( ) =

We repeat this once more to nd y ( ). We again use implicit dieren a on;


this me the Product Rule is also required.
d ( )
d ( )
y =
yy
dx
dx

y = y y + y y .
Now evaluate both sides at x = :
y ( ) = y ( ) + y( )y ( )
y ( ) = + =

In summary, we have:
y( ) =

y ( ) =

y ( ) =

y ( ) = .

We can now form p (x):

.
p (x) =
=

+x+

x +

!
!
+x+x +x .

. Figure .

x
shows this func on plo ed with p (x). Note how similar they are near x = .

Notes:

x
y = p (x)

It turns out that the dieren al equa on we started with, y = y , where


y( ) = , can be solved without too much diculty: y =

y=

x
.5

.5

Figure . : A graph of y = /(x )


and y = p (x) from Example
.

Chapter

Sequences and Series


It is beyond the scope of this text to pursue error analysis when using Taylor polynomials to approximate solu ons to dieren al equa ons. This topic is
o en broached in introductory Dieren al Equa ons courses and usually covered in depth in Numerical Analysis courses. Such an analysis is very important;
one needs to know how good their approxima on is. We explored this example
simply to demonstrate the usefulness of Taylor polynomials.
Most of this chapter has been devoted to the study of innite series. This
sec on has taken a step back from this study, focusing instead on nite summaon of terms. In the next sec on, we explore Taylor Series, where we represent
a func on with an innite series.

Notes:

Exercises .

Terms and Concepts

. f(x) = x cos x,

. What is the dierence between a Taylor polynomial and a


Maclaurin polynomial?
. T/F: In general, pn (x) approximates f(x) be er and be er
as n gets larger.
. For some func on f(x), the Maclaurin polynomial of degree
is p (x) = + x x + x x . What is p (x)?
. For some func on f(x), the Maclaurin polynomial of degree
is p (x) = + x x + x x . What is f ( )?

Problems
In Exercises , nd the Maclaurin polynomial of degree
n for the given func on.
. f(x) = e ,

n=

. f(x) = sin x,

n=

. f(x) = x ex ,

n=

. f(x) = tan x,

n=

. f(x) =

n=

+x

n=

+x

n=

x,

n= ,

. f(x) = ln(x + ),

n= ,

c=

c = /

. f(x) = sin x,

n= ,

c = /

. f(x) =
. f(x) =

n= ,

,
,

x +

n= ,
,

with the Maclaurin polynomial of de-

with the Taylor polynomial of degree


. Approximate
centered at x = .
. Approximate ln . with the Taylor polynomial of degree
centered at x = .
Exercises ask for an n to be found such that pn (x) approximates f(x) within a certain bound of accuracy.

c=

, nd the nth term of the indicated Taylor

. Find a formula for the nth term of the Maclaurin polynomial


for f(x) = cos x.
. Find a formula for the nth term of the Maclaurin polynomial
for f(x) =

. Find a formula for the nth term of the Maclaurin polynomial


for f(x) =

c=

n= ,

In Exercises
polynomial.

. Find a formula for the nth term of the Maclaurin polynomial


for f(x) = ex .

c=

n= ,

. Approximate cos
gree .

. Find n such that the Maclaurin polynomial of degree n of


f(x) = sin x approximates cos within .
of the actual
value.

. f(x) = cos x,

. f(x) =

. Approximate sin . with the Maclaurin polynomial of degree .

. Find n such that the Maclaurin polynomial of degree n of


f(x) = cos x approximates cos / within .
of the actual value.

In Exercises
, nd the Taylor polynomial of degree n,
at x = c, for the given func on.
. f(x) =

In Exercises , approximate the func on value with the


indicated Taylor polynomial and give approximate bounds on
the error.

.
Find n such that the Taylor polynomial of
degree n of f(x) =
x, centered at x = , approximates
within .
of
the actual value.

n=

. f(x) =

c=

. Find n such that the Maclaurin polynomial of degree n of


f(x) = ex approximates e within .
of the actual value.

. f(x) = e x ,

. f(x) =

n= ,

c=

+x

. Find a formula for the nth term of the Taylor polynomial for
f(x) = ln x.

In Exercises
, approximate the solu on to the given
dieren al equa on with a degree Maclaurin polynomial.
. y = y,

y( ) =

. y = y,

y( ) =

. y =

y( ) =

Taylor Series

Taylor Series

In Sec on . , we showed how certain func ons can be represented by a power


series func on. In . , we showed how we can approximate func ons with polynomials, given that enough deriva ve informa on is available. In this sec on we
combine these concepts: if a func on f(x) is innitely dieren able, we show
how to represent it with a power series func on.
Deni on

Taylor and Maclaurin Series

Let f(x) have deriva ves of all orders at x = c.


. The Taylor Series of f(x), centered at c is

f (n) (c)

n!

n=

. Se ng c =

(x c)n .

gives the Maclaurin Series of f(x):

f (n) ( ) n
x .
n!
n=

The dierence between a Taylor polynomial and a Taylor series is the former
is a polynomial, containing only a nite number of terms, whereas the la er is
a series, a summa on of an innite set of terms. When crea ng the Taylor polynomial of degree n for a func on f(x) at x = c, we needed to evaluate f, and the
rst n deriva ves of f, at x = c. When crea ng the Taylor series of f, it helps to
nd a pa ern that describes the nth deriva ve of f at x = c. We demonstrate
this in the next two examples.
The Maclaurin series of f(x) = cos x
Example
Find the Maclaurin series of f(x) = cos x.
In Example
we found the th degree Maclaurin polynomial of cos x. In doing so, we created the table shown in Figure . . No ce how
f (n) ( ) = when n is odd, f (n) ( ) = when n is divisible by , and f (n) ( ) =
when n is even but not divisible by . Thus the Maclaurin series of cos x is
S

Notes:

x
!

x
!

x
!

f(x) = cos x
f (x) = sin x
f (x) = cos x
f (x) = sin x
f ( ) (x) = cos x
f ( ) (x) = sin x
f ( ) (x) = cos x
f ( ) (x) = sin x
f ( ) (x) = cos x
f ( ) (x) = sin x

f( ) =
f ( ) =
f ( ) =
f ( ) =
f ( )( ) =
f ( )( ) =
f ( )( ) =
f ( )( ) =
f ( )( ) =
f ( )( ) =

Figure . : A table of the deriva ves of


f(x) = cos x evaluated at x = .

Chapter

Sequences and Series


We can go further and write this as a summa on. Since we only need the terms
where the power of x is even, we write the power series in terms of x n :

n=

( )n

( n)!

Example
The Taylor series of f(x) = ln x at x =
Find the Taylor series of f(x) = ln x centered at x = .
f(x) = ln x
f (x) = /x
f (x) = /x
f (x) = /x
f ( ) (x) = /x
f ( ) (x) = /x
..
.
f (n) (x) =
( )n+ (n )!
xn

f( ) =
f ( ) =
f ( ) =
f ( ) =
f ( )( ) =
f ( )( ) =
..
.
f (n) ( ) =
( )

n+

(n )!

Figure . : Deriva ves of ln x evaluated


at x = .

Figure . shows the nth deriva ve of ln x evaluated at x =


for n = , . . . , , along with an expression for the nth term:
S

f (n) ( ) = ( )n+ (n )!

for n .

Remember that this is what dis nguishes Taylor series from Taylor polynomials;
we are very interested in nding a pa ern for the nth term, not just nding a
nite set of coecients for a polynomial. Since f( ) = ln = , we skip the
rst term and start the summa on with n = , giving the Taylor series for ln x,
centered at x = , as

n=

( )n+ (n )!

n!

(x )n =

n=

( )n+

(x )n
.
n

It is important to note that Deni on denes a Taylor series given a funcon f(x); however, we cannot yet state that f(x) is equal to its Taylor series. We
will nd that most of the me they are equal, but we need to consider the
condi ons that allow us to conclude this.
Theorem states that the error between a func on f(x) and its nth degree
Taylor polynomial pn (x) is Rn (x), where
(n+ )




(z)
Rn (x) max f
(x c)(n+ ) .
(n + )!
If Rn (x) goes to for each x in an interval I as n approaches innity, we conclude that the func on is equal to its Taylor series expansion.

Notes:

Theorem

Func on and Taylor Series Equality

Let f(x) have deriva ves of all orders at x = c, let Rn (x) be as stated in
Theorem , and let I be an interval on which the Taylor series of f(x)
converges. If lim Rn (x) = for all x in I, then
n

f(x) =

f (n) (c)
n=

n!

(x c)n on I.

We demonstrate the use of this theorem in an example.


Example
Establishing equality of a func on and its Taylor series
Show that f(x) = cos x is equal to its Maclaurin series, as found in Example
for all x.
S

bounded by

Given a value x, the magnitude of the error term Rn (x) is




max f (n+ ) (z) n+

x .
Rn (x)
(n + )!

Since all deriva ves of cos x are sin x or cos x, whose magnitudes are bounded
by , we can state


n+
Rn (x)
x
(n + )!
which implies

|xn+ |
|xn+ |
Rn (x)
.
(n + )!
(n + )!

( . )

xn+
= . Applying the Squeeze Theorem to Equa on ( . ),
n (n + )!
we conclude that lim Rn (x) = for all x, and hence

For any x, lim

cos x =

n=

( )n

( n)!

for all x.

It is natural to assume that a func on is equal to its Taylor series on the


series interval of convergence, but this is not the case. In order to properly
establish equality, one must use Theorem . This is a bit disappoin ng, as we
developed beau ful techniques for determining the interval of convergence of
a power series, and proving that Rn (x) can be cumbersome as it deals with
high order deriva ves of the func on.

Notes:

Taylor Series

Chapter

Sequences and Series


There is good news. A func on f(x) that is equal to its Taylor series, centered
at any point the domain of f(x), is said to be an analy c func on, and most, if
not all, func ons that we encounter within this course are analy c func ons.
Generally speaking, any func on that one creates with elementary func ons
(polynomials, exponen als, trigonometric func ons, etc.) that is not piecewise
dened is probably analy c. For most func ons, we assume the func on is equal
to its Taylor series on the series interval of convergence and only use Theorem
when we suspect something may not work as expected.
We develop the Taylor series for one more important func on, then give a
table of the Taylor series for a number of common func ons.
The Binomial Series
Example
Find the Maclaurin series of f(x) = ( + x)k , k = .
When k is a posi ve integer, the Maclaurin series is nite.
S
For instance, when k = , we have
f(x) = ( + x) =

+ x+ x + x +x .

The coecients of x when k is a posi ve integer are known as the binomial coecients, giving the series we are developing
its name.

+ x. Knowing a
series representa on of
When k = / , we have f(x) =
this func on would give a useful way of approxima ng
. , for instance.
k
To develop the Maclaurin series for f(x) = ( + x) for any value of k = ,
we consider the deriva ves of f evaluated at x = :
f(x) = ( + x)k
f (x) = k( + x)

f( ) =
k

f (x) = k(k )( + x)k

f ( ) = k
f ( ) = k(k )

f (x) = k(k )(k )( + x)k

f ( ) = k(k )(k )

(
)
f (n) (x) = k(k ) k (n ) ( + x)kn

(
)
f (n) ( ) = k(k ) k (n )

..
.

Thus the Maclaurin series for f(x) = ( + x)k is

..
.

(
)
k(k ) k (n )
k(k ) k(k )(k )
+k+
+
+ ... +
+ ...
!
!
n!
It is important to determine the interval of convergence of this series. With
(
)
k(k ) k (n ) n
x ,
an =
n!

Notes:

.
we apply the Ra o Test:

k(k ) (k n) n+
|an+ |
lim
= lim
x
n |an |
n
(n + )!


k n
= lim
x
n
n

(
)
/

k(k

(n

)



xn




n!

= |x|.

The series converges absolutely when the limit of the Ra o Test is less than
; therefore, we have absolute convergence when |x| < .
While outside the scope of this text, the interval of convergence depends
on the value of k. When k > , the interval of convergence is [ , ]. When
< k < , the interval of convergence is [ , ). If k , the interval of
convergence is ( , ).
We learned that Taylor polynomials oer a way of approxima ng a dicult
to compute func on with a polynomial. Taylor series oer a way of exactly
represen ng a func on with a series. One probably can see the use of a good
approxima on; is there any use of represen ng a func on exactly as a series?
While we should not overlook the mathema cal beauty of Taylor series (which
is reason enough to study them), there are prac cal uses as well. They provide
a valuable tool for solving a variety of problems, including problems rela ng to
integra on and dieren al equa ons.
In Key Idea
(on the following page) we give a table of the Taylor series
of a number of common func ons. We then give a theorem about the algebra
of power series, that is, how we can combine power series to create power
series of new func ons. This allows us to nd the Taylor series of func ons like
f(x) = ex cos x by knowing the Taylor series of ex and cos x.
Before we inves gate combining func ons, consider the Taylor series for the
arctangent func on (see Key Idea ). Knowing that tan ( ) = / , we can
use this series to approximate the value of :

= tan ( ) = + +
)
(
=
+ +

Unfortunately, this par cular expansion of converges very slowly. The rst
terms approximate as .
, which is not par cularly good.

Notes:

Taylor Series

Chapter

Sequences and Series

Key Idea

Important Taylor Series Expansions

Func on and Series


ex =

sin x =

+x+

n!

n=

( )n

n=

cos x =

x n+
( n + )!

( )n

n=

ln x =

tan x =

xn

( )n

n=

a Convergence

!
x
!

+ kx +

x n+
n+

(, )

(, )

(x )

(x )

k(k )
x +
!
x

(, )

+ x + x + x +

(
)

k(k ) k (n ) n
x
n!
n=

(x )

n=

( + x)k =

( , ]

( , )
( , )a

[ , ]

at x = depends on the value of k.

Theorem
Let f(x) =

(x )n
n

n=

( n)!

( )n+

Interval of
Convergence

First Few Terms

Algebra of Power Series

an xn and g(x) =

n=

n=

. f(x) g(x) =
. f(x)g(x) =

n=

n=

(
)
. f h(x) =

n=

bn xn converge absolutely for |x| < R, and let h(x) be con nuous.

(an bn )xn

a n xn

)(

(
)n
an h(x)
Notes:

n=

for |x| < R.


b n xn

(
n=

for |h(x)| < R.

)
a bn + a bn + . . . an b xn for |x| < R.

.
Example
Combining Taylor series
Write out the rst terms of the Taylor Series for f(x) = ex cos x using Key Idea
and Theorem .
Key Idea

ex =

+x+

x
!

informs us that
+

and

cos x =

x
!

x
!

+ .

Applying Theorem , we nd that


(
)(
)
x
x
x
x
x
e cos x =
+x+
+
+

+
+ .
!
!
!
!
Distribute the right hand expression across the le :
(
)
(
)
)
x
x
x
x
x
x
x

+
+ + x

+
+ +
+
+

!
!
!
!
!
!
!
(
(
)
)
x
x
x
x
x
x
+

+
+ +
+
+ +
!
!
!
!
!
!

Distribute again and collect like terms.


=

+x

While this process is a bit tedious, it is much faster than evalua ng all the necessary deriva ves of ex cos x and compu ng the Taylor series directly.
Because the series for ex and cos x both converge on (, ), so does the
series expansion for ex cos x.
Example
Use Theorem

Crea ng new Taylor series

to create series for y = sin(x ) and y = ln( x).


Given that

sin x =

( )n

n=

x n+
x
x
x
=x
+

+ ,
( n + )!
!
!
!

we simply subs tute x for x in the series, giving


sin(x ) =

n=

Notes:

( )n

(x ) n+
x
x
x
=x
+

.
( n + )!
!
!
!

Taylor Series

Chapter

Sequences and Series


Since the Taylor series for sin x has an innite radius of convergence, so does the
Taylor series for sin(x ).
The Taylor expansion forln x given in Key Idea is centered at x = , so we
will center the series for ln( x) at x = as well. With

Note: In Example
, one could create

a series forln( x) by simply recognizing that ln( x) = ln(x / ) = / ln x,


and hence mul plying the Taylor series
for ln x by / . This example was chosen to demonstrate other aspects of series, such as the fact that the interval of
convergence changes.

ln x =

( )n+

n=

(x )n
(x )
(x )
= (x )
+
,
n

x for x to obtain

( x )
( x )
n+ ( x )
( )
= ( x )
+
.
ln( x) =
n
n=

we subs tute

While this isnot strictly a power series, it is a series that allows us to study the
func
on ln( x). Since the interval of convergence of ln x is ( , ], and the range
of x on ( , ] is ( , ], the interval of convergence of this series expansion of
ln( x) is ( , ].
Example

Using Taylor series to evaluate denite integrals


ex dx.

Use the Taylor series of ex to evaluate

We learned, when studying Numerical Integra on, that ex


does not have an an deriva ve expressible in terms of elementary func ons.
This means any denite integral of this func on must have its value approximated, and not computed exactly.
We can quickly write out the Taylor series for ex using the Taylor series of
x
e:
S

ex =

x
n=

n!

+x+

x
!

and so
ex =
=

(x )n
n!
n=

( )n

n=

Notes:

x +

xn
n!

x
!

x
!

+ .

.
We use Theorem to integrate:

x
x
x
x n+
ex dx = C + x
+

+ + ( )n
+
!
!
( n + )n!
This is the an deriva ve of ex ; while we can write it out as a series, we cannot write it out in terms of elementary func ons. We can evaluate the denite
integral

ex dx using this an deriva ve; subs tu ng and for x and sub-

trac ng gives

ex dx =

Summing the terms shown above give the approxima on of .


. Since
this is an alterna ng series, we can use the Alterna ng Series Approxima on
Theorem, (Theorem ), to determine how accurate this approxima on is. The
next term of the series is /( !) .
. Thus we know our approxima on is within .
of the actual value of the integral. This is arguably
much less work than using Simpsons Rule to approximate the value of the integral.
Example
Using Taylor series to solve dieren al equa ons
Solve the dieren al equa on y = y in terms of a power series, and use the
theory of Taylor series to recognize the solu on in terms of an elementary funcon.
We found the rst
S
this dieren al equa on in Example

terms of the power series solu on to


in Sec on . . These are:

= , a =
= .


We include the unsimplied expressions for the coecients found in Example
as we are looking for a pa ern. It can be shown that an = n /n!. Thus the
solu on, wri en as a power series, is
a = ,

a = ,

a =

y=

n=

Using Key Idea

and Theorem
ex =

x
n=

Notes:

a =

= ,

n!

n!

xn =

( x)n
.
n!
n=

, we recognize f(x) = e x :

ex=

( x)n
.
n!
n=

Taylor Series

Chapter

Sequences and Series


Finding a pa ern in the coecients that match the series expansion of a
known func on, such as those shown in Key Idea , can be dicult. What if
the coecients in the previous example were given in their reduced form; how
could we s ll recover the func on y = e x ?
Suppose that all we know is that
a = ,

a = ,

a = ,

a =

a =

Deni on states that each term of the Taylor expansion of a func on includes
an n!. This allows us to say that
a =

b
!

a =

b
!

and

a =

b
!

for some values b , b and b . Solving for these values, we see that b = ,
b = and b = . That is, we are recovering the pa ern we had previously
seen, allowing us to write
f(x) =

an x =

bn
n=

n=

n!

xn

+ x+

x +

x +

x +

From here it is easier to recognize that the series is describing an exponen al


func on.
There are simpler, more direct ways of solving the dieren al equa on y =
y. We applied power series techniques to this equa on to demonstrate its u lity, and went on to show how some mes we are able to recover the solu on in
terms of elementary func ons using the theory of Taylor series. Most dierenal equa ons faced in real scien c and engineering situa ons are much more
complicated than this one, but power series can oer a valuable tool in nding,
or at least approxima ng, the solu on.
This chapter introduced sequences, which are ordered lists of numbers, followed by series, wherein we add up the terms of a sequence. We quickly saw
that such sums do not always add up to innity, but rather converge. We studied tests for convergence, then ended the chapter with a formal way of dening
func ons based on series. Such seriesdened func ons are a valuable tool
in solving a number of dierent problems throughout science and engineering.
Coming in the next chapters are new ways of dening curves in the plane
apart from using func ons of the form y = f(x). Curves created by these new
methods can be beau ful, useful, and important.

Notes:

Exercises .

Terms and Concepts


. What is the dierence between a Taylor polynomial and a
Taylor series?
. What theorem must we use to show that a func on is equal
to its Taylor series?

. f(x) = ln x
. f(x) = /( x) (show equality only on ( , ))
In Exercises , use the Taylor series given in Key Idea
to verify the given iden ty.
. cos(x) = cos x

Problems

. sin(x) = sin x

Key Idea gives the nth term of the Taylor series of common
func ons. In Exercises , verify the formula given in the
Key Idea by nding the rst few terms of the Taylor series of
the given func on and iden fying a pa ern.

d
dx

d
dx

. f(x) = ex ;

c=

. f(x) = sin x;

(
(

)
sin x = cos x

)
cos x = sin x

In Exercises , write out the rst terms of the Binomial


series with the given k-value.
. k= /

c=

. f(x) = /( x); c =

. k= /

. f(x) = tan x; c =

. k= /

In Exercises , nd a formula for the nth term of the Taylor


series of f(x), centered at c, by nding the coecients of the
rst few powers of x and looking for a pa ern. (The formulas for several of these are found in Key Idea ; show work
verifying these formula.)
. f(x) = cos x;

c = /

. f(x) = /x;

c=

. f(x) = ex ;

c=

( )
. f(x) = cos x
. f(x) = ex

. f(x) = x/(x + );

c=

c = /

In Exercises
, show that the Taylor series for f(x), as
given in Key Idea , is equal to f(x) by applying Theorem ;
that is, show lim Rn (x) = .

x+

( )
x/

. f(x) = ex sin x

. f(x) = ( + x)

(only nd the rst terms)

cos x

(only nd the rst terms)

In Exercises
, approximate the value of the given definite integral by using the rst nonzero terms of the integrands Taylor series.
.

. f(x) = ex
. f(x) = sin x

. f(x) = tan
c=

In Exercises , use the Taylor series given in Key Idea


to create the Taylor series of the given func ons.

. f(x) = sin

. f(x) = ln( + x);

. f(x) = sin x;

. k=

( )
sin x dx
cos

( )
x dx

:C

We have explored func ons of the form y = f(x) closely throughout this text.
We have explored their limits, their deriva ves and their an deriva ves; we
have learned to iden fy key features of their graphs, such as rela ve maxima
and minima, inec on points and asymptotes; we have found equa ons of their
tangent lines, the areas between por ons of their graphs and the x-axis, and the
volumes of solids generated by revolving por ons of their graphs about a horizontal or ver cal axis.
Despite all this, the graphs created by func ons of the form y = f(x) are
limited. Since each x-value can correspond to only y-value, common shapes
like circles cannot be fully described by a func on in this form. Fi ngly, the
ver cal line test excludes ver cal lines from being func ons of x, even though
these lines are important in mathema cs.
In this chapter well explore new ways of drawing curves in the plane. Well
s ll work within the framework of func ons, as an input will s ll only correspond
to one output. However, our new techniques of drawing curves will render the
ver cal line test pointless, and allow us to create important and beau ful
new curves. Once these curves are dened, well apply the concepts of calculus
to them, con nuing to nd equa ons of tangent lines and the areas of enclosed
regions.

Conic Sec ons

The ancient Greeks recognized that interes ng shapes can be formed by intersec ng a plane with a double napped cone (i.e., two iden cal cones placed p
to p as shown in the following gures). As these shapes are formed as sec ons
of conics, they have earned the ocial name conic sec ons.
The three most interes ng conic sec ons are given in the top row of Figure
. . They are the parabola, the ellipse (which includes circles) and the hyperbola. In each of these cases, the plane does not intersect the ps of the cones
(usually taken to be the origin).

Parabola

Ellipse

Point

Circle

Line
Figure . : Conic Sec ons

Crossed Lines

Hyperbola

Chapter

Curves in the Plane


When the plane does contain the origin, three degenerate cones can be
formed as shown the bo om row of Figure . : a point, a line, and crossed
lines. We focus here on the nondegenerate cases.
While the above geometric constructs dene the conics in an intui ve, visual
way, these constructs are not very helpful when trying to analyze the shapes
algebraically or consider them as the graph of a func on. It can be shown that
all conics can be dened by the general seconddegree equa on
Ax + Bxy + Cy + Dx + Ey + F = .
While this algebraic deni on has its uses, most nd another geometric perspec ve of the conics more benecial.
Each nondegenerate conic can be dened as the locus, or set, of points that
sa sfy a certain distance property. These distance proper es can be used to
generate an algebraic formula, allowing us to study each conic as the graph of a
func on.

Focus
}

Vertex
Directrix

Deni on

Symmetry

Axis of

Parabolas

.}

(x, y)

Figure . : Illustra ng the deni on of


the parabola and establishing an algebraic formula.

Parabola

A parabola is the locus of all points equidistant from a point (called a


focus) and a line (called the directrix) that does not contain the focus.
Figure . illustrates this deni on. The point halfway between the focus
and the directrix is the vertex. The line through the focus, perpendicular to the
directrix, is the axis of symmetry, as the por on of the parabola on one side of
this line is the mirrorimage of the por on on the opposite side.
The deni on leads us to an algebraic formula for the parabola. Let P =
(x, y) be a point on a parabola whose focus is at F = ( , p) and whose directrix
is at y = p. (Well assume for now that the focus lies on the y-axis; by placing
the focus p units above the x-axis and the directrix p units below this axis, the
vertex will be at ( , ).)
We use the Distance Formula to nd the distance d between F and P:

d = (x ) + (y p) .
The distance d from P to the directrix is more straigh orward:
d = y (p) = y + p.

Notes:

.
These two distances are equal. Se ng d = d , we can solve for y in terms of x:

Now square both sides.

d =d

x + (y p) = y + p
x + (y p) = (y + p)

x + y yp + p = y + yp + p
x = yp
y=

x .

The geometric deni on of the parabola has led us to the familiar quadra c
func on whose graph is a parabola with vertex at the origin. When we allow the
vertex to not be at ( , ), we get the following standard form of the parabola.
Key Idea

General Equa on of a Parabola

. Ver cal Axis of Symmetry: The equa on of the parabola with vertex at (h, k) and directrix y = k p in standard form is
y=

(x h) + k.

The focus is at (h, k + p).


. Horizontal Axis of Symmetry: The equa on of the parabola with
vertex at (h, k) and directrix x = h p in standard form is
x=

(y k) + h.

The focus is at (h + p, k).


Note: p is not necessarily a posi ve number.
Finding the equa on of a parabola
Example
Give the equa on of the parabola with focus at ( , ) and directrix at y = .
The vertex is located halfway between the focus and direcS
trix, so (h, k) = ( , . ). This gives p = . . Using Key Idea
we have the
Notes:

Conic Sec ons

Chapter

Curves in the Plane


equa on of the parabola as

y=

Figure . : The parabola described in Example


.

The parabola is sketched in Figure . .

We need to put the equa on of the parabola in its general


S
form. This requires us to complete the square:
x=

=
10

10
10

10

(x ) + . = (x ) + . .

Finding the focus and directrix of a parabola


Example
Find the focus and directrix of the parabola x = y y + . The point ( , )
lies on the graph of this parabola; verify that it is equidistant from the focus and
directrix.

( . )

10

Figure . : The parabola described in Example


. The distances from a point on
the parabola to the focus and directrix is
given.

y y+
(
)
y y+
(
y y+

(
)
(y )

(y ) .

Hence the vertex is located at ( , ). We have = p , so p = . We conclude


that the focus is located at ( , ) and the directrix is x = . The parabola is
graphed in Figure . , along with its focus and directrix.
The point ( , ) lies on the graph and is ( ) =
directrix. The distance from ( , ) to the focus is:

( ) +( ) =
= .

units from the

Indeed, the point on the parabola is equidistant from the focus and directrix.

Reec ve Property

.
Figure . : Illustra ng the parabolas reec ve property.

One of the fascina ng things about the nondegenerate conic sec ons is their
reec ve proper es. Parabolas have the following reec ve property:
Any ray emana ng from the focus that intersects the parabola
reects o along a line perpendicular to the directrix.
This is illustrated in Figure . . The following theorem states this more rigorously.

Notes:

Theorem

Conic Sec ons

Reec ve Property of the Parabola

Let P be a point on a parabola. The tangent line to the parabola at P


makes equal angles with the following two lines:
. The line containing P and the focus F, and
. The line perpendicular to the directrix through P.
Because of this reec ve property, paraboloids (the D analogue of parabolas) make for useful ashlight reectors as the light from the bulb, ideally located
at the focus, is reected along parallel rays. Satellite dishes also have paraboloid
shapes. Signals coming from satellites eec vely approach the dish along parallel rays. The dish then focuses these rays at the focus, where the sensor is
located.

Ellipses
Deni on

Ellipse

An ellipse is the locus of all points whose sum of distances from two xed
points, each a focus of the ellipse, is constant.
An easy way to visualize this construc on of an ellipse is to pin both ends of
a string to a board. The pins become the foci. Holding a pencil ght against the
string places the pencil on the ellipse; the sum of distances from the pencil to
the pins is constant: the length of the string. See Figure . .
We can again nd an algebraic equa on for an ellipse using this geometric
deni on. Let the foci be located along the x-axis, c units from the origin. Let
these foci be labeled as F = (c, ) and F = (c, ). Let P = (x, y) be a point
on the ellipse. The sum of distances from F to P (d ) and from F to P (d ) is a
constant d. That is, d + d = d. Using the Distance Formula, we have

(x + c) + y + (x c) + y = d.
Using a fair amount of algebra can produce the following equa on of an ellipse
(note that the equa on is an implicitly dened func on; it has to be, as an ellipse
fails the Ver cal Line Test):
y
x
(d) + (d)
Notes:

= .

d2

d1

.
d1 + d2 = constant

Figure . : Illustra ng the construc on of


an ellipse with pins, pencil and string.

Chapter

Curves in the Plane


This is
not par cularly illumina ng, but by making the subs tu on a = d/ and
b = a c , we can rewrite the above equa on as

Ver ces

y
x
+
= .
a
b

Foci

{z

} |

{z
c

Major axis

Minor axis

Figure . : Labeling the signicant features of an ellipse.

This choice of a and b is not without reason; as shown in Figure . , the values
of a and b have geometric meaning in the graph of the ellipse.
In general, the two foci of an ellipse lie on the major axis of the ellipse, and
the midpoint of the segment joining the two foci is the center. The major axis
intersects the ellipse at two points, each of which is a vertex. The line segment
through the center and perpendicular to the major axis is the minor axis. The
constant sum of distances that denes the ellipse is the length of the major
axis, i.e., a.
Allowing for the shi ing of the ellipse gives the following standard equa ons.
Key Idea

Standard Equa on of the Ellipse

The equa on of an ellipse centered at (h, k) with major axis of length a


and minor axis of length b in standard form is:
. Horizontal major axis:
. Ver cal major axis:

Example
Finding the equa on of an ellipse
Find the general equa on of the ellipse graphed in Figure . .

4
2

2
2

(x h)
(y k)
+
= .
b
a

The foci lie along the major axis, c units from the center, where c =
a b .

(y k)
(x h)
+
= .
a
b

The center is located at ( , ). The distance from the cenS


ter to a vertex is units, hence a = . The minor axis seems to have length ,
so b = . Thus the equa on of the ellipse is
(x + )

(y )

= .

Figure . : The ellipse used in Example


.

Graphing an ellipse
Example
Graph the ellipse dened by x + y x

Notes:

y= .

.
y

It is simple to graph an ellipse once it is in standard form. In


order to put the given equa on in standard form, we must complete the square
with both the x and y terms. We rst rewrite the equa on by regrouping:
S

x + y x

y=

( x x) + ( y

Conic Sec ons

y) = .

Now we complete the squares.


( x x) + ( y

y) =

(x x) + (y y) =

(x x + ) + (y y + ) =
(
)
(
)
(x ) + (y ) =
(x ) + (y )

(x )

= .

.
Figure . : Graphing the ellipse in Example
.
y
1

(x ) + (y ) =
+

(y )

We see the center of the ellipse is at ( , ). We have a = and b = ; the major axis
so the ver ces are located at ( , ) and ( , ). We nd
is horizontal,
=
. . The foci are located along the major axis, approxic=
mately . units from the center, at ( . , ). This is all graphed in Figure
. .

e=0
1

(a)
y
1

Eccentricity
When a = b, we have a circle. The general equa on becomes
(x h)
(y k)
+
=
a
a

(x h) + (y k) = a ,

the familiar
equa on of the circle centered at (h, k) with radius a. Since a = b,

c = a b = . The circle has two foci, but they lie on the same point, the
center of the circle.
Consider Figure . , where several ellipses are graphed with a = . In (a),
we have c = and the ellipse is a circle. As c grows, the resul ng ellipses look
less and less circular. A measure of this noncircularness is eccentricity.

e = 0.3
1

(b)
y
1

e = 0.8

Deni on

Eccentricity of an Ellipse
c
The eccentricity e of an ellipse is e = .
a

(c)
y
1

Notes:

e = 0.99
1

(d)
Figure . : Understanding the eccentricity of an ellipse.

Chapter

Curves in the Plane


The eccentricity of a circle is ; that is, a circle has no noncircularness. As
c approaches a, e approaches , giving rise to a very noncircular ellipse, as seen
in Figure . (d).
It was long assumed that planets had circular orbits. This is known to be
incorrect; the orbits are ellip cal. Earth has an eccentricity of .
it has
a nearly circular orbit. Mercurys orbit is the most eccentric, with e = .
.
(Plutos eccentricity is greater, at e = .
, the greatest of all the currently
known dwarf planets.) The planet with the most circular orbit is Venus, with
e= .
. The Earths moon has an eccentricity of e = .
, also very circular.

Reec ve Property
The ellipse also possesses an interes ng reec ve property. Any ray emana ng from one focus of an ellipse reects o the ellipse along a line through
the other focus, as illustrated in Figure . . This property is given formally in
the following theorem.
F1

F2

Theorem

Reec ve Property of an Ellipse

Let P be a point on a ellipse with foci F and F . The tangent line to the
ellipse at P makes equal angles with the following two lines:
Figure . : Illustra ng the reec ve
property of an ellipse.

. The line through F and P, and


. The line through F and P.
This reec ve property is useful in op cs and is the basis of the phenomena
experienced in whispering halls.

Hyperbolas
The deni on of a hyperbola is very similar to the deni on of an ellipse; we
essen ally just change the word sum to dierence.
Deni on

Hyperbola

A hyperbola is the locus of all points where the absolute value of dierence of distances from two xed points, each a focus of the hyperbola,
is constant.

Notes:

The two foci lie on the transverse axis of the hyperbola; the midpoint of
the line segment joining the foci is the center of the hyperbola. The transverse
axis intersects the hyperbola at two points, each a vertex of the hyperbola. The
line through the center and perpendicular to the transverse axis is the conjugate axis. This is illustrated in Figure . . It is easy to show that the constant
dierence of distances used in the deni on of the hyperbola is the distance
between the ver ces, i.e., a.
Key Idea

Conjugate

We do not have a convenient way of visualizing the construc on of a hyperbola as we did for the ellipse. The geometric deni on does allow us to nd an
algebraic expression that describes it. It will be useful to dene some terms rst.

Conic Sec ons

axis

z }| { z }| {

Ver ces

Transverse
axis

Foci

Figure . : Labeling the signicant features of a hyperbola.

Standard Equa on of a Hyperbola

The equa on of a hyperbola centered at (h, k) in standard form is:


. Horizontal Transverse Axis:

(y k)
(x h)

= .
a
b

. Ver cal Transverse Axis:

(x h)
(y k)

= .
a
b

The ver ces are located a units from the center and the foci are located
c units from the center, where c = a + b .

..

Graphing Hyperbolas

Figure .
x

Consider the hyperbola = . Solving for y, we nd y = x / .


As x grows
large, the part of the equa on for y becomes less signicant and
y x / = x/ . That is, as x gets large, the graph of the hyperbola looks
very much like the lines y = x/ . These lines are asymptotes of the hyperbola,
as shown in Figure . .

along with its asymptotes, y =

=
x/ .

This is a valuable tool in sketching. Given the equa on of a hyperbola in


general form, draw a rectangle centered at (h, k) with sides of length a parallel
to the transverse axis and sides of length b parallel to the conjugate axis. (See
Figure . for an example with a horizontal transverse axis.) The diagonals of
the rectangle lie on the asymptotes.

k+b
k

These lines pass through (h, k). When the transverse axis is horizontal, the
slopes are b/a; when the transverse axis is ver cal, their slopes are a/b. This
gives equa ons:

kb

..
Notes:

: Graphing the hyperbola

ha

h+a

Figure . : Using the asymptotes of a hyperbola as a graphing aid.

Chapter

Curves in the Plane

b
y = (x h) + k
a

a
y = (x h) + k.
b

Graphing a hyperbola
(y )
(x )
Sketch the hyperbola given by

= .

10

Figure .
Example

Ver cal
Transverse Axis

Example

Horizontal
Transverse Axis

: Graphing the hyperbola in


.

The hyperbola is centered at ( , ); a = and b = . In


S
Figure . we draw the prescribed rectangle centered at ( , ) along with the
asymptotes dened by its diagonals. The hyperbola has a ver cal transverse
axis, so the ver ces are located at ( , ) and ( , ). This is enough to make a
good sketch.

We also nd the loca on of the foci: as c = a + b , we have c =


. . Thus the foci are located at ( , . ) as shown in the gure.
Graphing a hyperbola
Example
Sketch the hyperbola given by x y + y =

We must complete the square to put the equa on in general


S
form. (We recognize this as a hyperbola since it is a general quadra c equa on
and the x and y terms have opposite signs.)
.

x y + y=

x (y y) =

x (y y + ) =
(
)
x (y ) =

x (y ) =

.
Figure .
Example

: Graphing the hyperbola in


.

(y )

We see the hyperbola is centered at ( , ), with a horizontal transverse axis,


where a = and b = . The appropriate rectangle is sketched in Figure .
along with the
asymptotes of the hyperbola. The ver ces are located at ( , ).
. , so the foci are located at ( . , ) as shown in the
We have c =
gure.

Notes:

Conic Sec ons

Eccentricity
y
10

Deni on

Eccentricity of a Hyperbola
c
The eccentricity of a hyperbola is e = .
a

10

10

(a)
y
10

Reec ve Property
Hyperbolas share a similar reec ve property with ellipses. However, in the
case of a hyperbola, a ray emana ng from a focus that intersects the hyperbola
reects along a line containing the other focus, but moving away from that focus. This is illustrated in Figure . (on the next page). Hyperbolic mirrors are
commonly used in telescopes because of this reec ve property. It is stated
formally in the following theorem.
Theorem

10

..

(b)
y
10
5

. The line through F and P, and


10

. The line through F and P.


.

10

e=3

10

(c)
y
10
5

.
10

10

e = 10

Notes:

10

Let P be a point on a hyperbola with foci F and F . The tangent line to


the hyperbola at P makes equal angles with the following two lines:

Determining the loca on of a known event has many prac cal uses (loca ng
the epicenter of an earthquake, an airplane crash site, the posi on of the person
speaking in a large room, etc.).
To determine the loca on of an earthquakes epicenter, seismologists use
trilatera on (not to be confused with triangula on). A seismograph allows one

10
e = 1.5

Reec ve Property of Hyperbolas

Loca on Determina on

e = 1.05

Note that this is the deni on of eccentricity as used for the ellipse. When
c is close in value to a (i.e., e ), the hyperbola is very narrow (looking almost like crossed lines). Figure . shows hyperbolas centered at the origin
with a = . The graph in (a) has c = . , giving an eccentricity of e = . ,
which is close to . As c grows larger, the hyperbola widens and begins to look
like parallel lines, as shown in part (d) of the gure.

10

10

(d)
Figure . : Understanding the eccentricity of a hyperbola.

Chapter

Curves in the Plane

F1

to determine how far away the epicenter was; using three separate readings,
the loca on of the epicenter can be approximated.
A key to this method is knowing distances. What if this informa on is not
available? Consider three microphones at posi ons A, B and C which all record
a noise (a persons voice, an explosion, etc.) created at unknown loca on D.
The microphone does not know when the sound was created, only when the
sound was detected. How can the loca on be determined in such a situa on?
If each loca on has a clock set to the same me, hyperbolas can be used
to determine the loca on. Suppose the microphone at posi on A records the
sound at exactly : , loca on B records the me exactly second later, and
loca on C records the noise exactly seconds a er that. We are interested in
the dierence of mes. Since the speed of sound is approximately
m/s, we
can conclude quickly that the sound was created
meters closer to posi on A
than posi on B. If A and B are a known distance apart (as shown in Figure .
(a)), then we can determine a hyperbola on which D must lie.
The dierence of distances is
; this is also the distance between ver ces
of the hyperbola. So we know a =
. Posi ons A and B lie on the foci, so
c=
. From this we can nd b
and can sketch the hyperbola, given
in part (b) of the gure. We only care about the side closest to A. (Why?)
We can also nd the hyperbola dened by posi ons B and C. In this case,
a =
as the sound traveled an extra seconds to get to C. We s ll have
c =
, centering this hyperbola at (
,
). We nd b
. This
hyperbola is sketched in part (c) of the gure. The intersec on point of the two
graphs is the loca on of the sound, at approximately (
,
. ).

F2

.
Figure . : Illustra ng the reec ve
property of a hyperbola.

1,000

1,000

500

1,000

500

500

500

1,000

1,000

500

1,000

500

500

500

1,000

1,000

500

1,000

(a)

1,000

(b)
Figure .

B
500

500

1,000

500

1,000

(c)

: Using hyperbolas in loca on detec on.

This chapter explores curves in the plane, in par cular curves that cannot
be described by func ons of the form y = f(x). In this sec on, we learned of
ellipses and hyperbolas that are dened implicitly, not explicitly. In the following
sec ons, we will learn completely new ways of describing curves in the plane,
using parametric equa ons and polar coordinates, then study these curves using
calculus techniques.
Notes:

..

Exercises .

Terms and Concepts

. What is the dierence between degenerate and nondegenerate conics?


. Use your own words to explain what the eccentricity of an
ellipse measures.

x +

(y + ) =

In Exercises , nd the equa on of the ellipse shown in


the graph. Give the loca on of the foci and the eccentricity
of the ellipse.
y

. What has the largest eccentricity: an ellipse or a hyperbola?


. Explain why the following is true: If the coecient of the x
term in the equa on of an ellipse in standard form is smaller
than the coecient of the y term, then the ellipse has a
horizontal major axis.

. Explain how one can quickly look at the equa on of a hyperbola in standard form and determine whether the transverse axis is horizontal or ver cal.

Problems

In Exercises , nd the equa on of the parabola dened


by the given informa on. Sketch the parabola.

In Exercises
, nd the equa on of the ellipse dened
by the given informa on. Sketch the elllipse.

. Focus: ( , ); directrix: y =
. Focus: ( , ); directrix: y =

. Foci: ( , ); ver ces: ( , )

. Focus: ( , ); directrix: x =

. Foci: ( , ) and ( , ); ver ces: ( , ) and ( , )

. Focus: ( / , ); directrix: x = /

. Foci: ( , ); ver ces: ( , )

. Focus: ( , ); vertex: ( , )

. Focus: ( , ); vertex: ( , ); center: ( , )

. Focus: ( , ); vertex: ( , )
. Vertex: ( , ); directrix: y = /

In Exercises
, the equa on of a parabola and a point
on its graph are given. Find the focus and directrix of the
parabola, and verify that the given point is equidistant from
the focus and directrix.
. y= x ,P=( , )
,

In Exercises
, sketch the ellipse dened by the given
equa on. Label the center, foci and ver ces.
.

(x )

(y )

, write the equa on of the given ellipse in

. x x+ y y=

. Vertex: ( , ); directrix: x =

. x = (y ) + , P = (

In Exercises
standard form.

x + y =

x + y

y+

. x +y x y+

=
=

. Consider the ellipse given by

(x )

(y )

= .

).
(a) Verify that the foci are located at ( ,

(b) The points P = ( , ) and P = ( + , + )


( .
, .
) lie on the ellipse. Verify that the sum
of distances from each point to the foci is the same.

In Exercises
in the graph.

, nd the equa on of the hyperbola shown

. Foci: ( , ); ver ces: ( , )

. Foci: ( , ); ver ces: ( , )

.
2

In Exercises

standard form.

.
.
y
5

.
5

, write the equa on of the hyperbola in

x y =

x y + y=

. x

y +

y=

. ( y x)( y + x) =

.
2
x

.
.
y

. Johannes Kepler discovered that the planets of our solar


system have ellip cal orbits with the Sun at one focus. The
Earths ellip cal orbit is used as a standard unit of distance;
the distance from the center of Earths ellip cal orbit to one
vertex is Astronomical Unit, or A.U.
The following table gives informa on about the orbits of
three planets.
Distance from
eccentricity
center to vertex
Mercury
.
A.U.
.
Earth
A.U.
.
Mars
.
A.U.
.
(a) In an ellipse, knowing c = a b and e = c/a
allows
us to nd b in terms of a and e. Show b =

e .
a

.
2
x

.
.

In Exercises
, sketch the hyperbola dened by the
given equa on. Label the center and foci.

. Foci: ( , ) and ( , ); ver ces: ( , ) and ( , )


. Foci: ( , ) and ( , ); ver ces: ( , ) and ( , )

In Exercises
, nd the equa on of the hyperbola dened by the given informa on. Sketch the hyperbola.

(x )

. (y )

(y + )

(x + )

(b) For each planet, nd equa ons of their ellip cal orbit
x
y
of the form
+
= . (This places the center at
a
b
( , ), but the Sun is in a dierent loca on for each
planet.)
(c) Shi the equa ons so that the Sun lies at the origin.
Plot the three ellip cal orbits.
. A loud sound is recorded at three sta ons that lie on a line
as shown in the gure below. Sta on A recorded the sound
second a er Sta on B, and Sta on C recorded the sound
seconds a er B. Using the speed of sound as
m/s,
determine the loca on of the sounds origina on.
A

m B

Parametric Equa ons

We are familiar with sketching shapes, such as parabolas, by following this basic
procedure:
Choose
x

Use a func on
(f to nd y)
y = f(x)

Plot point
(x, y)

The rectangular equa on y = f(x) works well for some shapes like a parabola
with a ver cal axis of symmetry, but in the previous sec on we encountered several shapes that could not be sketched in this manner. (To plot an ellipse using
the above procedure, we need to plot the top and bo om separately.)
In this sec on we introduce a new sketching procedure:

Choose
t

Use a func on
(f to nd x)
x = f(t)
Use a func on
(g to nd y)
y = g(t)

Plot point
(x, y)

Here, x and y are found separately but then plo ed together. This leads us
to a deni on.
Deni on

Parametric Equa ons and Curves

Let
( f)and (g be con )nuous func ons on an interval I. The set of all points
x, y = f(t), g(t) in the Cartesian plane, as t varies over I, is the graph
of the parametric equa ons x = f(t) and y = g(t), where t is the parameter. A curve is a graph along with the parametric equa ons that dene
it.
This is a formal deni on of the word curve. When a curve lies in a plane
(such as the Cartesian plane), it is o en referred to as a plane curve. Examples
will help us understand the concepts introduced in the deni on.
Example

Plo ng parametric func ons

Plot the graph of the parametric equa ons x = t , y = t + for t in [ , ].

Notes:

Parametric Equa ons

Chapter

Curves in the Plane


t

We plot the graphs of parametric equa ons in much the


S
same manner as we plo ed graphs of func ons like y = f(x): we make a table of
values, plot points, then connect these points with a reasonable looking curve.
Figure . (a) shows such a table of values; note how we have columns.
The points (x, y) from the table are plo ed in Figure . (b). The points have
been connected with a smooth curve. Each point has been labeled with its corresponding t-value. These values, along with the two arrows along the curve,
are used to indicate the orienta on of the graph. This informa on helps us determine the direc on in which the graph is moving.

(a)
y
4
t=
t=

We o en use the le er t as the parameter as we o en regard t as representing me. Certainly there are many contexts in which the parameter is not me,
but it can be helpful to think in terms of me as one makes sense of parametric
plots and their orienta on (for instance, At me t = the posi on is ( , ) and
at me t = the posi on is ( , ).).

t=
x

t=

t=

Example

(b)

Sketch the graph of the parametric equa ons x = cos t, y = cos t +


in [ , ].

Figure . : A table of values of the parametric equa ons in Example


along
with a sketch of their graph.

/
/
/

(a)
y
t=
t = /
.
t = /

t = /

t=

Technology Note: Most graphing u li es can graph func ons given in parametric form. O en the word parametric is abbreviated as PAR or PARAM in
the op ons. The user usually needs to determine the graphing window (i.e, the
minimum and maximum x- and y-values), along with the values of t that are to
be plo ed. The user is o en prompted to give a t minimum, a t maximum, and
a t-step or t. Graphing u li es eec vely plot parametric func ons just as
weve shown here: they plots lots of points. A smaller t-step plots more points,
making for a smoother graph (but may take longer). In Figure . , the t-step is

(b)
Figure . : A table of values of the parametric equa ons in Example
along
with a sketch of their graph.

for t

We again start by making a table of values in Figure . (a),


S
then plot the points (x, y) on the Cartesian plane in Figure . (b).
It is not dicult to show that the curves in Examples
and
are por ons
of the same parabola. While the parabola is the same, the curves are dierent.
In Example
, if we let t vary over all real numbers, wed obtain the en re
parabola. In this example, le ng t vary over all real numbers would s ll produce
the same graph; this por on of the parabola would be traced, and retraced,
innitely. The orienta on shown in Figure . shows the orienta on on [ , ],
but this orienta on is reversed on [, ].
These examples begin to illustrate the powerful nature of parametric equaons. Their graphs are far more diverse than the graphs of func ons produced
by y = f(x) func ons.

Plo ng parametric func ons

Notes:

.
; in Figure .

Parametric Equa ons

, the t-step is / .

One nice feature of parametric equa ons is that their graphs are easy to
shi . While this is not too dicult in the y = f(x) context, the resul ng funcon can look rather messy. (Plus, to shi to the right by two, we replace x with
x , which is counterintui ve.) The following example demonstrates this.
Shi ing the graph of parametric func ons
Example
Sketch the graph of the parametric equa ons x = t + t, y = t t. Find new
parametric equa ons that shi this graph to the right places and down .
The graph of the parametric equa ons is given in Figure .
(a). It is a parabola with a axis of symmetry along the line y = x; the vertex is at
( , ).
In order to shi the graph to the right units, we need to increase the xvalue by for every point. The straigh orward way to accomplish this is simply
to add to the func on dening x: x = t + t + . To shi the graph down by
units, we wish to decrease each y-value by , so we subtract from the func on
dening y: y = t t . Thus our parametric equa ons for the shi ed graph
are x = t + t + , y = t t . This is graphed in Figure . (b). No ce how
the vertex is now at ( , ).

y=t t

(a)
y

Because the x- and y-values of a graph are determined independently, the


graphs of parametric func ons o en possess features not seen on y = f(x)
type graphs. The next example demonstrates how such graphs can arrive at the
same point more than once.

x=t +t

x=t +t+
y=t t

(b)

Figure . : Illustra ng how to shi


graphs in Example
.

Graphs that cross themselves


Example
Plot the parametric func ons x = t t + t +
and y = t t + and
determine the t-values where the graph crosses itself.
Using the methods developed in this sec on, we again plot
S
points and graph the parametric equa ons as shown in Figure . . It appears
that the graph crosses itself at the point ( , ), but well need to analy cally
determine this.
. = x(t) and
We are looking for two dierent values, say, s and t, where x(s)
y(s) = y(t). That is, the x-values are the same precisely when the y-values are
the same. This gives us a system of equa ons with unknowns:
s s + s+

s s+

Notes:

=t t + t+

=t t+

x = t 5t + t +
y=t t+

.5

Figure . : A graph of the parametric


equa ons from Example
.

Chapter

Curves in the Plane


Solving this system is not trivial but involves only algebra. Using the quadra c
formula,
one can solve for t in the second equa on and nd that t =

s s + . This can be subs tuted into the rst equa on, revealing that the
graph crosses itself at t = and t = . We conrm our result by compu ng
x( ) = x( ) = and y( ) = y( ) = .

Conver ng between rectangular and parametric equa ons


It is some mes useful to rewrite equa ons in rectangular form (i.e., y = f(x))
into parametric form, and viceversa. Conver ng from rectangular to parametric can be very simple: given y = f(x), the parametric equa ons x = t, y = f(t)
produce the same graph. As an example, given y = x , the parametric equa ons
x = t, y = t produce the familiar parabola. However, other parametriza ons
can be used. The following example demonstrates one possible alterna ve.
Example
Conver ng from rectangular to parametric
Consider y = x . Find parametric equa ons x = f(t), y = g(t) for the parabola
where t = dy
dx . That is, t = a corresponds to the point on the graph whose
tangent line has slope a.

We start by compu ng dy
dx : y = x. Thus we set t = x. We
can solve for x and nd x = t/ . Knowing that y = x , we have y = t / . Thus
parametric equa ons for the parabola y = x are

x = t/

y=t / .

To nd the point where the tangent line has a slope of , we set t = . This
gives the point ( , ). We can verify that the slope of the line tangent to the
curve at this point indeed has a slope of .
We some mes chose the parameter to accurately model physical behavior.
Conver ng from rectangular to parametric
Example
An object is red from a height of
and lands seconds later,
away. Assuming ideal projec le mo on, the height, in feet, of the object can be described
by h(x) = x / + x, where x is the distance in feet from the ini al loca on.
(Thus h( ) = h(
) =
.) Find parametric equa ons x = f(t), y = g(t)
for the path of the projec le where x is the horizontal distance the object has
traveled at me t (in seconds) and y is the height at me t.
Physics tells us that the horizontal mo on of the projec le
S
is linear; that is, the horizontal speed of the projec le is constant. Since the
object travels
in s, we deduce that the object is moving horizontally at

Notes:

.
a rate of
/s, giving the equa on x = t. As y = x / + x, we nd
y = t + t. We can quickly verify that y =
/s , the accelera on
due to gravity, and that the projec le reaches its maximum at t = , halfway
along its path.
These parametric equa ons make certain determina ons about
( the objects
)
loca
on
easy:
seconds
into
the
ight
the
object
is
at
the
point
x( ), y( ) =
(
)
,
. That is, it has traveled horizontally
and is at a height of
, as
shown in Figure . .
It is some mes necessary to convert given parametric equa ons into rectangular form. This can be decidedly more dicult, as some simple looking
parametric equa ons can have very complicated rectangular equa ons. This
conversion is o en referred to as elimina ng the parameter, as we are looking
for a rela onship between x and y that does not involve the parameter t.

Parametric Equa ons

y
5

t=

x=

y = 6t + 96t

.
5

Figure .
Example

: Graphing projec le mo on in
.

Elimina ng the parameter


Example
Find a rectangular equa on for the curve described by
x=

and

t +

y=

t
.
t +

There is not a set way to eliminate a parameter. One method


S
is to solve for t in one equa on and then subs tute that value in the second. We
use that technique here, then show a second, simplermethod.
Star ng with x = /(t + ), solve for t: t =
/x . Subs tute this
value for t in the equa on for y:
t
t +
/x
=
/x +
/x
=
/x
(
)
=

x
x

y
2
y=1x
1

y=

x.

Thus y = x. One may have recognized this earlier by manipula ng the


equa on for y:
t
=
= x.
y=
t +
t +

Notes:

1
t2 + 1
t2
y= 2
t +1

x=

Figure . : Graphing parametric and


rectangular equa ons for a graph in Example
.

Chapter

Curves in the Plane


This is a shortcut that is very specic to this problem; some mes shortcuts exist
and are worth looking for.
We should be careful to limit the domain of the func on y = x. The
parametric equa ons limit x to values in ( , ], thus to produce the same graph
we should limit the domain of y = x to the same.
The graphs of these func ons is given in Figure . . The por on of the graph
dened by the parametric equa ons is given in a thick line; the graph dened
by y = x with unrestricted domain is given in a thin line.
Elimina ng the parameter
Example
Eliminate the parameter in x = cos t + , y = sin t +

We should not try to solve for t in this situa on as the reS


sul ng algebra/trig would be messy. Rather, we solve for cos t and sin t in each
equa on, respec vely. This gives

4
2

cos t =

and

sin t =

.
Figure . : Graphing the parametric
equa ons x = cos t + , y = sin t +
in Example
.

The Pythagorean Theorem gives cos t + sin t = , so:

cos t + sin t =
)
(
y
+
=

(x )

(y )

This nal equa on should look familiar it is the equa on of an ellipse! Figure
. plots the parametric equa ons, demonstra ng that the graph is indeed of
an ellipse with a horizontal major axis and center at ( , ).
The Pythagorean Theorem can also be used to iden fy parametric equa ons
for hyperbolas. We give the parametric equa ons for ellipses and hyperbolas in
the following Key Idea.

Notes:

Parametric Equa ons


y
1

Key Idea

Parametric Equa ons of Ellipses and Hyperbolas

The parametric equa ons


x = a cos t + h,

y = b sin t + k

dene an ellipse with horizontal axis of length a and ver cal axis
of length b, centered at (h, k).
The parametric equa ons
x = a tan t + h,

Astroid
x = cos t
y = sin t

y = b sec t + k

dene a hyperbola with ver cal transverse axis centered at (h, k),
and
x = a sec t + h, y = b tan t + k
denes a hyperbola with horizontal transverse axis. Each has
asymptotes at y = b/a(x h) + k.

Special Curves

One might note a feature shared by two of these graphs: sharp corners,
or cusps. We have seen graphs with cusps before and determined that such
func ons are not dieren able at these points. This leads us to a deni on.

y
5

Smooth

A curve C dened by x = f(t), y = g(t) is smooth on an interval I if f and


g are con nuous on I and not simultaneously (except possibly at the
endpoints of I). A curve is piecewise smooth on I if I can be par oned
into subintervals where C is smooth on each subinterval.

Hypotrochoid
x = cos(t) + cos( t/ )
y = sin(t) sin( t/ )

Consider the astroid, given by x = cos t, y = sin t. Taking deriva ves, we


have:
x = cos t sin t and

Rose Curve
x = cos(t) sin( t)
y = sin(t) sin( t)

Figure . gives a small gallery of interes ng and famous curves along


with parametric equa ons that produce them. Interested readers can begin
learning more about these curves through internet searches.

Deni on

y = sin t cos t.

Notes:

Figure .
curves.

Epicycloid
x = cos(t) cos( t)
y = sin(t) sin( t)
: A gallery of interes ng planar

Chapter

Curves in the Plane


It is clear that each is when t = , / , , . . .. Thus the astroid is not smooth
at these points, corresponding to the cusps seen in the gure.
We demonstrate this once more.
Determine where a curve is not smooth
Example
Let a curve C be dened by the parametric equa ons x = t t +
y = t t + . Determine the points, if any, where it is not smooth.

and

We begin by taking deriva ves.


x = t

y = t .

We set each equal to :


x =
y =
.

Figure . : Graphing the curve in Example


; note it is not smooth at ( , ).

t
= t=
t = t=

We see at t = both x and y are ; thus C is not smooth at t = , corresponding to the point ( , ). The curve is graphed in Figure . , illustra ng the cusp
at ( , ).
If a curve is not smooth at t = t , it means that x (t ) = y (t ) = as
dened. This, in turn, means that rate of change of x (and y) is ; that is, at
that instant, neither x nor y is changing. If the parametric equa ons describe
the path of some object, this means the object is at rest at t . An object at rest
can make a sharp change in direc on, whereas moving objects tend to change
direc on in a smooth fashion.
One should be careful to note that a sharp corner does not have to occur
when a curve is not smooth. For instance, one can verify that x = t , y = t produce the familiar y = x parabola. However, in this parametriza on, the curve
is not smooth. A par cle traveling along the parabola according to the given
parametric equa ons comes to rest at t = , though no sharp point is created.
Our previous experience with cusps taught us that a func on was not dieren able at a cusp. This can lead us to wonder about deriva ves in the context
of parametric equa ons and the applica on of other calculus concepts. Given a
curve dened parametrically, how do we nd the slopes of tangent lines? Can
we determine concavity? We explore these concepts and more in the next secon.

Notes:

Exercises .

Terms and Concepts


. T/F: When sketching the graph of parametric equa ons, the
x and y values are found separately, then plo ed together.

In Exercises
, four sets of parametric equa ons are
given. Describe how their graphs are similar and dierent.
Be sure to discuss orienta on and ranges.
.

(a) x = t y = t ,

< t <

. The direc on in which a graph is moving is called the


of the graph.

(b) x = sin t y = sin t,


y = e t,

. An equa on wri en as y = f(x) is wri en in

< t <

(d) x = t y = t ,

< t <

(c) x = et

form.

. Create parametric equa ons x = f(t), y = g(t) and sketch


their graph. Explain any interes ng features of your graph
based on the func ons f and g.

(a) x = cos t y = sin t,


(b) x = cos(t )

t , t

. x = , y = sin t,
. x=t ,

. x = sec t,

/ t /

. x=t t+ , y=t + , t
In Exercises , sketch the graph of the given parametric equa ons; using a graphing u lity is advisable. Be sure to
indicate the orienta on of the graph.

y = tan t

. x=t ,

y=t

. x=

t+

. x = et ,

. x=t t , y=t ,

. x = cot t,

. x = /t, y = sin t,

<t

. x = cosh t,

. x = cos t + , y = sin t + ,
. x = cos t, y = cos( t),

. x = cos t, y = sin( t),

y=

t+
t+

y=t

. x = cos( t),

y = cos t

y=et

. x = ln t,

. x = cos t, y = sin t,

y= t+

. x = sin t + ,

y= , t

y = sin(cos t),

<t<

In Exercises
, eliminate the parameter in the given
parametric equa ons.
. x= t+ ,

. x = t + t, y =

y = sin( /t),

(d) x = cos(cos t)

In Exercises , sketch the graph of the given parametric


equa ons by hand, making a table of points to plot. Be sure
to indicate the orienta on of the graph.

y = sin(t ),

(c) x = cos( /t)

Problems

< t <

y = csc t
y = sinh t
y = sin t

In Exercises
, eliminate the parameter in the given
parametric equa ons. Describe the curve dened by the
parametric equa ons based on its rectangular form.
. x = at + x ,

. x = sec t, y = tan t, / < t < /

. x = r cos t,

y = bt + y
y = r sin t

. x = cos t + cos( t),

y = sin t + sin( t),

. x = a cos t + h,

y = b sin t + k

. x = cos t + sin( t),

y = sin t + cos( t),

. x = a sec t + h,

y = b tan t + k

In Exercises

, nd parametric equa ons for the given


dy
rectangular equa on using the parameter t =
. Verify that
dx
at t = , the point on the graph has a tangent line with slope
of .
. y= x

x+

. y = sin x on [ , ]

x on [ , )

In Exercises
, nd the values of t where the graph of
the parametric equa ons crosses itself.
. x=t t+ , y=t
. x=t t +t+ , y=t t
. x = cos t, y = sin( t) on [ , ]
. x = cos t cos( t),

y = sin t cos( t) on [ , ]

In Exercises
, nd the value(s) of t where the curve
dened by the parametric equa ons is not smooth.
. x = t + t t,
. x = t t,

y = sin t sin( t)

In Exercises
, nd parametric equa ons that describe
the given situa on.
. A projec le is red from a height of
in s.

. y = ex

. y=

. x = cos t cos( t),

y=t + t+

y=t t t

. x = cos t, y = cos t

, landing

away

. A projec le is red from a height of


in s.

, landing

away

. A projec le is red from a height of


in s.

, landing

away

. A circle of radius , centered at the origin, that is traced


clockwise once on [ , ].
. A circle of radius , centered at ( , ), that is traced once
counterclockwise on [ , ].
. An ellipse centered at ( , ) with ver cal major axis of
length and minor axis of length .
. An ellipse with foci at ( , ) and ver ces at ( , ).
. A hyperbola with foci at ( , ) and ( , ), and with
ver ces at ( , ) and ( , ).
. A hyperbola with ver ces at ( , ) and asymptotes y =
x.

Calculus and Parametric Equa ons

The previous sec on dened curves based on parametric equa ons. In this secon well employ the techniques of calculus to study these curves.
We are s ll interested in lines tangent to points on a curve. They describe
how the y-values are changing with respect to the x-values, they are useful in
making approxima ons, and they indicate instantaneous direc on of travel.
The slope of the tangent line is s ll dy
dx , and the Chain Rule allows us to calculate this in the context of parametric equa ons. If x = f(t) and y = g(t), the
Chain Rule states that
dy dx
dy
=
.
dt
dx dt
Solving for dy
dx , we get
/
dy
dy
=
dx
dt

g (t)
dx
= ,
dt
f (t)

provided that f (t) = . This is important so we label it a Key Idea.


Key Idea

Finding

dy
dx

with Parametric Equa ons.

Let x = f(t) and y = g(t), where f and g are dieren able on some open
interval I and f (t) = on I. Then
g (t)
dy
= .
dx
f (t)
We use this to dene the tangent line.
Deni on

Tangent and Normal Lines

Let a curve C be parametrized by x = f(t) and y = g(t), where f and g


are dieren able func ons on some interval I (containing t) = t . The
tangent line to C at t = t is the line through f(t ), g(t ) with slope
m = g (t )/f (t ), provided f (t ) = .
(
)
The normal line to C at t = t is the line through f(t ), g(t ) with slope
m = f (t )/g (t ), provided g (t ) = .
The deni on leaves two special cases to consider. When the tangent line is
horizontal, the normal line is undened by the above deni on as g (t ) = .

Notes:

Calculus and Parametric Equa ons

Chapter

Curves in the Plane


Likewise, when the normal line is horizontal, the tangent line is undened. It
seems reasonable that these lines be dened (one can draw a line tangent to
the right side of a circle, for instance), so we add the following to the above
deni on.
. If the tangent line at t = t has a slope of , the normal line to C at t = t
is the line x = f(t ).
. If the normal line at t = t has a slope of , the tangent line to C at t = t
is the line x = f(t ).
Tangent and Normal Lines to Curves
Example
Let x = t t + and y = t + t , and let C be the curve dened by these
equa ons.
. Find the equa ons of the tangent and normal lines to C at t = .
. Find where C has ver cal and horizontal tangent lines.
S

. We start by compu ng f (t) =

dy
=
dx

40

t+
.
t

Make note of something that might seem unusual: dy


dx is a func on of t,
not x. Just as points on the curve are found in terms of t, so are the slopes
of the tangent lines.

20

20

t and g (t) = t + . Thus

40

60

80

The point on C at t = is ( , ). The slope of the tangent line is m = /


and the slope of the normal line is m = . Thus,

20.

the equa on of the tangent line is y =

Figure . : Graphing tangent and normal lines in Example


.

(x

)+

the equa on of the normal line is y = (x


This is illustrated in Figure .

)+

, and
.

. To nd where C has a horizontal tangent line, we set dy


dx = and solve for
t. In this case, this amounts to se ng g (t) = and solving for t (and
making sure that f (t) = ).
g (t) =

t+

t= .

The point on C corresponding to t = is ( , ); the tangent line at


that point is horizontal (hence with equa on y = ).

Notes:

Calculus and Parametric Equa ons

To nd where C has a ver cal tangent line, we nd where it has a horizontal

and
normal line, and set gf (t)
(t) = . This amounts to se ng f (t) =

solving for t (and making sure that g (t) = ).


f (t) =

t= . .

The point on C corresponding to t = . is ( . , .


that point is x = . .

). The tangent line at

The points where the tangent lines are ver cal and horizontal are indicated on the graph in Figure . .
Example

Tangent and Normal Lines to a Circle

. Find where the unit circle, dened by x = cos t and y = sin t on [ , ],


has ver cal and horizontal tangent lines.
. Find the equa on of the normal line at t = t .
S

. We compute the deriva ve following Key Idea

dy
g (t)
cos t
=
=
.
dx
f (t)
sin t
The deriva ve is when cos t = ; that is, when t = / , / . These
are the points ( , ) and ( , ) on the circle.

The normal line is horizontal (and hence, the tangent line is ver cal) when
sin t = ; that is, when t = , , , corresponding to the points ( , )
and ( , ) on the circle. These results should make intui ve sense.

y
1

sin t
= tan t . This normal
cos t
line goes through the point (cos t , sin t ), giving the line

. The slope of the normal line at t = t is m =

sin t
(x cos t ) + sin t
cos t
= (tan t )x,

y=

as long as cos t = . It is an important fact to recognize that the normal lines to a circle pass through its center, as illustrated in Figure . .
Stated in another way, any line that passes through the center of a circle
intersects the circle at right angles.

Notes:

Figure . : Illustra ng how a circles normal lines pass through its center.

Chapter

Curves in the Plane


Example
Tangent lines when dy
dx is not dened
Find the equa on of the tangent line to the astroid x = cos t, y = sin t at
t = , shown in Figure . .

y
1

We start by nding x (t) and y (t):

S
1

Figure .

: A graph of an astroid.

x (t) = sin t cos t,

y (t) = cos t sin t.

Note that both of these are at t = ; the curve is not smooth at t = forming
a cusp on the graph. Evalua ng dy
dx at this point returns the indeterminate form
of / .
We can, however, examine the slopes of tangent lines near t = , and take
the limit as t .
lim

cos t sin t
y (t)
= lim

t
x (t)
sin t cos t
sin t
= lim
t
cos t
= .

(We can cancel as t = .)

We have accomplished something signicant. When the deriva ve dy


dx returns an
indeterminate form at t = t , we can dene its value by se ng it to be lim dy
dx ,
tt

if that limit exists. This allows us to nd slopes of tangent lines at cusps, which
can be very benecial.
We found the slope of the tangent line at t = to be ; therefore the tangent line is y = , the x-axis.

Concavity
We con nue to analyze curves in the plane by considering their concavity;
that is, we are interested in ddxy , the second deriva ve of y with respect to x.
To nd this, we need to nd the deriva ve of dy
dx with respect to x; that is,
[ ]
d y
d dy
,
=
dx
dx dx
but recall that dy
dx is a func on of t, not x, making this computa on not straightforward.
To make the upcoming nota on a bit simpler, let h(t) = dy
dx . We want
dh
d
[h(t)];
that
is,
we
want
.
We
again
appeal
to
the
Chain
Rule.
Note:
dx
dx
/
dh dx
dh
dh
dx
dh
=

=
.
dt
dx dt
dx
dt
dt

Notes:

.
In words, to nd ddxy , we rst take the deriva ve of
divide by x (t). We restate this as a Key Idea.
Key Idea

Finding

d y
dx

dy
dx

Calculus and Parametric Equa ons

with respect to t, then

with Parametric Equa ons

Let x = f(t) and y = g(t) be twice dieren able func ons on an open
interval I, where f (t) = on I. Then
[ ]/
[ ]/
d y
dx
d dy
d dy
=
f (t).
=
dx
dt dx
dt
dt dx
Examples will help us understand this Key Idea.
Concavity of Plane Curves
Example
Let x = t t + and y = t + t as in Example
t-intervals on which the graph is concave up/down.

. Determine the

Concavity is determined by the second deriva ve of y with


respect to x,
so we compute that here following Key Idea .
t+
dy
=
and f (t) = t . So:
In Example
, we found
dx
t
[
]/
d y
t+
d
( t )
=
dx
dt
t
/
S

t )

t )

t )

d y
dx

d y
dx

>

and con-

< . We determine the intervals when the second derivacave down when
ve is greater/less than by rst nding when it is or undened.
is never , ddxy = for all t. It is undened
( t )
when t = ; that is, when t = / . Following the work established in
Sec on . , we look at values of t greater/less than / on a number line:
As the numerator of

n
ow

d
ve
nca
; co
t>

0
0.

80

t < / ; concave up

Figure . : Graphing the parametric


equa ons in Example
to demonstrate
concavity.

( t )

The graph of the parametric func ons is concave up when

Notes:

d y
dx ,

Chapter

Curves in the Plane


d y
<
dx

d y
>
dx
c. up

c. down

Reviewing Example
, we see that when t = / = . , the graph of the
parametric equa ons has a ver cal tangent line. This point is also a point of inec on for the graph, illustrated in Figure . .

Concavity of Plane Curves


Example

Find the points of inec on of the graph of the parametric equa ons x = t,
y = sin t, for t .

y=

(a)
y

.
x

t= .

Figure . : In (a), a graph of ddx y , showing


where it is approximately . In (b), graph
of the parametric equa ons in Example
along with the points of inec on.

and

d y
dx .

The points of inec on are found by se ng ddxy = . This is not trivial, as equaons that mix polynomials and trigonometric func ons generally do not have
nice solu ons.
In Figure . (a) we see a plot of the second deriva ve. It shows that it has
zeros at approximately t = . , . , . , . , . and . These approximaons are not very good, made only by looking at the graph. Newtons Method
provides more accurate approxima ons. Accurate to decimal places, we have:

(b)

dy
dx

dy
y (t)
cos t
=
=
=
t cos t.
dx
x (t)
/( t)
[ ]

d dy
cos t/ t
d y
t sin t
dt dx

=
=
= cos t t sin t.
dx
x (t)
/( t)

cos t 4t sin t

We need to compute

, .

, .

, .

and

The corresponding points have been plo ed on the graph of the parametric
equa ons in Figure . (b). Note how most occur near the x-axis, but not exactly on the axis.

Arc Length
We con nue our study of the features of the graphs of parametric equa ons
by compu ng their arc length.
Recall in Sec on . we found the arc length of the graph of a func on, from
x = a to x = b, to be

( )
b
dy
dx.
L=
+
dx
a

Notes:

.
We can use this equa on and convert it to the parametric equa on context.

Le ng x = f(t) and y = g(t), we know that dy


dx = g (t)/f (t). It will also be
useful to calculate the dieren al of x:
dx = f (t)dt

dt =

f (t)

dx.

Star ng with the arc length formula above, consider:

( )
b
dy
+
dx
L=
dx
a

b
g (t)
dx.
+
=
f (t)
a
Factor out the f (t) :
=

b
a
t
t

f (t) + g (t)

dx
f (t)
| {z }
=dt

f (t) + g (t) dt.

Note the new bounds (no longer x bounds, but t bounds). They are found
by nding t and t such that a = f(t ) and b = f(t ). This formula is important,
so we restate it as a theorem.
Theorem

Arc Length of Parametric Curves

Let x = f(t) and y = g(t) be parametric equa ons with f and g connuous on some open interval I containing t and t on which the graph
traces itself only once. The arc length of the graph, from t = t to t = t ,
is
t
f (t) + g (t) dt.
L=
t

As before, these integrals are o en not easy to compute. We start with a


simple example, then give another where we approximate the solu on.

Notes:

Calculus and Parametric Equa ons

Chapter

Curves in the Plane


Example
Arc Length of a Circle
Find the arc length of the circle parametrized by x =
[ , / ].
S

By direct applica on of Theorem


L=

cos t, y =

sin t on

, we have

( sin t) + ( cos t) dt.

Apply the Pythagorean Theorem.


=



= t

dt
= / .

This should make sense; we know from geometry that the circumference of
a circle with radius is ; since we are nding the arc length of / of a circle,
the arc length is / = / .
Arc Length of a Parametric Curve
Example
The graph of the parametric equa ons x = t(t ), y = t crosses itself as
shown in Figure . , forming a teardrop. Find the arc length of the teardrop.
y
1

We can see by the parametriza ons of x and y that when


S
t = , x = and y = . This means well integrate from t = to t = .
Applying Theorem , we have

L=
( t ) + ( t) dt
=

Figure . : A graph of the parametric


equa ons in Example
, where the arc
length of the teardrop is calculated.

t t + dt.

Unfortunately, the integrand does not have an an deriva ve expressible by elementary func ons. We turn to numerical integra on to approximate its value.
Using subintervals, Simpsons Rule approximates the value of the integral as
.
. Using a computer, more subintervals are easy to employ, and n =
gives a value of .
. Increasing n shows that this value is stable and a good
approxima on of the actual value.

Notes:

Calculus and Parametric Equa ons

Surface Area of a Solid of Revolu on


Related to the formula for nding arc length is the formula for nding surface
area. We can adapt the formula found in Key Idea from Sec on . in a similar
way as done to produce the formula for arc length done before.
Key Idea

Surface Area of a Solid of Revolu on

Consider the graph of the parametric equa ons x = f(t) and y = g(t),
where f and g are con nuous on an open interval I containing t and
t on which the graph does not cross itself.
. The surface area of the solid formed by revolving the graph about
the x-axis is (where g(t) on [t , t ]):
Surface Area =

t
t

g(t)

f (t) + g (t) dt.

. The surface area of the solid formed by revolving the graph about
the y-axis is (where f(t) on [t , t ]):
Surface Area =

t
t

f(t)

f (t) + g (t) dt.

Surface Area of a Solid of Revolu on


Example
Consider the teardrop shape formed by the parametric equa ons x = t(t ),
y = t as seen in Example
. Find the surface area if this shape is rotated
about the x-axis, as shown in Figure . .
The teardrop shape is formed between t = and t = .
S
Using Key Idea , we see we need for g(t) on [ , ], and this is not the
case. To x this, we simplify replace g(t) with g(t), which ips the whole graph
about the x-axis (and does not change the surface area of the resul ng solid).
The surface area is:

( t ) ( t ) + ( t) dt
Area S =
=

( t )

t t + dt.

Once again we arrive at an integral that we cannot compute in terms of elementary func ons. Using Simpsons Rule with n = , we nd the area to be

Notes:

Figure . : Rota ng a teardrop shape


about the x-axis in Example
.

Chapter

Curves in the Plane


S = . . Using larger values of n shows this is accurate to
decimal.

places a er the

A er dening a new way of crea ng curves in the plane, in this sec on


we have applied calculus techniques to the parametric equa on dening these
curves to study their proper es. In the next sec on, we dene another way of
forming curves in the plane. To do so, we create a new coordinate system, called
polar coordinates, that iden es points in the plane in a manner dierent than
from measuring distances from the y- and x- axes.

Notes:

Exercises .

Terms and Concepts

. x = cos t, y = sin( t) on [ , ]

. T/F: Given parametric equa ons x = f(t) and y = g(t),


dy
= f (t)/g (t), as long as g (t) = .
dx
. Given parametric equa ons x = f(t) and y = g(t),
the deriva ve dy
as given in Key Idea
is a func on of
dx
?

. x = cos t sin( t), y = sin t sin( t) on [ , ]


. x = et/ cos t, y = et/ sin t
, nd t = t where the graph of the given
dy
parametric equa ons is not smooth, then nd lim
.
tt dx
In Exercises

. T/F: Given parametric equa ons (x =)f(t) and y = g(t), to


nd ddx y , one simply computes dtd dy
.
dx

. x=

= at t = t , then the normal line to the curve at


. T/F: If dy
dx
t = t is a ver cal line.

. x = t + t

Problems
In Exercises

, parametric equa ons for a curve are given.

(c) Sketch the graph of the parametric func ons along


with the found tangent and normal lines.
. x = t, y = t ; t =
. x=

t+

y=t t + t

sin t

In Exercises
, parametric equa ons for a curve are
given. Find ddx y , then determine the intervals on which the
graph of the curve is concave up/down. Note: these are the
same equa ons as in Exercises .
. x = t, y = t
. x=

t, y = t +

. x = t t, y = t + t

t, y = t + ; t =

. x=t , y=t t

. x = t t, y = t + t; t =
. x = t , y = t t; t =

. x=t t + t , y=t t+
. x = cos t, y =

dy
(a) Find .
dx
(b) Find the equa ons of the tangent and normal line(s)
at the point(s) given.

y=t

t +

. x = sec t,
and t =

y = tan t on (/ , / )

. x = cos t, y = sin( t) on [ , ]

. x = sec t, y = tan t on (/ , / ); t = /

. x = cos t sin( t), y = sin t sin( t) on [/ , / ]

. x = cos t, y = sin( t) on [ , ]; t = /

. x = et/ cos t,

. x = cos t sin( t), y = sin t sin( t) on [ , ];


. x = et/ cos t, y = et/ sin t;

t = /

t = /

In Exercises , nd t-values where the curve dened by


the given parametric equa ons has a horizontal tangent line.
Note: these are the same equa ons as in Exercises .
. x = t, y = t
. x=

t, y = t +

. x = t t, y = t + t
. x=t ,y=t t
. x = sec t, y = tan t on (/ , / )

y = et/ sin t

In Exercises
, nd the arc length of the graph of the
parametric equa ons on the given interval(s).
. x = sin( t), y = cos( t) on [ , ]
. x = et/ cos t,

y = et/ sin t on [ , ] and [ , ]

. x= t+ , y=
. x= t

In Exercises
length.

t on [ , ]

, y = t on [ , ]

, numerically approximate the given arc

. Approximate the arc length of one petal of the rose curve


x = cos t cos( t), y = sin t cos( t) using Simpsons Rule
and n = .

. Approximate the arc length of the bow e curve x =


cos t, y = sin( t) using Simpsons Rule and n = .
. Approximate the arc length of the parabola x = t t,
y = t + t on [ , ] using Simpsons Rule and n = .
. A common approximate of the circumference of
an ellipse
a +b
given by x = a cos t, y = b sin t is C
.

Use this formula to approximate the circumference of x =


cos t, y = sin t and compare this to the approximaon given by Simpsons Rule and n = .

In Exercises , a solid of revolu on is described. Find or


approximate its surface area as specied.
. Find the surface area of the sphere formed by rota ng the
circle x = cos t, y = sin t about:

(a) the x-axis and


(b) the y-axis.
. Find the surface area of the torus (or donut) formed by
rota ng the circle x = cos t + , y = sin t about the yaxis.
. Approximate the surface area of the solid formed by rotating the upper right half of the bow e curve x = cos t,
y = sin( t) on [ , / ] about the x-axis, using Simpsons
Rule and n = .
. Approximate the surface area of the solid formed by rota ng the one petal of the rose curve x = cos t cos( t),
y = sin t cos( t) on [ , / ] about the x-axis, using Simpsons Rule and n = .

Introduc on to Polar Coordinates

Introduc on to Polar Coordinates

We are generally introduced to the idea of graphing curves by rela ng x-values


to y-values through a func on f. That is, we set y = f(x), and plot lots of point
pairs (x, y) to get a good no on of how the curve looks. This method is useful
but has limita ons, not least of which is that curves that fail the ver cal line
test cannot be graphed without using mul ple func ons.
The previous two sec ons introduced and studied a new way of plo ng
points in the x, y-plane. Using parametric equa ons, x and y values are computed independently and then plo ed together. This method allows us to graph
an extraordinary range of curves. This sec on introduces yet another way to plot
points in the plane: using polar coordinates.

P = P(r, )

Polar Coordinates

Start with a point O in the plane called the pole (we will always iden fy this
point with the origin). From the pole, draw a ray, called the ini al ray (we will
always draw this ray horizontally, iden fying it with the posi ve x-axis). A point
P in the plane is determined by the distance r that P is from O, and the angle formed between the ini al ray and the segment OP (measured counterclockwise). We record the distance and angle as an ordered pair (r, ). To avoid
confusion with rectangular coordinates, we will denote polar coordinates with
the le er P, as in P(r, ). This is illustrated in Figure .
Prac ce will make this process more clear.

O
Figure
nates.

ini al ray
.

Illustra ng polar coordi-

Plo ng Polar Coordinates


Example
Plot the following polar coordinates:
A = P( , / )

B = P( . , )

C = P( , / )

D = P( , / )
A

To aid in the drawing, a polar grid is provided at the bo om


of this page. To place the point A, go out unit along the ini al ray (pu ng
you on the inner circle shown on the grid), then rotate counter-clockwise /
radians (or ). Alternately, one can consider the rota on rst: think about the
ray from O that forms an angle of / with the ini al ray, then move out unit
along this ray (again placing you on the inner circle of the grid).

To plot B, go out . units along the ini al ray and rotate radians (
).
To plot C, go out units along the ini al ray then rotate clockwise / radians, as the angle given is nega ve.
To plot D, move along the ini al ray units in other words, back up
unit, then rotate counter-clockwise by / . The results are given in Figure . .

Notes:

.
O
D
C

Figure .
ple
.

: Plo ng polar points in Exam-

Chapter

Curves in the Plane

Consider the following two points: A = P( , ) and B = P( , ). To locate


A, go out unit on the ini al ray then rotate radians; to locate B, go out
units on the ini al ray and dont rotate. One should see that A and B are located
at the same point in the plane. We can also consider C = P( , ), or D =
P( , ); all four of these points share the same loca on.
This ability to iden fy a point in the plane with mul ple polar coordinates is
both a blessing and a curse. We will see that it is benecial as we can plot
beau ful func ons that intersect themselves (much like we saw with parametric
func ons). The unfortunate part of this is that it can be dicult to determine
when this happens. Well explore this more later in this sec on.

Polar to Rectangular Conversion


It is useful to recognize both the rectangular (or, Cartesian) coordinates of a
point in the plane and its polar coordinates. Figure . shows a point P in the
plane with rectangular coordinates (x, y) and polar coordinates P(r, ). Using
trigonometry, we can make the iden es given in the following Key Idea.
Key Idea

Conver ng Between Rectangular and Polar


Coordinates
Given the polar point P(r, ), the rectangular coordinates are determined
by
x = r cos
y = r sin .

P
r

.
O

Given the rectangular coordinates (x, y), the polar coordinates are determined by
y
r =x +y
tan = .
x

Figure . : Conver ng between rectangular and polar coordinates.

Example

Conver ng Between Polar and Rectangular Coordinates

. Convert the polar coordinates P( , / ) and P( , / ) to rectangular


coordinates.
. Convert the rectangular coordinates ( , ) and ( , ) to polar coordinates.
S

Notes:

.
(a) We start with P( , / ). Using Key Idea

Introduc on to Polar Coordinates

, we have

y = sin( / ) =
.

So the rectangular coordinates are ( ,


) ( , .
).
(b) The polar point P( , / ) is converted to rectangular with:

/
y = sin( / ) =
/ .
x = cos( / ) =

So the rectangular coordinates are ( / ,


/ )( .
, .
).
x = cos( / ) =

P( ,

)
P( ,

These points are plo ed in Figure . (a). The rectangular coordinate


system is drawn lightly under the polar coordinate system so that the rela onship between the two can be seen.
.

.
O

(a) To convert the rectangular point ( , ) to polar coordinates, we use


the Key Idea to form the following two equa ons:
+

=r

The rst equa on tells us that r =


func on, we nd
tan =

tan =

= tan

.
(a)

. Using the inverse tangent


.

( , )

Thus polar coordinates of ( , ) are P( , . ).


(b) To convert ( , ) to polar coordinates, we form the equa ons

( , )
.

( ) +

=r

tan =

.
( , )

. We need to be careful in compu ng : using the


Thus r =
inverse tangent func on, we have
tan =

= tan ( ) = / =

This is not the angle we desire. The range of tan x is (/ , / );


that is, it returns angles that lie in the st and th quadrants. To
nd loca ons in the nd and rd quadrants, add to the result of
tan x. So
+ (/ ) puts the angle at / . Thus the polar
point is P( , / ).
An alternate method is to use the angle given by arctangent, but
change
the sign of r. Thus we could also refer to ( , ) as
P( , / ).

These points are plo ed in Figure . (b). The polar system is drawn
lightly under the rectangular grid with rays to demonstrate the angles
used.

Notes:

(b)
Figure . : Plo ng rectangular and polar points in Example
.

Chapter

Curves in the Plane

Polar Func ons and Polar Graphs


Dening a new coordinate system allows us to create a new kind of funcon, a polar func on. Rectangular coordinates lent themselves well to crea ng
func ons that related x and y, such as y = x . Polar coordinates allow us to create func ons that relate r and . Normally these func ons look like r = f(),
although we can create func ons of the form = f(r). The following examples
introduce us to this concept.
Introduc on to Graphing Polar Func ons
Example
Describe the graphs of the following polar func ons.
. r= .
. = /
S

. The equa on r = . describes all points that are . units from the pole;
as the angle is not specied, any is allowable. All points . units from
the pole describes a circle of radius . .
=

r = 1.5

.
O

Figure .

We can consider the rectangular equivalent of this equa on; using r =


x + y , we see that . = x + y , which we recognize as the equa on of
a circle centered at ( , ) with radius . . This is sketched in Figure . .
2

. The equa on = / describes all points such that the line through them
and the pole make an angle of / with the ini al ray. As the radius r is not
specied, it can be any value (even nega ve). Thus = / describes the
line through the pole that makes an angle of / = with the ini al
ray.
We can again consider the rectangular equivalent of this equa on. Combine tan = y/x and = / :

: Plo ng standard polar plots.

tan / = y/x

x tan / = y

This graph is also plo ed in Figure .

y = x.

The basic rectangular equa ons of the form x = h and y = k create ver cal
and horizontal lines, respec vely; the basic polar equa ons r = h and =
create circles and lines through the pole, respec vely. With this as a founda on,
we can create more complicated polar func ons of the form r = f(). The input
is an angle; the output is a length, how far in the direc on of the angle to go out.

Notes:

Introduc on to Polar Coordinates

We sketch these func ons much like we sketch rectangular and parametric
func ons: we plot lots of points and connect the dots with curves. We demonstrate this in the following example.
Sketching Polar Func ons
Example
Sketch the polar func on r = + cos on [ , ] by plo ng points.

A common ques on when sketching curves by plo ng points


is Which points should I plot? With rectangular equa ons, we o en choose
easy values integers, then added more if needed. When plo ng polar equaons, start with the common angles mul ples of / and / . Figure .
gives a table of just a few values of in [ , ].
Consider the point P( , ) determined by the rst line of the table. The angle
is radians we do not rotate from the ini al ray then
we go out units from
/ ); so rotate by /
the pole. When = / , r = .
(actually, it is +
radians and go out .
units.
The graph shown uses more points, connected with straight lines. (The points
on the graph that correspond to points in the table are signied with larger dots.)
Such a sketch is likely good enough to give one an idea of what the graph looks
like.
S

Technology Note: Plo ng func ons in this way can be tedious, just as it was
with rectangular func ons. To obtain very accurate graphs, technology is a great
aid. Most graphing calculators can plot polar func ons; in the menu, set the
plo ng mode to something like polar or POL, depending on ones calculator.
As with plo ng parametric func ons, the viewing window no longer determines the x-values that are plo ed, so addi onal informa on needs to be provided. O en with the window se ngs are the se ngs for the beginning and
ending values (o en called min and max ) as well as the step that is, how far
apart the values are spaced. The smaller the step value, the more accurate
the graph (which also increases plo ng me). Using technology, we graphed
the polar func on r = + cos from Example
in Figure . .
Sketching Polar Func ons
Example
Sketch the polar func on r = cos( ) on [ , ] by plo ng points.
We start by making a table of cos( ) evaluated at common
angles , as shown in Figure . . These points are then plo ed in Figure .
(a). This par cular graph moves around quite a bit and one can easily forget
which points should be connected to each other. To help us with this, we numbered each point in the table and on the graph.
S

r=

/
/
/
/

.
.
.

.
O

Figure .
Example

+ cos

: Graphing a polar func on in


by plo ng points.

.
O

Figure . : Using technology to graph a


polar func on.

Notes:

Chapter

Curves in the Plane

Pt.

/
/
/
/
/
/
/

(a)

cos( )
.
.
.
.
.
.
.
.
.

Figure .

.
O

/
/
/
/
/
/
/

cos( )
.
.
.
.
.
.
.
.

: Tables of points for plo ng a polar curve.

Using more points (and the aid of technology) a smoother plot can be made as
shown in Figure . (b). This plot is an example of a rose curve.
1

(b)
Figure .
.

Pt.

: Polar plots from Example

It is some mes desirable to refer to a graph via a polar equa on, and other
mes by a rectangular equa on. Therefore it is necessary to be able to convert
between polar and rectangular func ons, which we prac ce in the following example. We will make frequent use of the iden es found in Key Idea .
Example

Conver ng between rectangular and polar equa ons.

Convert from rectangular to polar.


. y=x
. xy =

Convert from polar to rectangular.


. r=

sin cos

. r = cos

. Replace y with r sin and replace x with r cos , giving:


y=x
r sin = r cos
sin
=r
cos
We have found that r = sin / cos = tan sec . The domain of this
polar func on is (/ , / ); plot a few points to see how the familiar
parabola is traced out by the polar equa on.

Notes:

.
. We again replace x and y using the standard iden
for r:

Introduc on to Polar Coordinates

es and work to solve

y
5

xy =
r cos r sin =
r =

cos sin

r=

cos sin
.

This func on is valid only when the product of cos sin is posi ve. This
occurs in the rst and third quadrants, meaning the domain of this polar
func on is ( , / ) (, / ).

We can rewrite the original rectangular equa on xy = as y = /x.


This is graphed in Figure . ; note how it only exists in the rst and third
quadrants.
. There is no set way to convert from polar to rectangular; in general, we
look to form the products r cos and r sin , and then replace these with
x and y, respec vely. We start in this problem by mul plying both sides
by sin cos :
r=

sin cos
r(sin cos ) =
r sin r cos = .
Now replace with y and x:
yx=
y=x+ .

The original polar equa on, r = /(sin cos ) does not easily reveal
that its graph is simply a line. However, our conversion shows that it is.
The upcoming gallery of polar curves gives the general equa ons of lines
in polar form.
. By mul plying both sides by r, we obtain both an r term and an r cos
term, which we replace with x + y and x, respec vely.
r = cos
r = r cos
x + y = x.

Notes:

Figure .
ple
.

: Graphing xy =

from Exam-

Chapter

Curves in the Plane


We recognize this as a circle; by comple ng the square we can nd its
radius and center.
x x+y =

(x ) + y = .

The circle is centered at ( , ) and has radius . The upcoming gallery


of polar curves gives the equa ons of some circles in polar form; circles
with arbitrary centers have a complicated polar equa on that we do not
consider here.
Some curves have very simple polar equa ons but rather complicated rectangular ones. For instance, the equa on r = + cos describes a cardiod
(a shape important the sensi vity of microphones, among other things; one is
graphed in the gallery in the Limaon sec on). Its rectangular form is not nearly
as simple; it is the implicit equa on x + y + x y xy x y = . The
conversion is not hard, but takes several steps, and is le as a problem in the
Exercise sec on.

Gallery of Polar Curves


There are a number of basic and classic polar curves, famous for their
beauty and/or applicability to the sciences. This sec on ends with a small gallery
of some of these graphs. We encourage the reader to understand how these
graphs are formed, and to inves gate with technology other types of polar funcons.

Lines
Through the origin:

Horizontal line:

Ver cal line:

Not through origin:

r = a csc

r = a sec

r=

slo
p

e=

b
sin m cos

Notes:

a
z }| {

Spiral

Circles
Centered on x-axis:
r = a cos

Centered on y-axis:
r = a sin

.z

a
}|

Limaons

Symmetric about x-axis: r = a b cos ;


With inner loop:
a
<
b

Archimedean spiral
r=

Centered on origin:
r=a

.z

a
}| {

Symmetric about y-axis: r = a b sin ;

Cardiod:
a
=
b

Dimpled:
a
< <
b

a, b >
Convex:
a
>
b

Rose Curves

Symmetric about x-axis: r = a cos(n); Symmetric about y-axis: r = a sin(n)


Curve contains n petals when n is even and n petals when n is odd.
r = a cos( )

r = a sin( )

r = a cos( )

r = a sin( )

Special Curves
Rose curves
r = a sin(/ )

r = a sin( / )

Lemniscate:

Eight Curve:

r = a cos( )

r = a sec cos( )

Chapter

Curves in the Plane


Earlier we discussed how each point in the plane does not have a unique
representa on in polar form. This can be a good thing, as it allows for the
beau ful and interes ng curves seen in the preceding gallery. However, it can
also be a bad thing, as it can be dicult to determine where two curves intersect.
Finding points of intersec on with polar curves
Example
Determine where the graphs of the polar equa ons r = + cos and r = cos
intersect.

y
2

As technology is generally readily available, it is usually a


S
good idea to start with a graph. We have graphed the two func ons in Figure
. (a); to be er discern the intersec on points, part (b) of the gure zooms in
around the origin. We start by se ng the two func ons equal to each other
and solving for :

+ cos = cos
cos =

(a)
y

cos =

0.5

0.5

0.5

0.5

(b)

Figure . : Graphs to help determine


the points of intersec on of the polar
func ons given in Example
.

(There are, of course, innite solu ons to the equa on cos = / ; as the
limaon is traced out once on [ , ], we restrict our solu ons to this interval.)
We need to analyze this solu on. When = / we obtain the point of
intersec on that lies in the th quadrant. When = / , we get the point of
intersec on that lies in the nd quadrant. There is more to say about this second
intersec on point, however. The circle dened by r = cos is traced out once on
[ , ], meaning that this point of intersec on occurs while tracing out the circle
a second me. It seems strange to pass by the point once and then recognize
it as a point of intersec on only when arriving there a second me. The rst
me the circle arrives at this point is when = / . It is key to understand that
these two points are the same: (cos / , / ) and (cos / , / ).
To summarize what we have done so far, we have found two points of intersec on: when = / and when = / . When referencing the circle
r = cos , the la er point is be er referenced as when = / .
There is yet another point of intersec on: the pole (or, the origin). We did
not recognize this intersec on point using our work above as each graph arrives
at the pole at a dierent value.
A graph intersects the pole when r = . Considering the circle r = cos ,
r = when = / (and odd mul ples thereof, as the circle is repeatedly

Notes:

.
traced). The limaon intersects the pole when + cos = ; this occurs when
cos = / , or for = cos ( / ). This is a nonstandard angle, approximately = .
=
. . The limaon intersects the pole twice in [ , ];
the other angle at which the limaon is at the pole is the reec on of the rst
angle across the x-axis. That is, = .
=
. .
If all one is concerned with is the (x, y) coordinates at which the graphs intersect, much of the above work is extraneous. We know they intersect at ( , );
we might not care at what value. Likewise, using = / and = /
can give us the needed rectangular coordinates. However, in the next sec on
we apply calculus concepts to polar func ons. When compu ng the area of a
region bounded by polar curves, understanding the nuances of the points of
intersec on becomes important.

Notes:

Introduc on to Polar Coordinates

Exercises .

Terms and Concepts


. In your own words, describe how to plot the polar point
P(r, ).
. T/F: When plo ng a point with polar coordinate P(r, ), r
must be posi ve.
. T/F: Every point in the Cartesian plane can be represented
by a polar coordinate.
. T/F: Every point in the Cartesian plane can be represented
uniquely by a polar coordinate.

Problems
. Plot the points with the given polar coordinates.
(a) A = P( , )

(c) C = P( , / )

(b) B = P( , )

(d) D = P( , / )

. Plot the points with the given polar coordinates.


(a) A = P( , )

(c) C = P( , )

(b) B = P( , )

(d) D = P( / , / )

. For each of the given points give two sets of polar coordinates that iden fy it, where .

. Convert each of the following polar coordinates to rectangular, and each of the following rectangular coordinates to
polar.
(a) A = P( , )

(c) C = ( , )

(d) D = ( , )

(b) B = P( , / )
In Exercises
interval.
. r= ,

, graph the polar func on on the given

. = / , r
. r=

cos ,

[ , ]

. r=

+ sin ,

[ , ]

. r=

sin ,

[ , ]

. r=

sin ,

[ , ]

. r=

+ sin ,

[ , ]

. r = cos( ),

[ , ]

. r = sin( ), [ , ]
. r = cos(/ ), [ , ]
. r = cos( / ),

[ , ]

. r = / , [ , ]

. r = sin(), [ , ]

. r = cos sin ,

[ , ]

. For each of the given points give two sets of polar coordinates that iden fy it, where .
C
D

. r = (/ ) ,

[, ]

. r=

. r=

sin cos

,
cos sin

[ , ]
[ , ]

. r = sec , (/ , / )

O
B

. r = csc , ( , )

. Convert each of the following polar coordinates to rectangular, and each of the following rectangular coordinates to
polar.
(a) A = P( , / )
(b) B = P( , / )

(c) C = ( , )

(d) D = ( , )

In Exercises

gular equa on.


. r = cos
. r = sin
. r = cos + sin

, convert the polar equa on to a rectan-

. r=
. r=
. r=

sin cos
cos

. (x + ) + y =
In Exercises
lar graphs.

, nd the points of intersec on of the po-

. r = sin( ) and r = cos on [ , ]

sin

. r = tan

. r = cos( ) and r = cos on [ , ]

. r=

. r = cos and r = sin on [ , ]

. = /

. r = sin and r =

In Exercises

polar equa on.


. y=x
. y= x+
. x=
. y=
. x=y
. x y=
. x +y =

, convert the rectangular equa on to a

+ sin on [ , ]

. r = sin( ) and r = cos( ) on [ , ]


. r = cos and r =

+ cos on [, ]

. r=

and r = sin( ) on [ , ]

. r=

cos and r =

+ sin on [ , ]

. Pick a integer value(for n,)where n = , , and use technolm


ogy to plot r = sin
for three dierent integer values
n
of m. Sketch these and determine a minimal interval on
which the en re graph is shown.
. Create your own polar func on, r = f() and sketch it. Describe why the graph looks as it does.

Chapter

Curves in the Plane

Calculus and Polar Func ons

The previous sec on dened polar coordinates, leading to polar func ons. We
inves gated plo ng these func ons and solving a fundamental ques on about
their graphs, namely, where do two polar graphs intersect?
We now turn our a en on to answering other ques ons, whose solu ons
require the use of calculus. A basis for much of what is done in this sec on is
the ability to turn a polar func on r = f() into a set of parametric equa ons.
Using the iden es x = r cos and y = r sin , we can create the parametric
equa ons x = f() cos , y = f() sin and apply the concepts of Sec on . .

Polar Func ons and

dy
dx

We are interested in the lines tangent a given graph, regardless of whether


that graph is produced by rectangular, parametric, or polar equa ons. In each
of these contexts, the slope of the tangent line is dy
dx . Given r = f(), we are
generally not concerned with r = f (); that describes how fast r changes with
respect to . Instead, we will use x = f() cos , y = f() sin to compute dy
dx .
Using Key Idea we have
dy / dx
dy
=
.
dx
d d

Each of the two deriva ves on the right hand side of the equality requires the
use of the Product Rule. We state the important result as a Key Idea.

Key Idea

Finding

dy
dx

with Polar Func ons

Let r = f() be a polar func on. With x = f() cos and y = f() sin ,
f () sin + f() cos
dy
=
.
dx
f () cos f() sin
Finding dy
Example
dx with polar func ons.
Consider the limaon r = + sin on [ , ].
. Find the equa ons of the tangent and normal lines to the graph at =
/ .
. Find where the graph has ver cal and horizontal tangent lines.

Notes:

Calculus and Polar Func ons

. We start by compu ng

dy
dx .

With f () = cos , we have

dy
=
dx

cos sin + cos ( + sin )


cos sin ( + sin )
cos ( sin + )
.
=
(cos sin ) sin

When = / , dy
(this requires a bit of simplica on).
dx =
In
rectangular
coordinates,
the
point on the graph at = / is ( +

/ , +
/ ). Thus the rectangular equa on of the line tangent to
the limaon at = / is

(
)
y = (
) x( +
/ ) + +
/ . x+ . .

The limaon and the tangent line are graphed in Figure .

The normal line has the oppositereciprocal slope as the tangent line, so
its equa on is
y

x+ .

. To nd the horizontal lines of tangency, we nd where


nd where the numerator of our equa on for dy
dx is .
cos ( sin + ) =
On [ , ], cos =

cos =

or

dy
dx

= ; thus we

sin +

= .

when = / , / .

Se ng sin + = gives = sin ( / ) .


= . .
We want the results in [ , ]; we also recognize there are two solu ons,
one in the rd quadrant and one in the th . Using reference angles, we
have our two solu ons as = . and . radians. The four points we
obtained where the limaon has a horizontal tangent line are given in Figure . with blacklled dots.
To nd the ver cal lines of tangency, we set the denominator of
(cos sin ) sin = .
Convert the cos term to sin :
( sin sin ) sin =
sin + sin

Notes:

= .

dy
dx

= .

Figure . : The limaon in Example


with its tangent line at = / and
points of ver cal and horizontal tangency.

Chapter

Curves in the Plane


Recognize this as a quadra c in the variable sin . Using the quadra c
formula, we have
sin =
We solve sin =
sin =

= sin
= .

and sin =

sin =

= sin

= .

In each of the solu ons above, we only get one of the possible two solu ons as sin x only returns solu ons in [/ , / ], the th and st
quadrants. Again using reference angles, we have:

+
sin =
= .
, .
radians
and
sin =

= .

These points are also shown in Figure .

.5

.5

radians.

with whitelled dots.

When the graph of the polar func on r = f() intersects the pole, it means
that f() = for some angle . Thus the formula for dy
dx in such instances is very
simple, reducing simply to
dy
= tan .
dx
This equa on makes an interes ng point. It tells us the slope of the tangent
line at the pole is tan ; some of our previous work (see, for instance, Example
) shows us that the line through the pole with slope tan has polar equa on
= . Thus when a polar graph touches the pole at = , the equa on of the
tangent line at the pole is = .

, .

.5
.5

Figure . : Graphing the tangent lines at


the pole in Example
.

Finding tangent lines at the pole.


Example
Let r = + sin , a limaon. Find the equa ons of the lines tangent to the
graph at the pole.

Notes:

.
S

Calculus and Polar Func ons

We need to know when r = .


+ sin =
sin = /

=
,

Note: Recall that the area of a sector of a


circle with radius r subtended by an angle
is A = r .

Thus the equa ons of the tangent lines, in polar, are = / and = / .
In rectangular form, the tangent lines are y = tan( / )x and y = tan( / )x.
The full limaon can be seen in Figure . ; we zoom in on the tangent lines in
Figure . .

Area

Area

f(ci ) .

i=

This is a Riemann sum. By taking the limit of the sum as n , we nd the

Notes:

r = f()

.5

.5

(a)
/

r = f()

When using rectangular coordinates, the equa ons x = h and y = k dened


ver cal and horzontal lines, respec vely, and combina ons of these lines create
rectangles (hence the name rectangular coordinates). It is then somewhat
natural to use rectangles to approximate area as we did when learning about
the denite integral.
When using polar coordinates, the equa ons = and r = c form lines
through the origin and circles centered at the origin, respec vely, and combina ons of these curves form sectors of circles. It is then somewhat natural to
calculate the area of regions dened by polar func ons by rst approxima ng
with sectors of circles.
Consider Figure . (a) where a region dened by r = f() on [, ] is given.
(Note how the sides of the region are the lines = and = , whereas in
rectangular coordinates the sides of regions were o en the ver cal lines x = a
and x = b.)
Par on the interval [, ] into n equally spaced subintervals as = <
< < n+ = . The length of each subinterval is = ( )/n,
represen ng a small change in angle. The area of the region dened by the i th
subinterval [i , i+ ] can be approximated with a sector of a circle with radius
f(ci ), for some ci in [i , i+ ]. The area of this sector is f(ci ) . This is shown
in part (b) of the gure, where [, ] has been divided into subintervals. We
approximate the area of the whole region by summing the areas of all sectors:

.5

.5

(b)
Figure . : Compu ng the area of a polar region.

Chapter

Curves in the Plane


exact area of the region in the form of a denite integral.

Theorem

Note: Example
requires the use of

the integral cos d. This is handled

well by using the power reducing formula


as found in the back of this text. Due to
the nature of the area formula, integrating cos and sin is required o en.
We oer here these indenite integrals
as a mesaving measure.

cos d = + sin( ) + C

sin d = sin( ) + C

Area of a Polar Region

Let f be con nuous and non-nega ve on [, ], where .


The area A of the region bounded by the curve r = f() and the lines
= and = is
A =

f() d =

r d

The theorem states that . This ensures that region does not
overlap itself, which would give a result that does not correspond directly to the
area.
Area of a polar region
Example
Find the area of the circle dened by r = cos . (Recall this circle has radius / .)
This is a direct applica on of Theorem
S
out on [ , ], leading to the integral

Area =
=

cos d

+ cos( )

d

)
sin( )

. The circle is traced

6
/

.
Figure . : Finding the area of the
.
shaded
region of a cardiod in Example
.

Of course, we already knew the area of a circle with radius / . We did this example to demonstrate that the area formula is correct.
Area of a polar region
Example
Find the area of the cardiod r = +cos bound between = / and = / ,
as shown in Figure . .
Notes:

.
This is again a direct applica on of Theorem

Area =

( + cos ) d

/
/

( + cos + cos ) d

Calculus and Polar Func ons

) /

sin( )

+ sin + +

Area Between Curves

Area =

Key Idea

r d

r d =

r r

.5

.5

Area Between Polar Curves

A=

r r

Area between polar curves


Example
Find the area bounded between the curves r =
shown in Figure . .

Notes:

Figure . : Illustra ng area bound between two polar curves.

d.

The area A of the region bounded by r = f () and r = f (), =


and = , where f () f () on [, ], is

r = f ()

r = f ()

Our study of area in the context of rectangular func ons led naturally to
nding area bounded between curves. We consider the same in the context of
polar func ons.
Consider the shaded region shown in Figure . . We can nd the area of
this region by compu ng the area bounded by r = f () and subtrac ng the
area bounded by r = f () on [, ]. Thus

d.

+ cos and r =

cos , as

We need to nd the points of intersec on between these

.
Figure . : Finding the area between polar curves in Example
.

Chapter

Curves in the Plane


two func ons. Se ng them equal to each other, we nd:
+ cos = cos
cos = /
= /
Thus we integrate

( cos ) ( + cos )

Area =

/
/
/

(
(

on [/ , / ].

( cos ) ( + cos )
)

cos cos

/
)
sin( ) sin +

= .

Amazingly enough, the area between these curves has a nice value.
/

Area dened by polar curves


Example
Find the area bounded between the polar curves r =
shown in Figure . (a).

and r =

cos( ), as

We need to nd the point of intersec on between the two


S
curves. Se ng the two func ons equal to each other, we have
cos( ) =

(a)
/

.5

cos( ) =

= /

= / .

In part (b) of the gure, we zoom in on the region and note that it is not really
bounded between two polar curves, but rather by two polar curves, along with
= . The dashed line breaks the region into its component parts. Below
the dashed line, the region is dened by r = , = and = / . (Note:
the dashed line lies on the line = / .) Above the dashed line the region is
bounded by r = cos( ) and = / . Since we have two separate regions,
we nd the area using two separate integrals.
Call the area below the dashed line A and the area above the dashed line
A . They are determined by the following integrals:
A =

.5

(b)
Figure . :
Graphing the region
bounded by the func ons in Example
.

Notes:

( ) d

A =

/
/

cos( )

d.

.
(The upper bound of the integral compu ng A is / as r = cos( ) is at the
pole when = / .)
We omit the
integra on details and let the reader
verify that A = / and
A = /
/ ; the total area is A = /
/ .

Arc Length
As we have already considered the arc length of curves dened by rectangular and parametric equa ons, we now consider it in the context of polar equaons. Recall that the arc length L of the graph dened by the parametric equaons x = f(t), y = g(t) on [a, b] is
L=

b
a

f (t) + g (t) dt =

b
a

x (t) + y (t) dt.

( . )

Now consider the polar func on r = f(). We again use the iden es x =
f() cos and y = f() sin to create parametric equa ons based on the polar
func on. We compute x () and y () as done before when compu ng dy
dx , then
apply Equa on ( . ).
The expression x () + y () can be simplied a great deal; we leave this
as an exercise and state that
x () + y () = f () + f() .
This leads us to the arc length formula.
Key Idea

Arc Length of Polar Curves

Let r = f() be a polar func on with f con nuous on an open interval


I containing [, ], on which the graph traces itself only once. The arc
length L of the graph on [, ] is
L=

f () + f() d =

(r ) + r d.

Arc length of a limaon


Example
Find the arc length of the limaon r = + sin t.
With r = + sin t, we have r = cos t. The limaon is
traced out once on [ , ], giving us our bounds of integra on. Applying Key
S

Notes:

Calculus and Polar Func ons

Chapter

Curves in the Plane

Idea

, we have
L=
=
=

Figure . : The limaon in Example


whose arc length is measured.

( cos ) + ( + sin ) d

cos + sin + sin + d

sin + d
.

The nal integral cannot be solved in terms of elementary func ons, so we resorted to a numerical approxima on. (Simpsons Rule, with n = , approximates
the value with .
. Using n =
gives the value above, which is accurate
to places a er the decimal.)

Surface Area
The formula for arc length leads us to a formula for surface area. The following Key Idea is based on Key Idea .
Key Idea

Surface Area of a Solid of Revolu on

Consider the graph of the polar equa on r = f(), where f is con nuous
on an open interval containing [, ] on which the graph does not cross
itself.
. The surface area of the solid formed by revolving the graph about
the ini al ray ( = ) is:
Surface Area =

f() sin

f () + f() d.

. The surface area of the solid formed by revolving the graph about
the line = / is:
Surface Area =

Notes:

f() cos

f () + f() d.

Calculus and Polar Func ons

Example
Surface area determined by a polar curve
Find the surface area formed by revolving one petal of the rose curve r = cos( )
about its central axis (see Figure . ).

We choose, as implied by the gure, to revolve the por on


S
of the curve that lies on [ , / ] about the ini al ray. Using Key Idea and the
fact that f () = sin( ), we have
Surface Area =
.

cos( ) sin()
.

sin( )

+ cos( )

The integral is another that cannot be evaluated in terms of elementary funcons. Simpsons Rule, with n = , approximates the value at .
.

(a)

This chapter has been about curves in the plane. While there is great mathema cs to be discovered in the two dimensions of a plane, we live in a three
dimensional world and hence we should also look to do mathema cs in D
that is, in space. The next chapter begins our explora on into space by introducing the topic of vectors, which are incredibly useful and powerful mathema cal
objects.
(b)
Figure . : Finding the surface area of a
rosecurve petal that is revolved around
its central axis.

Notes:

Exercises .

Terms and Concepts


. Given polar equa on r = f(), how can one create parametric equa ons of the same curve?
. With rectangular coordinates, it is natural to approximate
area with
; with polar coordinates, it is natural to
.
approximate area with

In Exercises

, nd the area of the described region.

. Enclosed by the circle: r = sin


. Enclosed by the circle r =
. Enclosed by one petal of r = sin( )

Problems

. Enclosed by the cardiod r =

In Exercises
(a)

, nd:
. Enclosed by the inner loop of the limaon r =

dy
dx

(b) the equa on of the tangent and normal lines to the


curve at the indicated value.
. r= ;

sin

= /

. r = cos ;

+ sin ;

. r=

cos ;

. Enclosed by the outer loop of the limaon r =


(including area enclosed by the inner loop)

+ cos t

. Enclosed between the inner and outer loop of the limaon


r = + cos t

= /

. r=

+ cos t

= /

. Enclosed by r = cos and r = sin , as shown:


y

. r = ;

= /

. r = cos( );

= /

. r = sin( );

= /

. r=

= /

sin cos

In Exercises
, nd the values of in the given interval where the graph of the polar func on has horizontal and
ver cal tangent lines.

. Enclosed by r = cos( ) and r = sin( ), as shown:


y

0.5

. r = ; [ , ]

. r = sin ;

[ , ]

. r = cos( );
. r=

+ cos ;

[ , ]

. Enclosed by r = cos and r = sin( ), as shown:

[ , ]

y
1

In Exercises , nd the equa on of the lines tangent to


the graph at the pole.
. r = sin ;

[ , ]

. r = sin( ); [ , ]

. Enclosed by r = cos and r =


y

cos , as shown:

. Approximate the arc length of the cardiod r =


with Simpsons Rule and n = .
1

. Approximate the arc length of one petal of the rose curve


r = sin( ) with Simpsons Rule and n = .
+ cos

In Exercises
area.

, answer the ques ons involving surface

, answer the ques ons involving arc

. Use Key Idea


to nd the surface area of the sphere
formed by revolving the circle r = about the ini al ray.

. Let x() = f() cos and y() = f() sin . Show, as suggested by the text, that

. Use Key Idea


to nd the surface area of the sphere
formed by revolving the circle r = cos about the ini al
ray.

In Exercises
length.

x () + y () = f () + f() .

. Use the arc length formula to compute the arc length of the
circle r = .
. Use the arc length formula to compute the arc length of the
circle r = sin .

. Find the surface area of the solid formed by revolving the


cardiod r = + cos about the ini al ray.
. Find the surface area of the solid formed by revolving the
circle r = cos about the line = / .
. Find the surface area of the solid formed by revolving the
line r = sec , / / , about the line
= / .

:V
This chapter introduces a new mathema cal object, the vector. Dened in Secon . , we will see that vectors provide a powerful language for describing
quan es that have magnitude and direc on aspects. A simple example of
such a quan ty is force: when applying a force, one is generally interested in
how much force is applied (i.e., the magnitude of the force) and the direc on in
which the force was applied. Vectors will play an important role in many of the
subsequent chapters in this text.
This chapter begins with moving our mathema cs out of the plane and into
space. That is, we begin to think mathema cally not only in two dimensions,
but in three. With this founda on, we can explore vectors both in the plane and
in space.

Introduc on to Cartesian Coordinates in Space

Up to this point in this text we have considered mathema cs in a dimensional


world. We have plo ed graphs on the x-y plane using rectangular and polar
coordinates and found the area of regions in the plane. We have considered
proper es of solid objects, such as volume and surface area, but only by rst
dening a curve in the plane and then rota ng it out of the plane.
While there is wonderful mathema cs to explore in D, we live in a D
world and eventually we will want to apply mathema cs involving this third dimension. In this sec on we introduce Cartesian coordinates in space and explore basic surfaces. This will lay a founda on for much of what we do in the
remainder of the text.
Each point P in space can be represented with an ordered triple, P = (a, b, c),
where a, b and c represent the rela ve posi on of P along the x-, y- and z-axes,
respec vely. Each axis is perpendicular to the other two.
Visualizing points in space on paper can be problema c, as we are trying
to represent a -dimensional concept on a dimensional medium. We cannot
draw three lines represen ng the three axes in which each line is perpendicular to the other two. Despite this issue, standard conven ons exist for plo ng
shapes in space that we will discuss that are more than adequate.
One conven on is that the axes must conform to the right hand rule. This
rule states that when the index nger of the right hand is extended in the direcon of the posi ve x-axis, and the middle nger (bent inward so it is perpendicular to the palm) points along the posi ve y-axis, then the extended thumb
will point in the direc on of the posi ve z-axis. (It may take some thought to

Chapter

Vectors

Figure
. : Plo ng the point P
( , , ) in space.

verify this, but this system is inherently dierent from the one created by using
the le hand rule.)
As long as the coordinate axes are posi oned so that they follow this rule,
it does not ma er how the axes are drawn on paper. There are two popular
methods that we briey discuss.
In Figure . we see the point P = ( , , ) plo ed on a set of axes. The
basic conven on here is that the x-y plane is drawn in its standard way, with the
z-axis down to the le . The perspec ve is that the paper represents the x-y plane
and the posi ve z axis is coming up, o the page. This method is preferred by
many engineers. Because it can be hard to tell where a single point lies in rela on
to all the axes, dashed lines have been added to let one see how far along each
axis the point lies.
One can also consider the x-y plane as being a horizontal plane in, say, a
room, where the posi ve z-axis is poin ng up. When one steps back and looks
at this room, one might draw the axes as shown in Figure . . The same point P
is drawn, again with dashed lines. This point of view is preferred by most mathema cians, and is the conven on adopted by this text.

Measuring Distances
It is of cri cal importance to know how to measure distances between points
in space. The formula for doing so is based on measuring distance in the plane,
and is known (in both contexts) as the Euclidean measure of distance.
Deni on

Distance In Space

Let P = (x , y , z ) and Q = (x , y , z ) be points in space. The distance


D between P and Q is

D = (x x ) + (y y ) + (z z ) .
Figure
. : Plo ng the point P =
( , , ) in space with a perspec ve used
in this text.

We refer to the line segment that connects points P and Q in space as PQ,
and refer to the length of this segment as ||PQ||. The above distance formula
allows us to compute the length of this segment.
Length of a line segment
Example
Let P = ( , , ) and let Q = ( , , ). Draw the line segment PQ and nd its
length.
The points P and Q are plo ed in Figure . ; no special conS
sidera on need be made to draw the line segment connec ng these two points;

Notes:

Introduc on to Cartesian Coordinates in Space

simply connect them with a straight line. One cannot actually measure this line
on the page and deduce anything meaningful; its true length must be measured
analy cally. Applying Deni on , we have

||PQ|| = ( ) + ( ) + ( ( )) =
. .

Spheres
Just as a circle is the set of all points in the plane equidistant from a given
point (its center), a sphere is the set of all points in space that are equidistant
from a given point. Deni on allows us to write an equa on of the sphere.
We start with a point C = (a, b, c) which is to be the center of a sphere with
radius r. If a point P = (x, y, z) lies on the sphere, then P is r units from C; that
is,

||PC|| = (x a) + (y b) + (z c) = r.
Squaring both sides, we get the standard equa on of a sphere in space with
center at C = (a, b, c) with radius r, as given in the following Key Idea.
Key Idea

Standard Equa on of a Sphere in Space

The standard equa on of the sphere with radius r, centered at C =


(a, b, c), is
(x a) + (y b) + (z c) = r .
Example
Equa on of a sphere
Find the center and radius of the sphere dened by x + x+y y+z z = .
To determine the center and radius, we must put the equaS
on in standard form. This requires us to complete the square (three mes).
x + x+y y+z z=

(x + x + ) + (y y + ) + (z z + )

(x + ) + (y ) + (z ) =

The sphere is centered at ( , , ) and has a radius of .


The equa on of a sphere is an example of an implicit func on dening a surface in space. In the case of a sphere, the variables x, y and z are all used. We
now consider situa ons where surfaces are dened where one or two of these

Notes:

Figure
ample

. : Plo ng points P and Q in Ex.

Chapter

Vectors
variables are absent.

Introduc on to Planes in Space


The coordinate axes naturally dene three planes (shown in Figure . ), the
coordinate planes: the x-y plane, the y-z plane and the x-z plane. The x-y plane
is characterized as the set of all points in space where the z-value is . This,
in fact, gives us an equa on that describes this plane: z = . Likewise, the x-z
plane is all points where the y-value is , characterized by y = .

the x-y plane

the y-z plane


Figure

the x-z plane

. : The coordinate planes.

The equa on x = describes all points in space where the x-value is . This
is a plane, parallel to the y-z coordinate plane, shown in Figure . .
Figure

. : The plane x = .

Example
Regions dened by planes
Sketch the region dened by the inequali es y .
The region is all points between the planes y = and
S
y = . These planes are sketched in Figure . , which are parallel to the x-z
plane. Thus the region extends innitely in the x and z direc ons, and is bounded
by planes in the y direc on.

Cylinders

Figure . : Sketching the boundaries of


a region in Example
.

The equa on x = obviously lacks the y and z variables, meaning it denes


points where the y and z coordinates can take on any value. Now consider the
equa on x + y = in space. In the plane, this equa on describes a circle
of radius , centered at the origin. In space, the z coordinate is not specied,
meaning it can take on any value. In Figure . (a), we show part of the graph
of the equa on x + y = by sketching circles: the bo om one has a constant z-value of . , the middle one has a z-value of and the top circle has a

Notes:

Introduc on to Cartesian Coordinates in Space

z-value of . By plo ng all possible z-values, we get the surface shown in Figure
. (b). This surface looks like a tube, or a cylinder; mathema cians call
this surface a cylinder for an en rely dierent reason.

Deni on

Cylinder

Let C be a curve in a plane and let L be a line not parallel to C. A cylinder


is the set of all lines parallel to L that pass through C. The curve C is the
directrix of the cylinder, and the lines are the rulings.
(a)

In this text, we consider curves C that lie in planes parallel to one of the
coordinate planes, and lines L that are perpendicular to these planes, forming
right cylinders. Thus the directrix can be dened using equa ons involving
variables, and the rulings will be parallel to the axis of the rd variable.
In the example preceding the deni on, the curve x + y = in the x-y
plane is the directrix and the rulings are lines parallel to the z-axis. (Any circle
shown in Figure . can be considered a directrix; we simply choose the one
where z = .) Sample rulings can also be viewed in part (b) of the gure. More
examples will help us understand this deni on.
Graphing cylinders
Example
Graph the cylinder following cylinders.
. z=y
. x = sin z

(b)

Figure

. We can view the equa on z = y as a parabola in the y-z plane, as illustrated in Figure . (a). As x does not appear in the equa on, the rulings
are lines through this parabola parallel to the x-axis, shown in (b). These
rulings give a general idea as to what the surface looks like, drawn in (c).

(a)
Figure

Notes:

(b)
. : Sketching the cylinder dened by z = y .

(c)

. : Sketching x + y = .

Chapter

Vectors
. We can view the equa on x = sin z as a sine curve that exists in the x-z
plane, as shown in Figure . (a). The rules are parallel to the y axis as
the variable y does not appear in the equa on x = sin z; some of these
are shown in part (b). The surface is shown in part (c) of the gure.

(a)
Figure

(b)

(c)

. : Sketching the cylinder dened by x = sin z.

Surfaces of Revolu on

(a)

One of the applica ons of integra on we learned previously was to nd the


volume of solids of revolu on solids formed by revolving a curve about a horizontal or ver cal axis. We now consider how to nd the equa on of the surface
of such a solid.

Consider the surface formed by revolving y = x about the x-axis. Cross


sec ons of this surface parallel to the y-z plane are circles, as shown in Figure
. (a). Each circle has equa on of the form y + z = r for some radius r.
The radius is a func on of x; in fact, it is r(x)= x. Thus the equa on of the
surface shown in Figure . b is y + z = ( x) .
We generalize the above principles to give the equa ons of surfaces formed
by revolving curves about the coordinate axes.
Key Idea

Surfaces of Revolu on, Part

Let r be a radius func on.


. The equa on of the surface formed by revolving y = r(x) or z =
r(x) about the x-axis is y + z = r(x) .

(b)
Figure .
olu on.

: Introducing surfaces of rev-

. The equa on of the surface formed by revolving x = r(y) or z =


r(y) about the y-axis is x + z = r(y) .
. The equa on of the surface formed by revolving x = r(z) or y =
r(z) about the z-axis is x + y = r(z) .

Notes:

Introduc on to Cartesian Coordinates in Space

Finding equa on of a surface of revolu on


Example
Let y = sin z on [ , ]. Find the equa on of the surface of revolu on formed by
revolving y = sin z about the z-axis.
Using Key Idea , we nd the surface has equa on x +y =
S
sin z. The curve is sketched in Figure . (a) and the surface is drawn in Figure
. (b).
Note how the surface (and hence the resul ng equa on) is the same if we
began with the curve x = sin z, which is also drawn in Figure . (a).
This par cular method of crea ng surfaces of revolu on is limited. For instance, in Example
of Sec on . we found the volume of the solid formed
by revolving y = sin x about the y-axis. Our current method of forming surfaces
can only rotate y = sin x about the x-axis. Trying to rewrite y = sin x as a funcon of y is not trivial, as simply wri ng x = sin y only gives part of the region
we desire.
What we desire is a way of wri ng the surface of revolu on formed by rota ng y = f(x) about the y-axis. We start by rst recognizing this surface is the
same as revolving z = f(x) about the z-axis. This will give us a more natural way
of viewing the surface.
A value of x is a measurement of distance from the z-axis. At the distance r,
we plot a z-height of f(r). When rota ng f(x) about the z-axis, we want all points
a distance of r from the z-axis in the x-y plane to havea z-height of f(r). All such
points sa sfy the equa on r = x + y ; hence r = x + y . Replacing r with

x + y in f(r) gives z = f( x + y ). This is the equa on of the surface.


Key Idea

(a)

(b)
Figure . : Revolving y = sin z about
the z-axis in Example
.

Surfaces of Revolu on, Part

Let z = f(x), x , be a curve in the x-z plane. The surface


formed by
)
(
revolving this curve about the z-axis has equa on z = f x + y .
Finding equa on of surface of revolu on
Example
Find the equa on of the surface found by revolving z = sin x about the z-axis.
Using Key Idea , the surface has equa on z = sin
The curve and surface are graphed in Figure . .
S

(a)

(
)
x +y .

(b)

Notes:

Figure . : Revolving z = sin x about


the z-axis in Example
.

Chapter

Vectors

Quadric Surfaces
Spheres, planes and cylinders are important surfaces to understand. We
now consider one last type of surface, a quadric surface. The deni on may
look in mida ng, but we will show how to analyze these surfaces in an illumina ng way.
Deni on

Quadric Surface

A quadric surface is the graph of the general seconddegree equa on in


three variables:
Ax + By + Cz + Dxy + Exz + Fyz + Gx + Hy + Iz + J = .
When the coecients D, E or F are not zero, the basic shapes of the quadric
surfaces are rotated in space. We will focus on quadric surfaces where these coeecients are ; we will not consider rota ons. There are six basic quadric surfaces: the ellip c paraboloid, ellip c cone, ellipsoid, hyperboloid of one sheet,
hyperboloid of two sheets, and the hyperbolic paraboloid.
We study each shape by considering traces, that is, intersec ons of each
surface with a plane parallel to a coordinate plane. For instance, consider the
ellip c paraboloid z = x / + y , shown in Figure . . If we intersect this
shape with the plane z = d (i.e., replace z with d), we have the equa on:
d=

+y .

Divide both sides by d:


=

Figure . : The ellip c paraboloid z =


x / +y .

y
x
+ .
d
d

This describes an ellipse so cross sec ons parallel to the x-y coordinate plane
are ellipses. This ellipse is drawn in the gure.
Now consider cross sec ons parallel to the x-z plane. For instance, le ng
y = gives the equa on z = x / , clearly a parabola. Intersec ng with the
plane x = gives a cross sec on dened by z = y , another parabola. These
parabolas are also sketched in the gure.
Thus we see where the ellip c paraboloid gets its name: some cross sec ons
are ellipses, and others are parabolas.
Such an analysis can be made with each of the quadric surfaces. We give a
sample equa on of each, provide a sketch with representa ve traces, and describe these traces.

Notes:

Ellip c Paraboloid,

z=

y
x
+
a
b
Plane
x=d
y=d
z=d

Trace
Parabola
Parabola
Ellipse

One variable in the equa on of the ellip c paraboloid will be raised to the rst power; above,
this is the z variable. The paraboloid will open in the direc on of this variables axis. Thus
x = y /a + z /b is an ellip c paraboloid that opens along the x-axis.
Mul plying the right hand side by ( ) denes an ellip c paraboloid that opens in the opposite
direc on.

Ellip c Cone,

z =

y
x
+
a
b
Plane
x=
y=

Trace
Crossed Lines
Crossed Lines

x=d
y=d
z=d

Hyperbola
Hyperbola
Ellipse

One can rewrite the equa on as z x /a y /b = . The one variable with a posi ve
coecient corresponds to the axis that the cones open along.

Ellipsoid,

y
z
x
+
+
=
a
b
c
Plane
x=d
y=d
z=d

Trace
Ellipse
Ellipse
Ellipse

If a = b = c = , the ellipsoid is a sphere with radius a; compare to Key Idea

Hyperboloid of One Sheet,

x
y
z
+

=
a
b
c
Plane
x=d
y=d
z=d

Trace
Hyperbola
Hyperbola
Ellipse

The one variable with a nega ve coecient corresponds to the axis that the hyperboloid opens
along.

Hyperboloid of Two Sheets,

x
y
z

=
c
a
b
Plane
x=d
y=d
z=d

Trace
Hyperbola
Hyperbola
Ellipse

The one variable with a posi ve coecient corresponds to the axis that the hyperboloid opens
along. In the case illustrated, when |d| < |c|, there is no trace.

Hyperbolic Paraboloid,

z=

x
y

a
b
Plane
x=d
y=d
z=d

Trace
Parabola
Parabola
Hyperbola

The parabolic traces will open along the axis of the one variable that is raised to the rst power.

Chapter

Vectors
Sketching quadric surfaces
Example
Sketch the quadric surface dened by the given equa on.
.y=

.x +

= .

. z=y x .

. y=

We rst iden fy the quadric by pa ernmatching with the equa ons given
previously. Only two surfaces have equa ons where one variable is raised
to the rst power, the ellip c paraboloid and the hyperbolic paraboloid.
In the la er case, the other variables have dierent signs, so we conclude
that this describes a hyperbolic paraboloid. As the variable with the rst
power is y, we note the paraboloid opens along the y-axis.

(a)

To make a decent sketch by hand, we need only draw a few traces. In this
case, the traces x = and z = form parabolas that outline the shape.
x = : The trace is the parabola y = z /
z = : The trace is the parabola y = x / .
Graphing each trace in the respec ve plane creates a sketch as shown in
Figure . (a). This is enough to give an idea of what the paraboloid
looks like. The surface is lled in in (b).

(b)
Figure
.
paraboloid.

Sketching an ellip c

. x +

This is an ellipsoid. We can get a good idea of its shape by drawing the
traces in the coordinate planes.
y
z
x = : The trace is the ellipse
+
= . The major axis is along the
yaxis with length (as b = , the length of the axis is ); the minor axis
is along the z-axis with length .
z
y = : The trace is the ellipse x +
= . The major axis is along the
z-axis, and the minor axis has length along the x-axis.
z = : The trace is the ellipse x +
y-axis.

(a)

Notes:

Figure

: Sketching an ellipsoid.

= , with major axis along the

Graphing each trace in the respec ve plane creates a sketch as shown in


Figure . (a). Filling in the surface gives Figure . (b).
. z=y x :

(b)

Introduc on to Cartesian Coordinates in Space

This denes a hyperbolic paraboloid, very similar to the one shown in the
gallery of quadric sec ons. Consider the traces in the yz and xz planes:
x = : The trace is z = y , a parabola opening up in the y z plane.

y = : The trace is z = x , a parabola opening down in the x z plane.


Sketching these two parabolas gives a sketch like that in Figure
and lling in the surface gives a sketch like (b).

Iden fying quadric surfaces


Example
Consider the quadric surface shown in Figure .
equa ons best ts this surface?
z

(a),

. Which of the following


(a)

(a)

x y

(c)

(b)

x y z =

(d)

z x y =
x y

The image clearly displays a hyperboloid of two sheets. The


gallery informs us that the equa on will have a form similar to cz ax by = .
We can immediately eliminate op on (a), as the constant in that equa on is
not .
The hyperboloid opens along the x-axis, meaning x must be the only variable with a posi ve coecient, elimina ng (c).
The hyperboloid is wider in the z-direc on than in the y-direc on, so we
need an equa on where c > b. This eliminates (b), leaving us with (d). We
should verify that the equa on given in (d), x y z = , ts.
We already established that this equa on describes a hyperboloid of two
sheets that opens in the x-direc on and is wider in the z-direc on than in the
y. Now note the coecient of the x-term. Rewri ng x in standard form, we
x
. Thus when y = and z = , x must be / ; i.e., each
have: x =
( / )
hyperboloid starts at x = / . This matches our gure.
z
= best ts the graph.
We conclude that x y
S

(b)
Figure
. :
paraboloid.

Sketching a hyperbolic

This sec on has introduced points in space and shown how equa ons can
describe surfaces. The next sec ons explore vectors, an important mathema cal
object that well use to explore curves in space.

Figure . : A possible equa on of this


quadric surface is found in Example
.

Notes:

Exercises

Terms and Concepts

. y=

. Axes drawn in space must conform to the


rule.
. In the plane, the equa on x = denes a
.
space, x = denes a

; in

. In the plane, the equa on y = x denes a


space, y = x denes a
.

; in

. Which quadric surface looks like a Pringles chip?


. Consider the hyperbola x y = in the plane. If this
hyperbola is rotated about the x-axis, what quadric surface
is formed?
. Consider the hyperbola x y = in the plane. If this
hyperbola is rotated about the y-axis, what quadric surface
is formed?

In Exercises , give the equa on of the surface of revolu on described.


. Revolve z =

+y

about the y-axis.

. Revolve y = x about the x-axis.


. Revolve z = x about the z-axis.
. Revolve z = /x about the z-axis.
In Exercises , a quadric surface is sketched. Determine
which of the given equa ons best ts the graph.

Problems
. The points A = ( , , ), B = ( , , ) and C = ( , , )
form a triangle in space. Find the distances between each
pair of points and determine if the triangle is a right triangle.

. The points A = ( , , ), B = ( , , ), C = ( , , ) and


D = ( , , ) form a quadrilateral ABCD in space. Is this
a parallelogram?

(a)

x=y +

(b)

x=y +

(b)

x y +z =

(b)

x +

. Find the center and radius of the sphere dened by


x x+y + y+z + = .
. Find the center and radius of the sphere dened by
x +y +z + x y z+ = .
In Exercises

the inequali es.

, describe the region in space dened by

. x +y +z <
(a)

x y z =

. x , y , z
. y
In Exercises

. z=x

, sketch the cylinder in space.

. y = cos z
.

(a)

x +

In Exercises

, sketch the quadric surface.

. zy +x =
.

. z =x +

. x = y z
(a)

y x z =

(b)

y +x z =

.
.
.

y +

z =
=

x + y +z =

Chapter

Vectors

An Introduc on to Vectors

Many quan es we think about daily can be described by a single number: temperature, speed, cost, weight and height. There are also many other concepts
we encounter daily that cannot be described with just one number. For instance,
a weather forecaster o en describes wind with its speed and its direc on (. . .
with winds from the southeast gus ng up to
mph . . .). When applying a
force, we are concerned with both the magnitude and direc on of that force.
In both of these examples, direc on is important. Because of this, we study
vectors, mathema cal objects that convey both magnitude and direc on informa on.
One barebones deni on of a vector is based on what we wrote above:
a vector is a mathema cal object with magnitude and direc on parameters.
This deni on leaves much to be desired, as it gives no indica on as to how
such an object is to be used. Several other deni ons exist; we choose here a
deni on rooted in a geometric visualiza on of vectors. It is very simplis c but
readily permits further inves ga on.

y
5

Deni on

Vector

A vector is a directed line segment.


.

Figure . : Drawing the same vector


with dierent ini al points.

Given points P and Q (either in the plane or in space), we denote with


#
PQ the vector from P to Q. The point P is said to be the ini al point of
the vector, and the point Q is the terminal point.
#
The magnitude, length or norm of PQ is the length of the line segment
#
PQ: || PQ || = || PQ ||.
Two vectors are equal if they have the same magnitude and direc on.

y
4
S
R

Q
P

Figure . : Illustra ng how equal vectors have the same displacement.

Figure . shows mul ple instances of the same vector. Each directed line
segment has the same direc on and length (magnitude), hence each is the same
vector.
We use R (pronounced r two) to represent all the vectors in the plane,
and use R (pronounced r three) to represent all the vectors in space.
#
#
Consider the vectors PQ and RS as shown in Figure . . The vectors look to
be equal; that is, they seem to have the same length and direc on. Indeed, they
are. Both vectors move units to the right and unit up from the ini al point
to reach the terminal point. One can analyze this movement to measure the

Notes:

.
magnitude of the vector, and the movement itself gives direc on informa on
(one could also measure the slope of the line passing through P and Q or R and
S). Since they have the same length and direc on, these two vectors are equal.
This demonstrates that inherently all we care about is displacement; that is,
how far in the x, y and possibly z direc ons the terminal point is from the ini al
#
#
point. Both the vectors PQ and RS in Figure . have an x-displacement of
and a y-displacement of . This suggests a standard way of describing vectors
in the plane. A vector whose x-displacement is a and whose y-displacement is
b will have terminal point (a, b) when the ini al point is the origin, ( , ). This
leads us to a deni on of a standard and concise way of referring to vectors.
Deni on

Component Form of a Vector

. The component form of a vector v in R , whose terminal point is


(a, b) when its ini al point is ( , ), is a, b .
. The component form of a vector v in R , whose terminal point is
(a, b, c) when its ini al point is ( , , ), is a, b, c .
The numbers a, b (and c, respec vely) are the components of v.
#
It follows from the deni on that the component form of the vector PQ,
where P = (x , y ) and Q = (x , y ) is
#
PQ = x x , y y ;
#
in space, where P = (x , y , z ) and Q = (x , y , z ), the component form of PQ
is
#
PQ = x x , y y , z z .

We prac ce using this nota on in the following example.


Example

Using component form nota on for vectors

. Sketch the vector v = , star ng at P = ( , ) and nd its magnitude.


whose ini al point is R = ( , )
. Find the component form of the vector w
and whose terminal point is S = ( , ).
. Sketch the vector u = , , star ng at the point Q = ( , , ) and
nd its magnitude.

Notes:

An Introduc on to Vectors

Chapter

Vectors
y

5
S

. Using P as the ini al point, we move units in the posi ve x-direc on and
units in the posi ve y-direc on to arrive at the terminal point P =
( , ), as drawn in Figure . (a).

P
P
5

The magnitude of v is determined directly from the component form:

|| v || =
+ ( ) =
.

(a)

. Using the note following Deni on

, we have

#
RS = ( ), ( ) = , .
One can readily see from Figure . (a) that the x- and y-displacement
#
of RS is and , respec vely, as the component form suggests.
. Using Q as the ini al point, we move units in the posi ve x-direc on,
unit in the posi ve y-direc on, and units in the posi ve z-direc on
to arrive at the terminal point Q = ( , , ), illustrated in Figure . (b).
The magnitude of u is:
|| u || =

(b)
Figure .
ple
.

: Graphing vectors in Exam-

+ ( ) +

Now that we have dened vectors, and have created a nice nota on by which
to describe them, we start considering how vectors interact with each other.
That is, we dene an algebra on vectors.

Notes:

Deni on

An Introduc on to Vectors

Vector Algebra

. Let u = u , u and v = v , v be vectors in R , and let c be a


scalar.
(a) The addi on, or sum, of the vectors u and v is the vector
u + v = u + v , u + v .
(b) The scalar product of c and v is the vector
cv = c v , v = cv , cv .
. Let u = u , u , u and v = v , v , v be vectors in R , and let c
be a scalar.
(a) The addi on, or sum, of the vectors u and v is the vector
u + v = u + v , u + v , u + v .
(b) The scalar product of c and v is the vector
cv = c v , v , v = cv , cv , cv .

In short, we say addi on and scalar mul plica on are computed component
wise.
Adding vectors
Example
Sketch the vectors u = , , v = , and u + v all with ini al point at the
origin.
We rst compute u + v.
v

u + v = , + ,

As vectors convey magnitude and direc on informa on, the sum of vectors
also convey length and magnitude informa on. Adding u + v suggests the following idea:

Notes:

= , .

These are all sketched in Figure

u
+

y
4

Figure . : Graphing the sum of vectors


in Example
.

Chapter

Vectors
Star ng at an ini al point, go out u, then go out v.

y
4

u
+

Figure . : Illustra ng how to add vectors using the Head to Tail Rule and Parallelogram Law.

This idea is sketched in Figure . , where the ini al point of v is the terminal point of u. This is known as the Head to Tail Rule of adding vectors. Vector
addi on is very important. For instance, if the vectors u and v represent forces
ac ng on a body, the sum u + v gives the resul ng force. Because of various
physical applica ons of vector addi on, the sum u +v is o en referred to as the
resultant vector, or just the resultant.
Analy cally, it is easy to see that u + v = v + u. Figure . also gives a
graphical representa on of this, using gray vectors. Note that the vectors u and
v, when arranged as in the gure, form a parallelogram. Because of this, the
Head to Tail Rule is also known as the Parallelogram Law: the vector u + v is
dened by forming the parallelogram dened by the vectors u and v; the ini al
point of u + v is the common ini al point of parallelogram, and the terminal
point of the sum is the common terminal point of the parallelogram.
While not illustrated here, the Head to Tail Rule and Parallelogram Law hold
for vectors in R as well.
It follows from the proper es of the real numbers and Deni on that
u v = u + ( )v.
The Parallelogram Law gives us a good way to visualize this subtrac on. We
demonstrate this in the following example.
Example
Vector Subtrac on
Let u = , and v = , . Compute and sketch u v.
The computa on of u v is straigh orward, and we show
S
all steps below. Usually the formal step of mul plying by ( ) is omi ed and
we just subtract.

u v = u + ( )v
= , + ,
= , .

u
4

.
Figure . : Illustra ng how to subtract
vectors graphically.

Figure . illustrates, using the Head to Tail Rule, how the subtrac on can be
viewed as the sum u + (v). The gure also illustrates how u v can be obtained by looking only at the terminal points of u andv (when their ini al points
are the same).
Example

Scaling vectors

. Sketch the vectors v = , and v with ini al point at the origin.

Notes:

.
. Compute the magnitudes of v and v.

An Introduc on to Vectors

. We compute v:

v = ,
= , .
Both v and v are sketched in Figure . . Make note that v does not
start at the terminal point of v; rather, its ini al point is also the origin.
. The gure suggests that v is twice as long as v. We compute their magnitudes to conrm this.

+
|| v || =

=
.

+
|| v || =

=
=
.
As we suspected, v is twice as long as v.
The zero vector is the vector whose ini al point is also its terminal point. It
is denoted by . Its component form, in R , is , ; in R , it is , , . Usually
the context makes is clear whether is referring to a vector in the plane or in
space.
Our examples have illustrated key principles in vector algebra: how to add
and subtract vectors and how to mul ply vectors by a scalar. The following theorem states formally the proper es of these opera ons.

Notes:

v
x

.
Figure . : Graphing vectors v and v
in Example
.

Chapter

Vectors

Theorem

Proper es of Vector Opera ons

The following are true for all scalars c and d, and for all vectors u, v and
, where u, v and w
are all in R or where u, v and w
are all in R :
w
. u + v = v + u

Commuta ve Property

= u + (v + w
)
. (u + v) + w

Associa ve Property

. v + = v

Addi ve Iden ty

. (cd)v = c(dv)
. c(u + v) = cu + cv

Distribu ve Property

. (c + d)v = cv + dv

Distribu ve Property

. v =
. || cv || = |c| || v ||
. || u || =

if, and only if, u = .

As stated before, each vector v conveys magnitude and direc on informaon. We have a method of extrac ng the magnitude, which we write as || v ||.
Unit vectors are a way of extrac ng just the direc on informa on from a vector.
Deni on

Unit Vector

A unit vector is a vector v with a magnitude of ; that is,


|| v || = .
Consider this scenario: you are given a vectorv and are told to create a vector
of length in the direc on of v. How does one do that? If we knew that u was
the unit vector in the direc on of v, the answer would be easy: u. So how do
we nd u ?
Property of Theorem
holds the key. If we divide v by its magnitude, it
becomes a vector of length . Consider:





v
(we can pull out
as it is a scalar)
|| v || = || v || || v ||
|| v ||
= .

Notes:

.
in the direc on ofv is

So the vector of length


this more clear.

|| v ||

An Introduc on to Vectors

v. An example will make

Using Unit Vectors


Example
= , , .
Let v = , and let w

. Find the unit vector in the direc on of v.

.
. Find the unit vector in the direc on of w
. Find the vector in the direc on of v with magnitude .
u

. We nd || v || =

. So the unit vector u in the direc on of v is

u = v = ,
.

|| = , so the unit vectorz in the direc on of w


is
. We nd || w

u = w
=
.
, ,
. To create a vector with magnitude
in the direc
on of v, we mul ply the

, /
is the vector we seek.
unit vector u by . Thus u =
/
This is sketched in Figure . .
The basic forma on of the unit vector u in the direc on of a vector v leads
to a interes ng equa on. It is:
v = || v ||

|| v ||

v.

We rewrite the equa on with parentheses to make a point:


)
(
v .
v = || v ||
|| v ||
| {z } | {z }
magnitude

direc on

This equa on illustrates the fact that a vector has both magnitude and direc on, where we view a unit vector as supplying only direc on informa on.
Iden fying unit vectors with direc on allows us to dene parallel vectors.

Notes:

v
u

.
Figure . : Graphing vectors in Example
. All vectors shown have their inial point at the origin.

Chapter

Vectors

Note: is direc onless; because || || =


, there is no unit vector in the direc on
of .
Some texts dene two vectors as being
parallel if one is a scalar mul ple of the
other. By this deni on, is parallel to
all vectors as = v for all v.
We prefer the given deni on of parallel
as it is grounded in the fact that unit vectors provide direc on informa on. One
may adopt the conven on that is parallel to all vectors if they desire. (See also
the marginal note on page
.)

Deni on

Parallel Vectors

. Unit vectors u and u are parallel if u = u .


. Nonzero vectors v and v are parallel if their respec ve unit vectors are parallel.
It is equivalent to say that vectors v and v are parallel if there is a scalar
such that v = cv (see marginal note).
If one graphed all unit vectors in R with the ini al point at the origin, then
the terminal points would all lie on the unit circle. Based on what we know from
trigonometry, we can then say that the component form of all unit vectors in R
is cos , sin for some angle .
A similar construc on in R shows that the terminal points all lie on the unit
sphere. These vectors also have a par cular component form, but its deriva on
is not as straigh orward as the one for unit vectors in R . Important concepts
about unit vectors are given in the following Key Idea.
c =

Key Idea

Unit Vectors

. The unit vector in the direc on of v is


u =

|| v ||

v.

. A vector u in R is a unit vector if, and only if, its component form
is cos , sin for some angle .
. A vector u in R is a unit vector if, and only if, its component form
is sin cos , sin sin , cos for some angles and .
These formulas can come in handy in a variety of situa ons, especially the
formula for unit vectors in the plane.
0

.
0lb

Finding Component Forces


Example
Consider a weight of lb hanging from two chains, as shown in Figure . .
One chain makes an angle of with the ver cal, and the other an angle of

. Find the force applied to each chain.


S

Figure . : A diagram of a weight hanging from chains in Example


.

Notes:

Knowing that gravity is pulling the

lb weight straight down,

An Introduc on to Vectors

we can create a vector F to represent this force.


F =

, = ,

We can view each chain as pulling the weight up, preven ng it from falling.
We can represent the force from each chain with a vector. Let F represent the
force from the chain making an angle of with the ver cal, and let F represent the force form the other chain. Convert all angles to be measured from the
horizontal (as shown in Figure . ), and apply Key Idea . As we do not yet
know the magnitudes of these vectors, (that is the problem at hand), we use m
and m to represent them.
F = m cos
F = m cos

, sin

, sin

As the weight is not moving, we know the sum of the forces is . This gives:

+ m cos

, sin

F + F + F =

+ m cos , sin =

The sum of the entries in the rst component is , and the sum of the entries
in the second component is also . This leads us to the following two equa ons:
m cos

+ m cos

m sin

+ m sin

This is a simple -equa on, -unkown system of linear equa ons. We leave it to
the reader to verify that the solu on is

. .
) . ;
m =
m = (
+
It might seem odd that the sum of the forces applied to the chains is more
than lb. We leave it to a physics class to discuss the full details, but oer this
short explana on. Our equa ons were established so that the ver cal components of each force sums to lb, thus suppor ng the weight. Since the chains
are at an angle, they also pull against each other, crea ng an addi onal horizontal force while holding the weight in place.
Unit vectors were very important in the previous calcula on; they allowed
us to dene a vector in the proper direc on but with an unknown magnitude.
Our computa ons were then computed componentwise. Because such calcula ons are o en necessary, the standard unit vectors can be useful.

Notes:

45

Figure . : A diagram of the force vectors from Example


.

Chapter

Vectors

Deni on

Standard Unit Vectors

. In R , the standard unit vectors are


i = ,

and j = , .

. In R , the standard unit vectors are


i = , ,

Example

and j = , ,

and k = , , .

Using standard unit vectors

. Rewrite v = , using the standard unit vectors.


= i j + k in component form.
. Rewrite w
S

v = ,

= , + ,
= , ,
= i j

.
Fw

25lb

= i j + k
w
= , , + , , + , ,
= , ,

These two examples demonstrate that conver ng between component form


and the standard unit vectors is rather straigh orward. Many mathema cians
prefer component form, and it is the preferred nota on in this text. Many engineers prefer using the standard unit vectors, and many engineering text use
that nota on.
Finding Component Force
Example
A weight of lb is suspended from a chain of length while a wind pushes the
weight to the right with constant force of lb as shown in Figure . . What
angle will the chain make with the ver cal as a result of the winds pushing?
How much higher will the weight be?

Figure . : A gure of a weight being


pushed by the wind in Example
.

Notes:

.
The force of the wind is represented by the vector Fw = i.
The force of gravity on the weight is represented by Fg = j. The direc on
and magnitude of the vector represen ng the force on the chain are both unknown. We represent this force with
S

Fc = m cos , sin = m cos i + m sin j


for some magnitude m and some angle with the horizontal . (Note: is the
angle the chain makes with the ver cal; is the angle with the horizontal.)
As the weight is at equilibrium, the sum of the forces is :
Fc + Fw + Fg =
m cos i + m sin j + i j =
Thus the sum of the i and j components are , leading us to the following
system of equa ons:

+ m cos =
+ m sin =

. )

This is enough to determine Fc already, as we know m cos = and


m sin =
. Thus Fc = , . We can use this to nd the magnitude
m:

=
. lb.
m = ( ) +

We can then use either equality from Equa on ( . ) to solve for . We choose
the rst equality as using arccosine will return an angle in the nd quadrant:
)
(

. .
+
cos =
= cos
Subtrac ng from this angle gives us an angle of . with the ver cal.
We can now use trigonometry to nd out how high the weight is li ed.
The diagram shows that a right triangle is formed with the
chain as the hypotenuse with an interior angle of . . The length of the adjacent side (in
the diagram, the dashed ver cal line) is cos . .
. Thus the weight
is li ed by about .
, almost / in.
The algebra we have applied to vectors is already demonstra ng itself to be
very useful. There are two more fundamental opera ons we can perform with
vectors, the dot product and the cross product. The next two sec ons explore
each in turn.

Notes:

An Introduc on to Vectors

Exercises

Terms and Concepts

. Name two dierent things that cannot be described with


just one number, but rather need or more numbers to
fully describe them.

. What is the dierence between ( , ) and , ?

. What is a unit vector?

. What does it mean for two vectors to be parallel?


. What eect does mul plying a vector by have?

Problems

In Exercises , points P and Q are given. Write the vector


#
PQ in component form and using the standard unit vectors.

Q = ( , )

. P = ( , , ), Q = ( , , )
. P = ( , , ),

Q=( , , )

. Let u = , and v = , .

(b) Sketch the above vectors on the same axes, along


with u and v.
(c) Find x where u + x = v x.

In Exercises

, nd || u ||, ||v ||, || u +v || and || u v ||.

. u = , , v = ,

. Let u = , , and v = , , .

(a) Find u + v, u v, u v.

. u = , , , v = , ,

(b) Sketch the above vectors on the same axes, along


with u and v.
(c) Find x where u + x = v + x.

(a) Find u + v, u v, u v.

In Exercises
axes.

. P = ( , ), Q = ( , )
. P = ( , ),

, sketch u, v, u + v and u v on the same

. u = , , v = ,
. u = , , , v =

. Under what condi ons is || u || + || v || = || u + v ||?


In Exercises
v.

, nd the unit vector u in the direc on of

. v = ,
. v = ,

. v = , ,
. v = , ,

. Find the unit vector in the rst quadrant of R that makes


a angle with the x-axis.

. Find the unit vector in the second quadrant of R that


makes a angle with the y-axis.

. =

, =

A weight of plb is suspended from a chain of length while


. Verify, from Key Idea , thatu = sin cos , sin sin , cos a constant force of Fw pushes the weight to the right, making
is a unit vector for all angles and .
an angle of with the ver cal, as shown in the gure below.
A weight of
lb is suspended from two chains, making angles with the ver cal of and as shown in the gure below.

.
100lb

In Exercises , angles and are given. Find the force


applied to each chain.

p lb

Fw

In Exercises
, a force Fw and length are given. Find
the angle and the height the weight is li ed as it moves to
the right.
. Fw = lb,

p = lb

p=

. =

. Fw = lb,

. =

. Fw = lb,

. =

. Fw =

lb,

lb

p = lb

,
,

p = lb

Chapter

Vectors

The Dot Product

The previous sec on introduced vectors and described how to add them together and how to mul ply them by scalars. This sec on introduces a mul plica on on vectors called the dot product.
Deni on

Dot Product

. Let u = u , u and v = v , v in R . The dot product of u and


v, denoted u v, is
u v = u v + u v .
. Let u = u , u , u and v = v , v , v in R . The dot product of
u and v, denoted u v, is
u v = u v + u v + u v .

Note how this product of vectors returns a scalar, not another vector. We
prac ce evalua ng a dot product in the following example, then we will discuss
why this product is useful.
Example

Evalua ng dot products

. Let u = , , v = , in R . Find u v.
. Let x = , , and y = , , in R . Find x y.
S

. Using Deni on

, we have
u v = ( ) + ( ) = .

. Using the deni on, we have


x y = ( ) ( ) + ( ) =

Notes:

The Dot Product

The dot product, as shown by the preceding example, is very simple to evaluate. It is only the sum of products. While the deni on gives no hint as to why
we would care about this opera on, there is an amazing connec on between
the dot product and angles formed by the vectors. Before sta ng this connecon, we give a theorem sta ng some of the proper es of the dot product.

Theorem

Proper es of the Dot Product

be vectors in R or R and let c be a scalar.


Let u, v and w
. u v = v u

) = u v + u w

. u (v + w

Commuta ve Property
Distribu ve Property

. c(u v) = (cu) v = u (cv)

. v =
. v v = || v ||

The last statement of the theorem makes a handy connec on between the
magnitude of a vector and the dot product with itself. Our deni on and theorem give proper es of the dot product, but we are s ll likely wondering What
does the dot product mean? It is helpful to understand that the dot product of
a vector with itself is connected to its magnitude.

(a)

The next theorem extends this understanding by connec ng the dot product
to magnitudes and angles. Given vectorsu andv in the plane, an angle is clearly
formed whenu andv are drawn with the same ini al point as illustrated in Figure
. (a). (We always take to be the angle in [ , ] as two angles are actually
created.)
The same is also true of vectors in space: given u andv in R with the same
ini al point, there is a plane that contains both u and v. (When u and v are colinear, there are innite planes that contain both vectors.) In that plane, we can
again nd an angle between them (and again, ). This is illustrated
in Figure . (b).
The following theorem connects this angle to the dot product of u and v.

Notes:

(b)
Figure
. : Illustra ng the angle
formed by two vectors with the same
ini al point.

Chapter

Vectors

Theorem

The Dot Product and Angles

Let u and v be vectors in R or R . Then


u v = || u || || v || cos ,
, is the angle between u and v.

where ,

When is an acute angle (i.e., < / ), cos is posi ve; when =


/ , cos = ; when is an obtuse angle (/ < ), cos is nega ve.
Thus the sign of the dot product gives a general indica on of the angle between
the vectors, illustrated in Figure . .
v

= /2

u v > 0

u v = 0

u v < 0

Figure . : Illustra ng the rela onship between the angle between vectors and the
sign of their dot product.

We can use Theorem to compute the dot product, but generally this theorem is used to nd the angle between known vectors (since the dot product is
generally easy to compute). To this end, we rewrite the theorems equa on as
cos =

u v
|| u |||| v ||

= cos

u v
|| u |||| v ||

We prac ce using this theorem in the following example.


y

Using the dot product to nd angles


Example
= , , as shown in Figure
Let u = , , v = , and w
angles , and .

Figure
.

: Vectors used in Example

Notes:

. Find the

We start by compu ng the magnitude of each vector.


|| u || =

|| v || =

|| = .
|| w

.
We now apply Theorem

The Dot Product

to nd the angles.
(

u v

(
)(
)

( ) = = .

= cos
= cos

v w

(
)( )
(
)

= cos
= cos
.

u w

(
)( )
(
)

= cos

= cos
.

. .

We see from our computa on that + = , as indicated by Figure . .


While we knew this should be the case, it is nice to see that this non-intui ve
formula indeed returns the results we expected.
We do a similar example next in the context of vectors in space.
Using the dot product to nd angles
Example
= , , , as illustrated in Figure
Let u = , , , v = , , and w
. . Find the angle between each pair of vectors.
S

. Between u and v:
= cos

= cos
=

Notes:

u v
|| u |||| v ||
(
)

)
Figure
.

: Vectors used in Example

Chapter

Vectors
:
. Between u and w
= cos
= cos
=

u w

||
|| u |||| w
(
)

:
. Between v and w
= cos
= cos
=

)
v w

||
|| v |||| w
)
(

While our work shows that each angle is / , i.e., , none of these angles
looks to be a right angle in Figure . . Such is the case when drawing three
dimensional objects on the page.

Note: The term perpendicular originally


referred to lines. As mathema cs progressed, the concept of being at right
angles to was applied to other objects,
such as vectors and planes, and the term
orthogonal was introduced. It is especially used when discussing objects that
are hard, or impossible, to visualize: two
vectors in -dimensional space are orthogonal if their dot product is . It is not
wrong to say they are perpendicular, but
common conven on gives preference to
the word orthogonal.

All three angles between these vectors was / , or . We know from


geometry and everyday life that angles are nice for a variety of reasons,
so it should seem signicant that these angles are all / . No ce the common
feature in each calcula on (and also the calcula on of in Example
): the
dot products of each pair of angles was . We use this as a basis for a deni on
of the term orthogonal, which is essen ally synonymous to perpendicular.
Deni on

Orthogonal

Vectors u and v are orthogonal if their dot product is .


Finding orthogonal vectors
Example
Let u = , and v = , , .
. Find two vectors in R that are orthogonal to u.
. Find two nonparallel vectors in R that are orthogonal to v.
S

Notes:

The Dot Product

. Recall that a line perpendicular to a line with slope m has slope /m,
the opposite reciprocal slope. We can think of the slope of u as / , its
rise over run. A vector orthogonal to u will have slope / . There are
many such choices, though all parallel:
,

or

or

or

, , etc.

. There are innite direc ons in space orthogonal to any given direc on,
so there are an innite number of nonparallel vectors orthogonal to v.
Since there are so many, we have great leeway in nding some.
One way is to arbitrarily pick values for the rst two components, leaving
the third unknown. For instance, letv = , , z. Ifv is to be orthogonal
to v, then v v = , so
+

+ z=

z=

Sov = , , / is orthogonal tov. We can apply a similar technique


by leaving the rst or second component unknown.
Another method of nding a vector orthogonal to v mirrors what we did
in part . Letv = , , . Here we switched the rst two components
ofv, changing the sign of one of them (similar to the opposite reciprocal
concept before). Le ng the third component be eec vely ignores the
third component of v, and it is easy to see that

v v = , , , , = .
.

Clearly v and v are not parallel.


An important construc on is illustrated in Figure . , where vectors u and
v are sketched. In part (a), a do ed line is drawn from the p of u to the line
containing v, where the do ed line is orthogonal to v. In part (b), the do ed
is formed, parallel to v. It is clear by the
line is replaced with the vectorz and w
+ z. What is important about this construc on is this: u is
diagram that u = w
decomposed as the sum of two vectors, one of which is parallel tov and one that
is perpendicular to v. It is hard to overstate the importance of this construc on
(as well see in upcoming examples).
, z and u as shown in Figure . (b) form a right triangle,
The vectors w
in terms of v and
where the angle between v and u is labeled . We can nd w
u.
Using trigonometry, we can state that
|| = || u || cos .
|| w

Notes:

. )

(a)
u

(b)

Figure . : Developing the construcon of the orthogonal projec on.

Chapter

Vectors

is parallel to to v ; that is, the direc on of w


is the
We also know that w
is the vector in
direc on of v, described by the unit vector || v ||v. The vector w
the direc on || v ||v with magnitude || u || cos :
(
)
= || u || cos
w
Replace cos using Theorem

|| v ||

u v
= || u ||
|| u |||| v ||
u v
v.
=
|| v ||
Now apply Theorem

v.

|| v ||

.
=

u v
v.
v v

Since this construc on is so important, it is given a special name.


Deni on

Orthogonal Projec on

Let u and v be given. The orthogonal projec on of u onto v, denoted


proj v u, is
u v
v.
proj v u =
v v
Example

Compu ng the orthogonal projec on

. Let u = , and v = , . Find proj v u, and sketch all three vectors


with ini al points at the origin.
= , , and x = , , . Find proj x w
, and sketch all three
. Let w
vectors with ini al points at the origin.
S

Notes:

.
. Applying Deni on

The Dot Product

, we have
u v
v
v v

,
=

.
= ,

proj v u =

v
x

projv u

Vectors u, v and proj v u are sketched in Figure . (a). Note how the
projec on is parallel tov; that is, it lies on the same line through the origin
asv, although it points in the opposite direc on. That is because the angle
between u and v is obtuse (i.e., greater than ).

(a)

. Apply the deni on:


=
projx w
=

x
w
x
x x
, ,

= , , .

(b)

These vectors are sketched in Figure . (b), and again in part (c) from
a dierent perspec ve. Because of the nature of graphing these vectors,
the sketch in part (b) makes it dicult to recognize that the drawn projecon has the geometric proper es it should. The graph shown in part (c)
illustrates these proper es be er.
Consider Figure . where the concept of the orthogonal projec on is
again illustrated. It is clear that
u = proj v u +z.

. )

(c)

As we know what u and proj v u are, we can solve forz and state that
z = u proj v u.
This leads us to rewrite Equa on (

. ) in a seemingly silly way:

Figure . : Graphing the vectors used


in Example
.

u = proj v u + (u proj v u).

This is not nonsense, as pointed out in the following Key Idea. (Nota on note:
the expression y means is parallel to y. We can use this nota on to state

Notes:

projv u

Figure . : Illustra ng the orthogonal


projec on.

Chapter

Vectors
x y which means x is parallel to y. The expression y means is orthogonal to y, and is used similarly.)
Key Idea

Orthogonal Decomposi on of Vectors

Let u and v be given. Then u can be wri en as the sum of two vectors,
one of which is parallel to v, and one of which is orthogonal to v:
u = proj v u + (u proj v u).
| {z }
| {z }
v

We illustrate the use of this equality in the following example.


Orthogonal decomposi on of vectors

Example

. Let u = , and v = , as in Example


. Decompose u as the
sum of a vector parallel to v and a vector orthogonal to v.
= , , and x = , , as in Example
as
. Let w
. Decompose w
the sum of a vector parallel to x and a vector orthogonal to x.
S

. In Example

, we found that proj v u = . , . . Let

z = u proj v u = , . , . = . , . .
Is z orthogonal to v ? (I.e, is z v ?) We check for orthogonality with the
dot product:
z v = . , . , = .
Since the dot product is , we knowz v. Thus:

u = proj v u + (u proj v u)
, = . , . + . , . .
|
{z
}
| {z }
v

. We found in Example
we have:

= , , . Applying the Key Idea,


that projx w

= , , , , = , , .
proj x w
z = w

Notes:

The Dot Product

We check to see ifz x:

z x = , , , , = .

Since the dot product is , we know the two vectors are orthogonal. We
as the sum of two vectors, one parallel and one orthogonal
now write w
to x:
= projx w
+ (
)
w
w projx w

, , = , , + , ,
| {z }
| {z }
x

We give an example of where this decomposi on is useful.


Orthogonally decomposing a force vector
Example
Consider Figure . (a), showing a box weighing lb on a ramp that rises
over a span of
. Find the components of force, and their magnitudes, ac ng
on the box (as sketched in part (b) of the gure):
. in the direc on of the ramp, and
. orthogonal to the ramp.
As the ramp rises
over a horizontal distance of
, we can
represent the direc on of the ramp with the vector r = , . Gravity pulls
down with a force of lb, which we represent with g = , .
S

. To nd the force of gravity in the direc on of the ramp, we compute projr g:


projr g =
=
=

g r
r
r r

,
,

, .

20

(a)

. To nd the component z of gravity orthogonal to the ramp, we use Key


Idea .

Notes:

The magnitude of projr g is || projr g || = /


. lb. Though
the box weighs lb, a force of about lb is enough to keep the box from
sliding down the ramp.

z = g projr g

,
=

z
20
g

projr g

(b)
Figure . : Sketching the ramp and box
in Example
. Note: The vectors are not
drawn to scale.

Chapter

Vectors
The magnitude of this force is ||z || . lb. In physics and engineering,
knowing this force is important when compu ng things like sta c fric onal
force. (For instance, we could easily compute if the sta c fric onal force
alone was enough to keep the box from sliding down the ramp.)

Applica on to Work
In physics, the applica on of a force F to move an object in a straight line a
distance d produces work; the amount of work W is W = Fd, (where F is in the
direc on of travel). The orthogonal projec on allows us to compute work when
the force is not in the direc on of travel.
Consider Figure . , where a force F is being applied to an object moving
in the direc on of d. (The distance the object travels is the magnitude of d.) The
work done is the amount of force in the direc on of d, || proj d F ||, mes || d ||:
F



F d


d || d ||
|| proj d F || || d || =
d d


F d


=
|| d || || d ||
|| d ||



F d
|| d ||
=
|| d ||




= F d .

proj d F

Figure
. : Finding work when the
force and direc on of travel are given as
vectors.

The expression F d will be posi ve if the angle between F and d is acute;


when the angle is obtuse (hence F d is nega ve), the force is causing mo on
in the opposite direc on of d, resul ng in nega ve work. We want to capture
this sign, so we drop the absolute value and nd that W = F d.
Deni on

Work

Let F be a constant force that moves an object in a straight line from point
#
P to point Q. Let d = PQ. The work W done by F along d is W = F d.
Compu ng work
Example
A man slides a box along a ramp that rises
over a distance of
lb of force as shown in Figure . . Compute the work done.

Notes:

by applying

The Dot Product

The gure indicates that the force applied makes a angle with the horizontal, so F =
cos , sin . , . The ramp is
represented by d = , . The work done is simply
S

F d =

cos

, sin

Note how we did not actually compute the distance the object traveled, nor
the magnitude of the force in the direc on of travel; this is all inherently computed by the dot product!
The dot product is a powerful way of evalua ng computa ons that depend
on angles without actually using angles. The next sec on explores another product on vectors, the cross product. Once again, angles play an important role,
though in a much dierent way.

Notes:

lb.

30

3
15

Figure . : Compu ng work when sliding a box up a ramp in Example


.

Exercises

Terms and Concepts


. The dot product of two vectors is a

, not a vector.

. How are the concepts of the dot product and vector magnitude related?

In Exercises
, vectors u and v are given. Find projv u,
the orthogonal projec on of u onto v, and sketch all three
vectors on the same axes.
. u = , , v = ,
. u = , , v = ,

. How can one quickly tell if the angle between two vectors
is acute or obtuse?

. u = , , v = ,

. Give a synonym for orthogonal.

. u = , , v = ,
. u = , , , v = , ,

Problems
In Exercises

. u = , , , v = , ,
, nd the dot product of the given vectors.

. u = , , v = ,

In Exercises , vectors u and v are given. Write u as the


sum of two vectors, one of which is parallel to v and one of
which is perpendicular to v. Note: these are the same pairs
of vectors as found in Exercises .

. u = , , v = ,

. u = , , v = ,

. u = , , , v = , ,

. u = , , v = ,

. u = , , , v = , ,

. u = , , v = ,

. u = , , v = , ,

. u = , , v = ,

. u = , , , v = , ,

. u = , , , v = , ,

in R and show that


. Create your own vectors u, v and w
u (v + w
) = u v + u w
.
. Create your own vectorsu andv in R and scalar c and show
that c(u v) = u (cv).
In Exercises
, nd the measure of the angle between
the two vectors in both radians and degrees.
. u = , , v = ,
. u = , , v = ,
. u = , , , v = , ,
. u = , , , v = , ,
In Exercises , a vectorv is given. Give two vectors that
are orthogonal to v.
. v = ,
. v = ,
. v = , ,
. v = , ,

. u = , , , v = , ,
. A

lb box sits on a ramp that rises


over a distance of
. How much force is required to keep the box from sliding down the ramp?

. A lb box sits on a
ramp that makes a angle with
the horizontal. How much force is required to keep the box
from sliding down the ramp?
. How much work is performed in moving a box horizontally
with a force of lb applied at an angle of to the
horizontal?
. How much work is performed in moving a box horizontally
with a force of lb applied at an angle of to the
horizontal?
. How much work is performed in moving a box up the length
of a ramp that rises
over a distance of
, with a force
of lb applied horizontally?
. How much work is performed in moving a box up the length
of a ramp that rises
over a distance of
, with a force
of lb applied at an angle of to the horizontal?
. How much work is performed in moving a box up the length
of a
ramp that makes a angle with the horizontal,
with lb of force applied in the direc on of the ramp?

The Cross Product

Orthogonality is immensely important. A quick scan of your current environment will undoubtedly reveal numerous surfaces and edges that are perpendicular to each other (including the edges of this page). The dot product provides
a quick test for orthogonality: vectors u and v are perpendicular if, and only if,
u v = .
Given two nonparallel, nonzero vectors u and v in space, it is very useful
that is perpendicular to both u and v. There is a opera on,
to nd a vector w
called the cross product, that creates such a vector. This sec on denes the
cross product, then explores its proper es and applica ons.
Deni on

Cross Product

Let u = u , u , u and v = v , v , v be vectors in R . The cross


product of u and v, denoted u v, is the vector
u v = u v u v , (u v u v ), u v u v .
This deni on can be a bit cumbersome to remember. A er an example we
will give a convenient method for compu ng the cross product. For now, careful
examina on of the products and dierences given in the deni on should reveal
a pa ern that is not too dicult to remember. (For instance, in the rst component only and appear as subscripts; in the second component, only and
appear as subscripts. Further study reveals the order in which they appear.)
Lets prac ce using this deni on by compu ng a cross product.
Example
Compu ng a cross product
Let u = , , and v = , , . Find u v, and verify that it is orthogonal
to both u and v.
Using Deni on , we have
(
)

u v = ( ) ( ) , ( ) ( ) , ( ) ( ) =
S

, , .

(We encourage the reader to compute this product on their own, then verify
their result.)
We test whether or not u v is orthogonal to u andv using the dot product:
(
)
u v u = , , , , = ,
(
)
u v v = , , , , = .

Since both dot products are zero, u v is indeed orthogonal to both u and v.
Notes:

The Cross Product

Chapter

Vectors
A convenient method of compu ng the cross product starts with forming a
par cular matrix, or rectangular array. The rst row comprises the standard unit vectorsi, j, and k. The second and third rows are the vectors u and v,
respec vely. Using u and v from Example
, we begin with:
i

Now repeat the rst two columns a er the original three:


i

k i

This gives three full upper le to lower right diagonals, and three full upper right to lower le diagonals, as shown. Compute the products along each
diagonal, then add the products on the right and subtract the products on the
le :

(
) (
)
u v = i + j + k k + i + j = i + j + k =
We prac ce using this method.

, , .

Compu ng a cross product


Example
Let u = , , and v = , , . Compute both u v and v u.
To compute u v, we form the matrix as prescribed above,
S
complete with repeated rst columns:
i

We let the reader compute the products of the diagonals; we give the result:
(
) (
)
u v = i j + k k + i +j = , , .
Notes:

.
To compute v u, we switch the second and third rows of the above matrix,
then mul ply along diagonals and subtract:
i

Note how with the rows being switched, the products that once appeared on
the right now appear on the le , and viceversa. Thus the result is:
v u =

) (
)
i +j k k + i j = , , ,

which is the opposite of u v. We leave it to the reader to verify that each of


these vectors is orthogonal to u and v.

Proper es of the Cross Product


It is not coincidence that v u = (u v) in the preceding example; one
can show using Deni on
that this will always be the case. The following
theorem states several useful proper es of the cross product, each of which can
be veried by referring to the deni on.
Theorem

Proper es of the Cross Product

be vectors in R and let c be a scalar. The following iden


Let u, v and w
hold:
. u v = (v u)
.

= u w
+ v w

(a) (u + v) w

An commuta ve Property
Distribu ve Proper es

) = u v + u w

(b) u (v + w
. c(u v) = (cu) v = u (cv)
.

(a) (u v) u =

Orthogonality Proper es

(b) (u v) v =
. u u =
. u =

) = (u v) w

. u (v w

Notes:

es

Triple Scalar Product

The Cross Product

Chapter

Vectors
We introduced the cross product as a way to nd a vector orthogonal to
two given vectors, but we did not give a proof that the construc on given in
Deni on sa ses this property. Theorem asserts this property holds; we
leave it as a problem in the Exercise sec on to verify this.
Property from the theorem is also le to the reader to prove in the Exercise
sec on, but it reveals something more interes ng than the cross product of a
vector with itself is . Let u andv be parallel vectors; that is, let there be a scalar
c such that v = cu. Consider their cross product:
u v = u (cu)
= c(u u)
= .

(by Property of Theorem


(by Property of Theorem

)
)

We have just shown that the cross product of parallel vectors is . This hints
at something deeper. Theorem
related the angle between two vectors and
their dot product; there is a similar rela onship rela ng the cross product of two
vectors and the angle between them, given by the following theorem.
Note: Deni on
(through Theorem
) denes u and v to be orthogonal if
u v = . We could use Theorem
to
dene u andv are parallel if u v = . By
such a deni on, would be both orthogonal and parallel to every vector. Apparent paradoxes such as this are not uncommon in mathema cs and can be very useful. (See also the marginal note on page
.)

Theorem

The Cross Product and Angles

Let u and v be vectors in R . Then


|| u v || = || u || || v || sin ,
where ,

, is the angle between u and v.

Note that this theorem makes a statement about the magnitude of the cross
product. When the angle between u andv is or (i.e., the vectors are parallel),
the magnitude of the cross product is . The only vector with a magnitude of is
(see Property of Theorem ), hence the cross product of parallel vectors is .
We demonstrate the truth of this theorem in the following example.
The cross product and angles
Example
Let u = , , and v = , , as in Example
. Verify Theorem
nding , the angle between u and v, and the magnitude of u v.

Notes:

by

.
We use Theorem

= cos
= cos
.

The Cross Product

to nd the angle between u and v.


)
u v
|| u || || v ||
(
)

(

showed thatu v = , , , hence ||u v || =


Our work in Example
. Is || u v || = || u || ||v || sin ? Using numerical approxima ons, we nd:

|| u v || =
|| u || || v || sin =
sin .

Numerically, they seem equal. Using a right triangle, one can show that

))
(
(


= ,
sin cos
which allows us to verify the theorem exactly.
Right Hand Rule
The an commuta ve property of the cross product demonstrates that u v
andvu dier only by a sign these vectors have the same magnitude but point
in the opposite direc on. When seeking a vector perpendicular to u and v, we
essen ally have two direc ons to choose from, one in the direc on of u v and
one in the direc on of v u. Does it ma er which we choose? How can we tell
which one we will get without graphing, etc.?
Another wonderful property of the cross product, as dened, is that it follows the right hand rule. Given u and v in R with the same ini al point, point
the index nger of your right hand in the direc on of u and let your middle nger
point in the direc on ofv (much as we did when establishing the right hand rule
for the -dimensional coordinate system). Your thumb will naturally extend in
the direc on of u v. One can prac ce this using Figure . . If you switch,
and point the index nder in the direc on of v and the middle nger in the direc on of u, your thumb will now point in the opposite direc on, allowing you
to visualize the an commuta ve property of the cross product.

Applica ons of the Cross Product


There are a number of ways in which the cross product is useful in mathema cs, physics and other areas of science beyond just nding a vector perpendicular to two others. We highlight a few here.

Notes:

Figure . : Illustra ng the Right Hand


Rule of the cross product.

Chapter

Vectors
Area of a Parallelogram
It is a standard geometry fact that the area of a parallelogram is A = bh,
where b is the length of the base and h is the height of the parallelogram, as
illustrated in Figure . (a). As shown when dening the Parallelogram Law of
vector addi on, two vectors u and v dene a parallelogram when drawn from
the same ini al point, as illustrated in Figure . (b). Trigonometry tells us that
h = || u || sin , hence the area of the parallelogram is

(a)

A = || u || || v || sin = || u v ||,

where the second equality comes from Theorem


on ( . ) in the following example.

Example

. )

. We illustrate using Equa-

Finding the area of a parallelogram

(b)

. Find the area of the parallelogram dened by the vectors u = , and


v = , .

Figure . : Using the cross product to


nd the area of a parallelogram.

. Verify that the points A = ( , , ), B = ( , , ), C = ( , , ) and


D = ( , , ) are the ver ces of a parallelogram. Find the area of the
parallelogram.

u
x

.
(a)

. Figure . (a) sketches the parallelogram dened by the vectors u and


v. We have a slight problem in that our vectors exist in R , not R , and
the cross product is only dened on vectors in R . We skirt this issue by
viewingu andv as vectors in the xy plane of R , and rewrite them asu =
, , and v = , , . We can now compute the cross product. It is
easy to show thatu v = , , ; therefore the area of the parallelogram
is A = || u v || = .
. To show that the quadrilateral ABCD is a parallelogram (shown in Figure
. (b)), we need to show that the opposite sides are parallel. We can
# #
# #
quickly show that AB = DC = , , and BC = AD = , , . We nd
#
#
the area by compu ng the magnitude of the cross product of AB and BC:

# #
# #
AB BC = , , || AB BC || =
.
.

This applica on is perhaps more useful in nding the area of a triangle (in
short, triangles are used more o en than parallelograms). We illustrate this in
the following example.

(b)
Figure
. : Sketching the parallelograms in Example
.

Notes:

.
Example
Area of a triangle
Find the area of the triangle with ver ces A = ( , ), B = ( , ) and C = ( , ),
as pictured in Figure . .
We found the area of this triangle in Example
to be .
S
using integra on. There we discussed the fact that nding the area of a triangle
can be inconvenient using the bh formula as one has to compute the height,
which generally involves nding angles, etc. Using a cross product is much more
direct.
We can choose any two sides of the triangle to use to form vectors; we
#
#
choose AB = , and AC = , . As in the previous example, we will
rewrite these vectors with a third component of so that we can apply the cross
product. The area of the triangle is
# #
|| AB AC || =

|| , , , , || =

|| , , || =

The Cross Product

y
B

Figure . : Finding the area of a triangle in Example


.

.
Note: The word parallelepiped is pronounced parallelehpipeed.

We arrive at the same answer as before with less work.


Volume of a Parallelepiped
The three dimensional analogue to the parallelogram is the parallelepiped.
Each face is parallel to the face opposite face, as illustrated in Figure . . By
, one gets a vector whose magnitude is the area of the base.
crossing v and w
Do ng this vector with u computes the volume of parallelepiped! (Up to a sign;
take the absolute value.)
is
Thus the volume of a parallelepiped dened by vectors u, v and w
)|.
V = |u (v w

. )

Note how this is the Triple Scalar Product, rst seen in Theorem . Applying
the iden es given in the theorem shows that we can apply the Triple Scalar
Product in any order we choose to nd the volume. That is,
)| = |u (
|,
V = |u (v w
w v)| = |(u v) w

Figure . : A parallelepiped is the three


dimensional analogue to the parallelogram.

etc.

Finding the volume of parallelepiped


Example
Find the volume of the parallepiped dened by the vectors u = , , , v =
= , , .
, , and w
S

Then

We apply Equa on (

= , , .
. ). We rst nd v w

)| = | , , , , | = .
|u (v w

So the volume of the parallelepiped is cubic units.

Notes:

Figure
.

: A parallelepiped in Example

Chapter

Vectors
While this applica on of the Triple Scalar Product is interes ng, it is not used
all that o en: parallelepipeds are not a common shape in physics and engineering. The last applica on of the cross product is very applicable in engineering.
Torque
Torque is a measure of the turning force applied to an object. A classic scenario involving torque is the applica on of a wrench to a bolt. When a force is
applied to the wrench, the bolt turns. When we represent the force and wrench
with vectors F and , we see that the bolt moves (because of the threads) in a direc on orthogonal to F and . Torque is usually represented by the Greek le er
, or tau, and has units of Nm, a Newtonmeter, or lb, a footpound.
While a full understanding of torque is beyond the purposes of this book,
when a force F is applied to a lever arm , the resul ng torque is
= F.

F
F
90

60

. )

Compu ng torque
Example
A lever of length makes an angle with the horizontal of . Find the resul ng
torque when a force of lb is applied to the end of the level where:
. the force is perpendicular to the lever, and

. the force makes an angle of


Figure . : Showing a force being applied to a lever in Example
.

with the lever, as shown in Figure

. We start by determining vectors for the force and lever arm. Since the
lever arm makes a angle with the
long, we can
and is
horizontal
state that = cos , sin =
,
.
Since the force vector is perpendicular to the lever arm (as seen in the
le hand side of Figure . ), we can conclude it is making an angle of
with the horizontal. As it has
a magnitude
of lb, we can state

F = cos( ), sin( ) =
,
.
Using Equa on ( . ) to nd the torque requires a cross product. We
again let the third component of each vector be and compute the cross
product:
= F


,
,

,
,
=
= , ,

Notes:

.
This clearly has a magnitude of

-lb.

We can view the force and lever arm vectors as lying on the page; our
computa on of shows that the torque goes into the page. This follows
the Right Hand Rule of the cross product, and it also matches well with the
example of the wrench turning the bolt. Turning a bolt clockwise moves
it in.

,
. As our force
. Our lever arm can s ll be represented by =
vector makes a angle with , we can see (referencing the right hand
side of the gure) that F makes a angle with the horizontal. Thus

( +
( +
)
)
F = cos , sin =

,
.

, .

We again make the third component and take the cross product to nd
the torque:
= F

( + )
)
( +

,
,

=
,
,

= , ,
, ,

As one might expect, when the force and lever arm vectors are orthogonal, the magnitude of force is greater than when the vectors are not orthogonal.
While the cross product has a variety of applica ons (as noted in this chapter), its fundamental use is nding a vector perpendicular to two others. Knowing a vector is orthogonal to two others is of incredible importance, as it allows
us to nd the equa ons of lines and planes in a variety of contexts. The importance of the cross product, in some sense, relies on the importance of lines and
planes, which see widespread use throughout engineering, physics and mathema cs. We study lines and planes in the next two sec ons.

Notes:

The Cross Product

Exercises

Terms and Concepts

In Exercises
, nd the area of the parallelogram dened by the given vectors.

. The cross product of two vectors is a


scalar.

, not a

. One can visualize the direc on of u v using the


.
. Give a synonym for orthogonal.
. T/F: A fundamental principle of the cross product is that
u v is orthogonal to u and v.
.

object.

is a measure of the turning force applied to an

. u = , , , v = , ,
. u = , , , v = , ,
. u = , , v = ,
. u = , , v = ,
In Exercises

given ver ces.

, nd the area of the triangle with the

. Ver ces: ( , , ), ( , , ) and ( , , ).


. Ver ces: ( , , ), ( , , ) and ( , , ).

Problems
In Exercises , vectors u and v are given. Compute u v
and show this is orthogonal to both u and v.
. u = , , , v = , ,

. Ver ces: ( , ), ( , ) and ( , ).


. Ver ces: ( , ), ( , ) and ( , ).
In Exercises
, nd the area of the quadrilateral with
the given ver ces. (Hint: break the quadrilateral into triangles.)

. u = , , , v = , ,
. u = , , , v = , ,

. Ver ces: ( , ), ( , ), ( , ) and ( , ).

. u = , ,

. Ver ces: ( , , ), ( , , ), ( , , ) and ( , , ).

, v = , ,

. u = , , , v = , ,
. u = , , , v = ,

In Exercises
, nd the volume of the parallelepiped
dened by the given vectors.
,

. u = i, v = j

. u = , , , v = , , ,
. u = , , , v = , , ,

. u = i, v = k

In Exercises
and v.

. u = j, v = k
in R and show thatu(v+
. Pick any vectorsu,v and w
w) =
u v + u w
.
in R and show thatu (v
. Pick any vectorsu,v and w
w) =
.
(u v) w
In Exercises , the magnitudes of vectors u and v in R
are given, along with the angle between them. Use this informa on to nd the magnitude of u v.
. || u || = , || v || = ,

. || u || = , || v || = ,

= /

. || u || = , || v || = ,

. || u || = , || v || = ,

= /

= , ,
w
= , ,
w

, nd a unit vector orthogonal to both u

. u = , , , v = , ,
. u = , , , v = , ,
. u = , , , v = , ,
. u = , , , v = , ,
. A bicycle rider applies
lb of force, straight down,
onto a pedal that extends in horizontally from the
cranksha . Find the magnitude of the torque applied to
the cranksha .
. A bicycle rider applies
lb of force, straight down, onto
a pedal that extends in from the cranksha , making a
angle with the horizontal. Find the magnitude of the torque
applied to the cranksha .

. To turn a stubborn bolt, lb of force is applied to a in


wrench. What is the maximum amount of torque that can
be applied to the bolt?

. Show, using the deni on of the Cross Product, that u (u


v) = ; that is, that u is orthogonal to the cross product of
u and v.

. To turn a stubborn bolt, lb of force is applied to a in


wrench in a conned space, where the direc on of applied force makes a angle with the wrench. How much
torque is subsequently applied to the wrench?

. Show, using the deni on of the Cross Product, thatuu =


.

Chapter

Vectors

Lines

To nd the equa on of a line in the x-y plane, we need two pieces of informa on:
a point and the slope. The slope conveys direc on informa on. As ver cal lines
have an undened slope, the following statement is more accurate:
To dene a line, one needs a point on the line and the direc on of
the line.
This holds true for lines in space.

Figure

: Dening a line in space.

Let P be a point in space, let p be the vector with ini al point at the origin
and terminal point at P (i.e., p points to P), and let d be a vector. Consider the
points on the line through P in the direc on of d.
Clearly one point on the line is P; we can say that the vector p lies at this
point on the line. To nd another point on the line, we can start at p and move
in a direc on parallel to d. For instance, star ng at p and traveling one length of
d places one at another point on the line. Consider Figure . where certain
points along the line are indicated.
The gure illustrates how every point on the line can be obtained by star ng
with p and moving a certain distance in the direc on of d. That is, we can dene
the line as a func on of t:
(t) = p + t d.
( . )
In many ways, this is not a new concept. Compare Equa on (
familiar y = mx + b equa on of a line:
Star ng
Point

. ) to the

Direc on

y = b + mx

(t) = p + t d
How Far To
Go In That
Direc on

Figure

: Understanding the vector equa on of a line.

The equa ons exhibit the same structure: they give a star ng point, dene
a direc on, and state how far in that direc on to travel.
Equa on ( . ) is an example of a vectorvalued func on; the input of the
func on is a real number and the output is a vector. We will cover vectorvalued
func ons extensively in the next chapter.

Notes:

.
There are other ways to represent a line. Let p = x , y , z and let d =
a, b, c. Then the equa on of the line through p in the direc on of d is:
(t) = p + td
= x , y , z + t a, b, c
= x + at, y + bt, z + ct .
The last line states the the x values of the line are given by x = x + at, the
y values are given by y = y + bt, and the z values are given by z = z + ct.
These three equa ons, taken together, are the parametric equa ons of the line
through p in the direc on of d.
Finally, each of the equa ons for x, y and z above contain the variable t. We
can solve for t in each equa on:
x = x + at

y = y + bt

z = z + ct

xx
,
a
yy
t=
,
b
zz
,
t=
c
t=

assuming a, b, c = . Since t is equal to each expression on the right, we can set


these equal to each other, forming the symmetric equa ons of the line through
p in the direc on of d:
xx
yy
zz
=
=
.
a
b
c
Each representa on has its own advantages, depending on the context. We
summarize these three forms in the following deni on, then give examples of
their use.

Notes:

Lines

Chapter

Vectors

Deni on

Equa ons of Lines in Space

Consider the line in space that passes through p = x , y , z in the


direc on of d = a, b, c .
. The vector equa on of the line is
(t) = p + td.
. The parametric equa ons of the line are
x = x + at,

y = y + bt,

z = z + ct.

. The symmetric equa ons of the line are


xx
yy
zz
=
=
.
a
b
c

Finding the equa on of a line


Example
Give all three equa ons, as given in Deni on , of the line through P = ( , , )
in the direc on of d = , , . Does the point Q = ( , , ) lie on this line?
We iden fy the point P = ( , , ) with the vector p =
S
, , . Following the deni on, we have
the vector equa on of the line is (t) = , , + t , , ;
the parametric equa ons of the line are
x=
Figure
.

: Graphing a line in Example

t,

y=

+ t,

z=

+ t; and

the symmetric equa ons of the line are


x

The rst two equa ons of the line are useful when a t value is given: one
can immediately nd the corresponding point on the line. These forms are good
when calcula ng with a computer; most so ware programs easily handle equaons in these formats. (For instance, to make Figure . , a certain graphics
program was given the input (2-x,3+x,1+2*x). This par cular program requires the variable always be x instead of t).

Notes:

Lines

Does the point Q = ( , , ) lie on the line? The graph in Figure .


makes it clear that it does not. We can answer this ques on without the graph
using any of the three equa on forms. Of the three, the symmetric equa ons
are probably best suited for this task. Simply plug in the values of x, y and z and
see if equality is maintained:
? ?
=
=

= = . .

We see that Q does not lie on the line as it did not sa sfy the symmetric equaons.
Finding the equa on of a line through two points
Example
Find the parametric equa ons of the line through the points P = ( , , ) and
Q = ( , , ).
Recall the statement made at the beginning of this sec on:
S
to nd the equa on of a line, we need a point and a direc on. We have two
points; either one will suce. The direc on of the line can be found by the
#
vector with ini al point P and terminal point Q: PQ = , , .
#
The parametric equa ons of the line through P in the direc on of PQ are:
:

x=

y= + t

z=

t.

A graph of the points and line are given in Figure . . Note how in the
given parametriza on of the line, t = corresponds to the point P, and t =
corresponds to the point Q. This relates to the understanding of the vector equaon of a line described in Figure . . The parametric equa ons start at the
#
point P, and t determines how far in the direc on of PQ to travel. When t = ,
#
#
we travel lengths of PQ; when t = , we travel one length of PQ, resul ng in
the point Q.

Parallel, Intersec ng and Skew Lines


In the plane, two dis nct lines can either be parallel or they will intersect
at exactly one point. In space, given equa ons of two lines, it can some mes
be dicult to tell whether the lines are dis nct or not (i.e., the same line can be
represented in dierent ways). Given lines (t) = p +td and (t) = p +td ,
we have four possibili es: and are
the same line
intersec ng lines
parallel lines
skew lines

Notes:

they share all points;


share only point;
d d , no points in common; or
d d , no points in common.

Figure .
ple
.

: A graph of the line in Exam-

Chapter

Vectors
The next two examples inves gate these possibili es.
Example
Comparing lines
Consider lines and , given in parametric equa on form:
x
: y
z

=
=
=

+ t
t
t

x =
: y =
z =

+ s
+s
+ s.

Determine whether and are the same line, intersect, are parallel, or skew.
We start by looking at the direc ons of each line. Line
S
has the direc on given by d = , , and line has the direc on given by
d = , , . It should be clear that d and d are not parallel, hence and
are not the same line, nor are they parallel. Figure . veries this fact (where
the points and direc ons indicated by the equa ons of each line are iden ed).
We next check to see if they intersect (if they do not, they are skew lines).
To nd if they intersect, we look for t and s values such that the respec ve x, y
and z values are the same. That is, we want s and t such that:
+ t =
t =
t
=

Figure
ample

: Sketching the lines from Ex-

+ s
+s
+ s.

This is a rela vely simple system of linear equa ons. Since the last equa on is
already solved for t, subs tute that value of t into the equa on above it:
( + s) =

+s

s= , t= .

A key to remember is that we have three equa ons; we need to check if s =


, t = sa ses the rst equa on as well:
+ ( ) = + ( ).
It does not. Therefore, we conclude that the lines and are skew.
Comparing lines
Example
Consider lines and , given in parametric equa on form:
x
: y
z

=
=
=

. + . t
. + . t
. . t

x
: y
z

=
=
=

. . s
. . s
. + . s.

Determine whether and are the same line, intersect, are parallel, or skew.

Notes:

Lines

It is obviously very dicult to simply look at these equa ons


S
and discern anything. This is done inten onally. In the real world, most equaons that are used do not have nice, integer coecients. Rather, there are lots
of digits a er the decimal and the equa ons can look messy.
We again start by deciding whether or not each line has the same direc on.
The direc on of is given by d = . , . , . and the direc on of
is given by d = . , . , . . When it is not clear through observa on
whether two vectors are parallel or not, the standard way of determining this is
by comparing their respec ve unit vectors. Using a calculator, we nd:
d
= .
|| d ||
d
u =
= .
|| d ||

u =

, .

, .

, .

, .

The two vectors seem to be parallel (at least, their components are equal to
decimal places). In most situa ons, it would suce to conclude that the lines
are at least parallel, if not the same. One way to be sure is to rewrite d and d
in terms of frac ons, not decimals. We have

d =

.
,
,
d = ,
,
One can then nd the magnitudes of each vector in terms of frac ons, then
compute the unit vectors likewise. A er a lot of manual arithme c (or a er
briey using a computer algebra system), one nds that

u =

u =

We can now say without equivoca on that these lines are parallel.
Are they the same line? The parametric equa ons for a line describe one
point that lies on the line, so we know that the point P = ( . , . , . ) lies
on . To determine if this point also lies on , plug in the x, y and z values of P
into the symmetric equa ons for :
( . ) . ? ( . )
=
.
.

( . ) ( .
.

= .

= .

The point P lies on both lines, so we conclude they are the same line, just
parametrized dierently. Figure . graphs this line along with the points and
vectors described by the parametric equa ons. Note how d and d are parallel,
though point in opposite direc ons (as indicated by their unit vectors above).

Notes:

Figure .
ple
.

: Graphing the lines in Exam-

Chapter

Vectors

Distances

#
PQ

Given a point Q and a line (t) = p + td in space, it is o en useful to know


the distance from the point to the line. (Here we use the standard deni on
of distance, i.e., the length of the shortest line segment from the point to the
line.) Iden fying p with the point P, Figure . will help establish a general
method of compu ng this distance h.
#
From trigonometry, we know h = || PQ || sin . We have a similar iden ty
#
#
involving the cross product: || PQ d || = || PQ || || d || sin . Divide both sides
of this la er equa on by || d || to obtain h:

h=

Figure . : Establishing the distance


from a point to a line.

#
|| PQ d ||
.
|| d ||

. )

It is also useful to determine the distance between lines, which we dene as


the length of the shortest line segment that connects the two lines (an argument
from geometry shows that this line segments is perpendicular to both lines). Let
lines (t) = p + td and (t) = p + td be given, as shown in Figure . .
To nd the direc on orthogonal to both d and d , we take the cross product:
#
c = d d . The magnitude of the orthogonal projec on of P P onto c is the
distance h we seek:

#
h = proj c P P

#
P P c


c
=
c c
#
|P P c|
|| c ||
=
|| c ||
#
|P P c|
.
=
|| c ||

Figure . : Establishing the distance


between lines.

A problem in the Exercise sec on is to show that this distance is when the lines
#
#
intersect. Note the use of the Triple Scalar Product: P P c = P P (d d ).
The following Key Idea restates these two distance formulas.

Notes:

Key Idea

Distances to Lines

. Let P be a point on a line that is parallel to d. The distance h from


a point Q to the line is:
#
|| PQ d ||
h=
.
|| d ||
. Let P be a point on line that is parallel to d , and let P be a
point on line parallel to d , and let c = d d , where lines
and are not parallel. The distance h between the two lines is:
#
|P P c|
h=
.
|| c ||

Example
Finding the distance from a point to a line
Find the distance from the point Q = ( , , ) to the line (t) = , , +
t , , .
The equa on of the line line gives us the point P = ( , , )
S
#
that lies on the line, hence PQ = , , . The equa on also gives d = , , .
Following Key Idea , we have the distance as
#
|| PQ d ||
h=
|| d ||
|| , , ||

= .
.
The point Q is approximately .

units from the line (t).

Finding the distance between lines


Example
Find the distance between the lines
x
: y
z

Notes:

=
=
=

+ t
t
t

x
: y
z

=
=
=

+ s
+s
+ s.

Lines

Chapter

Vectors
These are the sames lines as given in Example
, where
S
we showed them to be skew. The equa ons allow us to iden fy the following
points and vectors:
P =( , , )
d = , ,

From Key Idea

P = ( , , )

d = , ,

#
P P = , , .

c = d d = , , .

we have the distance h between the two lines is


h=

#
|P P c|
|| c ||

=
The lines are approximately .

units apart.

One of the key points to understand from this sec on is this: to describe a
line, we need a point and a direc on. Whenever a problem is posed concerning a line, one needs to take whatever informa on is oered and glean point
and direc on informa on. Many ques ons can be asked (and are asked in the
Exercise sec on) whose answer immediately follows from this understanding.
Lines are one of two fundamental objects of study in space. The other fundamental object is the plane, which we study in detail in the next sec on. Many
complex three dimensional objects are studied by approxima ng their surfaces
with lines and planes.

Notes:

Exercises

Terms and Concepts


. To nd an equa on of a line, what two pieces of informaon are needed?
. Two dis nct lines in the plane can intersect or be
.
. Two dis nct lines in space can intersect, be
.

or be

. Use your own words to describe what it means for two lines
in space to be skew.

Problems
In Exercises , write the vector, parametric and symmetric equa ons of the lines described.
. Passes through P = ( , , ), parallel to d = , , .
. Passes through P = ( , , ), parallel to d = , , .
. Passes through P = ( , , ) and Q = ( , , ).
. Passes through P = ( , , ) and Q = ( , , ).
. Passes through P = ( , , ) and orthogonal to both
d = , , and d = , , .
. Passes through P = ( , , ) and orthogonal to both
d = , , and d = , , .
. Passes through the point of intersec on of
(t) and
(t)
and orthogonal to both lines, where

(t) = , , + t , , and

(t) = , , + t , , .
. Passes through the point of intersec on of (t) and (t)
and orthogonal
to both lines, where

x
=
t

x = +t

= y = + t and = y = t .

z= + t
z= +t
. Passes through P = ( , ), parallel to d = , .

. Passes through P = ( , ), parallel to d = , .


In Exercises
, determine if the described lines are the
same line, parallel lines, intersec ng or skew lines. If intersec ng, give the point of intersec on.
.
(t) = , , + t , , ,

(t) = , , + t , , .

.
(t) = , , + t , , ,

(t) = , , + t , , .
.
(t) = , , + t , , ,

(t) = , , + t , , .
.
(t) = , , + t , , ,

(t) = , , + t , , .

x = + t
. = y= t

z=t

x =
= y=

z=

and

+ t
+ t

x = . + . t
x = . + . t
. = y = . + . t and = y = + . t

z= . + . t
z= . + . t

x = . + . t
x = . + . t
. t
. = y = . . t and = y = .

z= .
+ . t
z= . + . t

x = . + . t
. = y= . . t

z= . + . t

In Exercises
line.

and

x = . t
= y= . + . t

z= . + . t

, nd the distance from the point to the

. P = ( , , ),

(t) = , , + t , ,

. P = ( , , ),

(t) = , , + t , ,

. P = ( , ),

(t) = , + t ,

. P = ( , ),

(t) = , + t ,

In Exercises

, nd the distance between the two lines.

.
(t) = , , + t , , ,

(t) = , , + t , , .
.
(t) = , , + t , , ,

(t) = , , + t , , .
Exercises explore special cases of the distance formulas found in Key Idea .
. Let Q be a point on the line (t). Show why the distance
formula correctly gives the distance from the point to the
line as .
. Let lines (t) and (t) be intersec ng lines. Show why
the distance formula correctly gives the distance between
these lines as .

. Let lines (t) and (t) be parallel.


(a) Show why the distance formula for distance between
lines cannot be used as stated to nd the distance between the lines.
#
(b) Show why le ng c = (P P d ) d allows one to

the use the formula.


(c) Show how one can use the formula for the distance
between a point and a line to nd the distance between parallel lines.

Planes

Any at surface, such as a wall, table top or s piece of cardboard can be


thought of as represen ng part of a plane. Consider a piece of cardboard with
a point P marked on it. One can take a nail and s ck it into the cardboard at P
such that the nail is perpendicular to the cardboard; see Figure .
This nail provides a handle for the cardboard. Moving the cardboard around
moves P to dierent loca ons in space. Til ng the nail (but keeping P xed) lts
the cardboard. Both moving and l ng the cardboard denes a dierent plane
in space. In fact, we can dene a plane by: ) the loca on of P in space, and )
the direc on of the nail.
The previous sec on showed that one can dene a line given a point on the
line and the direc on of the line (usually given by a vector). One can make a
similar statement about planes: we can dene a plane in space given a point on
the plane and the direc on the plane faces (using the descrip on above, the
direc on of the nail). Once again, the direc on informa on will be supplied by
a vector, called a normal vector, that is orthogonal to the plane.
What exactly does orthogonal to the plane mean? Choose any two points P
#
and Q in the plane, and consider the vector PQ. We say a vector n is orthogonal
#
to the plane if n is perpendicular to PQ for all choices of P and Q; that is, if
#
n PQ = for all P and Q.
This gives us way of wri ng an equa on describing the plane. Let P =
(x , y , z ) be a point in the plane and let n = a, b, c be a normal vector to
the plane. A point Q = (x, y, z) lies in the plane dened by P and n if, and only
#
#
if, PQ is orthogonal to n. Knowing PQ = x x , y y , z z , consider:
#
PQ n =

x x , y y , z z a, b, c =
a(x x ) + b(y y ) + c(z z ) =
Equa on (
produces:

. )

. ) denes an implicit func on describing the plane. More algebra


ax + by + cz = ax + by + cz .

The right hand side is just a number, so we replace it with d:


ax + by + cz = d.

As long as c = , we can solve for z:


z=

Notes:

Planes

(d ax by).

Figure . : Illustra ng dening a plane


with a sheet of cardboard and a nail.

Chapter

Vectors
Equa on ( . ) is especially useful as many computer programs can graph funcons in this form. Equa ons ( . ) and ( . ) have specic names, given next.
Deni on

Equa ons of a Plane in Standard and General Forms

The plane passing through the point P = (x , y , z ) with normal vector


n = a, b, c can be described by an equa on with standard form
a(x x ) + b(y y ) + c(z z ) = ;
the equa ons general form is
ax + by + cz = d.
A key to remember throughout this sec on is this: to nd the equa on of a
plane, we need a point and a normal vector. We will give several examples of
nding the equa on of a plane, and in each one dierent types of informa on
are given. In each case, we need to use the given informa on to nd a point on
the plane and a normal vector.
Finding the equa on of a plane.
Example
Write the equa on of the plane that passes through the points P = ( , , ),
Q = ( , , ) and R = ( , , ) in standard form.
We need a vector n that is orthogonal to the plane. Since P,
S
#
# # #
Q and R are in the plane, so are the vectors PQ and PR; PQ PR is orthogonal
#
#
to PQ and PR and hence the plane itself.
# #
It is straigh orward to compute n = PQ PR = , , . We can use any
point we wish in the plane (any of P, Q or R will do) and we arbitrarily choose P.
Following Deni on , the equa on of the plane in standard form is
(x ) + (y ) + z = .
The plane is sketched in Figure
Figure
ample

: Sketching the plane in Ex-

We have just demonstrated the fact that any three non-collinear points dene a plane. (This is why a three-legged stool does not rock; its three feet
always lie in a plane. A four-legged stool will rock unless all four feet lie in the
same plane.)
Finding the equa on of a plane.
Example
Verify that lines and , whose parametric equa ons are given below, inter-

Notes:

Planes

sect, then give the equa on of the plane that contains these two lines in general
form.
x = + s
x =
+ t
+s
t
: y =
: y =
z = + s
z =
+t
The lines clearly are not parallel. If they do not intersect,
S
they are skew, meaning there is not a plane that contains them both. If they do
intersect, there is such a plane.
To nd their point of intersec on, we set the x, y and z equa ons equal to
each other and solve for s and t:
+ s
+s
+ s

=
=
=

+ t
t
+t

s= ,

t= .

When s = and t = , the lines intersect at the point P = ( , , ).


Let d = , , and d = , , be the direc ons of lines and ,
respec vely. A normal vector to the plane containing these the two lines will
also be orthogonal to d and d . Thus we nd a normal vector n by compu ng
n = d d = , .
We can pick any point in the plane with which to write our equa on; each
line gives us innite choices of points. We choose P, the point of intersec on.
We follow Deni on to write the planes equa on in general form:

Figure
ample

: Sketching the plane in Ex-

(x + ) + (y ) z =
x+ + y z=

x+ y z= .

The planes equa on in general form is x + y z = ; it is sketched in Figure


. .
Finding the equa on of a plane
Example
Give the equa on, in standard form, of the plane that passes through the point
P = ( , , ) and is orthogonal to the line with vector equa on (t) = , , +
t , , .
As the plane is to be orthogonal to the line, the plane must
S
be orthogonal to the direc on of the line given by d = , , . We use this as
our normal vector. Thus the planes equa on, in standard form, is
(x + ) + y + (z ) = .
The line and plane are sketched in Figure

Notes:

Figure .
ple
.

: The line and plane in Exam-

Chapter

Vectors
Example
Finding the intersec on of two planes
Give the parametric equa ons of the line that is the intersec on of the planes
p and p , where:
p : x (y ) + (z ) =
p : (x ) + (y + ) + (z ) =
To nd an equa on of a line, we need a point on the line and
S
the direc on of the line.
We can nd a point on the line by solving each equa on of the planes for z:
p : z = x + y
p :z= xy

We can now set these two equa ons equal to each other (i.e., we are nding
values of x and y where the planes have the same z value):
x + y

= xy
y= x
y=

Figure . : Graphing the planes and


their line of intersec on in Example
.

( x )

We can choose any value for x; we choose x = . This determines that y = .


We can now use the equa ons of either plane to nd z: when x = and y = ,
z = on both planes. We have found a point P on the line: P = ( , , ).
We now need the direc on of the line. Since the line lies in each plane,
its direc on is orthogonal to a normal vector for each plane. Considering the
equa ons for p and p , we can quickly determine their normal vectors. For p ,
n = , , and for p , n = , , . A direc on orthogonal to both of
these direc ons is their cross product: d = n n = , , .
The parametric equa ons of the line through P = ( , , ) in the direc on
of d = , , is:
:

x= t+

y= t+

The planes and line are graphed in Figure

z = t .
.

Finding the intersec on of a plane and a line


Example
Find the point of intersec on, if any, of the line (t) = , , + t , ,
and the plane with equa on in general form x + y + z = .
The equa on of the plane shows that the vectorn = , ,
S
is a normal vector to the plane, and the equa on of the line shows that the line

Notes:

Planes

moves parallel to d = , , . Since these are not orthogonal, we know


there is a point of intersec on. (If there were orthogonal, it would mean that
the plane and line were parallel to each other, either never intersec ng or the
line was in the plane itself.)
To nd the point of intersec on, we need to nd a t value such that (t)
sa ses the equa on of the plane. Rewri ng the equa on of the line with parametric equa ons will help:

x = t
(t) = y = + t .

z= +t

Replacing x, y and z in the equa on of the plane with the expressions containing
t found in the equa on of the line allows us to determine a t value that indicates
the point of intersec on:
x+y+z=
( t) + ( + t) + ( + t) =

Figure . : Illustra ng the intersec on


of a line and a plane in Example
.

t= .

When t = , the point on the line sa ses the equa on of the plane; that point
is ( ) = , , . Thus the point ( , , ) is the point of intersec on between
the plane and the line, illustrated in Figure . .

Distances
Just as it was useful to nd distances between points and lines in the previous
sec on, it is also o en necessary to nd the distance from a point to a plane.
Consider Figure . , where a plane with normal vector n is sketched containing a point P and a point Q, not on the plane, is given. We measure the
#
distance from Q to the plane by measuring the length of the projec on of PQ
onto n. That is, we want:

#
#
n PQ

#

|n PQ|
proj n PQ =
n =
|| n ||
|| n ||

Equa on ( . ) is important as it does more than just give the distance between
a point and a plane. We will see how it allows us to nd several other distances
as well: the distance between parallel planes and the distance from a line and a
plane. Because Equa on ( . ) is important, we restate it as a Key Idea.

Notes:

Figure . : Illustra ng nding the distance from a point to a plane.

Chapter

Vectors

Key Idea

Distance from a Point to a Plane

Let a plane with normal vector n be given, and let Q be a point. The
distance h from Q to the plane is
h=

#
|n PQ|
,
|| n ||

where P is any point in the plane.


Distance between a point and a plane
Example
Find the distance bewteen the point Q = ( , , ) and the plane with equa on
x y+ z= .
Using the equa on of the plane, we nd the normal vector
S
n = , , . To nd a point on the plane, we can let x and y be anything we
choose, then let z be whatever sa ses the equa on. Le ng x and y be seems
#
simple; this makes z = . . Thus we let P = , , . , and PQ = , , . .
The distance h from Q to the plane is given by Key Idea :
#
|n PQ|
h=
|| n ||
| , , , , . |
=
|| , , ||
| |
=
.

We can use Key Idea


to nd other distances. Given two parallel planes,
we can nd the distance between these planes by le ng P be a point on one
plane and Q a point on the other. If is a line parallel to a plane, we can use the
Key Idea to nd the distance between them as well: again, let P be a point in the
plane and let Q be any point on the line. (One can also use Key Idea .) The
Exercise sec on contains problems of these types.
These past two sec ons have not explored lines and planes in space as an exercise of mathema cal curiosity. However, there are many, many applica ons
of these fundamental concepts. Complex shapes can be modeled (or, approximated) using planes. For instance, part of the exterior of an aircra may have
a complex, yet smooth, shape, and engineers will want to know how air ows
across this piece as well as how heat might build up due to air fric on. Many
equa ons that help determine air ow and heat dissipa on are dicult to apply
to arbitrary surfaces, but simple to apply to planes. By approxima ng a surface
with millions of small planes one can more readily model the needed behavior.
Notes:

Exercises

Terms and Concepts


. In order to nd the equa on of a plane, what two pieces of
informa on must one have?
. What is the rela onship between a plane and one of its normal vectors?

Problems
In Exercises , give any two points in the given plane.
.

x y+ z=

(x + ) + (y ) z =

. x=
.

(y + ) (z ) =

In Exercises , give the equa on of the described plane


in standard and general forms.
. Passes through ( , , ) and has normal vector
n = , , .
. Passes through ( , , ) and has normal vector
n = , , .
. Passes through the points ( , , ), ( , , ) and ( , , ).
. Passes through the points ( , , ), ( , , ) and ( , , ).
. Contains the intersec ng lines
(t) = , , + t , , and
(t) = , , + t , , .
. Contains the intersec ng lines
(t) = , , + t , , and
(t) = , , + t , , .
. Contains the parallel lines
(t) = , , + t , , and
(t) = , , + t , , .
. Contains the parallel lines
(t) = , , + t , , and
(t) = , , + t , , .
. Contains
the point ( , , ) and the line

x = + t
(t) = y = + t

z= + t

. Contains
the point ( , , ) and the line

x = t
(t) = y = t

z=t

. Contains the point ( , , ) and is orthogonal to the line


(t) = , , + t , , .
. Contains
the point ( , , ) and is orthogonal to the line
x = + t

(t) = y = + t

z= + t

. Contains the point ( , , ) and is parallel to the plane


(x ) + (y + ) z = .
. Contains the point ( , , ) and is parallel to the plane
x= .
In Exercises
, give the equa on of the line that is the
intersec on of the given planes.
. p :
p :

(x ) + (y ) + z = , and
(x ) (y + ) + (z ) = .

. p :
p :

(x ) + (y + ) + (z ) = , and
x (y ) + (z ) = .

In Exercises
, nd the point of intersec on between
the line and the plane.
. line: , , + t , , ,
plane: x y z =
. line: , , + t , , ,
plane: x + y z =
. line: , , + t , , ,
plane: x y z =
. line: , , + t , , ,
plane: x y z =
In Exercises

, nd the given distances.

. The distance from the point ( , , ) to the plane


(x ) + (y ) + (z ) = .
. The distance from the point ( , , ) to the plane
(x ) y + (z + ) = .
. The distance between the parallel planes
x + y + z = and
(x ) + (y ) + (z + ) =

. The distance between the parallel planes


(x ) + (y + ) + (z ) = and
(x ) + (y ) + (z ) =
. Show why if the point Q lies in a plane, then the distance

formula correctly gives the distance from the point to the


plane as .
. How is Exercise in Sec on . easier to answer once we
have an understanding of planes?

:V

In the previous chapter, we learned about vectors and were introduced to the
power of vectors within mathema cs. In this chapter, well build on this founda on to dene func ons whose input is a real number and whose output is a
vector. Well see how to graph these func ons and apply calculus techniques
to analyze their behavior. Most importantly, well see why we are interested in
doing this: well see beau ful applica ons to the study of moving objects.

VectorValued Func ons

We are very familiar with real valued func ons, that is, func ons whose output
is a real number. This sec on introduces vectorvalued func ons func ons
whose output is a vector.
Deni on

VectorValued Func ons

r( )

A vectorvalued func on is a func on of the form


r(t) = f(t), g(t)

or r(t) = f(t), g(t), h(t) ,

where f, g and h are real valued func ons.

The domain ofr is the set of all values of t for whichr(t) is dened. The
range ofr is the set of all possible output vectorsr(t).
Evalua ng and Graphing VectorValued Func ons
Evalua ng a vectorvalued func on at a specic value of t is straigh orward;
simply evaluate
each component func on at that value of t. For instance, if

r(t) = t , t + t , then r( ) = , . We can sketch this vector, as is


done in Figure . (a). Plo ng lots of vectors is cumbersome, though, so generally we do not sketch the whole vector but just the terminal point. The graph
of a vectorvalued func on is the set of all terminal points of r(t), where the
ini al point of each vector is always the origin. In Figure . (b) we sketch the
graph ofr ; we can indicate individual points on the graph with their respec ve
vector, as shown.
Vectorvalued func ons are closely( related to) parametric
equa )ons of graphs.
(
While in both methods we plot points x(t), y(t) or x(t), y(t), z(t) to produce
a graph, in the context of vectorvalued func ons each such point represents a
vector. The implica ons of this will be more fully realized in the next sec on as
we apply calculus ideas to these func ons.

(a)
y

r( )
x

(b)

Figure . : Sketching the graph of a


vectorvalued func on.

Chapter

Vector Valued Func ons

t t

Example

Graphingvectorvalued func ons

Graphr(t) = t t,
, for t . Sketchr( ) andr( ).
t +

t +
/
/

We start by making a table of t, x and y values as shown in


S
Figure . (a). Plo ng these points gives an indica on of what the graph looks
like. In Figure . (b), we indicate these points and sketch the full graph. We
also highlightr( ) andr( ) on the graph.

/
/
(a)

Graphing vectorvalued func ons.


Example
Graphr(t) = cos t, sin t, t for t .

r( )

r( )
x

(b)
Figure . : Sketching the vectorvalued
func on of Example
.

We can again plot points, but careful considera on of this


S
func on is very revealing. Momentarily ignoring the third component, we see
the x and y components trace out a circle of radius centered at the origin.
No cing that the z component is t, we see that as the graph winds around the
z-axis, it is also increasing at a constant rate in the posi ve z direc on, forming a
spiral. This is graphed in Figure . . In the graphr( / ) ( .
, .
, .
is highlighted to help us understand the graph.

Algebra of VectorValued Func ons

Deni on

Opera ons on VectorValued Func ons

Let r (t) = f (t), g (t) and r (t) = f (t), g (t) be vectorvalued


func ons in R and let c be a scalar. Then:
. r (t) r (t) = f (t) f (t), g (t) g (t) .
. cr (t) = cf (t), cg (t) .
A similar deni on holds for vectorvalued func ons in R .
This deni on states that we add, subtract and scale vector-valued func ons
componentwise. Combining vectorvalued func ons in this way can be very
useful (as well as create interes ng graphs).
Figure
. : Viewing a vectorvalued
func on, and its deriva ve at one point.

Adding and scaling vectorvalued func ons.


Example
Let r (t) = . t, . t , r (t) = cos t, sin t and r(t) = r (t) + r (t). Graph
r (t),r (t),r(t) and r(t) on t .

Notes:

.
We can graphr andr easily by plo ng points (or just using
technology). Lets think about each for a moment to be er understand how
vectorvalued func ons work.
We can rewrite r (t) = . t, . t as r (t) = t . , . . That is, the
func on r scales the vector . , . by t. This scaling of a vector produces a
line in the direc on of . , . .
We are familiar withr (t) = cos t, sin t ; it traces out a circle, centered at
the origin, of radius . Figure . (a) graphsr (t) andr (t).
Adding r (t) to r (t) produces r(t) = cos t + . t, sin t + . t , graphed
in Figure . (b). The linear movement of the line combines with the circle to
create loops that move in the direc on of . , . . (We encourage the reader
to experiment by changingr (t) to t, t, etc., and observe the eects on the
loops.)
Mul plying r(t) by scales the func on by , producing r(t) = cos t +
, sin t + . , which is graphed in Figure . (c) along with r(t). The new
func on is mes bigger than r(t). Note how the graph of r(t) in (c) looks
iden cal to the graph ofr(t) in (b). This is due to the fact that the x and y bounds
of the plot in (c) are exactly
mes larger than the bounds in (b).

VectorValued Func ons

(a)
4

Adding and scaling vectorvalued func ons.


Example
A cycloid is a graph traced by a point p on a rolling circle, as shown in Figure
. . Find an equa on describing the cycloid, where the circle has radius .

(b)
y

.
Figure

. : Tracing a cycloid.
x

This problem is not very dicult if we approach it in a clever


way. We start by le ng p(t) describe the posi on of the point p on the circle,
where the circle is centered at the origin and only rotates clockwise (i.e., it does
not roll). This is rela vely simple given our previous experiences with parametric
equa ons; p(t) = cos t, sin t.
We now want the circle to roll. We represent this by le ng c(t) represent
the loca on of the center of the circle. It should be clear that the y component
of c(t) should be ; the center of the circle is always going to be if it rolls on a
horizontal surface.
The x component ofc(t) is a linear func on of t: f(t) = mt for some scalar m.
When t = , f(t) = (the circle starts centered on the y-axis). When t = ,
the circle has made one complete revolu on, traveling a distance equal to its

Notes:

(c)
Figure
ample

. : Graphing the func ons in Ex.

Chapter

Vector Valued Func ons


circumference, which is also . This gives us a point on our line f(t) = mt, the
point ( , ). It should be clear that m = and f(t) = t. So c(t) = t, .
We now combinep andc together to form the equa on of the cycloid: r(t) =
p(t) + c(t) = cos t + t, sin t + , which is graphed in Figure . .

10

Displacement
5

Figure

15

10

. : The cycloid in Example

A vectorvalued func onr(t) is o en used to describe the posi on of a moving object at me t. At t = t , the object is at r(t ); at t = t , the object is at
r(t ). Knowing the loca onsr(t ) andr(t ) give no indica on of the path taken
between them, but o en we only care about the dierence of the loca ons,
r(t ) r(t ), the displacement.
Deni on

Displacement

Let r(t) be a vectorvalued func on and let t < t be values in the


domain. The displacement d ofr, from t = t to t = t , is
d = r(t ) r(t ).
When the displacement vector is drawn with ini al point atr(t ), its terminal
point isr(t ). We think of it as the vector which points from a star ng posi on
to an ending posi on.

y
1

d
1

Figure . : Graphing the displacement


of a posi on func on in Example
.

Finding and graphing displacement vectors


Example
Letr(t) = cos( t), sin( t) . Graphr(t) on t , and nd the displacement ofr(t) on this interval.
The func on r(t) traces out the unit circle, though at a difS
ferent rate than the usual cos t, sin t parametriza on. At t = , we have
r(t ) = , ; at t = , we haver(t ) = , . The displacement ofr(t) on
[ , ] is thus d = , , = , .
A graph ofr(t) on [ , ] is given in Figure . , along with the displacement
vector d on this interval.
Measuring displacement makes us contemplate related, yet very dierent,
concepts. Considering the semicircular path the object in Example
took,
we can quickly verify that the object ended up a distance of units from its ini al
loca on. That is, we can compute || d || = . However, measuring distance from
the star ng point is dierent from measuring distance traveled. Being a semi

Notes:

.
circle, we can measure the distance traveled by this object as . units.
Knowing distance from the star ng point allows us to compute average rate of
change.
Deni on

Average Rate of Change

Letr(t) be a vectorvalued func on, where each of its component funcons is con nuous on its domain, and let t < t . The average rate of
change ofr(t) on [t , t ] is
average rate of change =

Averagerate of change
Example
Letr(t) = cos( t), sin( t) as in Example
ofr(t) on [ , ] and on [ , ].

r(t ) r(t )
.
t t

. Find the average rate of change

We computed in Example
that the displacement ofr(t)
S
on [ , ] was d = , . Thus the average rate of change ofr(t) on [ , ] is:
r( ) r( )
,
=
= , .
( )
We interpret this as follows: the object followed a semicircular path, meaning
it moved towards the right then moved back to the le , while climbing slowly,
then quickly, then slowly again. On average, however, it progressed straight up
at a constant rate of , per unit of me.
We can quickly see that the displacement on [ , ] is the same as on [ , ],
so d = , . The average rate of change is dierent, though:
r( ) r( )
,
=
= , / .
( )
As it took mes as long to arrive at the same place, this average rate of
change on [ , ] is / the average rate of change on [ , ].
We considered average rates of change in Sec ons . and . as we studied
limits and deriva ves. The same is true here; in the following sec on we apply
calculus concepts to vectorvalued func ons as we nd limits, deriva ves, and
integrals. Understanding the average rate of change will give us an understanding of the deriva ve; displacement gives us one applica on of integra on.

Notes:

VectorValued Func ons

Exercises

Terms and Concepts

In Exercises

, nd ||r(t) ||.

. Vectorvalued func ons are closely related to


of graphs.

. r(t) = t, t .

. When sketching vectorvalued func ons, technically one


isnt graphing points, but rather
.

. r(t) = cos t, sin t, t.

. It can be useful to think of


as a vector that points
from a star ng posi on to an ending posi on.

. r(t) = cos t, t, t .

Problems
In Exercises
given interval.

, sketch the vectorvalued func on on the

. r(t) = t , t , for t .
. r(t) =
. r(t) =
. r(t) =

In Exercises
, create a vectorvalued func on whose
graph matches the given descrip on.
. A circle of radius , centered at ( , ), traced counter
clockwise once on [ , ].
. A circle of radius , centered at ( , ), traced clockwise
once on [ , ].

. r(t) = t , t , for t .

. r(t) = cos t, sin t.

/t, /t , for t .

t , sin t , for t .

t , sin t , for t .

. An ellipse, centered at ( , ) with ver cal major axis of


length and minor axis of length , traced once counter
clockwise on [ , ].
. An ellipse, centered at ( , ) with horizontal major axis of
length and minor axis of length , traced once clockwise
on [ , ].
. A line through ( , ) with a slope of .

. r(t) = sin(t), cos(t), on [ , ].

. A line through ( , ) with a slope of / .

. r(t) = cos t, sin( t), on [ , ].

. A ver cally oriented helix with radius of that starts at


( , , ) and ends at ( , , ) a er revolu on on [ , ].

. r(t) = sec t, tan t, on [, ].


In Exercises , sketch the vectorvalued func on on the
given interval in R . Technology may be useful in crea ng the
sketch.
. r(t) = cos t, t, sin t, on [ , ].
. r(t) = cos t, sin t, t/ on [ , ].

. A ver cally oriented helix with radius of that starts at


( , , ) and ends at ( , , ) a er revolu ons on [ , ].
In Exercises , nd the average rate of change ofr(t) on
the given interval.

. r(t) = t, t on [ , ].

. r(t) = t, t + sin t on [ , ].

. r(t) = cos t, sin t, sin t on [ , ].

. r(t) = cos t, sin t, t on [ , ].

. r(t) = cos t, sin t, sin( t) on [ , ].

. r(t) = t, t , t on [ , ].

Calculus and VectorValued Func ons

The previous sec on introduced us to a new mathema cal object, the vector
valued func on. We now apply calculus concepts to these func ons. We start
with the limit, then work our way through deriva ves to integrals.

Limits of VectorValued Func ons


The ini al deni on of the limit of a vectorvalued func on is a bit in mida ng, as was the deni on of the limit in Deni on . The theorem following
the deni on shows that in prac ce, taking limits of vectorvalued func ons is
no more dicult than taking limits of realvalued func ons.
Deni on

Limits of VectorValued Func ons

Let I be an open interval containing c, and let r(t) be a vectorvalued


func on dened on I, except possibly at c. The limit of r(t), as t approaches c, is L, expressed as
lim r(t) = L,

tc

means that given any > , there exists a >


if |t c| < , we have ||r(t) L || < .

such that for all t = c,

Note how the measurement of distance between real numbers is the absolute value of their dierence; the measure of distance between vectors is the
vector norm, or magnitude, of their dierence.
Theorem

Limits of VectorValued Func ons

. Let r(t) = f(t), g(t) be a vectorvalued func on in R dened


on an open interval I containing c. Then

lim r(t) = lim f(t) , lim g(t) .


tc

tc

tc

. Letr(t) = f(t), g(t), h(t) be a vectorvalued func on in R dened on an open interval I containing c. Then

lim r(t) = lim f(t) , lim g(t) , lim h(t)


tc

Notes:

tc

tc

tc

Calculus and VectorValued Func ons

Chapter

Vector Valued Func ons


Theorem

states that we compute limits componentwise.

Example
Finding limits ofvectorvalued func ons
sin t
Letr(t) =
, t t + , cos t . Find lim r(t).
t
t
We apply the theorem and compute limits componentwise.

sin t
lim r(t) = lim
, lim t t + , lim cos t
t
t
t
t
t

= , , .

Con nuity
Deni on

Con nuity of VectorValued Func ons

Let r(t) be a vectorvalued func on dened on an open interval I containing c.


. r(t) is con nuous at c if lim r(t) = r(c).
tc

. Ifr(t) is con nuous at all c in I, thenr(t) is con nuous on I.


We again have a theorem that lets us evaluate con nuity componentwise.

Theorem

Con nuity of VectorValued Func ons

Let r(t) be a vectorvalued func on dened on an open interval I containing c. r(t) is con nuous at c if, and only if, each of its component
func ons is con nuous at c.
Evalua ng con nuity
of vectorvalued func ons
sin t
Let r(t) =
, t t + , cos t . Determine whether r is con nuous at
t
t = and t = .

Example

While the second and third components ofr(t) are dened


S
at t = , the rst component, (sin t)/t, is not. Since the rst component is not
even dened at t = ,r(t) is not dened at t = , and hence it is not con nuous
at t = .
Notes:

Calculus and VectorValued Func ons

At t = each of the component func ons is con nuous. Therefore r(t) is


con nuous at t = .

Deriva ves
Consider a vectorvalued func onr dened on an open interval I containing
t and t . We can compute the displacement ofr on [t , t ], as shown in Figure
. (a). Recall that dividing the displacement vector by t t gives the average
rate of change on [t , t ], as shown in (b).
r (t0 )

r(t1 )

r(t0 )

r(t0 )

r(t0 )

r(t1 )

r(t1 )

(a)
Figure

r(t1 ) r(t0 )
t1 t0

(b)

. : Illustra ng displacement, leading to an understanding of the deriva ve of vectorvalued func ons.

The deriva ve of a vectorvalued func on is a measure of the instantaneous


rate of change, measured by taking the limit as the length of [t , t ] goes to .
Instead of thinking of an interval as [t , t ], we think of it as [c, c + h] for some
value of h (hence the interval has length h). The average rate of change is
r(c + h) r(c)
h
for any value of h = . We take the limit as h to measure the instantaneous
rate of change; this is the deriva ve ofr.
Deni on

Deriva ve of a VectorValued Func on

Letr(t) be con nuous on an open interval I containing c.


. The deriva ve ofr at t = c is
r (c) = lim
h

r(c + h) r(c)
.
h

. The deriva ve ofr is


r (t) = lim
h

Notes:

r(t + h) r(t)
.
h

Alternate nota ons for the deriva ve ofr


include:
r (t) =

)
dr
d(
r(t) = .
dt
dt

Chapter

Vector Valued Func ons


If a vectorvalued func on has a deriva ve for all c in an open interval I, we
say thatr(t) is dieren able on I.
Once again we might view this deni on as in mida ng, but recall that we
can evaluate limits componentwise. The following theorem veries that this
means we can compute deriva ves componentwise as well, making the task
not too dicult.
Theorem

Deriva ves of VectorValued Func ons

. Letr(t) = f(t), g(t) . Then


r (t) = f (t), g (t) .
. Letr(t) = f(t), g(t), h(t) . Then
r (t) = f (t), g (t), h (t) .

Example

Letr(t) = t , t .

y
2

r(t)

. Sketchr(t) andr (t) on the same axes.

. Computer ( ) and sketch this vector with its ini al point at the origin and
atr( ).

r (t)

. Theorem

(a)

r(t) and r (t) are graphed together in Figure . (a). Note how plo ng
the two of these together, in this way, is not very illumina ng. When
dealing with realvalued func ons, plo ng f(x) with f (x) gave us useful
informa on as we were able to compare f and f at the same x-values.
When dealing with vectorvalued func ons, it is hard to tell which points
on the graph ofr correspond to which points on the graph ofr.

r (1)

1
r (1)
2

. We easily computer ( ) = , , which is drawn in Figure . with its


ini al point at the origin, as well as at r( ) = , . These are sketched
in Figure . (b).

allows us to compute deriva ves componentwise, so


r (t) = t, .

y
2

Deriva ves of vectorvalued func ons

(b)
Figure . : Graphing the deriva ve of a
vectorvalued func on in Example
.

Notes:

Calculus and VectorValued Func ons

Example
Deriva ves of vectorvalued func ons
Let r(t) = cos t, sin t, t. Compute r (t) and r (/ ). Sketch r (/ ) with its
ini al point at the origin and atr(/ ).
We computer asr (t) = sin t, cos t, . At t = / , we
have r (/ ) = , , . Figure . shows a graph of r(t), with r (/ )
plo ed with its ini al point at the origin and atr(/ ).
S

In Examples
and
, sketching a par cular deriva ve with its ini al
point at the origin did not seem to reveal anything signicant. However, when
we sketched the vector with its ini al point on the corresponding point on the
graph, we did see something signicant: the vector appeared to be tangent to
the graph. We have not yet dened what tangent means in terms of curves in
space; in fact, we use the deriva ve to dene this term.
Deni on

Figure . : Viewing a vectorvalued


func on and its deriva ve at one point.

Tangent Vector, Tangent Line

Letr(t) be a dieren able vectorvalued func on on an open interval I


containing c, wherer (c) = .
. A vector v is tangent to the graph ofr(t) at t = c if v is parallel to
r (c).
. The tangent line to the graph of r(t) at t = c is the line through
r(c) with direc on parallel to r (c). An equa on of the tangent
line is
(t) = r(c) + tr (c).

Example
Finding tangent lines to curves in space
Letr(t) = t, t , t on [ . , . ]. Find the vector equa on of the line tangent
to the graph ofr at t = .
To nd the equa on of a line, we need a point on the line
S
and the lines direc on. The point is given byr( ) = , , . (To be clear,
, , is a vector, not a point, but we use the point pointed to by this
vector.)
fromr ( ). We compute, componentwise, r (t) =
on comes
The direc

, t, t . Thusr ( ) = , , .
The vector equa on of the line is (t) = , , + t , , . This line
andr(t) are sketched in Figure . .

Notes:

Figure . : Graphing a curve in space


with its tangent line.

Chapter

Vector Valued Func ons


Example
Finding tangent lines to curves

Find the equa ons of the lines tangent tor(t) = t , t at t = and t = .


S

r( ) = ,

y
2

t , t . At t = , we have

and r ( ) = , ,

r(t)

We nd thatr (t) =

so the equa on of the line tangent to the graph ofr(t) at t = is


(t) = , + t , .

(t)
2

. .
Figure . : Graphing r(t) and its tangent line in Example
.

This line is graphed withr(t) in Figure

At t = , we haver ( ) = , = ! This implies that the tangent line has


no direc on. We cannot apply Deni on , hence cannot nd the equa on of
the tangent line.
the equa on of the tangent line to r(t) =
We were unable to compute
t , t at t = becauser ( ) = . The graph in Figure . shows that there
is a cusp at this point. This leads us to another deni on of smooth, previously
dened by Deni on in Sec on . .

Deni on

Smooth VectorValued Func ons

Letr(t) be a dieren able vectorvalued func on on an open interval I.


r(t) is smooth on I ifr (t) = on I.

Having established deriva ves of vectorvalued func ons, we now explore


the rela onships between the deriva ve and other vector opera ons. The following theorem states how the deriva ve interacts with vector addi on and the
various vector products.

Notes:

Theorem

Calculus and VectorValued Func ons

Properies of Deriva ves of VectorValued Func ons

Letr and s be dieren able vectorvalued func ons, let f be a dierenable realvalued func on, and let c be a real number.
)
d(
r(t) s(t) = r (t) s (t)
.
dt
)
d(
cr(t) = cr (t)
.
dt
)
d(
.
Product Rule
f(t)r(t) = f (t)r(t) + f(t)r (t)
dt
)
d(
.
Product Rule
r(t) s(t) = r (t) s(t) +r(t) s (t)
dt
)
d(
Product Rule
.
r(t) s(t) = r (t) s(t) +r(t) s (t)
dt
( )
d ( ( ))
.
Chain Rule
r f(t) = r f(t) f (t)
dt
y

Using
deriva ve proper es of vectorvalued func ons
Example

Letr(t) = t, t and let u(t) be the unit vector that points in the direc on
ofr(t).

r(t)

. Graphr(t) and u(t) on the same axes, on [ , ].

u(t)

. Find u (t) and sketch u ( ), u ( ) and u ( ). Sketch each with ini al


point the corresponding point on the graph of u.

.
S

. To form the unit vector that points in the direc on ofr, we need to divide
r(t) by its magnitude.
||r(t) || =

t + (t )

u(t) =

t + (t )

t, t

r(t) and u(t) are graphed in Figure . . Note how the graph of u(t)
forms part of a circle; this must be the case, as the length of u(t) is for
all t.

Notes:

Figure
ample

: Graphingr(t) andu(t) in Ex-

Chapter

Vector Valued Func ons


. To compute u (t), we use Theorem
u(t) = f(t)r(t),
y

(We could write

)
(
= t +(t )
where f(t) =
t + (t )

u(t)

u(t) =

, wri ng

,
t + (t )
t + (t )

and then take the deriva ve. It is a ma er of preference; this la er method


requires two applica ons of the Quo ent Rule where our method uses the
Product and Chain Rules.)

We nd f (t) using the Chain Rule:

Figure
. : Graphing some of the
deriva ves of u(t) in Example
.

f (t) =

t + (t )

t( t )
= (
)
t + (t )

We now nd u (t) using part of Theorem

t + (t )( t)

u (t) = f (t)u(t) + f(t)u (t)

t( t )
= (
, t .
) t, t +
t + (t )
t + (t )

This is admi edly very messy; such is usually the case when we deal
with unit vectors. We can use this formula to compute u ( ), u ( )
and u ( ):

u ( ) = ,
.
, .

u ( ) = ,
u ( ) = ,

Each of these is sketched in Figure . . Note how the length of the


vector gives an indica on of how quickly the circle is being traced at that
point. When t = , the circle is being drawn rela vely slow; when t =
, the circle is being traced much more quickly.
It is a basic geometric fact that a line tangent to a circle at a point P is perpendicular to the line passing through the center of the circle and P. This is

Notes:

.
illustrated in Figure . ; each tangent vector is perpendicular to the line that
passes through its ini al point and the center of the circle. Since the center of
the circle is the origin, we can state this another way: u (t) is orthogonal to u(t).
Recall that the dot product serves as a test for orthogonality: if u v = ,
then u is orthogonal to v. Thus in the above example, u(t) u (t) = .
This is true of any vectorvalued func on that has a constant length, that is,
that traces out part of a circle. It has important implica ons later on, so we state
it as a theorem (and leave its formal proof as an Exercise.)

Theorem

VectorValued Func ons of Constant Length

Letr(t) be a dieren able vectorvalued func on on an open interval I of


constant length. That is, ||r(t) || = c for all t in I (equivalently,r(t)r(t) =
c for all t in I). Thenr(t) r (t) = for all t in I.

Integra on
Indenite and denite integrals of vectorvalued func ons are also evaluated componentwise.

Theorem

Indenite and Denite Integrals of VectorValued


Func ons

Letr(t) = f(t), g(t) be a vectorvalued func on in R .

r(t) dt =
.
f(t) dt, g(t) dt
.

b
a

r(t) dt =

b
a

f(t) dt,

b
a

g(t) dt

A similar statement holds for vectorvalued func ons in R .

Evalua ng a denite integral of a vectorvalued func on


t

r(t) dt.
Letr(t) = e , sin t . Evaluate

Example

Notes:

Calculus and VectorValued Func ons

Chapter

Vector Valued Func ons


We follow Theorem .

r(t) dt =
e , sin t dt

t
sin t dt
=
e dt ,




t
=
e , cos t

=
(e ) , cos( ) +

, .

Solving an ini al value problem


Example
Letr (t) = , cos t, t. Findr(t) where:
r( ) = , , and
r ( ) = , , .

Knowingr (t) = , cos t, t, we ndr (t) by evalua ng the


indenite integral.

r (t) dt =
dt , cos t dt ,
t dt

= t + C , sin t + C , t + C

= t, sin t, t + C , C , C

= t, sin t, t + C.
S

Note how each indenite integral creates its own constant which we collect as
one constant vector C. Knowingr ( ) = , , allows us to solve for C:

r (t) = t, sin t, t + C
r ( ) = , , + C
, , = C.

Sor (t) = t, sin t, t + , , = t + , sin t + , t . To nd r(t),


we integrate once more.

sin t + dt,
t dt

= t + t, cos t + t, t + C.

r (t) dt =

t + dt,

Withr( ) = , , , we solve for C:


Notes:

Calculus and VectorValued Func ons

r(t) = t + t, cos t + t, t + C
r( ) = , , + C

, , = , , + C
, , = C.

Sor(t) = t + t, cos t + t, t + , , = t + t , cos t + t, t + .

What does the integra on of a vectorvalued func on mean? There are


many applica ons, but none as direct as the area under the curve that we
used in understanding the integral of a realvalued func on.
A key understanding for us comes from considering the integral of a derivave:
b
b

r (t) dt = r(t) = r(b) r(a).
a

Integra ng a rate of change func on gives displacement.


No ng that vectorvalued func ons are closely related to parametric equaons, we can describe the arc length of the graph of a vectorvalued func on
as an integral. Given parametric equa ons x = f(t), y = g(t), the arc length on
[a, b] of the graph is
b
f (t) + g (t) dt,
Arc Length =
a

as stated in Theorem in Sec on . . Ifr(t) = f(t), g(t), note that f (t) + g (t) =
|| r (t) ||. Therefore we can express the arc length of the graph of a vector
valued func on as an integral of the magnitude of its deriva ve.
Theorem

Arc Length of a VectorValued Func on

Letr(t) be a vectorvalued func on wherer (t) is con nuous on [a, b].


The arc length L of the graph ofr(t) is
L=

b
a

||r (t) || dt.

Note that we are actually integra ng a scalarfunc on here, not a vector


valued func on.
The next sec on takes what we have established thus far and applies it to
objects in mo on. We will letr(t) describe the path of an object in the plane or
in space and will discover the informa on provided byr (t) andr (t).

Notes:

Exercises

Terms and Concepts

In Exercises
, give the equa on of the line tangent to
the graph ofr(t) at the given t value.

. Limits, deriva ves and integrals of vectorvalued func ons


are all evaluated
wise.
. The denite integral of a rate of change func on gives
.
. Why is it generally not useful to graph both r(t) and r (t)
on the same axes?

Problems
. lim

t + , t , sin t

. lim

et ,

t
t+

, nd the deriva ve of the given func on.

. r(t) = cos t, et , ln t
t

Exercises

ask you to verify


parts of Theorem .

In
each
let
f(t)
=
t
,
r(t)
=
t ,t ,
and s(t) =

sin t, et , t . Compute the various deriva ves as indicated.


. Simplify f(t)r(t), then nd its deriva ve; show this is the
same as f (t)r(t) + f(t)r (t).
. Simplifyr(t) s(t), then nd its deriva ve; show this is the
same asr (t) s(t) +r(t) s (t).

. r(t) = cos t, et , ln t

t
, tan t
t+

. Simplifyr(t) s(t), then nd its deriva ve; show this is the


same asr (t) s(t) +r(t) s (t).
In Exercises
integral.

. r(t) = (t ) sin t, t +
. r(t) = t + , t

, nd the value(s) of t for whichr(t) is not

. r(t) = t t + , cos(t), sin (t)

. r(t) = t , /t

. r(t) = cos t sin t, sin t cos t, cos( t)

In Exercises , iden fy the interval(s) on whichr(t) is connuous.

. r(t) = et , tan t, t at t = .

. r(t) = t t + , t + t t +

r(t + h) r(t)
, wherer(t) = t , t, .
. lim
h
h

. r(t) =

. r(t) = cos t, sin t, t at t = .

. r(t) = cos t, sin t t

t
. lim
, ( + t) t
t
sin t

In Exercises

. r(t) = cos t, sin t at t = / .

In Exercises
smooth.

In Exercises , evaluate the given limit.


t

. r(t) = t + t, t t at t = .

sin t, t +

. r(t) = t + , t , sin t, t + ,

In Exercises , ndr (t). Sketchr(t) andr ( ), with the


ini al point ofr ( ) atr( ).

. r(t) = t + t, t t

. r(t) = t t + , , t t +

, evaluate the given denite or indenite

t , cos t, tet dt
+t

, sec t

dt

sin t, cos t dt
t + , t dt

In Exercises

. r(t) = t t + , t t + t

. r(t) = t + , t t

, solve the given ini al value problems.

. Findr(t), given thatr (t) = t, sin t andr( ) = , .


t

. Findr(t), given thatr (t) = /(t + ), tan t and


r( ) = , .

. r(t) = cos t, sin t, sin t on [ , ].

. Findr(t), given thatr (t) = t , t, ,

r ( ) = , , andr( ) = , , .
t

. Findr(t), given thatr (t) = cos t, sin t, e ,


r ( ) = , , andr( ) = , , .

In Exercises

cated interval.

, nd the arc length of r(t) on the indi-

. r(t) = cos t, sin t, t on [ , ].

. r(t) = t , t , t on [ , ].

. r(t) = et cos t, et sin t on [ , ].

. Prove Theorem ; that is, show ifr(t) has constant length

and is dieren able, then r(t)


( r (t) )= . (Hint: use the
d
Product Rule to compute dt r(t) r(t) .)

Chapter

Vector Valued Func ons

The Calculus of Mo on

A common use of vectorvalued func ons is to describe the mo on of an object


in the plane or in space. A posi on func onr(t) gives the posi on of an object
at me t. This sec on explores how deriva ves and integrals are used to study
the mo on described by such a func on.
Deni on

Velocity, Speed and Accelera on

Letr(t) be a posi on func on in R or R .


. Velocity, denoted v(t), is the instantaneous rate of posi on
change; that is, v(t) = r (t).
. Speed is the magnitude of velocity, || v(t) ||.
. Accelera on, denoted a(t), is the instantaneous rate of velocity
change; that is, a(t) = v (t) = r (t).
Finding velocity and accelera on
Example

An object is moving with posi on func onr(t) = t t, t + t , t ,


where distances are measured in feet and me is measured in seconds.

10

. Find v(t) and a(t).


. Sketchr(t); plot v( ), a( ), v( ) and a( ), each with their ini al point
at their corresponding point on the graph ofr(t).

. When is the objects speed minimized?


5

10

(a)

. Taking deriva ves, we nd

v(t) = r (t) = t , t +

10

and a(t) = r (t) = , .

Note that accelera on is constant.


. v( ) = , , a( ) = , ; v( ) = , , a( ) = , .
These are plo ed withr(t) in Figure . (a).

10

We can think of accelera on as pulling the velocity vector in a certain


direc on. At t = , the velocity vector points down and to the le ; at
t = , the velocity vector has been pulled in the , direc on and is

(b)

Figure . : Graphing the posi on, velocity and accelera on of an object in Example
.

Notes:

The Calculus of Mo on

now poin ng up and to the right. In Figure . (b) we plot more velocity/accelera on vectors, making more clear the eect accelera on has on
velocity.
Sincea(t) is constant in this example, as t grows largev(t) becomes almost
parallel to a(t). For instance, when t = , v( ) = , , which is
nearly parallel to , .
. The objects speed is given by

|| v(t) || = ( t ) + ( t + ) =
t + .

To nd the minimal speed, we could apply calculus techniques (such as


set the deriva ve equal to and solve for t, etc.) but we can nd it by
inspec on. Inside the square root we have a quadra c which is minimized

when t = . Thus the speed is minimized at t = , with a speed of


/s.
The graph in Figure . (b) also implies speed is minimized here. The
lled dots on the graph are located at integer values of t between
and . Dots that are far apart imply the object traveled a far distance in
second, indica ng high speed; dots that are close together imply the
object did not travel far in second, indica ng a low speed. The dots are
closest together near t = , implying the speed is minimized near that
value.

Analyzing Mo on
Example
Two objects follow an iden cal path
atdierent rates on [ , ]. The posi on
r (t) = t, t ; the posi on func on for Object is
func on for Object
is

r (t) = t , t , where distances are measured in feet and me is measured


in seconds. Compare the velocity, speed and accelera on of the two objects on
the path.

We begin by compu ng the velocity and accelera on funcS


on for each object:

v (t) = , t
v (t) = t , t

a (t) = ,
a (t) = t, t

We immediately see that Object has constant accelera on, whereas Object
does not.
At t = , we have v ( ) = , and v ( ) = , ; the velocity
of Object is three mes that of Object and so it follows that
the speed of
/s compared to
/s.)
Object is three mes that of Object (

Notes:

Figure . : Plo ng velocity and accelera on vectors for Object in Example


.

Chapter

Vector Valued Func ons


y
1

r1 (t)

0.5

.1

0.5

0.5

(a)
y

r (t)
.5

x
.5

.5

(b)
Figure . : Comparing the posi ons of
Objects and in Example
.

At t = , the velocity of Object is v( ) = , and the velocity of Object


is ! This tells us that Object comes to a complete stop at t = .
In Figure . , we see the velocity and accelera on vectors for Object
plo ed for t = , / , , / and t = . Note again how the constant accelera on vector seems to pull the velocity vector from poin ng down, right to
up, right. We could plot the analogous picture for Object , but the velocity and
accelera on vectors are rather large (a ( ) = , !)
Instead, we simply plot the loca ons of Object and on intervals of / th
of a second, shown in Figure . (a) and (b). Note how the x-values of Object
increase at a steady rate. This is because the x-component of a(t) is ; there is
no accelera on in the x-component. The dots are not evenly spaced; the object
is moving faster near t = and t = than near t = .
In part (b) of the Figure, we see the points plo ed for Object . Note the
large change in posi on from t = to t = . ; the object starts moving very
quickly. However, it slows considerably at it approaches the origin, and comes
to a complete stop at t = . While it looks like there are points near the origin,
there are in reality points there.
Since the objects begin and end at the same loca on, the have the same displacement. Since they begin and end at the same me, with the same displacement, they have they have the same average rate of change (i.e, they have the
same average velocity). Since they follow the same path, they have the same
distance traveled. Even though these three measurements are the same, the
objects obviously travel the path in very dierent ways.
Analyzing the mo on of a whirling ball on a string
Example
A young boy whirls a ball, a ached to a string, above his head in a counterclockwise circle. The ball follows a circular path and makes revolu ons per
second. The string has length .
. Find the posi on func onr(t) that describes this situa on.
. Find the accelera on of the ball and derive a physical interpreta on of it.
. A tree stands
in front of the boy. At what t-values should the boy
release the string so that the ball hits the tree?
S

. The ball whirls in a circle. Since the string is


long, the radius of the
circle is . The posi on func on r(t) = cos t, sin t describes a circle
with radius , centered at the origin, but makes a full revolu on every
seconds, not two revolu ons per second. We modify the period of the

Notes:

The Calculus of Mo on

trigonometric func ons to be / by mul plying t by . The nal posi on


func on is thus
r(t) = cos( t), sin( t) .

(Plot this for


second.)

t / to verify that one revolu on is made in / a

. To nd a(t), we deriver(t) twice.


v(t) = r (t) = sin( t), cos( t)

a(t) = r (t) = cos( t), sin( t)


=

cos( t), sin( t) .

Note how a(t) is parallel tor(t), but has a dierent magnitude and points
in the opposite direc on. Why is this?
Recall the classic physics equa on, Force = mass accelera on. A force
ac ng on a mass induces accelera on (i.e., the mass moves); accelera on
ac ng on a mass induces a force (gravity gives our mass a weight). Thus
force and accelera on are closely related. A moving ball wants to travel
in a straight line. Why does the ball in our example move in a circle? It is
a ached to the boys hand by a string. The string applies a force to the ball,
aec ng its mo on: the string accelerates the ball. This is not acceleraon in the sense of it travels faster; rather, this accelera on is changing
the velocity of the ball. In what direc on is this force/accelera on being
applied? In the direc on of the string, towards the boys hand.
The magnitude of the accelera on is related to the speed at which the ball
is traveling. A ball whirling quickly is rapidly changing direc on/velocity.
When velocity is changing rapidly, the accelera on must be large.

There are many ways to nd this me value. We choose one that is relavely simple computa onally. As shown in Figure . , the vector from
the release point to the tree is , r(t ). This line segment is tangent
to the circle, which means it is also perpendicular to r(t ) itself, so their
dot product is .

Notes:

t )
r(

Let t = t be the me when the boy lets go of the string. The ball will be
atr(t ), traveling in the direc on of v(t ). We want to nd t so that this
line contains the point ( , ) (since the tree is
directly in front of the
boy).

. When the boy releases the string, the string no longer applies a force to
the ball, meaning accelera on is and the ball can now move in a straight
line in the direc on of v(t).

Figure . : Modeling the ight of a ball


in Example
.

Chapter

Vector Valued Func ons

(
)
r(t ) , r(t ) =
cos( t ), sin( t ) cos( t ), sin( t ) =
cos ( t ) +

sin( t ) sin ( t ) =
sin( t ) =

sin( t ) = /
t = sin ( / )
t . + n,

where n is an integer. Solving for t we have:


t .

+ n/

This is a wonderful formula. Every / second a er t = .


s the boy
can release the string (since the ball makes revolu ons per second, he
has two chances each second to release the ball).
Analyzing mo on in space
Example
An object moves in a spiral with posi on func on r(t) = cos t, sin t, t, where
distances are measured in meters and me is in minutes. Describe the objects
speed and accelera on at me t.
S

Withr(t) = cos t, sin t, t, we have:

v(t) = sin t, cos t, and


a(t) = cos t, sin t, .

The speed of the object is ||v(t) || = ( sin t) + cos t + =


m/min;
it moves at a constant speed. Note that the object does not accelerate in the
z-direc on, but rather moves up at a constant rate of m/min.
The objects in Examples
and
traveled at a constant speed. That is,
|| v(t) || = c for some constant c. Recall Theorem , which states that if a
vectorvalued func on r(t) has constant length, then r(t) is perpendicular to
its deriva ve: r(t) r (t) = . In these examples, the velocity func on has
constant length, therefore we can conclude that the velocity is perpendicular to
the accelera on: v(t) a(t) = . A quick check veries this.
There is an intui ve understanding of this. If accelera on is parallel to velocity, then it is only aec ng the objects speed; it does not change the direc on
of travel. (For example, consider a dropped stone. Accelera on and velocity are

Notes:

The Calculus of Mo on

parallel straight down and the direc on of velocity never changes, though
speed does increase.) If accelera on is not perpendicular to velocity, then there
is some accelera on in the direc on of travel, inuencing the speed. If speed
is constant, then accelera on must be orthogonal to velocity, as it then only
aects direc on, and not speed.
Key Idea

Objects With Constant Speed

If an object moves with constant speed, then its velocity and accelera on
vectors are orthogonal. That is, v(t) a(t) = .

Projec le Mo on
An important applica on of vectorvalued posi on func ons is projec le
mo on: the mo on of objects under only the inuence of gravity. We will measure me in seconds, and distances will either be in meters or feet. We will show
that we can completely describe the path of such an object knowing its ini al
posi on and ini al velocity (i.e., where it is and where it is going.)
Suppose an object has ini al posi on r( ) = x , y and ini al velocity
v( ) = vx , vy . It is customary to rewrite v( ) in terms of its speed v and
direc on u, where u is a unit vector. Recall all unit vectors in R can be wri en
as cos , sin , where is an angle measure counterclockwise from the x-axis.
(We refer to as the angle of eleva on.) Thus v( ) = v cos , sin .

Since the accelera on of the object is known, namely a(t) = , g, where


g is the gravita onal constant, we can nd r(t) knowing our two ini al condions. We rst nd v(t):
v(t) =
v(t) =

a(t) dt
, g dt

v(t) = , gt + C.
Knowing v( ) = v cos , sin , we have C = v cos t, sin t and so

v(t) = v cos , gt + v sin .


Notes:

Note: In this text we use g =


/s when
using Imperial units, and g = . m/s
when using SI units.

Chapter

Vector Valued Func ons


We integrate once more to ndr(t):

r(t) = v(t) dt

r(t) =
v cos , gt + v sin dt

(
)
(
)
r(t) = v cos t, gt + v sin t + C.
Knowingr( ) = x , y , we conclude C = x , y and
r(t) =

Key Idea

)
(
)
v cos t + x , gt + v sin t + y

Projec le Mo on

The posi on func on of a projec le propelled from an ini al posi on of


r = x , y , with ini al speed v , with angle of eleva on and neglecting all accelera ons but gravity is

(
)
(
)
r(t) = v cos t + x , gt + v sin t + y
.
Le ng v = v cos , sin , r(t) can be wri en as

r(t) =
, gt + v t +r .

We demonstrate how to use this posi on func on in the next two examples.
Projec le Mo on
Example
Sydney shoots her Red Ryder bb gun across level ground from an eleva on of
, where the barrel of the gun makes a angle with the horizontal. Find how
far the bb travels before landing, assuming the bb is red at the adver sed rate
of
/s and ignoring air resistance.
S

Notes:

A direct applica on of Key Idea gives

r(t) = (
cos )t, t + (
sin )t +

. t, t + . t + ,

.
where we set her ini al posi on to be , . We need to nd when the bb lands,
then we can nd where. We accomplish this by se ng the y-component equal
to and solving for t:

t +

t+

t .

s.

t=

)( )

(We discarded a nega ve solu on that resulted from our quadra c equa on.)
We have found that the bb lands . s a er ring; with t = . , we nd
the x-component of our posi on func on is
. ( . )=
.
. The bb
lands about
feet away.
Projec le Mo on
Example
Alex holds his sisters bb gun at a height of
and wants to shoot a target that
is
above the ground,
away. At what angle should he hold the gun to hit
his target? (We s ll assume the muzzle velocity is
/s.)
The posi on func on for the path of Alexs bb is

r(t) = (
cos )t, t + (
sin )t + .

We need to nd so thatr(t) =
nd and t such that
(

cos )t =

, for some value of t. That is, we want to

and

t +(

sin )t +

= .

This is not trivial (though not hard). We start by solving each equa on for cos
and sin , respec vely.
cos =

and

sin =

+
t

Using the Pythagorean Iden ty cos + sin = , we have


(
)
)
(
+ t
+
=
t
t
Mul ply both sides by (
+( +
t

Notes:

t +

t) :
t ) =
= .

The Calculus of Mo on

Chapter

Vector Valued Func ons


This is a quadra c in t . That is, we can apply the quadra c formula to nd t ,
then solve for t itself.

,
(
)(
)
,

t =
t = .

t= .

Clearly the nega ve t values do not t our context, so we have t = .


t= .
. Using cos = /(
t), we can solve for :
(
)
)
(
and cos
= cos
.
.
= .

and

and

. .

Alex has two choices of angle. He can hold the rie at an angle of about with
the horizontal and hit his target . s a er ring, or he can hold his rie almost
straight up, with an angle of . , where hell hit his target about s later. The
rst op on is clearly the op on he should choose.

Distance Traveled
Consider a driver who sets her cruisecontrol to
speed for an hour. We can ask:

mph, and travels at this

. How far did the driver travel?


. How far from her star ng posi on is the driver?
The rst is easy to answer: she traveled
miles. The second is impossible to
answer with the given informa on. We do not know if she traveled in a straight
line, on an oval racetrack, or along a slowlywinding highway.
This highlights an important fact: to compute distance traveled, we need
only to know the speed, given by || v(t) ||.
Theorem

Distance Traveled

Letv(t) be a velocity func on for a moving object. The distance traveled


by the object on [a, b] is:
distance traveled =

b
a

|| v(t) || dt.

Note that this is just a restatement of Theorem : arc length is the same as distance traveled, just viewed in a dierent context.

Notes:

The Calculus of Mo on

Example
Distance Traveled, Displacement, and Average Speed
A par cle moves in space with posi on func onr(t) = t, t , sin(t) on [ , ],
where t is measured in seconds and distances are in meters. Find:
. The distance traveled by the par cle on [ , ].

. The displacement of the par cle on [ , ].


. The par cles average speed.
S

. We use Theorem

to establish the integral:

|| v(t) || dt
distance traveled =
=

Figure .
Example

+ ( t) + cos (t) dt.

This cannot be solved in terms of elementary func ons so we turn to numerical integra on, nding the distance to be . m.
. The displacement is the vector
r( ) r( ) = , , , , = , , .
That is, the par cle ends with an x-value increased by
z-values the same (see Figure . ).

and with y- and

. We found above that the par cle traveled . m over seconds. We can
compute average speed by dividing: . / = . m/s.
We should also consider Deni on of Sec on . , which says that the
b
average value of a func on f on [a, b] is ba a f(x) dx. In our context, the
average value of the speed is

average speed =
|| v(t) || dt
. = . m/s.
( )
Note how the physical context of a par cle traveling gives meaning to a
more abstract concept learned earlier.
In Deni on
on [a, b] to be

of Chapter we dened the average value of a func on f(x)

ba

Notes:

b
a

f(x) dx.

: The path of the par cle in


.

Chapter

Vector Valued Func ons


Note how in Example

we computed the average speed as

distance traveled
=
|| v(t) || dt;
travel me
( )

that is, we just found the average value of || v(t) || on [ , ].


Likewise, given posi on func onr(t), the average velocity on [a, b] is
displacement
=
travel me
ba

b
a

r (t) dt =

r(b) r(a)
;
ba

that is, it is the average value ofr (t), or v(t), on [a, b].

Key Idea

Average Speed, Average Velocity

Letr(t) be a con nuous posi on func on on an open interval I containing a < b.


The average speed is:
distance traveled
=
travel me

b
a

|| v(t) || dt
=
ba
ba

b
a

|| v(t) || dt.

The average velocity is:


displacement
=
travel me

b
a

r (t) dt
=
ba
ba

b
a

r (t) dt.

The next two sec ons inves gate more proper es of the graphs of vector
valued func ons and well apply these new ideas to what we just learned about
mo on.

Notes:

Exercises

Terms and Concepts

. r(t) = cos t, sin t on [ , ]

. How is velocity dierent from speed?

. r(t) = cos t, sin t on [ , ]

. What is the dierence between displacement and distance


traveled?

. r(t) = sec t, tan t on [ , / ]


. r(t) = t + cos t, sin t on [ , ]

. What is the dierence between average velocity and average speed?

. r(t) =

. Distance traveled is the same as


viewed in a dierent context.

. r(t) = t t, t + t, t on [ , ]

, just

. Describe a scenario where an objects average speed is a


large number, but the magnitude of the average velocity is
not a large number.
. Explain why it is not possible to have an average velocity
with a large magnitude but a small average speed.

, a posi on func onr(t) is given. Find v(t)

. r(t) =

t t + , t + t +

. r(t) = cos t, sin t


. r(t) = t/

, cos t, sin t

In Exercises
, a posi on func onr(t) is given. Sketch
r(t) on the indicated interval. Find v(t) and a(t), then add
v(t ) and a(t ) to your sketch, with their ini al points atr(t ),
for the given value of t .
. r(t) = t, sin t on [ , / ]; t = /

. r(t) = t , sin t on [ , / ]; t = /

. r(t) = t + t, t + t on [ , ]; t =

. r(t) =

. Projec le Mo on: r(t) =


]
[
v sin
on ,
g

(v cos )t, gt + (v sin )t

t+
,t
t +

on [ , ]; t =

In Exercises

, a posi on func on r(t) of an object is


given. Find the speed of the object in terms of t, and nd
where the speed is minimized/maximized on the indicated
interval.

. r(t) = t , t on [ , ]

. r(t) = t , t t on [ , ]

(a) Show that the posi ons are the same at the indicated
t and s values; i.e., showr (t ) = r (s ).
(b) Find the velocity, speed and accelera on of the two
objects at t and s , respec vely.

. r(t) = t + , t ,

t on [ , ]
. r(t) = t, t ,

In Exercises , posi on func onsr (t) andr (s) for two


objects are given that follow the same path on the respec ve
intervals.

Problems
In Exercises
and a(t).

t, cos t, sin t on [ , ]

. r (t) = t, t on [ , ]; t =
r (s) = s , s on [ , ]; s =

. r (t) = cos t, sin t on [ , ]; t = /


r (s) = cos( s), sin( s) on [ , / ]; s = /
. r (t) = t, t on [ , ]; t =
r (s) = t , t on [ , ]; s =


. r (t) = t, t on [ , ]; t =

r (s) = sin t, sin t on [ , / ]; s = /

In Exercises

, nd the posi on func on of an object


given its accelera on and ini al velocity and posi on.
. a(t) = , ; v( ) = , , r( ) = ,
. a(t) = , ; v( ) = , , r( ) = ,
. a(t) = cos t, sin t; v( ) = , , r( ) = ,
. a(t) = ,

; v( ) =

, r( ) = ,

In Exercises , nd the displacement, distance traveled,


average velocity and average speed of the described object
on the given interval.
. An object with posi on func onr(t) = cos t, sin t, t,
where distances are measured in feet and me is in seconds, on [ , ].

. An object with posi on func on r(t) = cos t, sin t,


where distances are measured in feet and me is in seconds, on [ , ].
. An object with velocity func onv(t) = cos t, sin t, where
distances are measured in feet and me is in seconds, on
[ , ].
. An object with velocity func on v(t) = , , , where
distances are measured in feet and me is in seconds, on
[ , ].
Exercises ask you to solve a variety of problems based
on the principles of projec le mo on.
. A boy whirls a ball, a ached to a
string, above his head
in a counterclockwise circle. The ball makes revolu ons
per second.
At what t-values should the boy release the string so that
the ball heads directly for a tree standing
in front of
him?
. David faces Goliath with only a stone in a
sling, which
he whirls above his head at revolu ons per second. They
stand
apart.
(a) At what t-values must David release the stone in his
sling in order to hit Goliath?
(b) What is the speed at which the stone is traveling
when released?
(c) Assume David releases the stone from a height of
and Goliaths forehead is above the ground. What
angle of eleva on must David apply to the stone to hit
Goliaths head?

. A hunter aims at a deer which is yards away. Her crossbow is at a height of , and she aims for a spot on the
deer
above the ground. The crossbow res her arrows
at
/s.
(a) At what angle of eleva on should she hold the crossbow to hit her target?
(b) If the deer is moving perpendicularly to her line of
sight at a rate of mph, by approximately how much
should she lead the deer in order to hit it in the desired loca on?
. A baseball player hits a ball at
mph, with an ini al height
of
and an angle of eleva on of , at Bostons Fenway
Park. The ball ies towards the famed Green Monster, a
wall
high located
from home plate.
(a) Show that as hit, the ball hits the wall.
(b) Show that if the angle of eleva on is
clears the Green Monster.

, the ball

. A Cessna ies at
at
mph and drops a box of supplies to the professor (and his wife) on an island. Ignoring
wind resistance, how far horizontally will the supplies travel
before they land?
. A football quarterback throws a pass from a height of ,
intending to hit his receiver yds away at a height of .
(a) If the ball is thrown at a rate of mph, what angle of
eleva on is needed to hit his intended target?
(b) If the ball is thrown at with an angle of eleva on of

, what ini al ball speed is needed to hit his target?

Unit Tangent and Normal Vectors

Unit Tangent and Normal Vectors

Unit Tangent Vector


Given a smooth vectorvalued func onr(t), we dened in Deni on that
any vector parallel to r (t ) is tangent to the graph of r(t) at t = t . It is o en
useful to consider just the direc on of r (t) and not its magnitude. Therefore
we are interested in the unit vector in the direc on of r (t). This leads to a
deni on.
Deni on

Unit Tangent Vector

Let r(t) be a smooth func on on an open interval I. The unit tangent


vector T(t) is
T(t) =
r (t).
||r (t) ||
Compu ng the unit tangent vector
Example
Letr(t) = cos t, sin t, t. Find T(t) and compute T( ) and T( ).
We apply Deni on

T(t) =

r (t)
||r (t) ||

= (

to nd T(t).

sin t

sin t,

cos t

.
cos t,
+

sin t, cos t,

We can now easily compute T( ) and T( ):

T( ) =
; T( ) = sin , cos ,
.
, ,

, .

, . .

These are plo ed in Figure . with their ini al points atr( ) andr( ), respecvely. (They look rather short since they are only length .)
The unit tangent vector T(t) always has a magnitude of , though it is somemes easy to doubt that is true. We can help solidify this thought in our minds
by compu ng || T( ) ||:

|| T( ) || ( .
) + .
+ . = .
.
Notes:

Figure . : Plo ng unit tangent vectors in Example


.

Chapter

Vector Valued Func ons


We have rounded in our computa on of T( ), so we dont get exactly. We
leave it to the reader to use the exact representa on of T( ) to verify it has
length .
In many ways, the previous example was too nice. It turned out thatr (t)
was always of length . In the next example the length ofr (t) is variable, leaving us with a formula that is not as clean.
y

Compu
Example
ng the unit tangent vector
Letr(t) = t t, t + t . Find T(t) and compute T( ) and T( ).

Therefore
.

Figure . : Plo ng unit tangent vectors in Example


.

We ndr (t) = t , t + , and

t + .
||r (t) || = ( t ) + ( t + ) =

t
t+
t , t+ =
,
.
t +
t +
t +

; when t = , we have T( ) =
When t = , we have T( ) = / , /


/
, /
. We leave it to the reader to verify each of these is a unit vector. They are plo ed in Figure .
T(t) =

Unit Normal Vector


y

Figure . : Given a direc on in the


plane, there are always two direc ons orthogonal to it.

Note: T(t) is a unit vector, by deni on.


This does not imply that T (t) is also a unit
vector.

Just as knowing the direc on tangent to a path is important, knowing a direcon orthogonal to a path is important. When dealing with real-valued func ons,
we dened the normal line at a point to the be the line through the point that
was perpendicular to the tangent line at that point. We can do a similar thing
with vectorvalued func ons. Givenr(t) in R , we have direc ons perpendicular to the tangent vector, as shown in Figure . . It is good to wonder Is one
of these two direc ons preferable over the other?
Given r(t) in R , there are innite vectors orthogonal to the tangent vector at a given point. Again, we might wonder Is one of these innite choices
preferable over the others? Is one of these the right choice?
The answer in both R and R is Yes, there is one vector that is not only
preferable, it is the right one to choose. Recall Theorem , which states that
if r(t) has constant length, then r(t) is orthogonal to r (t) for all t. We know
T(t), the unit tangent vector, has constant length. Therefore T(t) is orthogonal
to T (t).
Well see that T (t) is more than just a convenient choice of vector that is
orthogonal tor (t); rather, it is the right choice. Since all we care about is the
direc on, we dene this newly found vector to be a unit vector.

Notes:

Deni on

Unit Tangent and Normal Vectors

Unit Normal Vector

Letr(t) be a vectorvalued func on where the unit tangent vector, T(t),


is smooth on an open interval I. The unit normal vector N(t) is
N(t) =

|| T (t) ||

T (t).

Example
Compu ng the unit normal vector
Letr(t) = cos t, sin t, t as in Example
. Sketch both T(/ ) and N(/ )
with ini al points atr(/ ).

, we found T(t) = ( / ) sin t, ( / ) cos t, / .

In Example

Therefore
T (t) =

cos t, sin t,

and

|| T (t) || =

Thus

N(t) = T (t) = cos t, sin t, .


/
We compute T(/ ) = / , , / and N(/ ) = , , . These are
sketched in Figure . .
The previous example was once again too nice. In general, the expression
for T(t) contains frac ons of squareroots, hence the expression of T (t) is very
messy. We demonstrate this in the next example.
Compu
Example
ng the unit normal vector
Letr(t) = t t, t + t as in Example
. Find N(t) and sketchr(t) with the
unit tangent and normal vectors at t = , and .
S

In Example
T(t) =

, we found

t
t+
,
t +
t +

Finding T (t) requires two applica ons of the Quo ent Rule:

Notes:

Figure . : Plo ng unit tangent and


normal vectors in Example . .

Chapter

Vector Valued Func ons

(
)
t + ( ) ( t ) ( t + ) / ( t)
T (t) =
,
t +

(
)
t + ( ) ( t + ) ( t + ) / ( t)
t +

( t+ )
( t)
=
,
/
/
( t + )
( t + )

This is not a unit vector; to nd N(t), we need to divide T (t) by its magnitude.
|| T (t) || =

( t)
( t+ )
+
( t + )
( t + )

( t + )
( t + )

t +

Finally,
4

N(t) =

Figure . : Plo ng unit tangent and


normal vectors in Example
.

( t+ )

( t)

,
/( t + ) ( t + ) / ( t + )

t
t+
,
.
=
t +
t +

Using this formula for N(t), we compute the unit tangent and normal vectors
for t = , and and sketch them in Figure . .
The nal result for N(t) in Example
is suspiciously similar to T(t). There
is a clear reason for this. If u = u , u is a unit vector in R , then the only unit
vectors orthogonal to u are u , u and u , u . Given T(t), we can quickly
determine N(t) if we know which term to mul ply by ( ).
Consider again Figure . , where we have plo ed some unit tangent and
normal vectors. Note how N(t) always points inside the curve, or to the concave side of the curve. This is not a coincidence; this is true in general. Knowing
the direc on thatr(t) turns allows us to quickly nd N(t).

Notes:

Theorem

Unit Tangent and Normal Vectors

Unit Normal Vectors in R

Let r(t) be a vectorvalued func on in R where T (t) is smooth on an


open interval I. Let t be in I and T(t ) = t , t Then N(t ) is either
N(t ) = t , t

or N(t ) = t , t ,

whichever is the vector that points to the concave side of the graph ofr.

Applica on to Accelera on
Let r(t) be a posi on func on. It is a fact (stated later in Theorem ) that
accelera on, a(t), lies in the plane dened by T and N. That is, there are scalars
aT and aN such that
a(t) = aTT(t) + aN N(t).
The scalar aT measures how much accelera on is in the direc on of travel, that
is, it measures the component of accelera on that aects the speed. The scalar
aN measures how much accelera on is perpendicular to the direc on of travel,
that is, it measures the component of accelera on that aects the direc on of
travel.
We can nd aT using the orthogonal projec on of a(t) onto T(t) (review Defini on
in Sec on . if needed). Recalling that since T(t) is a unit vector,
T(t) T(t) = , so we have
proj T(t) a(t) =

)
(
a(t) T(t)
T(t) = a(t) T(t) T(t).
T(t) T(t)
|
{z
}
aT

Thus the amount of a(t) in the direc on of T(t) is aT = a(t) T(t). The same
logic gives aN = a(t) N(t).

While this is a ne way of compu ng aT , there are simpler ways of nding aN


(as nding N itself can be complicated). The following theorem gives alternate
formulas for aT and aN .

Notes:

Note: Keep in mind that both aT and


aN are func ons of t; that is, the scalar
changes depending on t. It is conven on
to drop the (t) nota on from aT (t) and
simply write aT .

Chapter

Vector Valued Func ons

Accelera on in the Plane Dened by T and N

Theorem

Let r(t) be a posi on func on with accelera on a(t) and unit tangent and
normal vectors T(t) and N(t). Then a(t) lies in the plane dened by T(t) and
N(t); that is, there exists scalars aT and aN such that
a(t) = aTT(t) + aN N(t).
Moreover,
)
d(
|| v(t) ||
dt

|| a(t) v(t) ||
= || v(t) || || T (t) ||
aN = a(t) N(t) = || a(t) || aT =
|| v(t) ||
aT = a(t) T(t) =

)
d(
Note the second formula for aT :
|| v(t) || . This measures the rate of
dt
change of speed, which again is the amount of accelera on in the direc on of
travel.
Compu ng aT and aN
Example
Letr(t) = cos t, sin t, t as in Examples

and

. Find aT and aN .

The previous examples give a(t) = cos t, sin t,

and

T(t) =

sin t,

cos t,

and N(t) = cos t, sin t, .

We can nd aT and aN directly with dot products:


aT = a(t) T(t) =

cos t sin t

cos t sin t +

aN = a(t) N(t) = cos t + sin t +

= .

= .

Thus a(t) = T(t) + N(t) = N(t), which is clearly the case.


What is the prac cal interpreta on of these numbers? aT = means the
object is moving at a constant speed, and hence all accelera on comes in the
form of direc on change.
Compu
Example
ng aT and aN
Letr(t) = t t, t + t as in Examples
Notes:

and

. Find aT and aN .

.
S

T(t) =

Unit Tangent and Normal Vectors

The previous examples give a(t) = , and

t
t+
t
t+

,
,
and N(t) =
.
t +
t +
t +
t +

While we can compute aN using N(t), we instead demonstrate using another


formula from Theorem .
t+
t
t
+
=
.
t +
t +
t +

)
(

t
aN = || a(t) || aT =
.
=

t +
t +

aT = a(t) T(t) =

When t = , aT =

and aN =

. We interpret this to

t=2

mean that at t = , the par cle is acclera ng mostly by increasing speed, not
by changing direc on. As the path near t = is rela vely straight, this should
make intui ve sense. Figure . gives a graph of the path
for reference.
Contrast this with t = , where aT = and aN = /
. . Here the
par cles speed is not changing and all accelera on is in the form of direc on
change.

r(t)
t=0
2

Figure
.

: Graphing r(t) in Example

Analyzing projec le mo on
Example
A ball is thrown from a height of
with an ini al speed of
/s and an angle
of eleva on of . Find the posi on func onr(t) of the ball and analyze aT and
aN .
S

on of the ball:
r(t) =

Using Key Idea

of Sec on

t=
t=

cos

)
t,

t +

sin

t+

With a(t) as simple as it is, nding aT is also simple:


aT = a(t) T(t) =

t
.
t t+

t=
t=

which we plot in Figure . .


From this we ndv(t) = cos , t + sin anda(t) = ,
Compu ng T(t) is not dicult, and with some simplica on we nd

t
T(t) =
,
.
t t+
t t+

Notes:

. we form the posi on func-

t=

.
.

t=

Figure . : Plo ng the posi on of a


thrown ball, with s increments shown.

Chapter

Vector Valued Func ons

aT

.
.
.

aN
.
.
.
.

Figure . : A table of values of aT and


aN in Example
.

We choose to not nd N(t) and nd aN through the formula aN = || a(t) || aT :

)
(
t
=
.
aN =

t t+
t t+
Figure . gives a table of values of aT and aN . When t = , we see the
balls speed is decreasing; when t = the speed of the ball is unchanged. This
corresponds to the fact that at t = the ball reaches its highest point.
A er t = we see that aN is decreasing in value. This is because as the ball
falls, its path becomes straighter and most of the accelera on is in the form of
speeding up the ball, and not in changing its direc on.
Our understanding of the unit tangent and normal vectors is aiding our understanding of mo on. The work in Example
gave quan ta ve analysis of
what we intui vely knew.
The next sec on provides two more important steps towards this analysis.
We currently describe posi on only in terms of me. In everyday life, though,
we o en describe posi on in terms of distance (The gas sta on is about miles
ahead, on the le .). The arc length parameter allows us to reference posi on
in terms of distance traveled.
We also intui vely know that some paths are straighter than others and
some are curvier than others, but we lack a measurement of curviness. The
arc length parameter provides a way for us to compute curvature, a quan ta ve
measurement of how curvy a curve is.

Notes:

Exercises

Terms and Concepts

In Exercises

, a posi on func on r(t) is given along


with its unit tangent vector T(t) evaluated at t = a, for some
value of a.

. If T(t) is a unit tangent vector, what is || T(t) ||?


. If
N(t) is a unit normal vector, what is
N(t) r (t)?
. The accelera on vector a(t) lies in the plane dened by
what two vectors?
. aT measures how much the accelera on is aec ng the
of an object.

Problems
In Exercises , givenr(t), nd T(t) and evaluate it at the
indicated value of t.
. r(t) =

t ,t t ,

. r(t) = t, cos t,

t=

. r(t) = t, cos t,

t=

t = /

. r(t) = cos t, sin t

. r(t) = t, t

. r(t) = cos t, sin t

. r(t) = et , et

t,

t +

; T( ) =

. r(t) = ( + sin t) cos t, sin t; T( ) =

. r(t) = cos t, sin t ; T(/ ) =

, nd
N(t).

. r(t) = a cos t, a sin t, bt;

a>

. r(t) = cos(at), sin(at), t


In Exercises , nd aT and aN givenr(t). Sketchr(t) on
the indicated interval, and comment on the rela ve sizes of
aT and aN at the indicated t values.

. r(t) = t, /t on ( , ]; consider t =

. r(t) = cos t, sin t, t =


In Exercises

, nd
N(t) using Deni on
the result using Theorem .

. r(t) = t, t on [ , ]; consider t =

. r(t) = cos t, sin t , t = /

. r(t) =

. r(t) = cos t, sin t, sin t

In Exercises
, nd the equa on of the line tangent to
the curve at the indicated t-value using the unit tangent vector. Note: these are the same problems as in Exercises
.
. r(t) =

. r(t) = cos t, sin t; T(/ ) =

, nd
N(a).

. r(t) = t, sin t, cos t

. r(t) = cos t, sin t, t =

t ,t t ,

(b) Using a graph ofr(t) and Theorem

In Exercises

t = /

. r(t) = cos t, sin t , t = /

(a) Conrm that T(a) is as stated.

. Conrm

and t = .
and t = .

. r(t) = cos t, sin t on [ , ]; consider t =


t = / .

and

/
. r(t) = cos(t ), sin(t ) on ( , ]; consider t =

and t = .
. r(t) = a cos t, a sin t, bt on [ , ], where a, b > ; consider t = and t = / .
. r(t) = cos t, sin t, sin t on [ , ]; consider t =
and t = / .

Chapter

Vector Valued Func ons

.
y
t=

t=
r(t)
t=
4

(a)
y
s=
s=
s=
s=
s=

r(s)

s=
s=

The Arc Length Parameter and Curvature

In normal conversa on we describe posi on in terms of both me and distance.


For instance, imagine driving to visit a friend. If she calls and asks where you
are, you might answer I am
minutes from your house, or you might say I
am miles from your house. Both answers provide your friend with a general
idea of where you are.
Currently, our vectorvalued func ons have dened points with a parameter
t,
which
we o en take to represent me. Consider Figure . (a), wherer(t) =

t t, t + t is graphed and the points corresponding to t = , and are


shown. Note how the arc length between t = and t = is smaller than the
arc length between t = and t = ; if the parameter t is me andr is posi on,
we can say that the par cle traveled faster on [ , ] than on [ , ].
Now consider Figure . (b), where the same graph is parametrized by a
dierent variable s. Points corresponding to s = through s = are plo ed.
The arc length of the graph between each adjacent pair of points is . We can
view this parameter s as distance; that is, the arc length of the graph from s =
to s = is , the arc length from s = to s = is , etc. If one wants to nd the
point . units from an ini al loca on (i.e., s = ), one would compute r( . ).
This parameter s is very useful, and is called the arc length parameter.
How do we nd the arc length parameter?
Start with any parametriza on of r. We can compute the arc length of the
graph ofr on the interval [ , t] with

arc length =

(b)

Figure . : Introducing the arc length


parameter.

||r (u) || du.

We can turn this into a func on: as t varies, we nd the arc length s from to t.
This func on is

s(t) =

||r (u) || du.

. )

This establishes a rela onship between s and t. Knowing this rela onship
explicitly, we can rewriter(t) as a func on of s: r(s). We demonstrate this in an
example.
Finding the arc length parameter
Example
Letr(t) = t , t + . Parametrizer with the arc length parameter s.
S

Using Equa on (
s(t) =

Notes:

. ), we write
t

||r (u) || du.

The Arc Length Parameter and Curvature

We can integrate this, explicitly nding a rela onship between s and t:


t
||r (u) || du
s(t) =
t
=
+ du
t
du
=

y
t=

= t.

Since s = t, we can write t = s/ and replace t inr(t) with s/ :

r(s) = (s/ ) , (s/ ) + =


s , s+
.

s=
s=
s=
t=

Clearly, as shown in Figure . , the graph of r is a line, where t = corresponds to the point ( , ). What point on the line is units away from this
ini al point? We nd it with s( ) = / , / .
Is the point ( / , / ) really units away from ( , )? We use the Distance Formula to check:
(

(
)
)

=
( ) +

+
=
= .
d=
Yes, s( ) is indeed units away, in the direc on of travel, from the ini al point.
Things worked out very nicely in Example . ; we were able to establish
directly that s = t. Usually, the arc length parameter is much more dicult to
describe in terms of t, a result of integra ng a squareroot. There are a number
of things that we can learn about the arc length parameter from Equa on ( . ),
though, that are incredibly useful.
First, take the deriva ve of s with respect to t. The Fundamental Theorem
of Calculus (see Theorem ) states that
ds
= s (t) = ||r (t) ||.
dt

. )

Le ng t represent me and r(t) represent posi on, we see that the rate of
change of s with respect to t is speed; that is, the rate of change of distance
traveled is speed, which should match our intui on.
The Chain Rule states that
dr ds
dr
=

dt
ds dt
r (t) = r (s) ||r (t) ||.

Notes:

s=

s=

s=
x

Figure . : Graphing r in Example


with parameters t and s.

Chapter

Vector Valued Func ons


Solving forr (s), we have
r (s) =

r (t)
= T(t),
||r (t) ||

. )

where T(t) is the unit tangent vector. Equa on . is o en misinterpreted, as


one is tempted to think it states r (t) = T(t), but there is a big dierence betweenr (s) andr (t). The key to take from it is thatr (s) is a unit vector. In fact,
the following theorem states that this characterizes the arc length parameter.
Theorem

Arc Length Parameter

Let r(s) be a vectorvalued func on. The parameter s is the arc length
parameter if, and only if, ||r (s) || = .

Curvature

(a)
y

Consider points A and B on the curve graphed in Figure . (a). One can
readily argue that the curve curves more sharply at A than at B. It is useful to use
a number to describe how sharply the curve bends; that number is the curvature
of the curve.
We derive this number in the following way. Consider Figure . (b), where
unit tangent vectors are graphed around points A and B. No ce how the direcon of the unit tangent vector changes quite a bit near A, whereas it does not
change as much around B. This leads to an important concept: measuring the
rate of change of the unit tangent vector with respect to arc length gives us a
measurement of curvature.
Deni on

(b)
Figure . : Establishing the concept of
curvature.

Curvature

Letr(s) be a vectorvalued func on where s is the arc length parameter.


The curvature of the graph ofr(s) is


dT


=
= T (s) .
ds
Ifr(s) is parametrized by the arc length parameter, then
T(s) =

Notes:

r (s)
||r (s) ||

and N(s) =

T (s)
.
|| T (s) ||

.
Having dened || T (s) || = , we can rewrite the second equa on as
T (s) = N(s).

. )

We already knew that T (s) is in the same direc on as N(s); that is, we can think
of T(s) as being pulled in the direc on of N(s). How hard is it being pulled?
By a factor of . When the curvature is large, T(s) is being pulled hard and the
direc on of T(s) changes rapidly. When is small, T(s) is not being pulled hard
and hence its direc on is not changing rapidly.
We use Deni on to nd the curvature of the line in Example
.
Example
Use Deni on

Finding the curvature of a line


to nd the curvature ofr(t) = t , t + .

In Example
, we found that the arc length parameter was
S
dened by s = t, so r(s) = t/ , t/ + parametrized r with the arc
length parameter. To nd , we need to nd T (s).
T(s) = r (s) (recall this is a unit vector)
= / , / .
Therefore
T (s) = ,
and


= T (s) = .

It probably comes as no surprise that the curvature of a line is . (How curvy


is a line? It is not curvy at all.)
While the deni on of curvature is a beau ful mathema cal concept, it is
nearly impossible to use most of the me; wri ng r in terms of the arc length
parameter is generally very hard. Fortunately, there are other methods of calcula ng this value that are much easier. There is a tradeo: the deni on is
easy to understand though hard to compute, whereas these other formulas
are easy to compute though it may be hard to understand why they work.

Notes:

The Arc Length Parameter and Curvature

Chapter

Vector Valued Func ons

Theorem

Formulas for Curvature

Let C be a smooth curve on an open interval I in the plane or in space.


. If C is dened by y = f(x), then
= (

|f (x)|
(
) )
+ f (x)

. If C is dened as a vectorvalued func on in the plane, r(t) =


x(t), y(t), then
= (

|x y x y |
)/ .
(x ) + (y )

. If C is dened in space by a vectorvalued func onr(t), then


=

a(t) N(t)
|| T (t) ||
||r (t) r (t) ||
=
.
=

||r (t) ||
||r (t) ||
|| v(t) ||

We prac ce using these formulas.


Finding the curvature of a circle
Example
Find the curvature of a circle with radius r, dened by c(t) = r cos t, r sin t.
Before we start, we should expect the curvature of a circle
S
to be constant, and not dependent on t. (Why?)
We compute using the second part of Theorem
.
=

|(r sin t)(r sin t) (r cos t)(r cos t)|


(
)/
(r sin t) + (r cos t)

=(
=

r (sin t + cos t)
)/
r (sin t + cos t)

r
= .
r
r

We have found that a circle with radius r has curvature = /r.

Notes:

The Arc Length Parameter and Curvature

Example
gives a great result. Before this example, if we were told The
curve has a curvature of at point A, we would have no idea what this really meant. Is big does is correspond to a really sharp turn, or a not-sosharp turn? Now we can think of in terms of a circle with radius / . Knowing
the units (inches vs. miles, for instance) allows us to determine how sharply the
curve is curving.
y

Let a point P on a smooth curve C be given, and let be the curvature of the
curve at P. A circle that:
passes through P,
lies on the concave side of C,
has a common tangent line as C at P and

has radius r = / (hence has curvature )

is the oscula ng circle, or circle of curvature, to C at P, and r is the radius of curvature. Figure . shows the graph of the curve seen earlier in Figure .
and its oscula ng circles at A and B. A sharp turn corresponds to a circle with
a small radius; a gradual turn corresponds to a circle with a large radius. Being
able to think of curvature in terms of the radius of a circle is very useful. (The
word oscula ng comes from a La n word related to kissing; an oscula ng circle kisses the graph at a par cular point. Many beau ful ideas in mathema cs
have come from studying the oscula ng circles to a curve.)
Finding curvature
Example
Find the curvature of the parabola dened by y = x at the vertex and at x = .
S

We use the rst formula found in Theorem


(x) = (
=(

| |

+ ( x)
+ x

)
/

Figure . : Illustra ng the oscula ng


circles for the curve seen in Figure . .

.
10

.
5

At the vertex (x = ), the curvature is = . At x = , the curvature


is = /( ) / .
. So at x = , the curvature of y = x is that of
a circle of radius / ; at x = , the curvature is that of a circle with radius
/ .
. . This is illustrated in Figure . . At x = , the curvature is
.
; the graph is nearly straight as the curvature is very close to .

Notes:

x
10

Figure .
y=x .

: Examining the curvature of

Chapter

Vector Valued Func ons


y

Example
Finding curvature

Find where the curvature ofr(t) = t, t , t is maximized.

We use the third formula in Theorem


asr(t) is dened
S
in space. We leave it to the reader to verify that

r (t) = , t, t , r (t) = , , t , and r (t)r (t) =


t , t, .

Thus

.
2

||r (t) r (t) ||


||r (t) ||

||
t , t,
||
=
|| , t, t ||

t +
t +
= (
)
+ t + t

(t) =

(a)

(b)
Figure . : Understanding the curvature of a curve in space.

While this is not a par cularly nice formula, it does explictly tell us what the
curvature is at a given t value. To maximize (t), we should solve (t) = for
t. This is doable, but very me consuming. Instead, consider the graph of (t)
as given in Figure . (a). We see that is maximized at two t values; using a
numerical solver, we nd these values are t .
. In part (b) of the gure
we graphr(t) and indicate the points where curvature is maximized.

Curvature and Mo on
Let r(t) be a posi on func on of an object, with velocity v(t) = r (t) and
accelera on a(t) = r (t). In Sec on . we established that accelera on is in
the plane formed by T(t) and N(t), and that we can nd scalars aT and aN such
that
a(t) = aTT(t) + aN N(t).
Theorem

gives formulas for aT and aN :


aT =

)
d(
|| v(t) ||
and
dt

aN =

|| v(t) a(t) ||
.
|| v(t) ||

We understood that the amount of accelera on in the direc on of T relates only


to how the speed of the object is changing, and that the amount of accelera on
in the direc on of N relates to how the direc on of travel of the object is changing. (That is, if the object travels at constant speed, aT = ; if the object travels
in a constant direc on, aN = .)

Notes:

The Arc Length Parameter and Curvature

In Equa on ( . ) at the beginning of this sec on, we found s (t) = ||v(t) ||.
We can combine this fact with the above formula for aT to write
)
d( )
d(
s (t) = s (t).
|| v(t) || =
aT =
dt
dt

Since s (t) is speed, s (t) is the rate at which speed is changing with respect to
me. We see once more that the component of accelera on in the direc on of
travel relates only to speed, not to a change in direc on.
Now compare the formula for aN above to the formula for curvature in Theorem
:
aN =
Thus

|| v(t) a(t) ||
|| v(t) ||

and

|| v(t) a(t) ||
||r (t) r (t) ||
=
.

||r (t) ||
|| v(t) ||

aN = || v(t) ||
(
)
= s (t)

. )

This last equa on shows that the component of accelera on that changes
the objects direc on is dependent on two things: the curvature of the path and
the speed of the object.
Imagine driving a car in a clockwise circle. You will naturally feel a force pushing you towards the door (more accurately, the door is pushing you as the car
is turning and you want to travel in a straight line). If you keep the radius of
the circle constant but speed up (i.e., increasing s (t)), the door pushes harder
against you (aN has increased). If you keep your speed constant but ghten the
turn (i.e., increase ), once again the door will push harder against you.
Pu ng our new formulas for aT and aN together, we have
a(t) = s (t)T(t) + || v(t) || N(t).

This is not a par cularly prac cal way of nding aT and aN , but it reveals some
great concepts about how accelera on interacts with speed and the shape of a
curve.
Curvature and road design
Example
The minimum radius of the curve in a highway cloverleaf is determined by the
opera ng speed, as given in the table in Figure . . For each curve and speed,
compute aN .
Using Equa on ( . ), we can compute the accelera on
S
normal to the curve in each case. We start by conver ng each speed from miles
per hour to feet per second by mul plying by
/
.

Notes:

Opera ng
Speed (mph)

Minimum
Radius ( )

Figure . : Opera ng speed and minimum radius in highway cloverleaf design.

Chapter

Vector Valued Func ons

mph,

/s,

= /

aN = || v(t) ||
(
)
=
.
/s .

= .

mph,

aN =
= .

mph,

/s,

aN =
= .

.
/s .

= /
)

/s, = /
( )
/s .

Note that each accelera on is similar; this is by design. Considering the classic
Force = mass accelera on formula, this accelera on must be kept small in
order for the res of a vehicle to keep a grip on the road. If one travels on a
turn of radius
at a rate of mph, the accelera on is double, at .
/s .
If the accelera on is too high, the fric onal force created by the res may not be
enough to keep the car from sliding. Civil engineers rou nely compute a safe
design speed, then subtract - mph to create the posted speed limit for addional safety.
We end this chapter with a reec on on what weve covered. We started
with vectorvalued func ons, which may have seemed at the me to be just
another way of wri ng parametric equa ons. However, we have seen that the
vector perspec ve has given us great insight into the behavior of func ons and
the study of mo on. Vectorvalued posi on func ons convey displacement,
distance traveled, speed, velocity, accelera on and curvature informa on, each
of which has great importance in science and engineering.

Notes:

Exercises

Terms and Concepts


. It is common to describe posi on in terms of both
and/or
.
. A measure of the curviness of a curve is

. Describe in your own words what an oscula ng circle is.

N(s).

. Given a posi on func on r(t), how are aT and aN aected


by the curvature?

Problems
In Exercises
, a posi on func on r(t) is given, where
t = corresponds to the ini al posi on. Find the arc length
parameter s, and rewriter(t) in terms of s; that is, ndr(s).
. r(t) = t, t, t

. C is dened byr(t) = t t, t , t ; points given


at t = and t = .
. C is dened by r(t) = cos t, sin t, t; points given at
t = and t = / .
. C is dened by r(t) = cos t,
given at t = and t = / .
In Exercises
is maximized.

sin t,

cos t

(a) Using a sketch, determine at which of these points the


curvature is greater.
(b) Find the curvature of C, and evaluate at each of the
given points.
. C is dened by y = x x; points given at x =
x= / .

and

; points given at x =

and

x= .

x +

. C is dened by y = cos x; points given at x =


x = / .
. C is dened by y =
x = and x = / .

cos t; points

, nd the value of x or t where curvature

. r(t) = t + t, t t

, a curve C is described along with points

. C is dened by y =

sin t,

. y = sin x

. r(t) = cos t, sin t, t

In Exercises
on C.

. C is dened byr(t) = t + , t , t + ; points given


at t = and t = .

. y=

. r(t) = cos t, sin t

. r(t) = cos t,

. C is dened byr(t) = t , t t ; points given at t =


and t = .
. C is dened by r(t) = tan t, sec t; points given at t =
and t = / .

. Give two shapes with constant curvature.

. Complete the iden ty: T (s) =

. C is dened by r(t) = cos t, sin t cos t ; points given at


t = and t = / .

. r(t) = t, /t, /t

In Exercises
cated value.

, nd the radius of curvature at the indi-

. y = tan x, at x = /
. y = x + x , at x = /
. r(t) = cos t, sin( t), at t =
. r(t) = cos( t), t, at t =
In Exercises , nd the equa on of the oscula ng circle
to the curve at the indicated t-value.

and

x on ( , ); points given at

. C is dened byr(t) = cos t, sin( t); points given at t =


and t = / .

. r(t) = t, t , at t =

. r(t) = cos t, sin t, at t =


. r(t) = cos t, sin t, at t = /

. r(t) = t t, t + t , at t =

:F

S 6

A func on of the form y = f(x) is a func on of a single variable; given a value


of x, we can nd a value y. Even the vectorvalued func ons of Chapter are
singlevariable func ons; the input is a single variable though the output is a
vector.
There are many situa ons where a desired quan ty is a func on of two or
more variables. For instance, wind chill is measured by knowing the temperature
and wind speed; the volume of a gas can be computed knowing the pressure
and temperature of the gas; to compute a baseball players ba ng average, one
needs to know the number of hits and the number of atbats.
This chapter studies mul variable func ons, that is, func ons with more
than one input.

Introduc on to Mul variable Func ons

Deni on

Func on of Two Variables

Let D be a subset of R . A func on f of two variables is a rule that assigns


each pair (x, y) in D a value z = f(x, y) in R. D is the domain of f; the set
of all outputs of f is the range.
Understanding a func on of two variables
Example
Let z = f(x, y) = x y. Evaluate f( , ), f( , ), and f( , ); nd the domain
and range of f.
S

Using the deni on f(x, y) = x y, we have:


f( , ) =
f( , ) =

=
=

f( , ) = ( )

The domain is not specied, so we take it to be all possible pairs in R for which
f is dened. In this example, f is dened for all pairs (x, y), so the domain D of f
is R .
The output of f can be made as large or small as possible; any real number r
can be the output. (In fact, given any real number r, f( , r) = r.) So the range
R of f is R.

Chapter

Func ons of Several Variables


Example
Let f(x, y) =

5
y2
x2
+
=1
9
4

Understanding a func on of two variables


x
y

. Find the domain and range of f.

The domain is all pairs (x, y) allowable as input in f. Because


S
of the squareroot, we need (x, y) such that x y :

Figure . : Illustra ng the domain of


f(x, y) in Example
.

The above equa on describes the interior of an ellipse as shown in Figure . .


We can represent the domain D graphically with the gure; in set nota on, we
can write D = {(x, y)| x + y }.
The range is the set of all possible output values. The squareroot ensures
that all output is . Since the x and y terms are squared, then subtracted, inside the squareroot, the largest output value comes at x = , y = : f( , ) =
. Thus the range R is the interval [ , ].

Graphing Func ons of Two Variables


(
)
The graph of a func on f of two variables is the set of all points x, y, f(x, y)
where (x, y) is in the domain of f. This creates a surface in space.
One can begin sketching a graph by plo ng points, but this has limita ons.
Consider Figure

(a)

. (a) where

points have been plo ed of f(x, y) =

x +y +
More points have been plo ed than one would reasonably want to do by hand,
yet it is not clear at all what the graph of the func on looks like. Technology allows us to plot lots of points, connect adjacent points with lines and add shading
to create a graph like Figure . b which does a far be er job of illustra ng the
behavior of f.
While technology is readily available to help us graph func ons of two variables, there is s ll a paperandpencil approach that is useful to understand and
master as it, combined with highquality graphics, gives one great insight into
the behavior of a func on. This technique is known as sketching level curves.

Level Curves
It may be surprising to nd that the problem of represen ng a three dimensional surface on paper is familiar to most people (they just dont realize it).
Topographical maps, like the one shown in Figure . , represent the surface
of Earth by indica ng points with the same eleva on with contour lines. The

(b)
Figure . : Graphing a func on of two
variables.

Notes:

Introduc on to Mul variable Func ons

eleva ons marked are equally spaced; in this example, each thin line indicates
an eleva on change in
increments and each thick line indicates a change
of
. When lines are drawn close together, eleva on changes rapidly (as
one does not have to travel far to rise
). When lines are far apart, such as
near Aspen Campground, eleva on changes more gradually as one has to walk
farther to rise
.
Given a func on z = f(x, y), we can draw a topographical map of f by
drawing level curves (or, contour lines). A level curve at z = c is a curve in the
x-y plane such that for all points (x, y) on the curve, f(x, y) = c.
When drawing level curves, it is important that the c values are spaced equally
apart as that gives the best insight to how quickly the eleva on is changing.
Examples will help one understand this concept.
Example
Let f(x, y) =
. and .

Drawing Level Curves


y
x
. Find the level curves of f for c = , . , . , . ,

Sample taken from the public domain USGS Digital Raster Graphics,
http://topmaps.usgs.gove/drg/.

Consider rst
S
c = . The level curve for c = is the set of
all points (x, y) such that =
x y . Squaring both sides gives us
x

= ,

an ellipse centered at ( , ) with horizontal major axis of length and minor axis
of length . Thus for any point (x, y) on this curve, f(x, y) = .
Now consider the level curve for c = .

. =
.
x
x
.

This is also an ellipse, where a =

Notes:

Figure . : A topographical map displays


eleva on by drawing contour lines, along
with the eleva on is constant.

=
y

= .

y
.

= .

and b =

Chapter

Func ons of Several Variables


y

In general, for z = c, the level curve is:

c = .6

c=
c=

c =

x
y
+
= ,
( c )
( c )

(a)

ellipses that are decreasing in size as c increases. A special case is when c = ;


there the ellipse is just the point ( , ).
The level curves are shown in Figure . (a). Note how the level curves for
c = and c = . are very, very close together: this indicates that f is growing
rapidly along those curves.
In Figure . (b), the curves are drawn on a graph of f in space. Note how
the eleva ons are evenly spaced. Near the level curves of c = and c = . we
can see that f indeed is growing quickly.
Example
Let f(x, y) =
(b)

Analyzing Level Curves


x+y
. Find the level curves for z = c.
x +y +

We begin by se ng f(x, y) = c for an arbitrary c and seeing


S
if algebraic manipula on of the equa on reveals anything signicant.

Figure . : Graphing the level curves in


Example
.

x+y
=c
x +y +
x + y = c(x + y + ).
We recognize this as a circle, though the center and radius are not yet clear. By
comple ng the square, we can obtain:
(
x

(
+ y

(
)
/( c ) , where |c| <
a circle centered at /( c), /( c) with radius

/ . The level curves for c = . , . and . are sketched in Figure


. (a). To help illustrate eleva on, we use thicker lines for c values near ,
and dashed lines indicate where c < .
There is one special level curve, when c = . The level curve in this situa on
is x + y = , the line y = x.

Notes:

Introduc on to Mul variable Func ons

In Figure . (b) we see a graph of the surface. Note how the y-axis is pointing away from the viewer to more closely resemble the orienta on of the level
curves in (a).
Seeing the level curves helps us understand the graph. For instance, the
graph does not make it clear that one can walk along the line y = x without
eleva on change, though the level curve does.

y
c = 0.2

4
c = 0.4
2

Func ons of Three Variables

We extend our study of mul variable func ons to func ons of three variables. (One can make a func on of as many variables as one likes; we limit our
study to three variables.)
Deni on

c=0

.
.

(a)

Func on of Three Variables

Let D be a subset of R . A func on f of three variables is a rule that


assigns each triple (x, y, z) in D a value w = f(x, y, z) in R. D is the domain
of f; the set of all outputs of f is the range.
Note how this deni on closely resembles that of Deni on

Example

Understanding a func on of three variables


x + z + sin y
. Evaluate f at the point ( , , ) and nd the
Let f(x, y, z) =
x+ yz
domain and range of f.
+ + sin
= .
+ ( )
As the domain of f is not specied, we take it to be the set of all triples (x, y, z)
for which f(x, y, z) is dened. As we cannot divide by , we nd the domain D is
S

f( , , ) =

D = {(x, y, z) | x + y z = }.
We recognize that the set of all points in R that are not in D form a plane in
space that passes through the origin (with normal vector , , ).
We determine the range R is R; that is, all real numbers are possible outputs
of f. There is no set way of establishing this. Rather, to get numbers near we
can let y = and choose z x . To get numbers of arbitrarily large magnitude, we can let z x + y.

Notes:

(b)
Figure . : Graphing the level curves in
Example
.

Chapter

Func ons of Several Variables

Level Surfaces
It is very dicult to produce a meaningful graph of a func on of three variables. A func on of one variable is a curve drawn in dimensions; a func on of
two variables is a surface drawn in dimensions; a func on of three variables is
a hypersurface drawn in dimensions.
There are a few techniques one can employ to try to picture a graph of
three variables. One is an analogue of level curves: level surfaces. Given w =
f(x, y, z), the level surface at w = c is the surface in space formed by all points
(x, y, z) where f(x, y, z) = c.
Finding level surfaces
Example
If a point source S is radia ng energy, the intensity I at a given point P in space
is inversely propor onal to the square of the distance between S and P. That is,
k
for some constant k.
when S = ( , , ), I(x, y, z) =
x +y +z
Let k = ; nd the level surfaces of I.
We can (mostly) answer this ques on using common sense.
S
If energy (say, in the form of light) is emana ng from the origin, its intensity will
be the same at all points equidistant from the origin. That is, at any point on
the surface of a sphere centered at the origin, the intensity should be the same.
Therefore, the level surfaces are spheres.
We now nd this mathema cally. The level surface at I = c is dened by
c=
c

.
.
.
.

r
.
.
.
.
.
.
.
.
.

Figure . : A table of c values and the


corresponding radius r of the spheres of
constant value in Example
.

x +y +z

A small amount of algebra reveals


x +y +z =

Given an intensity c, the level surface I = c is a sphere of radius / c, centered


at the origin.
Figure . gives a table of the radii of the spheres for given c values. Normally one would use equally spaced c values, but these values have been chosen
purposefully. At a distance of . from the point source, the intensity is ; to
move to a point of half that intensity, one just moves out . to . not much
at all. To again halve the intensity, one moves . , a li le more than before.
Note how each me the intensity if halved, the distance required to move
away grows. We conclude that the closer one is to the source, the more rapidly
the intensity changes.
In the next sec on we apply the concepts of limits to func ons of two or
more variables.
Notes:

Exercises

Terms and Concepts

. f(x, y) = x y ; c = , ,

. Give two examples (other than those given in the text) of


real world func ons that require more than one input.

. f(x, y) = x y ; c = , ,
. f(x, y) =

x y
;c= , ,
y x

. Most people are familiar with the concept of level curves in


the context of
maps.

. f(x, y) =

x y
;c= , ,
x +y +

. T/F: Along a level curve, the output of a func on does not


change.

. f(x, y) =

yx
; c = , , , ,
x

. The analogue of a level curve for func ons of three variables is a level
.

. f(x, y) =

. The graph of a func on of two variables is a

. What does it mean when level curves are close together?


Far apart?

Problems

. f(x, y) = x + y ; c = , , ,
In Exercises
, give the domain and range of the funcons of three variables.
. f(x, y, z) =

In Exercises , give the domain and range of the mul variable func on.

. f(x, y, z) =

. f(x, y) = x + y +
. f(x, y) = x + y
. f(x, y) = x y
. f(x, y) =
. f(x, y) =

x+ y
x +y +

x + y ;c= , , ,

. f(x, y, z) =

x
x+ y z
x y z

zx +y

. f(x, y, z) = z sin x cos y


In Exercises
, describe the level surfaces of the given
func ons of three variables.
. f(x, y, z) = x + y + z

. f(x, y) = sin x cos y

. f(x, y, z) = z x + y

. f(x, y) =

. f(x, y, z) =

x +y
z

. f(x, y, z) =

z
xy

. f(x, y) =

x y

x +y

In Exercises
, describe in words and sketch the level
curves for the func on and given c values.
. f(x, y) = x y; c = , ,

. Compare the level curves of Exercises


and . How are
they similar, and how are they dierent? Each surface is a
quadric surface; describe how the level curves are consistent with what we know about each surface.

Chapter

Func ons of Several Variables

y
P1

P2

(a)

We con nue with the pa ern we have established in this text: a er dening a
new kind of func on, we apply calculus ideas to it. The previous sec on dened
func ons of two and three variables; this sec on inves gates what it means for
these func ons to be con nuous.
We begin with a series of deni ons. We are used to open intervals such
as ( , ), which represents the set of all x such that < x < , and closed
intervals such as [ , ], which represents the set of all x such that x .
We need analogous deni ons for open and closed sets in the x-y plane.
Deni on

Let S be a set of points in R . A point P in R is a boundary point of S


if all open disks centered at P contain both points in S and points not in S.

P2

(b)

A point P in S is an interior point of S if there is an open disk centered at


P that contains only points in S.
A set S is open if every point in S is an interior point.

A set S is closed if it contains all of its boundary points.

P1

A set S is bounded if there is an M > such that the open disk, centered at the origin with radius M, contains S. A set that is not bounded
is unbounded.

P2

Open Disk, Boundary and Interior Points, Open and


Closed Sets, Bounded Sets

An open disk B in R
centered at (x , y ) with radius r is the set of all
points (x, y) such that (x x ) + (y y ) < r.

P1

Limits and Con nuity of Mul variable Func ons

(c)

Figure . : Illustra ng open and closed


sets in the x-y plane.

Figure . shows several sets in the x-y plane. In each set, point P lies on
the boundary of the set as all open disks centered there contain both points in,
and not in, the set. In contrast, point P is an interior point for there is an open
disk centered there that lies en rely within the set.
The set depicted in Figure . (a) is a closed set as it contains all of its boundary points. The set in (b) is open, for all of its points are interior points (or, equivalently, it does not contain any of its boundary points). The set in (c) is neither
open nor closed as it contains some of its boundary points.

Notes:

.
Example

Limits and Con nuity of Mul variable Func ons

Determining open/closed, bounded/unbounded

Determine if the domain of the func on f(x, y) =


or neither, and if it is bounded.

is open, closed,

This domain of this func on was found in Example


to be
y
x
D = {(x, y) | + }, the region bounded by the ellipse + = . Since
the region includes the boundary (indicated by the use of ), the set contains
all of its boundary points and hence is closed. The region is bounded as a disk
of radius , centered at the origin, contains D.
S

Determining open/closed, bounded/unbounded


Example
Determine if the domain of f(x, y) = xy is open, closed, or neither.
As we cannot divide by , we nd the domain to be D =
S
{(x, y) | x y = }. In other words, the domain is the set of all points (x, y) not
on the line y = x.
The domain is sketched in Figure . . Note how we can draw an open disk
around any point in the domain that lies en rely inside the domain, and also
note how the only boundary points of the domain are the points on the line
y = x. We conclude the domain is an open set. The set is unbounded.

Limits
Recall a pseudodeni on of the limit of a func on of one variable: lim f(x) =
xc

L means that if x is really close to c, then f(x) is really close to L. A similar


pseudodeni on holds for func ons of two variables. Well say that

lim

(x,y)(x ,y )

f(x, y) = L

means if the point (x, y) is really close to the point (x , y ), then f(x, y) is really
close to L. The formal deni on is given below.
Deni on

Limit of a Func on of Two Variables

Let S be an open set containing (x , y ), and let f be a func on of two


variables dened on S, except possibly at (x , y ). The limit of f(x, y) as
(x, y) approaches (x , y ) is L, denoted
lim

(x,y)(x ,y )

f(x, y) = L,

means that given any > , there exists > such that for all (x, y) =
(x , y ), if (x, y) is in the open disk centered at (x , y ) with radius , then
|f(x, y) L| < .
Notes:

.
Figure . : Sketching the domain of the
func on in Example
.

Chapter

Func ons of Several Variables

The concept behind Deni on


is sketched in Figure . . Given > ,
nd > such that if (x, y) is any point in the open disk centered at (x , y ) in
the x-y plane with radius , then f(x, y) should be within of L.
Compu ng limits using this deni on is rather cumbersome. The following
theorem allows us to evaluate limits much more easily.
Theorem

Basic Limit Proper es of Func ons of Two Variables

Let b, x , y , L and K be real numbers, let n be a posi ve integer, and let


f and g be func ons with the following limits:
lim

(x,y)(x ,y )

Figure . : Illustra ng the deni on of


a limit. The open disk in the x-y plane has
radius . Let (x, y) be any point in this
disk; f(x, y) is within of L.

f(x, y) = L

and

lim

(x,y)(x ,y )

g(x, y) = K.

The following limits hold.


. Constants:

lim

b=b

lim

x=x ;

lim

(
)
f(x, y) g(x, y) = L K

(x,y)(x ,y )

. Iden ty

(x,y)(x ,y )

. Sums/Dierences:
. Scalar Mul ples:
. Products:
. Quo ents:
. Powers:

(x,y)(x ,y )

lim

(x,y)(x ,y )

y=y

lim

b f(x, y) = bL

lim

f(x, y) g(x, y) = LK

lim

f(x, y)/g(x, y) = L/K, (K = )

lim

f(x, y)n = Ln

(x,y)(x ,y )

(x,y)(x ,y )

(x,y)(x ,y )

(x,y)(x ,y )

This theorem, combined with Theorems and of Sec on . , allows us to


evaluate many limits.
Evalua ng a limit
Example
Evaluate the following limits:
.

Notes:

lim

(x,y)( ,)

y
+ cos(xy)
x

lim

(x,y)( , )

xy
x +y

Limits and Con nuity of Mul variable Func ons

. The aforemen oned theorems allow us to simply evaluate y/x + cos(xy)


when x = and y = . If an indeterminate form is returned, we must do
more work to evaluate the limit; otherwise, the result is the limit. Therefore
lim

(x,y)( ,)

+ cos(xy) = + cos
x
= .

. We a empt to evaluate the limit by subs tu ng in for x and y, but the


result is the indeterminate form / . To evaluate this limit, we must
do more work, but we have not yet learned what kind of work to do.
Therefore we cannot yet evaluate this limit.
When dealing with func ons of a single variable we also considered one
sided limits and stated
lim f(x) = L

xc

if, and only if,

lim f(x) = L

xc+

and

lim f(x) = L.

xc

That is, the limit is L if and only if f(x) approaches L when x approaches c from
either direc on, the le or the right.
In the plane, there are innite direc ons from which (x, y) might approach
(x , y ). In fact, we do not have to restrict ourselves to approaching (x , y ) from
a par cular direc on, but rather we can approach that point along a path that is
not a straight line. It is possible to arrive at dierent limi ng values by approaching (x , y ) along dierent paths. If this happens, we say that
lim
f(x, y)
(x,y)(x ,y )

does not exist (this is analogous to the le and right hand limits of single variable
func ons not being equal).
Our theorems tell us that we can evaluate most limits quite simply, without
worrying about paths. When indeterminate forms arise, the limit may or may
not exist. If it does exist, it can be dicult to prove this as we need to show the
same limi ng value is obtained regardless of the path chosen. The case where
the limit does not exist is o en easier to deal with, for we can o en pick two
paths along which the limit is dierent.
Showing limits do not exist

Example
. Show

y = mx.

Notes:

lim

(x,y)( , )

xy
does not exist by nding the limits along the lines
x +y

Chapter

Func ons of Several Variables


. Show

sin(xy)
does not exist by nding the limit along the path
) x+y

lim

(x,y)( ,

y = sin x.
S

xy
along the lines y = mx means replace all ys
(x,y)( , ) x + y
with mx and evalua ng the resul ng limit:

. Evalua ng

lim

mx
x(mx)
= lim
x x (m + )
(x,mx)( , ) x + (mx)
m
= lim
x m +
m
.
=
m +
While the limit exists for each choice of m, we get a dierent limit for each
choice of m. That is, along dierent lines we get diering limi ng values,
meaning the limit does not exist.
lim

. Let f(x, y) =

sin(xy)
x+y .

We are to show that

lim

(x,y)( , )

f(x, y) does not exist

by nding the limit along the path y = sin x. First, however, consider
the limits found along the lines y = mx as done above.
(
)
sin x(mx)
sin(mx )
= lim
lim
x x(m + )
x + mx
(x,mx)( , )
sin(mx )
= lim

.
x
x
m+
By applying LHpitals Rule, we can show this limit is except when m =
, that is, along the line y = x. This line is not in the domain of f, so
we have found the following fact: along every line y = mx in the domain
of f,
lim f(x, y) = .
(x,y)( , )

Now consider the limit along the path y = sin x:


(
)
(
)
sin x sin x
sin x sin x
lim
= lim
x
x sin x
x sin x
(x, sin x)( , )

Now apply LHpitals Rule twice:

(
)
cos x sin x ( sin x x cos x)
= lim
( = / )
x
cos x
(
)
(
)
sin x sin x ( sin x x cos x) + cos x sin x ( cos x + x sin x)
= lim
x
sin x
= / the limit does not exist.

Notes:

Limits and Con nuity of Mul variable Func ons

Step back and consider what we have just discovered. Along any line y =
mx in the domain of the f(x, y), the limit is . However, along the path
y = sin x, which lies in the domain of f(x, y) for all x = , the limit does
not exist. Since the limit is not the same along every path to ( , ), we say
sin(xy)
does not exist.
lim
(x,y)( , ) x + y
Finding a limit
x y
Let f(x, y) =
. Find
lim f(x, y).
x +y
(x,y)( , )
Example

It is rela vely easy to show that along any line y = mx, the
S
limit is . This is not enough to prove that the limit exists, as demonstrated in
the previous example, but it tells us that if the limit does exist then it must be .
To prove the limit is
, we apply Deni on . Let > be given. We want
to nd > such that if (x ) + (y ) < , then |f(x, y) | < .

y
< for all (x, y) = ( , ), and that if
Set < / . Note that
x +y

x + y < , then x < .

Let (x ) + (y ) = x + y < . Consider |f(x, y) |:



x y

|f(x, y) | =
x +y


y
= x
x +y
<

<

= .

(x ) + (y ) < then |f(x, y) | < , which is what we


x y
= .
wanted to show. Thus
lim
(x,y)( , ) x + y
Thus if

Con nuity
Deni on denes what it means for a func on of one variable to be connuous. In brief, it meant that the graph of the func on did not have breaks,
holes, jumps, etc. We dene con nuity for func ons of two variables in a similar
way as we did for func ons of one variable.

Notes:

Chapter

Func ons of Several Variables

Deni on

Con nuous

Let a func on f(x, y) be dened on an open disk B containing the point


(x , y ).
. f is con nuous at (x , y ) if

lim

(x,y)(x ,y )

f(x, y) = f(x , y ).

. f is con nuous on B if f is con nuous at all points in B. If f is con nuous at all points in R , we say that f is con nuous everywhere.
Example
Let f(x, y) =
everywhere?
S

lim

(x,y)( , )

Con nuity of a func on of two variables


x =
. Is f con nuous at ( , )? Is f con nuous
cos y x =

cos y sin x
x

To determine if f is con nuous at ( , ), we need to compare


f(x, y) to f( , ).

Applying the deni on of f, we see that f( , ) = cos = .


We now consider the limit
lim f(x, y). Subs tu ng for x and y in
(x,y)( , )

(cos y sin x)/x returns the indeterminate form / , so we need to do more


work to evaluate this limit.
sin x
. The rst
Consider two related limits:
lim cos y and
lim
(x,y)( , )
(x,y)( , ) x
limit does not contain x, and since cos y is con nuous,
lim

(x,y)( , )

cos y = lim cos y = cos = .


y

The second limit does not contain y. By Theorem we can say


lim

(x,y)( ,

Finally, Theorem
as follows:
lim

(x,y)( ,

sin x
sin x
= lim
= .
x
x
) x

of this sec on states that we can combine these two limits


)
(
sin x
cos y sin x
=
lim (cos y)
x
x
(x,y)( , )
)
(
)(
=
lim cos y
lim
(x,y)( , )

= ( )( )
= .

Notes:

(x,y)( ,

sin x
) x

.
We have found that

lim

(x,y)( ,

Limits and Con nuity of Mul variable Func ons

cos y sin x
= f( , ), so f is con nuous at
x
)

( , ).
A similar analysis shows that f is con nuous at all points in R . As long as
x = , we can evaluate the limit directly; when x = , a similar analysis shows
that the limit is cos y. Thus we can say that f is con nuous everywhere. A graph
of f is given in Figure . . No ce how it has no breaks, jumps, etc.
The following theorem is very similar to Theorem , giving us ways to combine con nuous func ons to create other con nuous func ons.
Theorem

Proper es of Con nuous Func ons

Let f and g be con nuous on an open disk B, let c be a real number, and
let n be a posi ve integer. The following func ons are con nuous on B.
. Sums/Dierences: f g
. Constant Mul ples: c f
. Products:

fg

. Quo ents:

f/g

. Powers:

fn

n
f

. Roots:
. Composi ons:

(as longs as g =

on B)

(if n is even then f on B; if n is odd,


then true for all values of f on B.)

Adjust the deni ons of f and g to: Let f be


con nuous on B, where the range of f on B is
J, and let g be a single variable func on that is
con nuous on J. Then g f, i.e., g(f(x, y)), is
con nuous on B.

Establishing con nuity of a func on


Example
Let f(x, y) = sin(x cos y). Show f is con nuous everywhere.
We will apply both Theorems and
. Let f (x, y) = x .
S
Since y is not actually used in the func on, and polynomials are con nuous (by
Theorem ), we conclude f is con nuous everywhere. A similar statement can
be made about f (x, y) = cos y. Part of Theorem
states that f = f f
is con nuous everywhere, and Part of the theorem states the composi on of
sine with f is con nuous: that is, sin(f ) = sin(x cos y) is con nuous everywhere.

Notes:

Figure
.

: A graph of f(x, y) in Example

Chapter

Func ons of Several Variables

Func ons of Three Variables


The deni ons and theorems given in this sec on can be extended in a natural way to deni ons and theorems about func ons of three (or more) variables.
We cover the key concepts here; some terms from Deni ons and are not
redened but their analogous meanings should be clear to the reader.
Deni on

Open Balls, Limit, Con nuous

. An open ball in R centered


at (x , y , z ) with radius r is the set of all
points (x, y, z) such that (x x ) + (y y ) + (z z ) = r.
. Let D be an open set in R containing (x , y , z ), and let f(x, y, z) be a
func on of three variables dened on D, except possibly at (x , y , z ).
The limit of f(x, y, z) as (x, y, z) approaches (x , y , z ) is L, denoted
lim

(x,y,z)(x ,y ,z )

f(x, y, z) = L,

means that given any > , there is a >


such that for all
(x, y, z) = (x , y , z ), if (x, y, z) is in the open ball centered at
(x , y , z ) with radius , then |f(x, y, z) L| < .
. Let f(x, y, z) be dened on an open ball B containing (x , y , z ). f is
con nuous at (x , y , z ) if
lim
f(x, y, z) = f(x , y , z ).
(x,y,z)(x ,y ,z )

These deni ons can also be extended naturally to apply to func ons of four
or more variables. Theorem
also applies to func on of three or more variables, allowing us to say that the func on

ex +y y + z +
f(x, y, z) =
sin(xyz) +
is con nuous everywhere.
When considering single variable func ons, we studied limits, then con nuity, then the deriva ve. In our current study of mul variable func ons, we have
studied limits and con nuity. In the next sec on we study deriva on, which
takes on a slight twist as we are in a mul varible context.

Notes:

Exercises

Terms and Concepts

. f(x, y) =

. Describe in your own words the dierence between boundary and interior point of a set.

. f(x, y) =

. Use your own words to describe (informally) what


lim f(x, y) =
means.

. f(x, y) =

(x,y)( , )

. Give an example of a closed, bounded set.


. Give an example of a closed, unbounded set.
. Give an example of a open, bounded set.

. S=

. S = {(x, y)|y > sin x}

In Exercises

. f(x, y) =

x y

xy y
y +x

lim

(x,y)( , )

sin(x )
y

(b) Along the path y = x .


.

lim

(x,y)( , )

x+y
x

(a) Along the path y = .


(b) Along the path y = x + .

(a) Find the domain D of the given func on.


(c) State whether D is bounded or unbounded.

lim

(x,y)( , )

(a) Along the path y = mx.

(b) State whether D is an open or closed set.

x+y
xy

(b) Along the path x = .

}
{
. S = (x, y) | y = x

{
}
. S = (x, y) | x + y =

lim

(x,y)( , )

(a) Along the path y = mx.

(c) State whether S is bounded or unbounded.



}
(x )
(y )
+

(x, y)

x y
x +y

(a) Along the path y = mx.

(b) State whether S is open, closed, or neither.

lim

(x,y)( , )

(b) Along the path x = .

, a set S is given.

(a) Give one boundary point and one interior point, when
possible, of S.

x y
x +y

(a) Along the path y = .

In Exercises

yx

In Exercises
, a limit is given. Evaluate the limit along
the paths given, then state why these results show the given
limit does not exist.

. Give an example of a open, unbounded set.

Problems

yx

lim

(x,y)(,/ )

sin x
cos y

(a) Along the path x = .


(b) Along the path y = x / .

Chapter

Func ons of Several Variables

Par al Deriva ves

Let y be a func on of x. We have studied in great detail the deriva ve of y with


respect to x, that is, dy
dx , which measures the rate at which y changes with respect
to x. Consider now z = f(x, y). It makes sense to want to know how z changes
with respect to x and/or y. This sec on begins our inves ga on into these rates
of change.
Consider the func on z = f(x, y) = x + y , as graphed in Figure . (a).
By xing y = , we focus our a en on to all points on the surface where the
y-value is , shown in both parts (a) and (b) of the gure. These points form a
curve in space: z = f(x, ) = x + which is a func on of just one variable. We
can take the deriva ve of z with respect to x along this curve and nd equa ons
of tangent lines, etc.

(a)

The key no on to extract from this example is: by trea ng y as constant (it
does not vary) we can consider how z changes with respect to x. In a similar
fashion, we can hold x constant and consider how z changes with respect to
y. This is the underlying principle of par al deriva ves. We state the formal,
limitbased deni on rst, then show how to compute these par al deriva ves
without directly taking limits.

Deni on
(b)

Par al Deriva ve

Let z = f(x, y) be a con nuous func on on an open set S in R .

Figure . : By xing y = , the surface


f(x, y) = x + y is a curve in space.

. The par al deriva ve of f with respect to x is:


fx (x, y) = lim
h

f(x + h, y) f(x, y)
.
h

. The par al deriva ve of f with respect to y is:

Alternate nota ons for fx (x, y) include:

fy (x, y) = lim

f z
f(x, y),
,
, and zx ,
x
x x

f(x, y + h) f(x, y)
.
h

with similar nota ons for fy (x, y). For


ease of nota on, fx (x, y) is o en abbreviated fx .

Compu ng par al deriva ves with the limit deni on


Example
Let f(x, y) = x y + x + y . Find fx (x, y) using the limit deni on.

Notes:

.
Using Deni on

fx (x, y) = lim
h

= lim
h

= lim
h

= lim
h

= lim
h

, we have:

f(x + h, y) f(x, y)
h
(x + h) y + (x + h) + y (x y + x + y )
h
(x y + xhy + h y + x + h + y (x y + x + y )
h
xhy + h y + h
h
xy + hy +

= xy + .
We have found fx (x, y) = xy + .
Example
found a par al deriva ve using the formal, limitbased deni on. Using limits is not necessary, though, as we can rely on our previous
knowledge of deriva ves to compute par al deriva ves easily. When computing fx (x, y), we hold y xed it does not vary. Therefore we can compute the
deriva ve with
to x by trea ng y as(a constant
or coecient.
( respect
)
)

d
x = x, we compute x
x y = xy. Here we are trea ng y
Just as dx
as a coecient.
( )
( )
d

Just as dx
= , we compute x
y = . Here we are trea ng y as a
constant. More examples will help make this clear.
Finding par al deriva ves
Example
Find fx (x, y) and fy (x, y) in each of the following.
. f(x, y) = x y + y x +
. f(x, y) = cos(xy ) + sin x

. f(x, y) = ex y x +
S

. We have f(x, y) = x y + y x + .
Begin with fx (x, y). Keep y xed, trea ng it as a constant or coecient, as
appropriate:
fx (x, y) = x y .
Note how the y and terms go to zero.

Notes:

Par al Deriva ves

Chapter

Func ons of Several Variables


To compute fy (x, y), we hold x xed:
fy (x, y) = x y +

y.

Note how the x and terms go to zero.

. We have f(x, y) = cos(xy ) + sin x.


Begin with fx (x, y). We need to apply the Chain Rule with the cosine term;
y is the coecient of the x-term inside the cosine func on.
fx (x, y) = sin(xy )(y ) + cos x = y sin(xy ) + cos x.
To nd fy (x, y), note that x is the coecient of the y term inside of the
cosine term; also note that since x is xed, sin x is also xed, and we treat
it as a constant.
fy (x, y) = sin(xy )( xy) = xy sin(xy ).

. We have f(x, y) = ex y x + .
Beginning with fx (x, y), note how we need to apply the Product Rule.
fx (x, y) = ex y ( xy )
= xy ex

x + + ex

ex y
.
+
x +

x +

Note that
when nding fy (x, y) we do not have to apply the Product Rule;
since x + does not contain y, we treat it as xed and hence becomes
a coecient of the ex y term.

fy (x, y) = ex y ( x y ) x + = x y ex y x + .
We have shown how to compute a par al deriva ve, but it may s ll not be
clear what a par al deriva ve means. Given z = f(x, y), fx (x, y) measures the
rate at which z changes as only x varies: y is held constant.
Imagine standing in a rolling meadow, then beginning to walk due east. Depending on your loca on, you might walk up, sharply down, or perhaps not
change eleva on at all. This is similar to measuring zx : you are moving only east
(in the x-direc on) and not north/south at all. Going back to your original loca on, imagine now walking due north (in the y-direc on). Perhaps walking
due north does not change your eleva on at all. This is analogous to zy = : z
does not change with respect to y. We can see that zx and zy do not have to be
the same, or even similar, as it is easy to imagine circumstances where walking
east means you walk downhill, though walking north makes you walk uphill.

Notes:

Par al Deriva ves

The following example helps us visualize this more.


Evalua ng par al deriva ves
Example
Let z = f(x, y) = x y + xy + . Find fx ( , ) and fy ( , ) and interpret
their meaning.
S

y + x. Thus

We begin by compu ng fx (x, y) = x + y and fy (x, y) =

fx ( , ) =

and

fy ( , ) = .
(a)

It is also useful to note that f( , ) = . . What does each of these numbers


mean?
Consider fx ( , ) = , along with Figure . (a). If one stands on the
surface at the point ( , , . ) and moves parallel to the x-axis (i.e., only the xvalue changes, not the y-value), then the instantaneous rate of change is .
Increasing the x-value will decrease the z-value; decreasing the x-value will increase the z-value.
Now consider fy ( , ) = , illustrated in Figure . (b). Moving along the
curve drawn on the surface, i.e., parallel to the y-axis and not changing the xvalues, increases the z-value instantaneously at a rate of . Increasing the yvalue by would increase the z-value by approximately .
Since the magnitude of fx is greater than the magnitude of fy at ( , ), it is
steeper in the x-direc on than in the y-direc on.

Second Par al Deriva ves


Let z = f(x, y). We have learned to nd the par al deriva ves fx (x, y) and
fy (x, y), which are each func ons of x and y. Therefore we can take par al derivaves of them, each with respect to x and y. We dene these second par als
along with the nota on, give examples, then discuss their meaning.

Notes:

(b)
Figure . : Illustra ng the meaning of
par al deriva ves.

Chapter

Func ons of Several Variables

Deni on

Second Par al Deriva ve, Mixed Par al Deriva ve

Let z = f(x, y) be con nuous on an open set S.


. The second par al deriva ve of f with respect to x then x is
( )
( )
f
f
= fx x = fxx
=
x x
x

. The second par al deriva ve of f with respect to x then y is


( )
( )
f
f
=
= fx y = fxy
y x
yx

Note: The terms in Deni on


all depend on limits, so each deni on comes
with the caveat where the limit exists.

Similar deni ons hold for

f
f
= fyy and
= fyx .
y
xy

The second par al deriva ves fxy and fyx are mixed par al deriva ves.
The nota on of second par al deriva ves gives some insight into the notaon of the second deriva ve of a func on of a single variable. If y = f(x), then
d y
. The d y por on means take the deriva ve of y twice, while
f (x) =
dx
dx means with respect to x both mes. When we only know of func ons of
a single variable, this la er phrase seems silly: there is only one variable to take
the deriva ve with respect to. Now that we understand func ons of mul ple
variables, we see the importance of specifying which variables we are referring
to.
Second par al deriva ves
Example
For each of the following, nd all six rst and second par al deriva ves. That is,
nd
fx , fy , fxx , fyy , fxy and fyx .
. f(x, y) = x y + xy + cos x
. f(x, y) =

x
y

. f(x, y) = ex sin(x y)
In each, we give fx and fy immediately and then spend me deriving the second par al deriva ves.

Notes:

.
. f(x, y) = x y + xy + cos x
fx (x, y) = x y + y sin x

fy (x, y) = x y + xy
( )
(
fxx (x, y) =
fx =
x
x
(
( )
fyy (x, y) =
fy =
y
y
(
( )
fx =
fxy (x, y) =
y
y
)
(

(
fyx (x, y) =
fx =
x
x
x
= x y
y
x
fx (x, y) =
y
x
fy (x, y) =
y
( )
fx =
fxx (x, y) =
x
( )
fyy (x, y) =
fy =
y
( )
fx =
fxy (x, y) =
y
( )
fyx (x, y) =
fx =
x

)
x y + y sin x = xy cos x
x y + xy

= x +

= x y+ y

xy

)
x y + y sin x = x y + y

x y + xy

. f(x, y) =

. f(x, y) = ex sin(x y)

( x
x y
(

y
( x
y y
(

x
y

x )
x
=
y
y
)
x
=
y
x )
x
=
y
y

Because the following par al deriva ves get rather long, we omit the extra
nota on and just give the results. In several cases, mul ple applica ons
of the Product and Chain Rules will be necessary, followed by some basic
combina on of like terms.
fx (x, y) = ex sin(x y) + xyex cos(x y)
fy (x, y) = x ex cos(x y)
fxx (x, y) = ex sin(x y) + xyex cos(x y) + yex cos(x y) x y ex sin(x y)
fyy (x, y) = x ex sin(x y)

fxy (x, y) = x ex cos(x y) + xex cos(x y) x yex sin(x y)

fyx (x, y) = x ex cos(x y) + xex cos(x y) x yex sin(x y)


Notes:

Par al Deriva ves

Chapter

Func ons of Several Variables


No ce how in each of the three func ons in Example
, fxy = fyx . Due to
the complexity of the examples, this likely is not a coincidence. The following
theorem states that it is not.
Theorem

Mixed Par al Deriva ves

Let f be dened such that fxy and fyx are con nuous on an open set S.
Then for each point (x, y) in S, fxy (x, y) = fyx (x, y).
Finding fxy and fyx independently and comparing the results provides a convenient way of checking our work.

Understanding Second Par al Deriva ves


us.

Now that we know how to nd second par als, we inves gate what they tell

Again we refer back to a func on y = f(x) of a single variable. The second


deriva ve of f is the deriva ve of the deriva ve, or the rate of change of the
rate of change. The second deriva ve measures how much the deriva ve is
changing. If f (x) < , then the deriva ve is ge ng smaller (so the graph of f is
concave down); if f (x) > , then the deriva ve is growing, making the graph
of f concave up.
Now consider z = f(x, y). Similar statements can be made about fxx and fyy
as could be made about f (x) above. When taking deriva ves with respect to
x twice, we measure how much fx changes with respect to x. If fxx (x, y) < ,
it means that as x increases, fx decreases, and the graph of f will be concave
down in the x-direc on. Using the analogy of standing in the rolling meadow
used earlier in this sec on, fxx measures whether ones path is concave up/down
when walking due east.
Similarly, fyy measures the concavity in the y-direc on. If fyy (x, y) > , then
fy is increasing with respect to y and the graph of f will be concave up in the ydirec on. Appealing to the rolling meadow analogy again, fyy measures whether
ones path is concave up/down when walking due north.
We now consider the mixed par als fxy and fyx . The mixed par al fxy measures how much fx changes with respect to y. Once again using the rolling meadow
analogy, fx measures the slope if one walks due east. Looking east, begin walking north (sidestepping). Is the path towards the east ge ng steeper? If so,
fxy > . Is the path towards the east not changing in steepness? If so, then
fxy = . A similar thing can be said about fyx : consider the steepness of paths
heading north while sidestepping to the east.
The following example examines these ideas with concrete numbers and

Notes:

Par al Deriva ves

graphs.
Understanding second par al deriva ves
Example
Let z = x y + xy. Evaluate the rst and second par al deriva ves at
( / , / ) and interpret what each of these numbers mean.
We nd that:
fx (x, y) = x + y, fy (x, y) = y + x, fxx (x, y) = ,
fxy (x, y) = fyx (x, y) = . Thus at ( / , / ) we have
S

fx ( / , / ) = / ,

fyy (x, y) = and

fy ( / , / ) = / .

The slope of the tangent line at ( / , / , / ) in the direc on of x is / :


if one moves from that point parallel to the x-axis, the instantaneous rate of
change will be / . The slope of the tangent line at this point in the direc on
of y is / : if one moves from this point parallel to the y-axis, the instantaneous
rate of change will be / . These tangents lines are graphed in Figure . (a)
and (b), respec vely, where the tangent lines are drawn in a solid line.
Now consider only Figure . (a). Three directed tangent lines are drawn
(two are dashed), each in the direc on of x; that is, each has a slope determined
by fx . Note how as y increases, the slope of these lines get closer to . Since the
slopes are all nega ve, ge ng closer to means the slopes are increasing. The
slopes given by fx are increasing as y increases, meaning fxy must be posi ve.
Since fxy = fyx , we also expect fy to increase as x increases. Consider Figure
. (b) where again three directed tangent lines are drawn, this me each in
the direc on of y with slopes determined by fy . As x increases, the slopes become less steep (closer to ). Since these are nega ve slopes, this means the
slopes are increasing.
Thus far we have a visual understanding of fx , fy , and fxy = fyx . We now interpret fxx and fyy . In Figure . (a), we see a curve drawn where x is held constant
at x = / : only y varies. This curve is clearly concave down, corresponding
to the fact that fyy < . In part (b) of the gure, we see a similar curve where y
is constant and only x varies. This curve is concave up, corresponding to the fact
that fxx > .

Par al Deriva ves and Func ons of Three Variables


The concepts underlying par al deriva ves can be easily extend to more
than two variables. We give some deni ons and examples in the case of three
variables and trust the reader can extend these deni ons to more variables if
needed.

Notes:

(a)

(b)
Figure . : Understanding the second
par al deriva ves in Example
.

Chapter

Func ons of Several Variables

Deni on

Par al Deriva ves with Three Variables

Let w = f(x, y, z) be a con nuous func on on an open set S in R .


The par al deriva ve of f with respect to x is:
fx (x, y, z) = lim
h

f(x + h, y, z) f(x, y, z)
.
h

Similar deni ons hold for fy (x, y, z) and fz (x, y, z).


By taking par al deriva ves of par al deriva ves, we can nd second par al
deriva ves of f with respect to z then y, for instance, just as before.
Par al deriva ves of func ons of three variables
Example
For each of the following, nd fx , fy , fz , fxz , fyz , and fzz .
. f(x, y, z) = x y z + x y + x z + y z
. f(x, y, z) = x sin(yz)
S

. fx = xy z + xy + x z ; fy = x y z + x y + y z ;
fz = x y z + x z + y z ; fxz = xy z + x z ;
fyz =

x y z +

y z ; fzz =

x y z + x z+

y z

. fx = sin(yz); fy = xz cos(yz); fz = xy cos(yz);


fxz = y cos(yz); fyz = x cos(yz) xyz sin(yz); fzz = xy sin(xy)

Higher Order Par al Deriva ves


We can con nue taking par al deriva ves of par al deriva ves of par al
deriva ves of ; we do not have to stop with second par al deriva ves. These
higher order par al deriva ves do not have a dy graphical interpreta on; nevertheless they are not hard to compute and worthy of some prac ce.
We do not formally dene each higher order deriva ve, but rather give just
a few examples of the nota on.
( ( ))
f

and
fxyx (x, y) =
x y x
( ( ))

f
fxyz (x, y, z) =
.
z y x

Notes:

.
Example

Higher order par al deriva ves

. Let f(x, y) = x y + sin(xy). Find fxxy and fyxx .


. Let f(x, y, z) = x exy + cos(z). Find fxyz .
S

. To nd fxxy , we rst nd fx , then fxx , then fxxy :


fx = xy + y cos(xy)

fxx = y y sin(xy)

fxxy = y y sin(xy) xy cos(xy).


To nd fyxx , we rst nd fy , then fyx , then fyxx :
fy = x y + x cos(xy)
fyxx

fyx = xy + cos(xy) xy sin(xy)


)
= y y sin(xy) y sin(xy) + xy cos(xy)
(

= y y sin(xy) xy cos(xy).
Note how fxxy = fyxx .
. To nd fxyz , we nd fx , then fxy , then fxyz :
fx = x exy + x yexy
fxyz = .

fxy = x exy + x exy + x yexy = x exy + x yexy

In the previous example we saw that fxxy = fyxx ; this is not a coincidence.
While we do not state this as a formal theorem, as long as each par al deriva ve
is con nuous, it does not ma er the order in which the par al deriva ves are
taken. For instance, fxxy = fxyx = fyxx .
This can be useful at mes. Had we known this, the second part of Example
would have been much simpler to compute. Instead of compu ng fxyz
in the x, y then z orders, we could have applied the z, then x then y order (as
fxyz = fzxy ). It is easy to see that fz = sin z; then fzx and fzxy are clearly as fz
does not contain an x or y.

Notes:

Par al Deriva ves

Chapter

Func ons of Several Variables


A brief review of this sec on: par al deriva ves measure the instantaneous
rate of change of a mul variable func on with respect to one variable. With
z = f(x, y), the par al deriva ves fx and fy measure the instantaneous rate of
change of z when moving parallel to the x- and y-axes, respec vely. How do we
measure the rate of change at a point when we do not move parallel to one of
these axes? What if we move in the direc on given by the vector , ? Can
we measure that rate of change? The answer is, of course, yes, we can. This is
the topic of Sec on . . First, we need to dene what it means for a func on
of two variables to be dieren able.

Notes:

Exercises

Terms and Concepts


. What is the dierence between a constant and a coecient?
. Given a func on z = f(x, y), explain in your own words how
to compute fx .
. In the mixed par al frac on fxy , which is computed rst, fx
or fy ?
. In the mixed par al frac on
fx or fy ?

f
, which is computed rst,
xy

. f(x, y) = sin( x + y )
. f(x, y) =

xy +

. f(x, y) = ( x + y) y
. f(x, y) =

x +y +

. f(x, y) = x y x + y + at ( , )
. f(x, y) = x x + y y at ( , )
. f(x, y) = sin y cos x at (/ , / )
. f(x, y) = ln(xy) at ( , )
, nd fx , fy , fxx , fyy , fxy and fyx .

In Exercises

. f(x, y) = x y + x + y
. f(x, y) = y + xy + x y + x
x
y

. f(x, y) = ln(x + y)
. f(x, y) =

ln x
y

. f(x, y) = ex sin y +
In Exercises
, form a func on z = f(x, y) such that fx
and fy match those given.
. fx = sin y + , fy = x cos y
. fx = x + y, fy = x + y
. fx = xy y , fy = x xy +
. fx =

x
y
, fy =
x +y
x +y

In Exercises

+y

. f(x, y) = ex+

. f(x, y) = sin x cos y

, nd fx , fy , fz , fyz and fzy .

. f(x, y, z) = x e

xy

. f(x, y) = ex

. f(x, y) = x +

In Exercises , evaluate fx (x, y) and fy (x, y) at the indicated


point.

. f(x, y) =

. f(x, y) = cos( xy )

. f(x, y) = x

Problems

. f(x, y) =

. f(x, y) = (x + y)

y z

. f(x, y, z) = x y + x z + y z
. f(x, y, z) =

x
yz

. f(x, y, z) = ln(xyz)

Chapter

Func ons of Several Variables

Dieren ability and the Total Dieren al

We studied dieren als in Sec on . , where Deni on states that if y = f(x)


and f is dieren able, then dy = f (x)dx. One important use of this dieren al is
in Integra on by Subs tu on. Another important applica on is approxima on.
Let x = dx represent a change in x. When dx is small, dy y, the change
in y resul ng from the change in x. Fundamental in this understanding is this:
as dx gets small, the dierence between y and dy goes to . Another way of
sta ng this: as dx goes to , the error in approxima ng y with dy goes to .
We extend this idea to func ons of two variables. Let z = f(x, y), and let
x = dx and y = dy represent changes in x and y, respec vely. Let z =
f(x + dx, y + dy) f(x, y) be the change in z over the change in x and y. Recalling
that fx and fy give the instantaneous rates of z-change in the x- and y-direc ons,
respec vely, we can approximate z with dz = fx dx + fy dy; in words, the total
change in z is approximately the change caused by changing x plus the change
caused by changing y. In a moment we give an indica on of whether or not this
approxima on is any good. First we give a name to dz.
Deni on

Total Dieren al

Let z = f(x, y) be con nuous on an open set S. Let dx and dy represent


changes in x and y, respec vely. Where the par al deriva ves fx and fy
exist, the total dieren al of z is
dz = fx (x, y)dx + fy (x, y)dy.
Finding the total dieren al
Example
Let z = x e y . Find dz.
We compute the par al deriva ves: fx = x e
x e . Following Deni on , we have
S

and fy =

dz = x e y dx + x e y dy.
We can approximate z with dz, but as with all approxima ons, there is
error involved. A good approxima on is one in which the error is small. At a
given point (x , y ), let Ex and Ey be func ons of dx and dy such that Ex dx + Ey dy
describes this error. Then
z = dz + Ex dx + Ey dy
= fx (x , y )dx + fy (x , y )dy + Ex dx + Ey dy.

Notes:

.
If the approxima on of z by dz is good, then as dx and dy get small, so does
Ex dx + Ey dy. The approxima on of z by dz is even be er if, as dx and dy go to
, so do Ex and Ey . This leads us to our deni on of dieren ability.
Deni on

Mul variable Dieren ability

Let z = f(x, y) be dened on an open set S containing (x , y ) where


fx (x , y ) and fy (x , y ) exist. Let dz be the total dieren al of z at (x , y ),
let z = f(x + dx, y + dy) f(x , y ), and let Ex and Ey be func ons of
dx and dy such that
z = dz + Ex dx + Ey dy.
. f is dieren able at (x , y ) if, given > , there is a > such
that if || dx, dy || < , then || Ex , Ey || < . That is, as dx and dy
go to , so do Ex and Ey .
. f is dieren able on S if f is dieren able at every point in S. If f is
dieren able on R , we say that f is dieren able everywhere.
Showing a func on is dieren able
Example
Show f(x, y) = xy + y is dieren able using Deni on

We begin by nding f(x + dx, y + dy), z, fx and fy .

f(x + dx, y + dy) = (x + dx)(y + dy) + (y + dy)


= xy + xdy + ydx + dxdy + y + ydy + dy .
z = f(x + dx, y + dy) f(x, y), so
z = xdy + ydx + dxdy + ydy + dy .
It is straigh orward to compute fx = y and fy = x + y. Consider once more z:
z = xdy + ydx + dxdy + ydy + dy

(now reorder)

= ydx + xdy + ydy + dxdy + dy


= (y) dx + (x + y) dy + (dy) dx + ( dy) dy
|{z}
| {z }
| {z }
|{z}
fx

fy

Ex

Ey

= fx dx + fy dy + Ex dx + Ey dy.

With Ex = dy and Ey = dy, it is clear that as dx and dy go to , Ex and Ey also go


to . Since this did not depend on a specic point (x , y ), we can say that f(x, y)

Notes:

Dieren ability and the Total Dieren al

Chapter

Func ons of Several Variables


is dieren able for all pairs (x, y) in R , or, equivalently, that f is dieren able
everywhere.
Our intui ve understanding of dieren ability of func ons y = f(x) of one
variable was that the graph of f was smooth. A similar intui ve understanding of func ons z = f(x, y) of two variables is that the surface dened by f is
also smooth, not containing cusps, edges, breaks, etc. The following theorem
states that dieren able func ons are con nuous, followed by another theorem that provides a more tangible way of determining whether a great number
of func ons are dieren able or not.
Theorem

Con nuity and Dieren ability of Mul variable


Func ons
Let z = f(x, y) be dened on an open set S containing (x , y ). If f is
dieren able at (x , y ), then f is con nuous at (x , y ).

Theorem

Dieren ability of Mul variable Func ons

Let z = f(x, y) be dened on an open set S containing (x , y ). If fx and


fy are both con nuous on S, then f is dieren able on S.

The theorems assure us that essen ally all func ons that we see in the course
of our studies here are dieren able (and hence con nuous) on their natural domains. There is a dierence between Deni on and Theorem
, though: it
is possible for a func on f to be dieren able yet fx and/or fy is not con nuous.
Such strange behavior of func ons is a source of delight for many mathema cians.
When fx and fy exist at a point but are not con nuous at that point, we need
to use other methods to determine whether or not f is dieren able at that
point.
For instance, consider the func on

f(x, y) =

Notes:

xy
x +y

(x, y) = ( , )
(x, y) = ( , )

.
We can nd fx ( , ) and fy ( , ) using Deni on
fx ( , ) = lim
h

f( + h, ) f( , )
h

= lim

= ;
h
f( , + h) f( , )
fy ( , ) = lim
h
h
h

= lim
h

= .

Both fx and fy exist at ( , ), but they are not con nuous at ( , ), as


fx (x, y) =

y(y x )
(x + y )

and

fy (x, y) =

x(x y )
(x + y )

are not con nuous at ( , ). (Take the limit of fx as (x, y) ( , ) along the
x- and y-axes; they give dierent results.) So even though fx and fy exist at every point in the x-y plane, they are not con nuous. Therefore it is possible, by
Theorem
, for f to not be dieren able.
Indeed, it is not. One can show that f is not con nuous at ( , ) (see Example
), and by Theorem
, this means f is not dieren able at ( , ).

Approxima ng with the Total Dieren al


By the deni on, when f is dieren able dz is a good approxima on for z
when dx and dy are small. We give some simple examples of how this is used
here.
Approxima ng with the total dieren al
Example
Let z = x sin y. Approximate f( . , . ).
Recognizing that / .
. , we can
S
approximate
sin(/ ) =
f((. , ). ) using f( , / ). We can easily compute f( , / ) =

.
. Without calculus, this is the best approxima on we
=
could reasonably come up with. The total dieren al gives us a way of adjus ng
this ini al approxima on to hopefully get a more accurate answer.
We let z = f( . , . )f( , / ). The total dieren al dz is approximately
equal to z, so
f( . , . ) f( , / ) dz
To nd dz, we need fx and fy .

Notes:

f( . , . ) dz + f( , / ).

. )

Dieren ability and the Total Dieren al

Chapter

Func ons of Several Variables

sin y
fx (x, y) =
x

fx ( , / ) =

=
fy (x, y) =

x cos y

Approxima ng . with
dy .
. Thus

gives dx =

sin /

fy ( , / ) =

Returning to Equa on (

/ .

. ; approxima ng . with / gives

dz( , / ) = fx ( , / )( . ) + fy ( , / )( .

( . )+
=
( .
)
.

. ), we have

f( . , . ) .

+ .

= .

We, of course, can compute the actual value of f( . , . ) with a calculator; the
actual value, accurate to places a er the decimal, is .
. Obviously our
approxima on is quite good.
The point of the previous example was not to develop an approxima on
method for known func ons. A er all, we can very easily compute f( . , . )
using readily available technology. Rather, it serves to illustrate how well this
method of approxima on works, and to reinforce the following concept:
New posi on = old posi on + amount of change, so
New posi on old posi on + approximate amount of change.

In the previous example, we could easily compute f( , / ) and could approximate the amount of z-change when compu ng f( . , . ), le ng us approximate the new z-value.
It may be surprising to learn that it is not uncommon to know the values of f,
fx and fy at a par cular point without actually knowing the func on f. The total
dieren al gives a good method of approxima ng f at nearby points.
Approxima ng an unknown func on
Example
Given that f( , ) = , fx ( , ) = . and fy ( , ) = . , approximate
f( . , . ).

Notes:

.
The total dieren al approximates how much f changes from
S
the point ( , ) to the point ( . , . ). With dx = . and dy = . , we
have
dz = fx ( , )dx + fy ( , )dy
= . ( . ) + ( . )( .
= .
.

The change in z is approximately .


.
.

, so we approximate f( . , .

Error/Sensi vity Analysis


The total dieren al gives an approxima on of the change in z given small
changes in x and y. We can use this to approximate error propaga on; that is,
if the input is a li le o from what it should be, how far from correct will the
output be? We demonstrate this in an example.
Sensi vity analysis
Example
A cylindrical steel storage tank is to be built that is
tall and across in diameter. It is known that the steel will expand/contract with temperature changes;
is the overall volume of the tank more sensi ve to changes in the diameter or in
the height of the tank?
A cylindrical solid with height h and radius r has volume V =
S
r h. We can view V as a func on of two variables, r and h. We can compute
par al deriva ves of V:
V
= Vr (r, h) = rh
r

and

V
= Vh (r, h) = r .
h

The total dieren al is dV = ( rh)dr + (r )dh. When h =


and r = , we
have dV = dr + dh. Note that the coecient of dr is
. ; the
coecient of dh is a tenth of that, approximately . . A small change in radius
will be mul plied by
. , whereas a small change in height will be mul plied
by . . Thus the volume of the tank is more sensi ve to changes in radius
than in height.
The previous example showed that the volume of a par cular tank was more
sensi ve to changes in radius than in height. Keep in mind that this analysis only
applies to a tank of those dimensions. A tank with a height of
and radius of
would be more sensi ve to changes in height than in radius.

Notes:

Dieren ability and the Total Dieren al

Chapter

Func ons of Several Variables


One could make a chart of small changes in radius and height and nd exact
changes in volume given specic changes. While this provides exact numbers, it
does not give as much insight as the error analysis using the total dieren al.

Dieren ability of Func ons of Three Variables


The deni on of dieren ability for func ons of three variables is very similar to that of func ons of two variables. We again start with the total dieren al.

Deni on

Total Dieren al

Let w = f(x, y, z) be con nuous on an open set S. Let dx, dy and dz represent changes in x, y and z, respec vely. Where the par al deriva ves
fx , fy and fz exist, the total dieren al of w is
dz = fx (x, y, z)dx + fy (x, y, z)dy + fz (x, y, z)dz.

This dieren al can be a good approxima on of the change in w when w =


f(x, y, z) is dieren able.

Deni on

Mul variable Dieren ability

Let w = f(x, y, z) be dened on an open ball B containing (x , y , z )


where fx (x , y , z ), fy (x , y , z ) and fz (x , y , z ) exist. Let dw be the
total dieren al of w at (x , y , z ), let w = f(x + dx, y + dy, z +
dz) f(x , y , z ), and let Ex , Ey and Ez be func ons of dx, dy and dz such
that
w = dw + Ex dx + Ey dy + Ez dz.
. f is dieren able at (x , y , z ) if, given > , there is a >
such that if || dx, dy, dz || < , then || Ex , Ey , Ez || < .
. f is dieren able on B if f is dieren able at every point in B. If f
is dieren able on R , we say that f is dieren able everywhere.

Just as before, this deni on gives a rigorous statement about what it means
to be dieren able that is not very intui ve. We follow it with a theorem similar
to Theorem
.

Notes:

Theorem

Con nuity and Dieren ability of Func ons of Three


Variables
Let w = f(x, y, z) be dened on an open ball B containing (x , y , z ).
. If f is dieren able at (x , y , z ), then f is con nuous at (x , y , z ).
. If fx , fy and fz are con nuous on B, then f is dieren able on B.
This set of deni on and theorem extends to func ons of any number of
variables. The theorem again gives us a simple way of verifying that most funcons that we enounter are dieren able on their natural domains.
This sec on has given us a formal deni on of what it means for a func ons
to be dieren able, along with a theorem that gives a more accessible understanding. The following sec ons return to no ons prompted by our study of
par al deriva ves that make use of the fact that most func ons we encounter
are dieren able.

Notes:

Dieren ability and the Total Dieren al

Exercises

Terms and Concepts


. T/F: If f(x, y) is dieren able on S, the f is con nuous on S.
. T/F: If fx and fy are con nuous on S, then f is dieren able
on S.
. T/F: If z = f(x, y) is dieren able, then the change in z over
small changes dx and dy in x and y is approximately dz.
. Finish the sentence: The new z-value is approximately the
old z-value plus the approximate
.

Is the projec le more sensi ve to errors in ini al speed or


angle of eleva on?
. The length of a long wall is to be approximated. The angle
, as shown in the diagram (not to scale), is measured to
be , and the distance x is measured to be . Assume
that the triangle formed is a right triangle.
Is the measurement of the length of more sensi ve to errors in the measurement of x or in ?

=?

Problems

In Exercises , nd the total dieren al dz.


. z = x sin y + x
. z = ( x + y)
. z= x y
. z = xex+y
In Exercises , a func on z = f(x, y) is given. Give the
indicated approxima on using the total dieren al.
. f(x, y) =
f( , ) = .

x + y. Approximate f( .

, . ) knowing

. f(x, y) = sin x cos y. Approximate f( . , . ) knowing


f( , ) = .
. f(x, y) = x y xy . Approximate f( .
f( , ) = .

, .

. f(x, y) = ln(x y). Approximate f( . , .


f( , ) = .

) knowing
) knowing

Exercises

ask a variety of ques ons dealing with approxima ng error and sensi vity analysis.
. A cylindrical storage tank is to be tall with a radius of .
Is the volume of the tank more sensi ve to changes in the
radius or the height?
. Projec le Mo on: The x-value of an object moving under the principles of projec le mo on is x(, v , t) =
(v cos )t. A par cular projec le is red with an ini al velocity of v =
/s and an angle of eleva on of = .
It travels a distance of
in seconds.

. It is common sense that it is far be er to measure a long


distance with a long measuring tape rather than a short
one. A measured distance D can be viewed as the product of the length of a measuring tape mes the number
n of mes it was used. For instance, using a tape
mes gives a length of . To measure the same distance
with a tape, we would use the tape . mes. (I.e.,
=
. .) Thus D = n.
Suppose each me a measurement is taken with the tape,
the recorded distance is within / of the actual distance.
(I.e., d = / .
). Using dieren als, show
why common sense proves correct in that it is be er to use
a long tape to measure long distances.
In Exercises

, nd the total dieren al dw.

. w = x yz
. w = ex sin y ln z
In Exercises
, use the informa on provided and the
total dieren al to make the given approxima on.
. f( , ) = , fx ( , ) = , fy ( , ) = . Approximate
f( . , . ).
. f( , ) = , fx ( , ) = . , fy ( , ) = . . Approximate f( . , . ).
. f( , , ) = , fx ( , , ) = , fy ( , , ) = ,
fz ( , , ) = . . Approximate f( . , . , . ).
. f( , , ) = , fx ( , , ) = , fy ( , , ) = , fz ( , , ) =
. Approximate f( . , . , . ).

The Mul variable Chain Rule

The Mul variable Chain Rule

The Chain Rule, as learned in Sec on . , states that


If t = g(x), we can express the Chain Rule as

))
(
)
d((
f g(x) = f g(x) g (x).
dx

df dt
df
=
.
dx
dt dx
In this sec on we extend the Chain Rule to func ons of more than one variable.
Theorem

Mul variable Chain Rule, Part I

Let z = f(x, y), x = g(t) and y =


( h(t), where
) f, g and h are dieren able
func ons. Then z = f(x, y) = f g(t), h(t) is a func on of t, and
dz
df
dx
dy
=
= fx (x, y) + fy (x, y)
dt
dt
dt
dt
=

f dy
f dx
+
.
x dt
y dt

It is good to understand what the situa on of z = f(x, y), x = g(t) and


y = h(t) describes. We know that z = f(x, y) describes a surface; we also
recognize that x = g(t) and y = h(t) are parametric equa ons for a curve in
the x-y plane. Combining these together, we are describing a curve that lies on
the surface described
( by f. The
) parametric equa ons for this curve are x = g(t),
y = h(t) and z = f g(t), h(t) .
Consider Figure . in which a surface is drawn, along with a dashed curve
in the x-y plane. Restric ng f to just the points on this circle gives the curve
df
gives the instantaneous rate of change
shown on the surface. The deriva ve dt
df
gives
of f with respect to t. If we consider an object traveling along this path, dt
the rate at which the object rises/falls.
We now prac ce applying the Mul variable Chain Rule.
Using the Mul variable Chain Rule
dz
using the Chain Rule.
Let z = x y + x, where x = sin t and y = e t . Find
dt
Example

Following Theorem

fx (x, y) = xy + ,

Notes:

fy (x, y) = x ,

, we nd
dx
= cos t,
dt

dy
= e t.
dt

Figure . : Understanding the applicaon of the Mul variable Chain Rule.

Chapter

Func ons of Several Variables


Applying the theorem, we have
dz
= ( xy + ) cos t + x e t .
dt
This may look odd, as it seems that dz
dt is a func on of x, y and t. Since x and y
are func ons of t, dz
is
really
just
a
func
on of t, and we can replace x with sin t
dt
and y with e t :
dz
= ( xy + ) cos t + x e t = ( sin(t)e t + ) cos t + e t sin t.
dt
The previous example can make us wonder: if we subs tuted for x and y at
the end to show that dz
dt is really just a func on of t, why not subs tute before
dieren a ng, showing clearly that z is a func on of t?
That is, z = x y + x = (sin t) e t + sin t. Applying the Chain and Product
Rules, we have
dz
= sin t cos t e t + sin t e t + cos t,
dt
which matches the result from the example.
This may now make one wonder Whats the point? If we could already nd
the deriva ve, why learn another way of nding it? In some cases, applying
this rule makes deriving simpler, but this is hardly the power of the Chain Rule.
Rather, in the case where z = f(x, y), x = g(t) and y = h(t), the Chain Rule is
extremely powerful when we do not know what f, g and/or h are. It may be hard
to believe, but o en in the real world we know rateofchange informa on
(i.e., informa on about deriva ves) without explicitly knowing the underlying
func ons. The Chain Rule allows us to combine several rates of change to nd
another rate of change. The Chain Rule also has theore c use, giving us insight
into the behavior of certain construc ons (as well see in the next sec on).
We demonstrate this in the next example.
Applying the Mul varible Chain Rule
Example
An object travels along a path on a surface. The exact path and surface are not
known, but at me t = t it is known that :
z
= ,
x
Find

dz
dt

Notes:

at me t .

z
= ,
y

dx
=
dt

and

dy
= .
dt

The Mul variable Chain Rule

The Mul variable Chain Rule states that

dz
z dx z dy
=
+
dt
x dt
y dt
= ( ) + ( )( )
= .
By knowing certain ratesofchange informa on about the surface and about
the path of the par cle in the x-y plane, we can determine how quickly the object is rising/falling.
We next apply the Chain Rule to solve a max/min problem.
Applying the Mul variable Chain Rule
Example
Consider the surface z = x + y xy, a paraboloid, on which a par cle moves
with x and y coordinates given by x = cos t and y = sin t. Find dz
dt when t = ,
and nd where the par cle reaches its maximum/minimum z-values.
It is straigh orward to compute

fx (x, y) = x y,

fy (x, y) = y x,

dx
= sin t,
dt

dy
= cos t.
dt

Combining these according to the Chain Rule gives:


dz
= ( x y) sin t + ( y x) cos t.
dt
dz
= ( )( ) + ( )( ) = . When
When t = , x = and y = . Thus
dt
t = , the par cle is moving down, as shown in Figure . .
To nd where z-value is maximized/minimized on the par cles path, we set
dz
=
and solve for t:
dt
dz
=
dt

= ( x y) sin t + ( y x) cos t

= ( cos t sin t) sin t + ( sin t cos t) cos t

= sin t cos t

cos t = sin t

(for odd n)
t=n
We can use the First Deriva ve Test to nd that on [ , ], z has reaches its
absolute minimum at t = / and / ; it reaches its absolute maximum at

Notes:

Figure . : Plo ng the path of a par cle on a surface in Example


.

Chapter

Func ons of Several Variables


t = / and / , as shown in Figure

We can extend the Chain Rule to include the situa on where z is a func on
of more than one variable, and each of these variables is also a func on of more
than one variable. The basic case of this is where z = f(x, y), and x and y are
func ons of two variables, say s and t.

Theorem

Mul variable Chain Rule, Part II

. Let z = f(x, y), x = g(s, t) and y = h(s, t), where f, g and h are
dieren able func ons. Then z is a func on of s and t, and
z
f x
f y
=
+
,
s
x s y s
z
f x
f y

=
+
.
t
x t
y t

and

. Let z = f(x , x , . . . , xm ) be a dieren able func on of m variables,


where each of the xi is a dieren able func on of the variables
t , t , . . . , tn . Then z is a func on of the ti , and
f x
f x
f xm
z
=
+
+ +
.
ti
x ti
x ti
xm ti

Using the Mul varible Chain Rule, Part II


Example
z
Let z = x y + x, x = s + t and y = s t. Find z
s and t , and evaluate each
when s = and t = .
Following Theorem

deriva ves:

f
= xy +
x

x
= s
s
Thus

x
=
t

, we compute the following par al


f
=x ,
y
y
=
s

y
= .
t

z
= ( xy + )( s) + (x )( ) = xys + s + x ,
s
z
= ( xy + )( ) + (x )( ) = xy x + .
t

Notes:

and

.
When s =

and t = , x =
z
=
s

and y = , so
z
=
t

and

Using the Mul varible Chain Rule, Part II


Example
Let w = xy + z , where x = t es , y = t cos s, and z = s sin t. Find
and t = .
S

deriva ves:

Following Theorem
f
=y
x

Thus
When s =

when s =

, we compute the following par al

f
=x
y

x
= tes
t

w
t

f
= z,
z

y
= cos s
t

z
= s cos t.
t

w
= y( tes ) + x(cos s) + z(s cos t).
t
and t = , we have x = , y = and z = . Thus
w
= ( ) + = .
t

Implicit Dieren a on
We studied nding dy
dx when y is given as an implicit func on of x in detail
in Sec on . . We nd here that the Mul variable Chain Rule gives a simpler
method of nding dy
dx .
For instance, consider the implicit func on x y xy = . We learned to use
the following steps to nd dy
dx :
)
d(
d( )
x y xy =
dx
dx
dy
dy
y xy
=
xy + x
dx
dx
dy
xy y
=
.
dx
x xy

. )

Instead of using this method, consider z = x y xy . The implicit func on


above describes the level curve z = . Considering x and y as func ons of x, the
Mul variable Chain Rule states that
z dx z dy
dz
=
+
.
dx
x dx y dx

Notes:

. )

The Mul variable Chain Rule

Chapter

Func ons of Several Variables


dz
dx

Since z is constant (in our example, z = ),


Equa on ( . ) becomes

z
z dy
( )+
x
y dx
z / z
dy
=
dx
x y
=

= . We also know

dx
dx

= .

fx
.
fy

Note how our solu on for dy


dx in Equa on ( . ) is just the par al deriva ve
of z with respect to x, divided by the par al deriva ve of z with respect to y.
We state the above as a theorem.
Theorem

Implicit Dieren a on

Let f be a dieren able func on of x and y, where f(x, y) = c denes y


as an implicit func on of x, for some constant c. Then
dy
fx (x, y)
=
.
dx
fy (x, y)
We prac ce using Theorem

by applying it to a problem from Sec on . .

Implicit Dieren a on
Example
Given the implicitly dened func on sin(x y ) + y = x + y, nd y . Note: this
is the same problem as given in Example
of Sec on . , where the solu on
took about a full page to nd.
Let f(x, y) = sin(x y ) + y x y; the implicitly dened
S
.
func on above is equivalent to f(x, y) = . We nd dy
dx by applying Theorem
We nd
fx (x, y) = xy cos(x y )
so

dy
xy cos(x y )
=
,
dx
x y cos(x y )

which matches our solu on from Example

Notes:

fy (x, y) = x y cos(x y ) ,

and

Exercises

Terms and Concepts

. z = x + y,

x = cos t + ,

x
,
y +

. Let a level curve of z = f(x, y) be described by x = g(t),


= .
y = h(t). Explain why dz
dt

. z=

. Fill in the blank: The single variable Chain Rule states


))
(
)
d((
f g(x) = f g(x)
.
dx

. z=x + y ,

x = sin t,

. z = cos x sin y,

x = t,

. Fill in the blank: The Mul variable Chain Rule states


f
dy
df
=

+
.
dt
x
dt

. T/F: The Mul variable Chain Rule is only useful when all the
related func ons are known explicitly.
. The Mul variable Chain Rule allows us to compute implicit
derivaderiva ves easily by just compu ng two
ves.

Problems
In Exercises , func ons z = f(x, y), x = g(t) and
y = h(t) are given.
dz
(a) Use the Mul variable Chain Rule to compute .
dt
dz
(b) Evaluate
at the indicated t-value.
dt

. z=x y ,

x = t,

. z = x + y,
t = /
. z=

x
,
y +

. z=x + y ,
. z = cos x sin y,

x =

x = cos t,
x = sin t,
x = t,

y = t;

. z = x y,

cos t + ,

t=

t = /

y = sin t;
y = t + / ;

. z=x +y ,

x=t ,

. z=x y ,

x = t,

y= t
y=t

y = s + t;

+y )

x = t, y = st ;
, nd

s= ,t=

y = s t; s = , t =

x = s cos t, y = s sin t;

In Exercises
Theorem
.

s = , t = /

s= ,t=

dy
using Implicit Dieren a on and
dx

. x tan y =
. ( x + y) =
x +y
=
x+y

In Exercises
informa on.

t=

, nd

z
z
dz
, or
and , using the supplied
dt
s
t
dx
= ,
dt

dy
=
dt

z
= ,
x

z
= ,
y

z
= ,
x

z
= ,
y

dx
= ,
dt

dy
=
dt

z
z
= ,
= ,
x
y
x
y
x
= ,
= ,
= ,
s
t
s

y
=
t

z
z
= ,
= ,
x
y
x
x
= ,
= ,
s
t

y
=
t

t = /

In Exercises
, func ons z = f(x, y), x = g(t) and
y = h(t) are given. Find the values of t where dz
= . Note:
dt
these are the same surfaces/curves as found in Exercises
.
. z = x + y,

x = s t,

. ln(x + xy + y ) =

y = sin t ;

y = sin t;

y = t + /

(
)
. z = cos x + y , x = st ,

t=

y=t ;

y = sin t

z
and
(a) Use the Mul variable Chain Rule to compute
s
z
.
t
z
z
and
at the indicated s and t values.
(b) Evaluate
s
t

. z = e(x

x=t ,

y = sin t

In Exercises
, func ons z = f(x, y), x = g(s, t) and
y = h(s, t) are given.

. If z = f(x, y), where x = g(t) and y = h(t), we can subs tute and write z as an explicit func on of t.
T/F: Using the Mul variable Chain Rule to nd dz
is somedt
mes easier than rst subs tu ng and then taking the
deriva ve.

. z = x + y,

x = cos t,

y = sin t

y
= ,
s

Chapter

Func ons of Several Variables

Direc onal Deriva ves

Par al deriva ves give us an understanding of how a surface changes when we


move in the x and y direc ons. We made the comparison to standing in a rolling
meadow and heading due east: the amount of rise/fall in doing so is comparable
to fx . Likewise, the rise/fall in moving due north is comparable to fy . The steeper
the slope, the greater in magnitude fy .
But what if we didnt move due north or east? What if we needed to move
northeast and wanted to measure the amount of rise/fall? Par al deriva ves
alone cannot measure this. This sec on inves gates direc onal deriva ves,
which do measure this rate of change.
We begin with a deni on.
Deni on

Direc onal Deriva ves

Let z = f(x, y) be con nuous on an open set S and let u = u , u be a


unit vector. For all points (x, y), the direc onal deriva ve of f at (x, y) in
the direc on of u is
Du f(x, y) = lim
h

f(x + hu , y + hu ) f(x, y)
.
h

The par al deriva ves fx and fy are dened with similar limits, but only x or
y varies with h, not both. Here both x and y vary with a weighted h, determined
by a par cular unit vector u. This may look a bit in mida ng but in reality it is
not too dicult to deal with; it o en just requires extra algebra. However, the
following theorem reduces this algebraic load.
Theorem

Direc onal Deriva ves

Let z = f(x, y) be dieren able on an open set S containing (x , y ), and


let u = u , u be a unit vector. The direc onal deriva ve of f at (x , y )
in the direc on of u is
Du f(x , y ) = fx (x , y )u + fy (x , y )u .
Compu ng direc onal deriva ves
Example
Let z = x y and let P = ( , ). Find the direc onal deriva ve of f, at P,
in the following direc ons:
. toward the point Q = ( , ),
. in the direc on of , , and
Notes:

Direc onal Deriva ves

. toward the origin.


The surface is plo ed in Figure . , where the point P =
( , ) is indicated in the x, y-plane as well as the point ( , , ) which lies on the
surface of f. We nd that fx (x, y) = x and fx ( , ) = ; fy (x, y) = y and
fy ( , ) = .

. Let u be the unit vector that points from the point ( , ) to the point
#
Q = ( , ), as shown in the
gure.
Thevector PQ = , ; the unit vector

. Thus the direc onal deriva ve of
in this direc on is u = / , /
f at ( , ) in the direc on of u is

. .
Du f( , ) = ( / ) + ( )( / ) = /
Thus the instantaneous rate of change in moving from the point ( , , )
on the surface in the direc on of u (which points toward the point Q) is
about . . Moving in this direc on moves one steeply downward.
. We seek the direc onal deriva ve

vector in this direc on is u =


deriva ve of f at ( , ) in the direc

Du f( , ) = ( /

inthe direc
on
of , . The unit
/ , /
. Thus the direc onal
on of u is

) + ( )( / ) = .

Star ng on the surface of f at ( , ) and moving in the direc on of ,


(or u ) results in no instantaneous change in z-value. This is analogous to
standing on the side of a hill and choosing a direc on to walk that does not
change the eleva on. One neither walks up nor down, rather just along
the side of the hill.
Finding these direc ons of no eleva on change is important.
. At P = ( , ), the direc on towards the origin is given
by the

, ; the unit vector in this direc on is u = / ,


The direc onal deriva ve of f at P in the direc on of the origin is

Du f( , ) = ( / ) + ( )( / ) = /
.

vector

/
.
.

Moving towards the origin means walking uphill quite steeply, with an
ini al slope of about . .
As we study direc onal deriva ves, it will help to make an important connec on between the unit vector u = u , u that describes the direc on and
the par al deriva ves fx and fy . We start with a deni on and follow this with a
Key Idea.

Notes:

Figure . : Understanding the direconal deriva ve in Example


.

Chapter

Func ons of Several Variables

Deni on

Gradient

Let z = f(x, y) be dieren able on an open set S that contains the point
(x , y ).
. The gradient of f is f(x, y) = fx (x, y), fy (x, y).
. The gradient of f at (x , y ) is f(x , y ) = fx (x , y ), fy (x , y ).
To simplify nota on, we o en express the gradient as f = fx , fy . The
gradient allows us to compute direc onal deriva ves in terms of a dot product.
Note: The symbol is named nabla,
derived from the Greek name of a Jewish
harp. Oddly enough, in mathema cs the
expression f is pronounced del f.

Key Idea

The Gradient and Direc onal Deriva ves

The direc onal deriva ve of z = f(x, y) in the direc on of u is


Du f = f u.
The proper es of the dot product previously studied allow us to inves gate
the proper es of the direc onal deriva ve. Given that the direc onal deriva ve
gives the instantaneous rate of change of z when moving in the direc on of u,
three ques ons naturally arise:
. In what direc on(s) is the change in z the greatest (i.e., the steepest uphill)?
. In what direc on(s) is the change in z the least (i.e., the steepest downhill)?
. In what direc on(s) is there no change in z?
Using the key property of the dot product, we have
f u = || f || || u || cos = || f || cos ,

. )

where is the angle between the gradient andu. (Sinceu is a unit vector, ||u || =
.) This equa on allows us to answer the three ques ons stated previously.
. Equa on . is maximized when cos = , i.e., when the gradient and u
have the same direc on. We conclude the gradient points in the direc on
of greatest z change.

Notes:

.
. Equa on . is minimized when cos = , i.e., when the gradient and
u have opposite direc ons. We conclude the gradient points in the opposite direc on of the least z change.
. Equa on . is when cos = , i.e., when the gradient and u are orthogonal to each other. We conclude the gradient is orthogonal to direcons of no z change.
This result is rather amazing. Once again imagine standing in a rolling meadow
and face the direc on that leads you steepest uphill. Then the direc on that
leads steepest downhill is directly behind you, and sidestepping either le or
right (i.e., moving perpendicularly to the direc on you face) does not change
your eleva on at all.
Recall that a level curve is dened as a curve in the x-y plane along which the
z-values of a func on do not change. Let a surface z = f(x, y) be given, and lets
represent one such level curve as a vectorvalued func
( on, r(t)
) = x(t), y(t).
As the output of f does not change along this curve, f x(t), y(t) = c for all t, for
some constant c.
df
= . By the Mul variable Chain Rule, we also
Since f is constant for all t, dt
know
df
= fx (x, y)x (t) + fy (x, y)y (t)
dt
= fx (x, y), fy (x, y) x (t), y (t)
= f r (t)
= .

This last equality states f r (t) = : the gradient is orthogonal to the


deriva ve ofr, meaning the gradient is orthogonal tor itself. Our conclusion: at
any point on a surface, the gradient at that point is orthogonal to the level curve
that passes through that point.
We restate these ideas in a theorem, then use them in an example.
Theorem

The Gradient and Direc onal Deriva ves

Let z = f(x, y) be dieren able on an open set S with gradient f, let


P = (x , y ) be a point in S and let u be a unit vector.
. The maximum value of Du f(x , y ) is || f(x , y ) ||; the direc on
of maximal z increase is f(x , y ).
. The minimum value of Du f(x , y ) is || f(x , y ) ||; the direc on
of minimal z increase is f(x , y ).
. (At P, f(x , y ))is orthogonal to the level curve passing through
x , y , f(x , y ) .
Notes:

Direc onal Deriva ves

Chapter

Func ons of Several Variables


Example
Finding direc ons of maximal and minimal increase
Let f(x, y) = sin x cos y and let P = (/ , / ). Find the direc ons of maximal/minimal increase, and nd a direc on where the instantaneous rate of z
change is .
S

sin x sin y, thus

We begin by nding the gradient. fx = cos x cos y and fy =

f = cos x cos y, sin x sin y


(a)

(b)
Figure . : Graphing the surface and
important direc ons in Example
.

and, at P, f

( )
,
,
=
.

Thus the direc on of maximal increase is / , / . In this direc on, the


instantaneous rate of z change is || / , / || =
/ . .
Figure . shows the surface plo ed from two dierent perspec ves. In
each, the gradient is drawn at P with a dashed line (because of the nature of
this surface, the gradient points into the surface). Let u = u , u be the
unit vector in the direc on of f at P. Each graph of the gure also contains
the vector u , u , ||f ||. This vector has a run of (because in the x-y plane
it moves unit) and a rise of ||f ||, hence we can think of it as a vector with
slope of ||f || in the direc on of f, helping us visualize how steep the surface
is in its steepest direc on.
The direc on of minimal increase
is / , / ; in this direc on the instan
/ . .
taneous rate of z change is
Any direc on orthogonal to f is a direc on of no z change. We have two
choices: the direc on of , and the direc on of , . The unit vector
in the direc onof , is shown in each graph of the gure as well. The level
curve at z =
/ is drawn: recall that along this curve the z-values do not
change. Since , is a direc on of no z-change, this vector is tangent to the
level curve at P.
Understanding when f =
Example
Let f(x, y) = x + x y + y + . Find the direc onal deriva ve of f in any
direc on at P = ( , ).
We nd f = x + , y + . At P, we have f( , ) =
S
, . According to Theorem
, this is the direc on of maximal increase.
However, , is direc onless; it has no displacement. And regardless of the
unit vector u chosen, Du f = .
Figure . helps us understand what this means. We can see that P lies at
the top of a paraboloid. In all direc ons, the instantaneous rate of change is .
So what is the direc on of maximal increase? It is ne to give an answer of
= , , as this indicates that all direc onal deriva ves are .

Figure . : At the top of a paraboloid,


all direc onal deriva ves are .

Notes:

.
The fact that the gradient of a surface always points in the direc on of steepest increase/decrease is very useful, as illustrated in the following example.
The ow of water downhill
Example
Consider the surface given by f(x, y) =
x y . Water is poured on the
surface at ( , / ). What path does it take as it ows downhill?
Let r(t) = x(t), y(t) be the vectorvalued func on deS
scribing the path of the water in the x-y plane; we seek x(t) and y(t). We know
that water will always ow downhill in the steepest direc on; therefore, at any
point on its path, it will be moving in the direc on of f. (We ignore the physical eects of momentum on the water.) Thus r (t) will be parallel to f, and
there is some constant c such that cf = r (t) = x (t), y (t).
dy

We nd f = x, y and write x (t) as dx


dt and y (t) as dt . Then
cf = x (t), y (t)

dx dy
cx, cy =
.
,
dt dt
This implies
cx =
c=

dx
dt

and

dx
x dt

and

cy =
c=

dy
, i.e.,
dt
dy
.
y dt

As c equals both expressions, we have


dx
dy
=
.
x dt
y dt
To nd an explicit rela onship between x and y, we can integrate both sides with
dx
respect to t. Recall from our study of dieren als that dt = dx. Thus:
dt

dx
dy
dt =
dt
x dt
y dt

dx =
dy
x
y
ln |x| =

ln |y| + C

ln |x| = ln |y| + C

ln |x | = ln |y| + C

Notes:

Direc onal Deriva ves

Chapter

Func ons of Several Variables


Now raise both sides as a power of e:
x = eln |y|+C
x = eln |y| eC

(Note that eC is just a constant.)

x = yC
C

x =y

(Note that /C is just a constant.)

Cx = y.
As the water started at the point ( , / ), we can solve for C:
C( ) =
(a)
y

(b)

Figure . : A graph of the surface described in Example


along with the
path in the x-y plane with the level curves.

C=

Thus the water follows the curve y = x / in the x-y plane. The surface and
the path of the water is graphed in Figure . (a). In part (b) of the gure,
the level curves of the surface are plo ed in the x-y plane, along with the curve
y = x / . No ce how the path intersects the level curves at right angles. As the
path follows the gradient downhill, this reinforces the fact that the gradient is
orthogonal to level curves.

Func ons of Three Variables


The concepts of direc onal deriva ves and the gradient are easily extended
to three (and more) variables. We combine the concepts behind Deni ons
and and Theorem
into one set of deni ons.
Deni on

Direc onal Deriva ves and Gradient with Three


Variables
Let w = F(x, y, z) be dieren able on an open ball B and let u be a unit
vector in R .
. The gradient of F is F = Fx , Fy , Fz .
. The direc onal deriva ve of F in the direc on of u is
Du F = F u.

The same proper es of the gradient given in Theorem

Notes:

, when f is a func-

.
on of two variables, hold for F, a func on of three variables.
Theorem

The Gradient and Direc onal Deriva ves with Three


Variables
Let w = F(x, y, z) be dieren able on an open ball B, let F be the gradient of F, and let u be a unit vector.
. The maximum value of Du F is || F ||, obtained when the angle
between F and u is , i.e., the direc on of maximal increase is
F.
. The minimum value of Du F is || F ||, obtained when the angle
between F and u is , i.e., the direc on of minimal increase is
F.
. Du F =

when F and u are orthogonal.

We interpret the third statement of the theorem as the gradient is orthogonal to level surfaces, the threevariable analogue to level curves.
Finding direc onal deriva ves with func ons of three
variables
If a point source S is radia ng energy, the intensity I at a given point P in space
is inversely propor onal to the square of the distance between S and P. That is,
k
when S = ( , , ), I(x, y, z) =
for some constant k.
x +y +z
Let k = , let u = / , / , / be a unit vector, and let P = ( , , ).
Measure distances in inches. Find the direc onal deriva ve of I at P in the direc on of u, and nd the direc on of greatest intensity increase at P.
Example

We need the gradient I, meaning we need Ix , Iy and Iz . Each


S
par al deriva ve requires a simple applica on of the Quo ent Rule, giving

y
z
x
,
,
I =
(x + y + z ) (x + y + z ) (x + y + z )

I( , , ) =
.
, .
, .

,
,
Du I = I( , , ) u
=

The direc onal deriva ve tells us that moving in the direc on of u from P results in a decrease in intensity of about .
units per inch. (The intensity is
decreasing as u moves one farther from the origin than P.)
Notes:

Direc onal Deriva ves

Chapter

Func ons of Several Variables


The gradient gives the direc on of greatest intensity increase. No ce that

I( , , ) =
,
,
=

, , .

That is, the gradient at ( , , ) is poin ng in the direc on of , , , that


is, towards the origin. That should make intui ve sense: the greatest increase
in intensity is found by moving towards to source of the energy.
The direc onal deriva ve allows us to nd the instantaneous rate of z change
in any direc on at a point. We can use these instantaneous rates of change to
dene lines and planes that are tangent to a surface at a point, which is the topic
of the next sec on.

Notes:

Exercises

Terms and Concepts

(b) In the direc on toward the point Q = ( , ).

. What is the dierence between a direc onal deriva ve and


a par al deriva ve?

. f(x, y) = x + y xy x, P = ( , )

. For what u is Du f = fy ?

(a) In the direc on of v = ,

to level curves.

. The gradient points in the direc on of

(b) In the direc on toward the point Q = ( , ).


increase.

. It is generally more informa ve to view the direc onal


deriva ve not as the result of a limit, but rather as the result
product.
of a

(a) In the direc on of v = ,

(b) In the direc on toward the point Q = ( , ).

, a func on z = f(x, y) and a point P are

(a) Find the direc on of maximal increase of f at P.


, a func on z = f(x, y) is given. Find f.

. f(x, y) = x y + xy + xy

(b) What is the maximal value of Du f at P?


(c) Find the direc on of minimal increase of f at P.
(d) Give a direc on u such that Du f =

. f(x, y) = x y + xy + xy, P = ( , )

x +y +

. f(x, y) = x + y

. f(x, y) = sin x cos y, P =

. f(x, y) = x + y xy x

. f(x, y) =

. f(x, y) = x y x
In Exercises
, a func on z = f(x, y) and a point P are
given. Find the direc onal deriva ve of f in the indicated direc ons. Note: these are the same func ons as in Exercises
through .
. f(x, y) = x y + xy + xy, P = ( , )
(a) In the direc on of v = ,

(b) In the direc on toward the point Q = ( , ).


. f(x, y) = sin x cos y, P =

( )
,

(a) In the direc on of v = , .

(b) In the direc on toward the point Q = ( , ).


. f(x, y) =

x +y +

at P.

Note: these are the same func ons and points as in Exercises
through .

. f(x, y) = sin x cos y


. f(x, y) =

. f(x, y) = x y x, P = ( , )

In Exercises
given.

Problems
In Exercises

(a) In the direc on of v = , .

(b) In the direc on toward the point Q = ( , ).

. For what u is Du f = fx ?

. The gradient is

. f(x, y) = x + y, P = ( , )

, P = ( , ).

(a) In the direc on of v = , .

x +y +

( )
,

, P = ( , ).

. f(x, y) = x + y, P = ( , ).
. f(x, y) = x + y xy x, P = ( , )
. f(x, y) = x y x, P = ( , )
In Exercises
, a func on w = F(x, y, z), a vector v and
a point P are given.
(a) Find F(x, y, z).

(b) Find Du F at P.

. F(x, y, z) = x z + xy z , v = , , , P = ( , , )
. F(x, y, z) = sin(x) cos(y)ez , v = , , , P = ( , , )
. F(x, y, z) = x y y z , v = , , , P = ( , , )
. F(x, y, z) =

x +y +z

, v = , , , P = ( , , )

Chapter

Func ons of Several Variables

Tangent Lines, Normal Lines, and Tangent Planes

Deriva ves and tangent lines go handinhand.( Given y)= f(x), the line tangent
to the graph of f at x = x is the line through x , f(x ) with slope f (x ); that
is, the slope of the tangent line is the instantaneous rate of change of f at x .
When dealing with func ons of two variables, the graph is no longer a curve
but a surface. At a given point on the surface, it seems there are many lines that
t our intui on of being tangent to the surface.
In Figures . we see lines that are tangent to curves in space. Since each
curve lies on a surface, it makes sense to say that the lines are also tangent to
the surface. The next deni on formally denes what it means to be tangent
to a surface.
Figure . : Showing various lines tangent to a surface.

Deni on

Direc onal Tangent Line

Let z = f(x, y) be dieren able on an open set S containing (x , y ) and let


u = u , u be a unit vector.
(
)
. The line x through x , y , f(x , y ) parallel to , , fx (x , y ) is the
tangent line to f in the direc on of x at (x , y ).
(
)
. The line y through x , y , f(x , y ) parallel to , , fy (x , y ) is the
tangent line to f in the direc on of y at (x , y ).
(
)
. The line u through x , y , f(x , y ) parallel to u , u , Du f(x , y )
is the tangent line to f in the direc on of u at (x , y ).
It is instruc ve to consider each of three direc ons given in the deni on in
terms of slope. The direc on of x is , , fx (x , y ); that is, the run is one
unit in the x-direc on and the rise is fx (x , y ) units in the z-direc on. Note
how the slope is just the par al deriva ve with respect to x. A similar statement
can be made for y . The direc on of u is u , u , Du f(x , y ); the run is one
unit in the u direc on (where u is a unit vector) and the rise is the direc onal
deriva ve of z in that direc on.
Deni on leads to the following parametric equa ons of direc onal tangent lines:

x=x
x=x +u t
x=x +t
y=y +u t
y=y
y=y +t
.
, y (t) =
and u (t) =
x (t) =

z = z + Du f(x , y )t
z = z + fx (x , y )t
z = z + fy (x , y )t

Notes:

Tangent Lines, Normal Lines, and Tangent Planes

Example
Finding direc onal tangent lines
Find the lines tangent to the surface z = sin x cos y at (/ , / ) in the x and y
direc ons and also in the direc on of v = , .
S

The par al deriva ves with respect to x and y are:


fx (x, y) = cos x cos y
fy (x, y) = sin x sin y

fx (/ , / ) =
fy (/ , / ) = .

At (/ , / ), the z-value is .
Thus the parametric equa ons of the line tangent to f at (/ , / ) in the
direc ons of x and y are:

x = /
x = / + t
y = / + t .
y
=
/
and y (t) =
x (t) =

z = t
z=

(a)

The two lines are shown with the surface in Figure . (a). To nd the equaon of the tangent line inthe direc
on of v, we rst nd the unit vector in the
direc on of v: u = / , /
. The direc onal deriva ve at (/ , , ) in
the direc on of u is

Du f(/ , , ) = , / , /
= / .
Thus the direc onal tangent line is

x = / t/
u (t) =
y = /
+ t/

z = t/

(b)

The curve through (/ , / , ) in the direc on ofv is shown in Figure


along with u (t).

Figure . : A surface and direc onal


tangent lines in Example
.

(b)

Finding direc onal tangent lines


Example
Let f(x, y) = xy x y . Find the equa ons of all direc onal tangent lines to
f at ( , ).
First note that f( , ) = . We need to compute direc onal
S
deriva ves, so we need f. We begin by compu ng par al deriva ves.
f x = y x fx ( , ) = ;

f y = x y fy ( , ) = .

Thus f( , ) = , . Let u = u , u be any unit vector. The direc onal


deriva ve of f at ( , ) will be Du f( , ) = , u , u = . It does not ma er

Notes:

Chapter

Func ons of Several Variables


what direc on we choose; the direc onal deriva ve is always . Therefore

x= +u t
y= +u t .
u (t) =

z=
Figure . shows a graph of f and the point ( , , ). Note that this point
comes at the top of a hill, and therefore every tangent line through this point
will have a slope of .
That is, consider any curve on the surface that goes through this point. Each
curve will have a rela ve maximum at this point, hence its tangent line will have
a slope of . The following sec on inves gates the points on surfaces where all
tangent lines have a slope of .

Figure

: Graphing f in Example

Normal Lines
When dealing with a func on y = f(x) of one variable, we stated that a line
through (c, f(c)) was tangent to f if the line had a slope of f (c) and was normal
(or, perpendicular, orthogonal) to f if it had a slope of /f (c). We extend the
concept of normal, or orthogonal, to func ons of two variables.
Let z = f(x, y) be a dieren able func on of two variables. By Deni on ,
at (x , y ), x (t) is a line parallel to the vector dx = , , fx (x , y ) and y (t) is

(a line parallel to) dy = , , fy (x , y ). Since lines in these direc ons through


x , y , f(x , y ) are tangent to the surface, a line through this point and orthogonal to these direc ons would be orthogonal, or normal, to the surface. We can
use this direc on to create a normal line.
The direc on of the normal line is orthogonal to dx and dy , hence the direcon is parallel to dn = dx dy . It turns out this cross product has a very simple
form:
dx dy = , , fx , , fy = fx , fy , .
It is o en more convenient to refer to the opposite of this direc on, namely
fx , fy , . This leads to a deni on.

Notes:

Deni on

Tangent Lines, Normal Lines, and Tangent Planes

Normal Line

Let z = f(x, y) be dieren able on an open set S containing (x , y )


where
a = fx (x , y ) and b = fy (x , y )
are dened.
. A nonzero
vector parallel
to n = a, b, is orthogonal to f at
(
)
P = x , y , f(x , y ) .
. The line n through P with direc on parallel to n is the normal line
to f at P.

(
)
Thus the parametric equa ons of the normal line to a surface f at x , y , f(x , y )
is:

x = x + at
y = y + bt
n (t) =
.

z = f(x , y ) t

Finding a normal line


Example
Find the equa on of the normal line to z = x y + at ( , ).

We nd zx (x, y) = x and zy (x, y) = y; at ( , ), we


S
have zx = and zy = . We take the direc on of the normal line, following
Deni on , to be n = , , . The line with this direc on going through
the point ( , , ) is

x=
y= t+
n (t) =
or n (t) = , , t + , , .

z = t +
The surface z = x + y , along with the found normal line, is graphed in
Figure . .

The direc on of the normal line has many uses, one of which is the denion of the tangent plane which we dene shortly. Another use is in measuring
distances from the surface to a point. Given a point Q in space, it is general geometric concept to dene the distance from Q to the surface as being the length
of the shortest line segment PQ over all points P on the surface. This, in turn,
#
implies that PQ will be orthogonal to the surface at P. Therefore we can measure the distance from Q to the surface f by nding a point P on the surface such

Notes:

Figure . : Graphing a surface with a


normal line from Example
.

Chapter

Func ons of Several Variables


#
that PQ is parallel to the normal line to f at P.
Finding the distance from a point to a surface
Example
Let f(x, y) = x y and let Q = ( , , ). Find the distance from Q to the
surface dened by f.
This surface is used in Example
, so we know that at
S
(x, y), the direc on of the normal line will be dn = x, y, . A point P on
#
the surface will have coordinates (x, y, x y ), so PQ = x, y, x + y .
#
#
To nd where PQ is parallel to dn , we need to nd x, y and c such that cPQ = dn .

This implies

#
cPQ = dn

c x, y, x + y = x, y, .
c( x) = x

c( y) = y

c(x + y ) =
In each equa on, we can solve for c:
c=

x
y

=
=
.
x
y
x +y

The rst two frac ons imply x = y, and so the last frac on can be rewri en as
c = /( x ). Then

x
=
x
x
x( x ) = ( x)
x =

x +x

= .

This last equa on is a cubic, which is not dicult to solve with a numeric solver.
We nd that x = .
, hence P = ( .
, .
, .
). We nd the distance
from Q to the surface of f is

#
) +( .
) +( .
) = .
.
|| PQ || = ( .

We can take the concept of measuring the distance from a point to a surface
to nd a point Q a par cular distance from a surface at a given point P on the

Notes:

Tangent Lines, Normal Lines, and Tangent Planes

surface.
Example
Finding a point
from a surface
( a set distance
)
Let f(x, y) = x y + . Let P = , , f( , ) = ( , , ). Find points Q in space
#
that are units from the surface of f at P. That is, nd Q such that || PQ || =
#
and PQ is orthogonal to f at P.
We begin by nding par al deriva ves:

fx (x, y) =

fy (x, y) = y

fx ( , ) =
fy ( , ) =

The vector n = , , is orthogonal to f at P. For reasons that will become


more clear in a moment, we nd the unit vector in the direc on of n:
u =


n
.
= / , / , /
|| n ||

, .

, .

Thus a the normal line to f at P can be wri en as


n (t) = , , + t .

, .

, .

An advantage of this parametriza on of the line is that le ng t = t gives a


point on the line that is |t | units from P. (This is because the direc on of the
line is given in terms of a unit vector.) There are thus two points in space units
from P:
Q = n ( )
.

Q = n ( )
, .

, .

, .

, .

The surface is graphed along with points P, Q , Q and a por on of the normal
line to f at P.

Tangent Planes
We can use the direc on of the
( normal line to
) dene a plane. With a =
fx (x , y ), b = fy (x , y ) and P = x , y , f(x , y ) , the vector n = a, b,
is orthogonal to f at P. The plane through P with normal vector n is therefore
tangent to f at P.

Notes:

Figure . : Graphing the surface in Example


along with points units from
the surface.

Chapter

Func ons of Several Variables

Deni on

Tangent Plane

Let z = f(x, y) be dieren able on an open set S containing


(x , y( ), where a =) fx (x , y ), b = fy (x , y ), n = a, b, and
P = x , y , f(x , y ) .

The plane through P with normal vector n is the tangent plane to f at P.


The standard form of this plane is
(
)
a(x x ) + b(y y ) z f(x , y ) = .

Figure . : Graphing a surface with tangent plane from Example


.

Finding tangent planes


Example
Find the equa on tangent plane to z = x y + at ( , ).
Note that this is the same surface and point used in Example
. There we found n = , , and P = ( , , ). Therefore the equa on
of the tangent plane is
S

(y ) (z ) = .
The surface z = x + y and tangent plane are graphed in Figure

Using the tangent plane to approximate func on values


Example
The point ( , , ) lies on the surface of an unknown dieren able func on f
where fx ( , ) = and fy ( , ) = / . Find the equa on of the tangent
plane to f at P, and use this to approximate the value of f( . , . ).
Knowing the par al deriva ves at ( , ) allows us to form
S
the normal vector to the tangent plane, n = , / , . Thus the equa on
of the tangent line to f at P is:
(x ) / (y+ )(z ) =

z = (x ) / (y+ )+ . (

. )

Just as tangent lines provide excellent approxima ons of curves near their point
of intersec on, tangent planes provide excellent approxima ons of surfaces near
their point of intersec on. So f( . , . ) z( . , . ) = . .
This is not a new method of approxima on. Compare the right hand expression for z in Equa on ( . ) to the total dieren al:
dz = fx dx + fy dy

and

z = |{z} (x ) + / (y + ) + .
| {z } | {z } | {z }
fx

Notes:

dx

fy

{z
dz

dy

Tangent Lines, Normal Lines, and Tangent Planes

Thus the new z-value is the sum of the change in z (i.e., dz) and the old zvalue ( ). As men oned when studying the total dieren al, it is not uncommon
to know par al deriva ve informa on about a unknown func on, and tangent
planes are used to give accurate approxima ons of the func on.

The Gradient and Normal Lines, Tangent Planes


The methods developed in this sec on so far give a straigh orward method
of nding equa ons of normal lines and tangent planes for surfaces with explicit
equa ons of the form z = f(x, y). However, they do not handle implicit equaons well, such as x + y + z = . There is a technique that allows us to nd
vectors orthogonal to these surfaces based on the gradient.
Deni on

Gradient

Let w = F(x, y, z) be dieren able on an open ball B that contains the


point (x , y , z ).
. The gradient of F is F(x, y, z) = fx (x, y, z), fy (x, y, z), fz (x, y, z).
. The gradient of F at (x , y , z ) is
F(x , y , z ) = fx (x , y , z ), fy (x , y , z ), fz (x , y , z ) .
Recall that when z = f(x, y), the gradient f = fx , fy is orthogonal to level
curves of f. An analogous statement can be made about the gradient F, where
w = F(x, y, z). Given a point (x , y , z ), let c = F(x , y , z ). Then F(x, y, z) =
c is a level surface that contains the point (x , y , z ). The following theorem
states that F(x , y , z ) is orthogonal to this level surface.
Theorem

The Gradient and Level Surfaces

Let w = F(x, y, z) be dieren able on an open ball B containing


(x , y , z ) with gradient F, where F(x , y , z ) = c.
The vector F(x , y , z ) is orthogonal to the level surface F(x, y, z) = c
at (x , y , z ).

The gradient at a point gives a vector orthogonal to the surface at that point.
This direc on can be used to nd tangent planes and normal lines.

Notes:

Chapter

Func ons of Several Variables

Using the gradient to nd a tangent plane


x
y
z
Find the equa on of the plane tangent to the ellipsoid
+
+
=
Example

at

P = ( , , ).
We consider the equa on of the ellipsoid as a level surface
of a func on F of three variables, where F(x, y, z) = x + y + z . The gradient
is:
S

Figure . : An ellipsoid and its tangent


plane at a point.

F(x, y, z) = Fx , Fy , Fz
x y z
=
, ,
.

At P, the gradient is F( , , ) = / , / , / . Thus the equa on of the


plane tangent to the ellipsoid at P is
(x ) + (y ) + (z ) = .
The ellipsoid and tangent plane are graphed in Figure

Tangent lines and planes to surfaces have many uses, including the study of
instantaneous rates of changes and making approxima ons. Normal lines also
have many uses. In this sec on we focused on using them to measure distances
from a surface. Another interes ng applica on is in computer graphics, where
the eects of light on a surface are determined using normal vectors.
The next sec on inves gates another use of par al deriva ves: determining
rela ve extrema. When dealing with func ons of the form y = f(x), we found
rela ve extrema by nding x where f (x) = . We can start nding rela ve
extrema of z = f(x, y) by se ng fx and fy to , but it turns out that there is more
to consider.

Notes:

Exercises

Terms and Concepts


. Explain how the vector v = , , can be thought of as
having a slope of .
. Explain how the vector v = . , . , can be thought
of as having a slope of .
. T/F: Let z = f(x, y) be dieren able at P. If n is a normal
vector to the tangent plane of f at P, then n is orthogonal
to fx and fy at P.
. Explain in your own words why we do not refer to the tangent line to a surface at a point, but rather to direc onal
tangent lines to a surface at a point.

Problems
In Exercises , a func on z = f(x, y), a vectorv and a point
P are given. Give the parametric equa ons of the following
direc onal tangent lines to f at P:
(a) x (t)

. f(x, y) = x x y + y, P = ( , ).
In Exercises
, a func on z = f(x, y) and a point P are
given. Find the two points that are units from the surface
f at P. Note: these are the same func ons as in Exercises
.
. f(x, y) = x y xy , P = ( , ).
. f(x, y) = cos x sin y, P = (/ , / ).
. f(x, y) = x y, P = ( , ).
. f(x, y) = x x y + y, P = ( , ).
In Exercises
, a func on z = f(x, y) and a point P are
given. Find the equa on of the tangent plane to f at P. Note:
these are the same func ons as in Exercises .
. f(x, y) = x y xy , P = ( , ).
. f(x, y) = cos x sin y, P = (/ , / ).

(b) y (t)

. f(x, y) = x y, P = ( , ).

(c) u (t), where u is the unit vector in the direc on of v.

. f(x, y) = x x y + y, P = ( , ).

. f(x, y) = x y xy , v = , , P = ( , ).
. f(x, y) = cos x sin y, v = , , P = (/ , / ).

In Exercises , an implicitly dened func on of x, y and


z is given along with a point P that lies on the surface. Use
the gradient F to:

(a) nd the equa on of the normal line to the surface at


P, and

. f(x, y) = x y, v = , , P = ( , ).

(b) nd the equa on of the plane tangent to the surface


at P.

. f(x, y) = x x y + y, v = , , P = ( , ).
In Exercises , a func on z = f(x, y) and a point P are
given. Find the equa on of the normal line to f at P. Note:
these are the same func ons as in Exercises .

y
x

z
y

= , at P = ( ,


,
)

. f(x, y) = x y xy , P = ( , ).

. z

. f(x, y) = cos x sin y, P = (/ , / ).

. xy xz = , at P = ( , , )

. f(x, y) = x y, P = ( , ).

. sin(xy) + cos(yz) = , at P = ( , /

= , at P = ( , ,

, )

Chapter

Func ons of Several Variables

Extreme Values

Given a func on z = f(x, y), we are o en interested in points where z takes on


the largest or smallest values. For instance, if z represents a cost func on, we
would likely want to know what (x, y) values minimize the cost. If z represents
the ra o of a volume to surface area, we would likely want to know where z is
greatest. This leads to the following deni on.

Deni on

Rela ve and Absolute Extrema

Let z = f(x, y) be dened on a set S containing the point P = (x , y ).


. If there is an open disk D containing P such that f(x , y ) f(x, y)
for all (x, y) in D, then f has a rela ve maximum at P; if f(x , y )
f(x, y) for all (x, y) in D, then f has a rela ve minimum at P.
. If f(x , y ) f(x, y) for all (x, y) in S, then f has an absolute maximum at P; if f(x , y ) f(x, y) for all (x, y) in S, then f has an
absolute minimum at P.
. If f has a rela ve maximum or minimum at P, then f has a rela ve
extrema at P; if f has an absolute maximum or minimum at P, then
f has a absolute extrema at P.

If f has a rela ve or absolute maximum at P = (x , y ), it means every curve


on the surface of f through P will also have a rela ve or absolute maximum at P.
Recalling what we learned in Sec on . , the slopes of the tangent lines to these
curves at P must be or undened. Since direc onal deriva ves are computed
using fx and fy , we are led to the following deni on and theorem.

Deni on

Cri cal Point

Let z = f(x, y) be con nuous on an open set S. A cri cal point P =


(x , y ) of f is a point in S such that
fx (x , y ) =

and fy (x , y ) = , or

fx (x , y ) and/or fy (x , y ) is undened.

Notes:

Theorem

Extreme Values

Cri cal Points and Rela ve Extrema

Let z = f(x, y) be dened on an open set S containing P = (x , y ). If f


has a rela ve extrema at P, then P is a cri cal point of f.
Therefore, to nd rela ve extrema, we nd the cri cal points of f and determine which correspond to rela ve maxima, rela ve minima, or neither. The
following examples demonstrate this process.
Example
Finding cri cal points and rela ve extrema
Let f(x, y) = x + y xy x . Find the rela ve extrema of f.
S

We start by compu ng the par al deriva ves of f:


fx (x, y) = x y

and

fy (x, y) = y x.

Each is never undened. A cri cal point occurs when fx and fy are simultaneously
, leading us to solve the following system of linear equa ons:
xy

and

x+ y= .

This solu on to this system is x = / , y = / . (Check that at ( / , / ), both


fx and fy are .)
The graph in Figure . shows f along with this cri cal point. It is clear from
the graph that this is a rela ve minimum; further considera on of the func on
shows that this is actually the absolute minimum.

Figure . : The surface in Example


with its absolute minimum indicated.

Example
Finding cri cal points and rela ve extrema
Let f(x, y) = x + y + . Find the rela ve extrema of f.
S

We start by compu ng the par al deriva ves of f:

fx (x, y) =

x +y

and

y
fy (x, y) =
.
x +y

It is clear that fx = when x = & y = , and that fy = when y = & x = .


At ( , ), both fx and fy are not , but rather undened. The point ( , ) is s ll a
cri cal point, though, because the par al deriva ves are undened. This is the
only cri cal point of f.
The surface of f is graphed in Figure . along with the point ( , , ). The
graph shows that this point is the absolute maximum of f.

Notes:

Figure . : The surface in Example


with its absolute maximum indicated.

Chapter

Func ons of Several Variables


In each of the previous two examples, we found a cri cal point of f and then
determined whether or not it was a rela ve (or absolute) maximum or minimum
by graphing. It would be nice to be able to determine whether a cri cal point
corresponded to a max or a min without a graph. Before we develop such a test,
we do one more example that sheds more light on the issues our test needs to
consider.
Example
Finding cri cal points and rela ve extrema
Let f(x, y) = x x y + y. Find the rela ve extrema of f.
S

Once again we start by nding the par al deriva ves of f:


fx (x, y) = x

and

fy (x, y) = y + .

Each is always dened. Se ng each equal to and solving for x and y, we nd


fx (x, y) =
fy (x, y) =

Figure . : The surface in Example


with both cri cal points marked.

x=
y= .

We have two cri cal points: ( , ) and ( , ). To determine if they correspond


to a rela ve maximum or minimum, we consider the graph of f in Figure . .
The cri cal point ( , ) clearly corresponds to a rela ve maximum. However, the cri cal point at ( , ) is neither a maximum nor a minimum, displaying
a dierent, interes ng characteris c.
If one walks parallel to the y-axis towards this cri cal point, then this point
becomes a rela ve maximum along this path. But if one walks towards this point
parallel to the x-axis, this point becomes a rela ve minimum along this path. A
point that seems to act as both a max and a min is a saddle point. A formal
deni on follows.

Deni on

Saddle Point

Let P = (x , y ) be in the domain of f where fx = and fy = at


P. P is a saddle point of f if, for every open disk D containing P, there
are points (x , y ) and (x , y ) in D such that f(x , y ) > f(x , y ) and
f(x , y ) < f(x , y ).

At a saddle point, the instantaneous rate of change in all direc ons is and
there are points nearby with z-values both less than and greater than the z-value
of the saddle point.

Notes:

.
Before Example
we men oned the need for a test to dieren ate between rela ve maxima and minima. We now recognize that our test also needs
to account for saddle points. To do so, we consider the second par al deriva ves
of f.
Recall that with single variable func ons, such as y = f(x), if f (c) > , then
f is concave up at c, and if f (c) = , then f has a rela ve minimum at x = c. (We
called this the Second Deriva ve Test.) Note that at a saddle point, it seems the
graph is both concave up and concave down, depending on which direc on
you are considering.
It would be nice if the following were true:
fxx and fyy >
fxx and fyy <
fxx and fyy have opposite signs

rela ve minimum
rela ve maximum
saddle point.

However, this is not the case. Func ons f exist where fxx and fyy are both
posi ve but a saddle point s ll exists. In such a case, while the concavity in the
x-direc on is up (i.e., fxx > ) and the concavity in the y-direc on is also up (i.e.,
fyy > ), the concavity switches somewhere in between the x- and y-direc ons.
To account for this, consider D = fxx fyy fxy fyx . Since fxy and fyx are equal
when con nuous (refer back to Theorem
), we can rewrite this as D = fxx fyy
fxy . D can be used to test whether the concavity at a point changes depending on
direc on. If D > , the concavity does not switch (i.e., at that point, the graph
is concave up or down in all direc ons). If D < , the concavity does switch. If
D = , our test fails to determine whether concavity switches or not. We state
the use of D in the following theorem.
Theorem

Second Deriva ve Test

Let z = f(x, y) be dieren able on an open set containing P = (x , y ),


and let
D = fxx (x , y )fyy (x , y ) fxy (x , y ).
. If D >

and fxx (x , y ) > , then P is a rela ve minimum of f.

. If D >

and fxx (x , y ) < , then P is a rela ve maximum of f.

. If D < , then P is a saddle point of f.


. If D = , the test is inconclusive.
We rst prac ce using this test with the func on in the previous example,
where we visually determined we had a rela ve maximum and a saddle point.

Notes:

Extreme Values

Chapter

Func ons of Several Variables


Example
Using the Second Deriva ve Test
Let f(x, y) = x xy + y as in Example
. Determine whether the func on
has a rela ve minimum, maximum, or saddle point at each cri cal point.
We determined previously that the cri cal points of f are
S
( , ) and ( , ). To use the Second Deriva ve Test, we must nd the second
par al deriva ves of f:
fxx = x;

fyy = ;

fxy = .

Thus D(x, y) = x.
At ( , ): D( , ) =
> , and fxx ( , ) = . By the Second Derivave Test, f has a rela ve maximum at ( , ).
At ( , ): D( , ) = < . The Second Deriva ve Test states that f has a
saddle point at ( , ).
The Second Deriva ve Test conrmed what we determined visually.
Using the Second Deriva ve Test
Example
Find the rela ve extrema of f(x, y) = x y + y + xy.
f:

We start by nding the rst and second par al deriva ves of

fx = xy + y
fxx = y
fxy = x +

fy = x + y + x
fyy =
fyx = x + .

We nd the cri cal points by nding where fx and fy are simultaneously (they
are both never undened). Se ng fx = , we have:
fx =

xy + y =

This implies that for fx = , either y = or x +


Assume y = then consider fy = :

y( x + ) = .
= .

fy =
x + y+x= ,

and since y = , we have

x +x=
x(x + ) = .
Thus if y = , we have either x = or x = , giving two cri cal points: ( , )
and ( , ).

Notes:

Extreme Values

Going back to fx , now assume x + = , i.e., that x = / , then consider


fy = :
fy =
x + y+x= ,
/ + y / =
y= / .

and since x = / , we have

Thus if x = / , y = / giving the cri cal point ( / , / ).


With D = y ( x + ) , we apply the Second Deriva ve Test to each cri cal
point.
At ( , ), D < , so ( , ) is a saddle point.
At ( , ), D < , so ( , ) is also a saddle point.
At ( / , / ), D > and fxx > , so ( / , / ) is a rela ve minimum.
Figure . shows a graph of f and the three cri cal points. Note how this
func on does not vary much near the cri cal points that is, visually it is dicult to determine whether a point is a saddle point or rela ve minimum (or even
a cri cal point at all!). This is one reason why the Second Deriva ve Test is so
important to have.

Constrained Op miza on
When op mizing func ons of one variable such as y = f(x), we made use
of Theorem , the Extreme Value Theorem, that said that over a closed interval I, a con nuous func on has both a maximum and minimum value. To nd
these maximum and minimum values, we evaluated f at all cri cal points in the
interval, as well as at the endpoints (the boundary) of the interval.
A similar theorem and procedure applies to func ons of two variables. A
con nuous func on over a closed set also a ains a maximum and minimum
value (see the following theorem). We can nd these values by evalua ng the
func on at the cri cal values in the set and over the boundary of the set. A er
formally sta ng this extreme value theorem, we give examples.
Theorem

Extreme Value Theorem

Let z = f(x, y) be a con nuous func on on a closed, bounded set S. Then


f has a maximum and minimum value on S.
Finding extrema on a closed set
Example
Let f(x, y) = x y + and let S be the triangle with ver ces ( , ), ( , )
and ( , ). Find the maximum and minimum values of f on S.

Notes:

Figure . : Graphing f from Example


and its rela ve extrema.

Chapter

Func ons of Several Variables


It can help to see a graph of f along with the set S. In Figure
S
. (a) the triangle dening S is shown in the x-y plane in a dashed line. Above
it is the surface of f; we are only concerned with the por on of f enclosed by the
triangle on its surface.
We begin by nding the cri cal points of f. With fx = x and fy = y, we
nd only one cri cal point, at ( , ).
We now nd the maximum and minimum values that f a ains along the
boundary of S, that is, along the edges of the triangle. In Figure . (b) we
see the triangle sketched in the plane with the equa ons of the lines forming its
edges labeled.
Start with the bo om edge, along the line y = . If y is , then on
the surface, we are considering points f(x, ); that is, our func on reduces to
f(x, ) = x ( ) + = x + = f (x). We want to maximize/minimize
f (x) = x + on the interval [ , ]. To do so, we evaluate f (x) at its cri cal
points and at the endpoints.
The cri cal points of f are found by se ng its deriva ve equal to :

(a)
y

y=

Evalua ng f at this cri cal point, and at the endpoints of [ , ] gives:


/

y=

f ( ) =
f ( )=

x+

x+

x= .

f (x) =

y=

(b)

Figure . : Plo ng the surface of f


along with the restricted domain S.

f ( )=

f( , ) =
f( , ) =

f( , ) = .

No ce how evalua ng f at a point is the same as evalua ng f at its corresponding point.


We need to do this process twice more, for the other two edges of the triangle.
Along the le edge, along the line y = x + , we subs tute x + in for y
in f(x, y):
f(x, y) = f(x, x + ) = x ( x + ) +

= x x+

= f (x).

We want the maximum and minimum values of f on the interval [ , ], so we


evaluate f at its cri cal points and the endpoints of the interval. We nd the
cri cal points:
f (x) =

x= / .

Evaluate f at its cri cal point and the endpoints of [ , ]:


f ( / ) =

Notes:

f ( ) =
/ = .
f ( )=

f( , ) =
f( / , .
f( , ) = .

)= .

Extreme Values

Finally, we evaluate f along the right edge of the triangle, where y = / x+

f(x, y) = f(x, / x + ) = x ( / x + ) +

= x + x+

= f (x).

The cri cal points of f (x) are:


f (x) =

x= / = . .

We evaluate f at this cri cal point and at the endpoints of the interval [ , ]:
f ( )=
f ( . )= .
f ( )=

f( , ) =
f( . , . ) = .
f( , ) = .

One last point to test: the cri cal point of f, ( , ). We nd f( , ) = .


We have evaluated f at a total of dierent places, all shown in Figure . .
We checked each vertex of the triangle twice, as each showed up as the endpoint of an interval twice. Of all the z-values found, the maximum is . , found
at ( . , . ); the minimum is , found at ( , ).
This por on of the text is en tled Constrained Op miza on because we
want to op mize a func on (i.e., nd its maximum and/or minimum values)
subject to a constraint some limit to what values the func on can a ain. In
the previous example, we constrained ourselves by considering a func on only
within the boundary of a triangle. This was largely arbitrary; the func on and
the boundary were chosen just as an example, with no real meaning behind
the func on or the chosen constraint.
However, solving constrained op miza on problems is a very important topic
in applied mathema cs. The techniques developed here are the basis for solving
larger problems, where more than two variables are involved.
We illustrate the technique once more with a classic problem.
Constrained Op miza on
Example
The U.S. Postal Service states that the girth+length of Standard Post Package
must not exceed
. Given a rectangular box, the length is the longest side,
and the girth is twice the width+height.
Given a rectangular box where the width and height are equal, what are the
dimensions of the box that give the maximum volume subject to the constraint
of the size of a Standard Post Package?
Let w, h and denote the width, height and length of a rectS
angular box; we assume here that w = h. The girth is then (w + h) = w. The

Notes:

Figure . : The surface of f along with


important points along the boundary of S
and the interior.

Chapter

Func ons of Several Variables


volume of the box is V(w, ) = wh = w . We wish to maximize this volume
subject to the constraint w +
, or
w. (Common sense also
indicates that > , w > .)
We begin by nding the cri cal values of V. We nd that Vw = w and
V = w ; these are simultaneously only at ( , ). This gives a volume of , so
we can ignore this cri cal point.
We now consider the volume along the constraint =
w. Along this
line, we have:
V(w) = V(w,

w) = w (

w) =

w w = V (w).

The constraint is applicable on the w-interval [ , . ] as indicated in the gure.


Thus we want to maximize V on [ , . ].
Finding the cri cal values of V , we take the deriva ve and set it equal to :
Figure . : Graphing the volume of a
box with girth w and length , subject to
a size constraint.

V (w) =

w =

w(

w) =

w= ,

We found two cri cal values: when w = and when w = . . We again


ignore the w = solu on; the maximum volume, subject to the constraint,
comes at w = h = . , =
( . ) = . . This gives a volume of
V( . , . ) ,
in .
The volume func on V(w, ) is shown in Figure . along with the constraint =
w. As done previously, the constraint is drawn dashed in
the x-y plane and also along the surface of the func on. The point where the
volume is maximized is indicated.
It is hard to overemphasize the importance of op miza on. In the real
world, we rou nely seek to make something be er. By expressing the something as a mathema cal func on, making something be er means op mize
some func on.
The techniques shown here are only the beginning of an incredibly important
eld. Many func ons that we seek to op mize are incredibly complex, making
the step of nd the gradient and set it equal to highly nontrivial. Mastery
of the principles here are key to being able to tackle these more complicated
problems.

Notes:

Exercises

Terms and Concepts

. f(x, y) = x + y y +

. T/F: Theorem
states that if f has a cri cal point at P,
then f has a rela ve extrema at P.

. f(x, y) =

. T/F: A point P is a cri cal point of f if fx and fy are both at


P.

. f(x, y) = x y

. T/F: A point P is a cri cal point of f if fx or fy are undened


at P.
. Explain what it means to solve a constrained op miza on
problem.

Problems
In Exercises , nd the cri cal points of the given funcon. Use the Second Deriva ve Test to determine if each critical point corresponds to a rela ve maximum, minimum, or
saddle point.
. f(x, y) = x + y y + x
. f(x, y) = x + x + y y + xy
. f(x, y) = x + y y + xy
. f(x, y) =

x +y +

x x+

y y

. f(x, y) = x x + y
. f(x, y) =
. f(x, y) =

(x ) y
x +y

In Exercises
, nd the absolute maximum and minimum of the func on subject to the given constraint.
. f(x, y) = x + y + y + , constrained to the triangle with
ver ces ( , ), ( , ) and ( , ).
. f(x, y) = x y, constrained to the region bounded by
y = x and y = .
. f(x, y) = x + x + y + y, constrained to the region
bounded by the circle x + y = .
. f(x, y) = y x , constrained to the region bounded by
the parabola y = x + x and the line y = x.

:M

The previous chapter introduced mul variable func ons and we applied concepts of dieren al calculus to these func ons. We learned how we can view a
func on of two variables as a surface in space, and learned how par al derivaves convey informa on about how the surface is changing in any direc on.
In this chapter we apply techniques of integral calculus to mul variable funcons. In Chapter we learned how the denite integral of a single variable funcon gave us area under the curve. In this chapter we will see that integra on
applied to a mul variable func on gives us volume under a surface. And just
as we learned applica ons of integra on beyond nding areas, we will nd applica ons of integra on in this chapter beyond nding volume.

Iterated Integrals and Area

In Chapter
we found that it was useful to dieren ate func ons of several
variables with respect to one variable, while trea ng all the other variables as
constants or coecients. We can integrate func ons of several variables in a
similar way. For instance, if we are told that fx (x, y) = xy, we can treat y as
staying constant and integrate to obtain f(x, y):

f(x, y) = fx (x, y) dx

=
xy dx
= x y + C.

Make a careful note about the constant of integra on, C. This constant is
something with a deriva ve of with respect to x, so it could be any expression that contains only constants and func ons of y. For instance, if f(x, y) =
x y + sin y + y + , then fx (x, y) = xy. To signify that C is actually a func on
of y, we write:

f(x, y) =

fx (x, y) dx = x y + C(y).

Using this process we can even evaluate denite integrals.

Integra
y ng func ons of more than one variable
xy dx.
Evaluate the integral
Example

We nd the indenite integral as before, then apply the FunS


damental Theorem of Calculus to evaluate the denite integral:



xy dx = x y

= ( y) y ( ) y

= y y.

Chapter

Mul ple Integra on


We can also integrate with respect to y. In general,

and

h (y)
h (y)

g (x)
g (x)

h

fx (x, y) dx = f(x, y)

(y)

g

fy (x, y) dy = f(x, y)

(x)

h (y)

g (x)

(
)
(
)
= f h (y), y f h (y), y ,
(
)
(
)
= f x, g (x) f x, g (x) .

Note that when integra ng with respect to x, the bounds are func ons of y
(of the form x = h (y) and x = h (y)) and the nal result is also a func on of y.
When integra ng with respect to y, the bounds are func ons of x (of the form
y = g (x) and y = g (x)) and the nal result is a func on of x. Another example
will help us understand this.
Example
Evaluate

Integra ng func ons of more than one variable


)
x y + y dy.

We consider x as staying constant and integrate with respect

to y:

x y + y

dy =
=
=

) x



) (
)
+ x x +

x y
y
+

(
x x

x x .

Note how the bounds of the integral are from y =


answer is a func on of x.

to y = x and that the nal

In the previous example, we integrated a func on with respect to y and


ended up with a func on of x. We can integrate this as well. This process is
known as iterated integra on, or mul ple integra on.
Example

Evaluate

Integra ng an integral
)
)
x y + y dy dx.

We follow a standard order of opera ons and perform the


S
opera ons inside parentheses rst (which is the integral evaluated in Example

Notes:

.
.)
x

x y

+ y

dy

dx =
=
=

([

x y
y
+

)
(
x x

dx

)

x x x

=
Note how the bounds of x were x =
ber.

] x )

dx

to x =

and the nal result was a num-

The previous example showed how we could perform something called an


iterated integral; we do not yet know why we would be interested in doing so
nor what the result, such as the number / , means. Before we inves gate
these ques ons, we oer some deni ons.
Deni on

Iterated Integra on

Iterated integra on is the process of repeatedly integra ng the results


of previous integra ons. Integra ng one integral is denoted as follows.
Let a, b, c and d be numbers and let g (x), g (x), h (y) and h (y) be
func ons of x and y, respec vely. Then:
)
d ( h (y)
d h (y)
f(x, y) dx dy.
f(x, y) dx dy =
.
c

b
a

g (x)
g (x)

h (y)

h (y)

f(x, y) dy dx =

b
a

g (x)
g (x)

f(x, y) dy dx.

Again make note of the bounds of these iterated integrals.


d h (y)
With
f(x, y) dx dy, x varies from h (y) to h (y), whereas y varies from
c

h (y)

c to d. That is, the bounds of x are curves, the curves x = h (y) and x = h (y),
whereas the bounds of y are constants, y = c and y = d. It is useful to remember
that when se ng up and evalua ng such iterated integrals, we integrate from

Notes:

Iterated Integrals and Area

Chapter

Mul ple Integra on


curve to curve, then from point to point.
We now begin to inves gate why we are interested in iterated integrals and
what they mean.

Area of a plane region


y

Consider the plane region R bounded by a x b and g (x) y g (x),


shown in Figure . . We learned in Sec on . that the area of R is given by

y = g2 (x)

y = g1 (x)

Figure . : Calcula ng the area of a


plane region R with an iterated integral.

b
a

)
g (x) g (x) dx.

(
)
We can view the expression g (x) g (x) as
(
)
g (x) g (x) =

g (x)

dy =

g (x)

g (x)
g (x)

dy,

meaning we can express the area of R as an iterated integral:

area of R =

y
d

b
a

(
)
g (x) g (x) dx =

b
a

g (x)
g (x)

dy dx =

b
a

g (x)
g (x)

dy dx.

x = h2 (y)

In short: a certain iterated integral can be viewed as giving the area of a


plane region.

A region R could also be dened by c y d and h (y) x h (y), as


shown in Figure . . Using a process similar to that above, we have

x = h1 (y)
c

the area of R =

Figure . : Calcula ng the area of a


plane region R with an iterated integral.

We state this formally in a theorem.

Notes:

d
c

h (y)
h (y)

dx dy.

Theorem

Iterated Integrals and Area

Area of a plane region

. Let R be a plane region bounded by a x b and g (x) y


g (x), where g and g are con nuous func ons on [a, b]. The area
A of R is

A=

g (x)

g (x)

dy dx.

. Let R be a plane region bounded by c y d and h (y) x


h (y), where h and h are con nuous func ons on [c, d]. The area
A of R is

h (y)

h (y)

dx dy.

The following examples should help us understand this theorem.


Area of a rectangle
Example
Find the area A of the rectangle with corners ( , ) and ( , ), as shown in
Figure . .
Mul ple integra on is obviously overkill in this situa on, but
S
we proceed to establish its use.
The region R is bounded by x = , x = , y = and y = . Choosing to
integrate with respect to y rst, we have
( )




dx = x = .
dx =
y
dy dx =
A=

We could also integrate with respect to x rst, giving:


( )



x
A=
dy = y = .
dy =
dx dy =

Figure . : Calcula ng the area of a rectangle with an iterated integral in Example


.

Clearly there are simpler ways to nd this area, but it is interes ng to note
that this method works.
Area of a triangle
Example
Find the area A of the triangle with ver ces at ( , ), ( , ) and ( , ), as shown
in Figure . .
The triangle is bounded by the lines as shown in the gure.
Choosing to integrate with respect to x rst gives that x is bounded by x = y

y=

Notes:

y=

A=

y=

Figure . : Calcula ng the area of a triangle with iterated integrals in Example


.

Chapter

Mul ple Integra on


to x = y+ , while y is bounded by y = to y = . (Recall that since x-values
increase from le to right, the le most curve, x = y, is the lower bound and the
rightmost curve, x = (y + )/ , is the upper bound.) The area is
A=

y+

dx dy

( y+ )

dy
x
y
)
(
dy
y+
=
(
)

= y + y
=

= .

We can also nd the area by integra ng with respect to y rst. In this situaon, though, we have two func ons that act as the lower bound for the region
R, y = and y = x . This requires us to use two iterated integrals. Note
how the x-bounds are dierent for each integral:
x
x
dy dx
dy dx
+
A=
x

=
=

( ) x
y dx

=
= .

(
)
x dx

+
+

( ) x
y
(

dx

x+

dx

As expected, we get the same answer both ways.


Area of a plane region
Example
Find the area of the region enclosed by y = x and y = x , as shown in Figure
. .

Figure . : Calcula ng the area of a


plane region with iterated integrals in Example
.

ders of integra on.


Using dy dx:
x
x

Notes:

Once again well nd the area of the region using both or-

dy dx =

)
(
( x x ) dx = x x =

.
Using dx dy:
y
y/

dx dy =

( y y/ ) dy =

Changing Order of Integra on

)

=

Iterated Integrals and Area

In each of the previous examples, we have been given a region R and found
the bounds needed to nd the area of R using both orders of integra on. We
integrated using both orders of integra on to demonstrate their equality.
We now approach the skill of describing a region using both orders of integra on from a dierent perspec ve. Instead of star ng with a region and crea ng iterated integrals, we will start with an iterated integral and rewrite it in
the other integra on order. To do so, well need to understand the region over
which we are integra ng.
The simplest of all cases is when both integrals are bound by constants. The
region described by these bounds is a rectangle (see Example
), and so:
d b
b d
dx dy.
dy dx =
a

When the inner integrals bounds are not constants, it is generally very useful
to sketch the bounds to determine what the region we are integra ng over looks
like. From the sketch we can then rewrite the integral with the other order of
integra on.
Examples will help us develop this skill.
Changing the order of integra on
x/
dy dx with the order of integra on dx dy.
Rewrite the iterated integral
Example

We need to use the bounds of integra on to determine the


region we are integra ng over.
The bounds tell us that y is bounded by and x/ ; x is bounded by and .
We plot these four curves: y = , y = x/ , x = and x = to nd the region
described by the bounds. Figure . shows these curves, indica ng that R is a
triangle.
To change the order of integra on, we need to consider the curves that
bound the x-values. We see that the lower bound is x = y and the upper
bound

is x = . The bounds on y are to . Thus we can rewrite the integral as

Notes:

dx dy.

y=

x/

R
.

Figure . : Sketching the region R described by the iterated integral in Example


.

Chapter

Mul ple Integra on


Example

Changing the order of integra on


(y+ )/
dx dy.
Change the order of integra on of
y /

y
4

R
x=

(y

2 4
y/

4)
/2

x=

We sketch the region described by the bounds to help us


S
change the integra on order. x is bounded below and above (i.e., to the le and
right) by x = y / and x = (y + )/ respec vely, and y is bounded between
and . Graphing the previous curves, we nd the region R to be that shown in
Figure . .
To change the order of integra on, we need to establish curves that bound
y. The gure makes it clear that there are two lower bounds for y: y = on
x , and y = x on x . Thus we need two double integrals.
The upper bound for each is y =
x. Thus we have

Figure . : Drawing the region determined by the bounds of integra on in Example


.

(y+ )/
y /

dx dy =

dy dx +

x
x

dy dx.

This sec on has introduced a new concept, the iterated integral. We developed one applica on for iterated integra on: area between curves. However,
this is not new, for we already know how to nd areas bounded by curves.
In the next sec on we apply iterated integra on to solve problems we currently do not know how to handle. The real goal of this sec on was not to
learn a new way of compu ng area. Rather, our goal was to learn how to dene
a region in the plane using the bounds of an iterated integral. That skill is very
important in the following sec ons.

Notes:

Exercises

Terms and Concepts


. When integra ng fx (x, y) with respect to x, the constant of
integra on C is really which: C(x) or C(y)? What does this
mean?

In Exercises , a graph of a planar region R is given. Give


the iterated integrals, with both orders of integra on dy dx
and dx dy, that give the area of R. Evaluate one of the iterated integrals to nd the area.

. Integra ng an integral is called

. When evalua ng an iterated integral, we integrate from


to
, then from
to
.
. One understanding of an iterated integral is that
b g (x)
dy dx gives the
of a plane region.
a

g (x)

Problems

In Exercises
erated integral.
.

(a)

(b)

(a)

(b)

(a)

(b)

(a)

(b)

(a)

(b)

(a)

x(

(b)

dy

x + xy y

dy dx

)
x cos y + sin x dx dy

(
)
x y y + dy

)
x cos y + sin x dx

, evaluate the integral and subsequent it-

x + xy y

(
)
x y y + dy dx

)
x y dx
y

)
x y dx dy

)
cos x sin y dx

6
x=

)
cos x sin y dx dy

+x
x(

)
+x

dy

dy dx

y/

y
1

In Exercises , iterated integrals are given that compute


the area of a region R in the x-y plane. Sketch the region R,
and give the iterated integral(s) that give the area of R with
the opposite order of integra on.

y=

0.5

R
.

y=
0.5

0.5

.
x

R
y=

dy dx

dy dx

8
6

.
0.5

dx dy

dy dx

dx dy +

( x)/
(x )/

dy dx

dx dy

Double Integra on and Volume

Double Integra on and Volume

b
The denite integral of f over [a, b], a f(x) dx, was introduced as the signed
area under the curve. We approximated the value of this area by rst subdividing [a, b] into n subintervals, where the i th subinterval has length xi , and le ng
ci be any value in the i th subinterval. We formed rectangles that approximated
part of the region under the curve with width xi , height f(ci ), and hence with
area f(ci )xi . Summing all the rectangles areas gave an approxima on of the
denite integral, and Theorem stated that
b

f(x) dx = lim
f(ci )xi ,
x

connec ng the area under the curve with sums of the areas of rectangles.

We use a similar approach in this sec on to nd volume under a surface.


Let R be a closed, bounded region in the x-y plane and let z = f(x, y) be
a con nuous func on dened on R. We wish to nd the signed volume under
the surface of f over R. (We use the term signed volume to denote that space
above the x-y plane, under f, will have a posi ve volume; space above f and
under the x-y plane will have a nega ve volume, similar to the no on of signed
area used before.)
We start by par oning R into n rectangular subregions as shown in Figure
. (a). For simplicitys sake, we let all widths be x and all heights be y.
Note that the sum of the areas of the rectangles is not equal to the area of R,
but rather is a close approxima on. Arbitrarily number the rectangles through
n, and pick a point (xi , yi ) in the i th subregion.
The volume of the rectangular solid whose base is the i th subregion and
whose height is f(xi , yi ) is Vi = f(xi , yi )xy. Such a solid is shown in Figure
. (b). Note how this rectangular solid only approximates the true volume under the surface; part of the solid is above the surface and part is below.
For each subregion Ri used to approximate R, create the rectangular solid
with base area xy and height f(xi , yi ). The sum of all rectangular solids is
n

.5

.5

.
(a)

f(xi , yi )xy.

i=

This approximates the signed volume under f over R. As we have done before,
to get a be er approxima on we can use more rectangles to approximate the
region R.
In general, each rectangle could have a dierent width xj and height yk ,
giving the i th rectangle an area Ai = xj yk and the i th rectangular solid a

Notes:

(b)
Figure . : Developing a method for
nding signed volume under a surface.

Chapter

Mul ple Integra on


volume of f(xi , yi )Ai . Let ||A|| denote the length of the longest diagonal of all
rectangles in the subdivision of R; ||A|| means each rectangles width and
height are both approaching . If f is a con nuous func on, as ||A|| shrinks
n

f(xi , yi )Ai approximates the signed


(and hence n ) the summa on
i=

volume be er and be er. This leads to a deni on.

Note:
Recall that the integra on symbol

is an elongated S, represen ng the


b
word sum. We interpreted a f(x) dx as
take the sum of the areas of rectangles
over the interval [a, b]. The double integral uses two integra on symbols to represent a double sum. When adding up
the volumes of rectangular solids over a
par on of a region R, as done in Figure
. , one could rst add up the volumes
across each row (one type of sum), then
add these totals together (another sum),
as in
m
n

j=

f(xi , yj )xi yj .

i=

One can rewrite this as


)
( m
n

f(xi , yj )xi yj .
j=

i=

The summa on inside the parenthesis


indicates the sum of heights widths,
which gives an area; mul plying these areas by the thickness yj gives a volume.
The illustra on in Figure . relates to
this understanding.

Deni on

Double Integral, Signed Volume

Let z = f(x, y) be a con nuous func on dened over a closed region R


in the x-y plane. The signed volume V under f over R is denoted by the
double integral

V=

f(x, y) dA.

Alternate nota ons for the double integral are

f(x, y) dy dx.
f(x, y) dx dy =
f(x, y) dA =
R

The deni on above does not state how to nd the signed volume, though
the nota on oers a hint. We need the next two theorems to evaluate double
integrals to nd volume.
Theorem

Double Integrals and Signed Volume

Let z = f(x, y) be a con nuous func on dened over a closed region R


in the x-y plane. Then the signed volume V under f over R is
V=

f(x, y) dA =

lim

||A||

f(xi , yi )Ai .

i=

This theorem states that we can nd the exact signed volume using a limit
of sums. The par on of the region R is not specied, so any par oning where
the diagonal of each rectangle shrinks to results in the same answer.
This does not oer a very sa sfying way of compu ng area, though. Our
experience has shown that evalua ng the limits of sums can be tedious. We
seek a more direct method.
Recall Theorem
in Sec on . . This stated that if A(x) gives the crossb
sec onal area of a solid at x, then a A(x) dx gave the volume of that solid over
Notes:

Double Integra on and Volume

[a, b].
Consider Figure . , where a surface z = f(x, y) is drawn over a region R.
Fixing a par cular x value, we can consider the area under f over R where x has
that xed value. That area can be found with a denite integral, namely
A(x) =

g (x)
g (x)

f(x, y) dy.

Remember that though the integrand contains x, we are viewing x as xed.


Also note that the bounds of integra on are func ons of x: the bounds depend
on the value of x.
As A(x) is a cross-sec onal area func on, we can nd the signed volume V
under f by integra ng it:
)

(

V=

A(x) dx =

g (x)

g (x)

f(x, y) dy dx =

g (x)

g (x)

f(x, y) dy dx.

This gives a concrete method for nding signed volume under a surface. We
could do a similar procedure where we started with y xed, resul ng in a iterated
integral with the order of integra on dx dy. The following theorem states that
both methods give the same result, which is the value of the double integral. It
is such an important theorem it has a name associated with it.
Theorem

Fubinis Theorem

Let R be a closed, bounded region in the x-y plane and let z = f(x, y) be
a con nuous func on on R.
. If R is bounded by a x b and g (x) y g (x), where g
and g are con nuous func ons on [a, b], then

f(x, y) dA =

b
a

g (x)
g (x)

f(x, y) dy dx.

. If R is bounded by c y d and h (y) x h (y), where h


and h are con nuous func ons on [c, d], then

Notes:

f(x, y) dA =

d
c

h (y)
h (y)

f(x, y) dx dy.

Figure . : Finding volume under a surface by sweeping out a crosssec onal


area.

Chapter

Mul ple Integra on


Note that once again the bounds of integra on follow the curve to curve,
point to point pa ern discussed in the previous sec on. In fact, one of the
main points of the previous sec on is developing the skill of describing a region
R with the bounds of an iterated integral. Once this skill is developed, we can
use double integrals to compute many quan es, not just signed volume under
a surface.
Evalua ng a double integral
Example
Let f(x, y) = xy+ey . Find the signed volume under f on the region R, which is the
rectangle with corners ( , ) and ( , ) pictured in Figure . , using Fubinis
Theorem and both orders of integra on.
)
(
We wish to evaluate R xy + ey dA. As R is a rectangle,
the bounds are easily described as x and y .
S

Using the order dy dx:

)
(
y
xy + e dA =
R

Figure . : Finding the signed volume


under a surface in Example
.

=
=
=

)
(
xy + ey dy dx
([
] )

y
dx
xy + e
)
(
x + e e dx
)
)
x + e e x
(

+e e .

Now we check the validity of Fubinis Theorem by using the order dx dy:

(
)
(
)
y
xy + ey dx dy
xy + e dA =
R
] )
([

y
=
x y + xe dy
)
(
=
y + ey dy
=
=

)

y + e
y

+e e .

Both orders of integra on return the same result, as expected.

Notes:

Double Integra on and Volume

Example (
Evalua ng a)double integral
Evaluate R xy x y + dA, where R is the triangle bounded by x = ,
y = and x/ + y = , as shown in Figure . .
While it is not specied which order we are to use, we will
S
evaluate the double integral using both orders to help drive home the point that
it does not ma er which order we use.
Using the order dy dx: The bounds on y go from curve to curve, i.e.,
y x/ , and the bounds on x go from point to point, i.e., x .
x+

)
)
( xy x y + dA =
( xy x y + dy dx
R

=
=
=

x x

Now lets consider the order dx dy. Here x goes from curve to curve,
y, and y goes from point to point, y :
y

)
)
( xy x y + dx dy
( xy x y + dA =
R

=
=
=
=

Figure . : Finding the signed volume


under the surface in Example
.

)

x

= . .

=
x

) x +

dx
xy x y y + y
(
)
dx
x x x
(

) y

x y x xy + x
dy
(
)
y y + y+
dy
(

y
= . .

y +y +

)

y

We obtained the same result using both orders of integra on.


Note how in these two examples that the bounds of integra on depend only
on R; the bounds of integra on have nothing to do with f(x, y). This is an important concept, so we include it as a Key Idea.

Notes:

Chapter

Mul ple Integra on

Key Idea

Double Integra on Bounds

When evalua ng R f(x, y) dA using an iterated integral, the bounds of


integra on depend only on R. The surface f does not determine the
bounds of integra on.

Before doing another example, we give some proper es of double integrals.


Each should make sense if we view them in the context of nding signed volume
under a surface, over a region.
Theorem

R1
.

R2

Proper es of Double Integrals

Let f and g be con nuous func ons over a closed, bounded plane region
R, and let c be a constant.

f(x, y) dA.
c f(x, y) dA = c
.
R

. If f(x, y)

Figure
. : R is the union of two
nonoverlapping regions, R and R .

)
f(x, y) g(x, y) dA =
on R, then

f(x, y) dA

g(x, y) dA

f(x, y) dA .

. If f(x, y) g(x, y) on R, then

f(x, y) dA

g(x, y) dA.

. Let R be the union of two nonoverlapping regions, R = R


(see Figure . ). Then

f(x, y) dA =
f(x, y) dA +
f(x, y) dA.
R

Evalua ng a double integral


Example
Let f(x, y) = sin x cos y and R be the triangle with verces ( , ), ( , ) and
( , ) (see Figure . ). Evaluate the double integral R f(x, y) dA.

If we a empt to integrate using an iterated integral with the


S
order dy dx, note how there are two upper bounds on R meaning well need to
use two iterated integrals. We would need to split the triangle into two regions

Figure . : Finding the signed volume


under a surface in Example
.

Notes:

.
along the y-axis, then use Theorem
, part .
Instead, lets use the order dx dy. The curves bounding x are y
y; the bounds on y are y . This gives us:
y

sin x cos y dx dy
f(x, y) dA =

Double Integra on and Volume

=
=

)

cos x cos y

dy

(
)
cos y cos( y) + cos(y ) dy.

Recall that the cosine func on is an even func on; that is, cos x = cos(x).
Therefore, from the last integral above, we have cos(y ) = cos( y). Thus
the integrand simplies to , and we have

dy
f(x, y) dA =
R

= .

It turns out that over R, there is just as much volume above the x-y plane as below (look again at Figure . ), giving a nal signed volume of .
Evalua ng a double integral
Example
Evaluate R ( y) dA, where R is the region bounded by the parabolas y = x
and x = y, graphed in Figure . .
Graphing each curve can help us nd their points of interS
sec on. Solving analy cally, the second equa on tells us that y = x / . Subs tu ng this value in for y in the rst equa on gives us x / = x. Solving for
x:
x
x

x(x

= x
x=
)=
x= , .

Thus weve found analy cally what was easy to approximate graphically: the
regions intersect at ( , ) and ( , ), as shown in Figure . .
We now choose an order of integra on: dy dx or dx dy? Either order works;
since the integrand does not contain x, choosing dx dy might be simpler at
least, the rst integral is very simple.

Notes:

Figure . : Finding the volume under


the surface in Example
.

Chapter

Mul ple Integra on

Thus we have the following curve to curve, point to point bounds: y /

y, and y .

( y) dA =
=
=
=
=

y /

( y) dx dy

(
)
x( y)
y

((

y
y /

dy

y )(
y

)
y) dy =
+

The signed volume under the surface f is about

(y

y y

+ y

. cubic units.

In the previous sec on we prac ced changing the order of integra on of a


given iterated integral, where the region R was not explicitly given. Changing
the bounds of an integral is more than just an test of understanding. Rather,
there are cases where integra ng in one order is really hard, if not impossible,
whereas integra ng with the other order is feasible.
Changingtheorder of integra on
ex dx dy with the order dy dx. Comment
Rewrite the iterated integral

Example

on the feasibility to evaluate each integral.

y=

R
x

.
Figure . : Determining the region R
determined by the bounds of integra on
in Example
.

Once again we make a sketch of the region over which we


S
are integra ng to facilitate changing the order. The bounds on x are from x = y
to x = ; the bounds on y are from y = to y = . These curves are sketched
in Figure . , enclosing the region R.
To change the bounds, note that the curves bounding y are y = up to
y = x; the triangle is enclosed between x = and x = . Thus the new
bounds ofintegra on are y x and x , giving the iterated integral

ex dy dx.

How easy is it to evaluate each iterated integral? Consider the order of integra ng dx dy, as given
in the original problem. The rst indenite integral we

need to evaluate is ex dx; we have stated before (see Sec on . ) that this
integral cannot be evaluated in terms of elementary func ons. We are stuck.
Changing the order of integra on makes a big dierence here. In the second

Notes:

dy

Double Integra on and Volume

iterated integral, we are faced with ex dy; integra ng with respect to y gives
us yex + C, and the rst denite integral evaluates to
x
ex dy = xex .
Thus

dy dx =

)
xex dx.

This last integral is easy to evaluate with subs tu on, giving a nal answer of
( e ) . . Figure . shows the surface over R.
In short, evalua ng one iterated integral is impossible; the other iterated integral is rela vely simple.
Deni on denes the average value of a singlevariable func on f(x) on
the interval [a, b] as
b
f(x) dx;
average value of f(x) on [a, b] =
ba a
that is, it is the area under f over an interval divided by the length of the interval. We make an analogous statement here: the average value of z = f(x, y)
over a region R is the volume under f over R divided by the area of R.
The Average Value of f on R

Deni on

Let z = f(x, y) be a con nuous func on dened over a closed region R


in the x-y plane. The average value of f on R is

f(x, y) dA
R

average value of f on R =
.
dA
R

Finding average value of a func on over a region R


Example
Find the average value of f(x, y) = y over the region R, which is bounded by
the parabolas y = x and x = y. Note: this is the same func on and region
as used in Example
.
In Example

Notes:

f(x, y) dA =

we found
y
( y) dx dy =
y /

Figure . : Showing the surface f dened in Example


over its region R.

Chapter

Mul ple Integra on


We nd the area of R by compu ng

dA =

dA:

y /

dx dy =

Dividing the volume under the surface by the area gives the average value:
average value of f on R =
While the surface, as shown in Figure
the average z-value on R is . .

Figure . : Finding the average value of


f in Example
.

/
/

= . .

, covers z-values from z = to z = ,

The previous sec on introduced the iterated integral in the context of nding the area of plane regions. This sec on has extended our understanding of
iterated integrals; now we see they can be used to nd the signed volume under
a surface.
This new understanding allows us to revisit what
we did in the previous secon. Given a region R in the plane, we computed R dA; again, our understanding at the me was that we were nding the area of R. However, we can
now view the func on
z = as a surface, a at surface with constant z-value of
. The double integral R dA nds the volume, under z = , over R, as shown
in Figure . . Basic geometry tells us that if the base of a general right cylinder
has area A, its volume is A h, where h is the height. In our case, the height is
. We were actually compu ng the volume of a solid, though we interpreted
the number as an area.
The next sec on extends our abili es to nd volumes under surfaces. Currently, some integrals are hard to compute because either the region R we are
integra ng over is hard to dene with rectangular curves, or the integrand itself is hard to deal with. Some of these problems can be solved by conver ng
everything into polar coordinates.

Figure . : Showing how an iterated integral used to nd area also nds a certain
volume.

Notes:

Exercises

Terms and Concepts


. An integral can be interpreted as giving the signed area over
an interval; a double integral can be interpreted as giving
the signed
over a region.
. Explain why the following statement is false: Fu b g (x)
binis Theorem states that
f(x, y) dy dx =
a
g (x)
b g (y)
f(x, y) dx dy.
a

g (y)

. Explain why if f(x, y) >

f(x, y) dA > .
R
. If R f(x, y) dA =
g(x, y)?

x
+
y

dx dy

(sin x cos y) dx dy

x y+

dy dx

)
(
x y xy dx dy

y/

In Exercises

(x + y + ) dx dy

(b) Set up the iterated integrals, in both orders, that evaluate the given double integral for the described region
R.
(c) Evaluate one of the iterated integrals to nd the signed
volume under the surface z = f(x, y) over the region
R.

x and y = x .

x y dA, where R is the rectangle with corners

yex dA, where R is bounded by x =

, x = y and

y= .
.

)
x y dA, where R is bounded by x = , y =

and x + y = .

y=

ey dA, where R is bounded by y = ln x and


(x ).

)
x yx dA, where R is the half of the circle x +y =

x y dA, where R is bounded by y =

)
y dA, where R is bounded by y = , y = x/e

and y = ln x.

In Exercises , state why it is dicult/impossible to integrate the iterated integral in the given order of integra on.
Change the order of integra on and evaluate the new iterated integral.

/ /

y/

ex dx dy

( )
cos y dy dx

y
dx dy
x +y
x tan y
dy dx
+ ln y

In Exercises
, nd the average value of f over the region R. No ce how these func ons and regions are related to
the iterated integrals given in Exercises .

( )
xy dx dy

(a) Sketch the region R given by the problem.

in the rst and second quadrants.

x/ +

x y dA, where R is bounded by y =

( , ), ( , ), ( , ) and ( , ).

g(x, y) dA, does this imply f(x, y) =

(b) rewrite the integral using the other order of integraon.


.

(a) Evaluate the given iterated integral, and

In Exercises

over a region R, then

Problems

x and y = x .

x
+ ; R is the rectangle with opposite corners
y
( , ) and ( , ).

. f(x, y) =

. f(x, y) = sin x cos y; R is bounded by x =


y = / and y = / .

, x = ,

. f(x, y) = x y + ; R is bounded by the lines y = ,


y = x/ and x = .
. f(x, y) = x y xy ;
x= .

R is bounded by y = x, y =

and

Chapter

Mul ple Integra on

.
/

.5

.5

}|r2

(a)

{z

r1

(b)

Figure . : Approxima ng a region R


with por ons of sectors of circles.

Double Integra on with Polar Coordinates

We have used iterated integrals to evaluate double integrals, which give the
signed volume under a surface, z = f(x, y), over a region R of the x-y plane.
The integrand is simply f(x, y), and the bounds of the integrals are determined
by the region R.
Some regions R are easy to describe using rectangular coordinates that is,
with equa ons of the form y = f(x), x = a, etc. However, some regions are
easier to handle if we represent their boundaries with polar equa ons of the
form r = f(), = , etc.

The basic form of the double integral is R f(x, y) dA. We interpret this integral as follows: over the region R, sum up lots of products of heights (given by
f(xi , yi )) and areas (given by Ai ). That is, dA represents a li le bit of area. In
rectangular coordinates, we can describe a small rectangle as having area dx dy
or dy dx the area of a rectangle is simply lengthwidth a small change in x
mes a small change in y. Thus we replace dA in the double integral with dx dy
or dy dx.
Now consider represen ng a region R with polar coordinates. Consider Figure . (a). Let R be the region in the rst quadrant bounded by the curve.
We can approximate this region using the natural shape of polar coordinates:
por ons of sectors of circles. In the gure, one such region is shaded, shown
again in part (b) of the gure.
As the area of a sector of a circle with radius r, subtended by an angle , is
A = r , we can nd the area of the shaded region. The whole sector has area
r , whereas the smaller, unshaded sector has area r . The area of the
shaded region is the dierence of these areas:
Ai =

r r =

r r

)(

) r +r (
)
=
r r .

Note that (r + r )/ is just the average of the two radii.


To approximate the region R, we use many such subregions; doing so shrinks
the dierence r r between radii to and shrinks the change in angle also
to . We represent these innitesimal changes in radius and angle as dr and d,
respec vely. Finally, as dr is small, r r , and so (r + r )/ r . Thus, when
dr and d are small,
Ai ri dr d.
Taking a limit, where the number of subregions goes to innity and both
r r and go to , we get
dA = r dr d.

So to evaluate R f(x, y) dA, replace dA with r dr d. Convert the func on


z = f(x, y) to a func on with polar coordinates with the subs tu ons x = r cos ,

Notes:

Double Integra on with Polar Coordinates

y = r sin . Finally, nd bounds g () r g () and that describe


R. This is the key principle of this sec on, so we restate it here as a Key Idea.
Key Idea

Evalua ng Double Integrals with Polar Coordinates

Let R be a plane region bounded by the polar equa ons and


g () r g (). Then

f(x, y) dA =

g ()
g ()

(
)
f r cos , r sin r dr d.

Examples will help us understand this Key Idea.


Evalua ng a double integral with polar coordinates
Example
Find the signed volume under the plane z = x y over the circle with
equa on x + y = .
The bounds of the integral are determined solely by the reS
gion R over which we are integra ng. In this case, it is a circle with equa on
x + y = . We need to nd polar bounds for this region. It may help to review
Sec on . ; bounds for this circle are r and .
We replace f(x, y) with f(r cos , r sin ). That means we make the following
subs tu ons:
x y
r cos r sin .

Finally, we replace dA in the double integral with r dr d. This gives the nal
iterated integral, which we evaluate:

(
)
r cos r sin r dr d
f(x, y) dA =
R

)
r r (cos sin ) dr d

sin + cos

)

r r (cos sin ) d
)
(
(
)
cos sin
d

The surface and region R are shown in Figure

Notes:

)
)

Figure . : Evalua ng a double integral


with polar coordinates in Example
.

Chapter

Mul ple Integra on

Evalua ng a double integral with polar coordinates


Example
Find the volume under the paraboloid z = (x ) y over the region
bounded by the circles (x ) + y = and (x ) + y = .
x

(a)

At rst glance, this seems like a very hard volume to comS


pute as the region R (shown in Figure . (a)) has a hole in it, cu ng out a
strange por on of the surface, as shown in part (b) of the gure. However, by
describing R in terms of polar equa ons, the volume is not very dicult to compute. It is straigh orward to show that the circle (x ) + y = has polar
equa on r = cos , and that the circle (x ) + y = has polar equa on
r = cos . Each of these circles is traced out on the interval . The
bounds on r are cos r cos .
Replacing x with r cos in the integrand,along with replacing y with r sin ,
prepares us to evaluate the double integral R f(x, y) dA:

f(x, y) dA =
=

cos

cos
cos
cos

(
)
(
) )
r cos
r sin r dr d

)
r + r cos dr d

) cos

d
r + r cos
cos
]
([
(
=
cos ) + ( cos )
[
])
( cos ) + ( cos )
d

cos d.
=
=

(b)
Figure . : Showing the region R and
surface used in Example
.

To integrate cos , rewrite it as cos cos and employ the power-reducing


formula twice:
cos = cos cos
(
) (
)
+ cos( )
+ cos( )
=
=

=
=

Notes:

+ cos( ) + cos ( )
+ cos( ) +
cos( ) +

+ cos( )

cos( ).

))

.
Picking up from where we le

=

=
(
=
=

Double Integra on with Polar Coordinates

o above, we have
cos d
(
+ cos( ) +
+

sin( ) +

cos( ) d
)

sin( )

While this example was not trivial, the double integral would have been much
harder to evaluate had we used rectangular coordinates.
Example

Evalua ng a double integral with polar coordinates

Find the volume under the surface f(x, y) =

over the sector of the


x +y +
circle with radius a centered at the origin in the rst quadrant, as shown in Figure
. .
The region R we are integra ng over is a circle with radius
S
a, restricted to the rst quadrant. Thus, in polar, the bounds on R are r a,
/ . The integrand is rewri en in polar as
x +y +

r cos + r sin +

We nd the volume as follows:

f(x, y) dA =

=
=
=
=

r
r +

r +

dr d

) a
ln |r + | d

ln(a + ) d

) /

ln(a + )

ln(a + ).

Figure . shows that f shrinks to near very quickly. Regardless, as a grows,


so does the volume, without bound.

Notes:

Figure . : The surface and region R


used in Example
.

Note: Previous work has shown that


there is nite area under x + over the
en re x-axis. However, Example
shows that there is innite volume under
over the en re x-y plane.
x +y +

Chapter

Mul ple Integra on


Example
Finding the volume of a sphere
Find the volume of a sphere with radius a.
The sphere of radius a,
centered at the origin, has equa on
S
x +y +z = a ; solving for z, we have z = a x y . This gives the upper
half of a sphere. We wish to nd the volume under this top half, then double it
to nd the total volume.
The region we need to integrate over is the circle of radius a, centered at the
origin. Polar bounds for this equa on are r a, .
All together, the volume of a sphere with radius a is:
a

a x y dA =
a (r cos ) (r sin ) r dr d
R

r a r dr d.

We can evaluate this inner integral with subs tu on. With u = a r , du =


r dr. The new bounds of integra on are u( ) = a to u(a) = . Thus we
have:

)
(
u / du d
=
a

=
=
=
=

(
(

u
a

)

a

)

d

a

a .

Generally, the formula for the volume of a sphere with radius r is given as / r ;
we have jus ed this formula with our calcula on.

Figure .
Example

: Visualizing the solid used in


.

Finding the volume of a solid


Example
A sculptor wants to make a solid bronze cast of the solid shown in Figure . ,
where the base of the solid has boundary, in polar coordinates, r = cos( ),
and the top is dened by the plane z = x + . y. Find the volume of the
solid.
From the outset, we should recognize that knowing how to
S
set up this problem is probably more important than knowing how to compute

Notes:

Double Integra on with Polar Coordinates

the integrals. The iterated integral to come is not hard to evaluate, though it is
long, requiring lots of algebra. Once the proper iterated integral is determined,
one can use readilyavailable technology to help compute the nal answer.
The region R that we are integra ng over is bound by r cos( ),
for (note that this rose curve is traced out on the interval [ , ], not
[ , ]). This gives us our bounds of integra on. The integrand is z = x+ . y;
conver ng to polar, we have that the volume V is:
V=

f(x, y) dA =

cos( )

)
r cos + . r sin r dr d.

Distribu ng the r, the inner integral is easy to evaluate, leading to


)
(
.
cos ( ) cos ( ) cos +
cos ( ) sin d.

This integral takes me to compute by hand; it is rather long and cumbersome.


The powers of cosine need to be reduced, and products like cos( ) cos need
to be turned to sums using the Product To Sum formulas in the back cover of
this text.
We rewrite cos ( ) as ( +cos( )). We can also rewrite cos ( ) cos
as:
cos ( ) cos =

cos ( ) cos( ) cos =

+ cos( ) (

)
cos( )+cos( ) .

This last expression s ll needs simplica on, but eventually all terms can be reduced to the form a cos(m) or a sin(m) for various values of a and m.
We forgo the algebra and recommend the reader employ technology, such
as WolframAlpha, to compute the numeric answer. Such technology gives:

cos( )

r cos + . r sin r dr d = .

u .

Since the units were not specied, we leave the result as almost . cubic units
(meters, feet, etc.) Should the ar st want to scale the piece uniformly, so that
each rose petal had a length other than , she should keep in mind that scaling
by a factor of k scales the volume by a factor of k .
We have used iterated integrals to nd areas of plane regions and volumes
under surfaces. Just as a single integral can be used to compute much more than
area under the curve, iterated integrals can be used to compute much more
than we have thus far seen. The next two sec ons show two, among many,
applica ons of iterated integrals.

Notes:

Exercises

Terms and Concepts

. When evalua ng R f(x, y) dA using polar coordinates,


and dA is replaced with
f(x, y) is replaced with
.
. Why would one be interested in evalua ng a double integral with polar coordinates?

In Exercises
, an iterated integral in rectangular coordinates is given. Rewrite the integral using polar coordinates
and evaluate the new double integral.
.

. f(x, y) = x y + ; R is the region enclosed by the circle


x +y = .
. f(x, y) = x + y; R is the region enclosed by the circle
x +y = .

Problems
In Exercises , a func on f(x, y) is given and
a region R of
the x-y plane is described. Set up and evaluate R f(x, y) dA
using polar coordinates.

x + y dy dx

)
y x dx dy

)
x + y dx dy
(

x+

dy dx +

(
)
x + dy dx

(
)
x + dy dx +

Hint: draw the region of each integral carefully and see how
they all connect.
In Exercises
, special double integrals are presented
that are especially well suited for evalua on in polar coordinates.

e(x +y ) dA.
. Consider
R

. f(x, y) = y; R is the region enclosed by the circles with


polar equa ons r = cos and r = cos .
. f(x, y) = ; R is the region enclosed by the petal of the rose
curve r = sin( ) in the rst quadrant.
(
. f(x, y) = ln x + y ); R is the annulus enclosed by the circles x + y = and x + y = .
. f(x, y) = x y ; R is the region enclosed by the circle
x +y = .
. f(x, y) = x y ; R is the region enclosed by the circle
x +y =
in the rst and fourth quadrants.

(a) Why is this integral dicult to evaluate in rectangular


coordinates, regardless of the region R?

(b) Let R be the region bounded by the circle of radius a


centered at the origin. Evaluate the double integral
using polar coordinates.
(c) Take the limit of your answer from (b), as a .
What does this imply about the volume under the
surface of e(x +y ) over the en re x-y plane?
. The surface of a right circular cone with height h and
base radius
a can be described by the equa on f(x, y) =

y
x
hh
+ , where the p of the cone lies at ( , , h)
a
a
and the circular base lies in the x-y plane, centered at the
origin.
Conrm that the volume of a right circular cone with

. f(x, y) = (x y)/(x + y); R is the region enclosed by the


lines y = x, y = and the circle x + y = in the rst
quadrant.

height h and base radius a is V =

f(x, y) dA in polar coordinates.


R

a h by evalua ng

Center of Mass

We have used iterated integrals to nd areas of plane regions and signed volumes under surfaces. A brief recap of these uses will be useful in this sec on as
we apply iterated integrals to compute the mass and center of mass of planar
regions.

To nd the area of a planar region, we evaluated the double integral R dA.


That is, summing up the areas
of lots of li le subregions of R gave us the total
area. Informally, we think of R dA as meaning sum up lots of li le areas over
R.
To nd the signed volume under a surface, we evaluated the double integral
f(x, y) dA. Recall that the dA is not just a bookend at the end of an inR
tegral; rather, it is mul plied by f(x, y). We regard f(x, y) as giving
a height, and
dA s ll giving an area: f(x, y) dA gives a volume. Thus, informally, R f(x, y) dA
means sum up lots of li le volumes over R.
We now extend these ideas to other contexts.

Mass and Weight

(a)

Consider a thin sheet of material with constant thickness and nite area.
Mathema cians (and physicists and engineers) call such a sheet a lamina. So
consider a lamina, as shown in Figure . (a), with the shape of some planar
region R, as shown in part (b).
We
can write a simple double integral that represents the mass of the lamina: R dm, where dm means a li le mass. That is, the double integral
states the total mass of the lamina can be found by summing up lots of li le
masses over R.
To evaluate this double integral, par on R into n subregions as we have
done in the past. The i th subregion has area Ai . A fundamental property of
mass is that mass=densityarea. If the lamina has a constant density , then
the mass of this i th subregion is mi = Ai . That is, we can compute a small
amount of mass by mul plying a small amount of area by the density.
If density is variable, with density func on = (x, y), then we can approximate the mass of the i th subregion of R by mul plying Ai by (xi , yi ), where
(xi , yi ) is a point in that subregion. That is, for a small enough subregion of R,
the density across that region is almost constant.
The total mass M of the lamina is approximately the sum of approximate
masses of subregions:
M

Notes:

Center of Mass

i=

mi =

i=

(xi , yi )Ai .

y = f (x)

y=

)
f (x

(b)

Figure .
lamina.

: Illustra ng the concept of a

Note: Mass and weight are dierent


measures. Since they are scalar mul ples of each other, it is o en easy to
treat them as the same measure. In this
sec on we eec vely treat them as the
same, as our technique for nding mass is
the same as for nding weight. The density func ons used will simply have dierent units.

Chapter

Mul ple Integra on


Taking the limit as the size of the subregions shrinks to gives us the actual
mass; that is, integra ng (x, y) over R gives the mass of the lamina.

Deni on

Mass of a Lamina with Vairable Density

Let (x, y) be a con nuous density func on of a lamina corresponding to


a plane region R. The mass M of the lamina is

(x, y) dA.
dm =
mass M =
R

Finding the mass of a lamina with constant density


Example
Find the mass of a square lamina, with side length , with a density of =
gm/cm .
y

We represent the lamina with a square region in the plane


S
as shown in Figure . . As the density is constant, it does not ma er where
we place the square.

Following Deni on

, the mass M of the lamina is

0.5

M=
.

0.5

Figure . : A region R represen ng a


lamina in Example
.

dA =

dx dy =

dx dy = gm.

This is all very straigh orward; note that all we really did was nd the area
of the lamina and mul ply it by the constant density of gm/cm .
Finding the mass of a lamina with variable density
Example
Find the mass of a square lamina, represented by the unit square with lower
le hand corner at the origin (see Figure . ), with variable density (x, y) =
(x + y + )gm/cm .
The variable density , in this example, is very uniform, givS
ing a density of in the center of the square and changing linearly. A graph
of (x, y) can be seen in Figure . ; no ce how same amount of density is
above z = as below. Well comment on the signicance of this momentarily.
The mass M is found by integra ng (x, y) over R. The order of integra on

Notes:

Center of Mass

is not important; we choose dx dy arbitrarily. Thus:


M=

(x + y + ) dA =
=
=
=

(x + y + ) dx dy

)

x + x(y + ) dy
)
(
+ y dy
)

y + y

= gm.

It turns out that since since the density of the lamina is so uniformly distributed
above and below z = that the mass of the lamina is the same as if it had
a constant density of . The density func ons in Examples
and
are
graphed in Figure . , which illustrates this concept.
Finding the weight of a lamina with variable density
Example
Find the weight of the lamina represented by the circle with radius , centered
at the origin, with density func on (x, y) = (x + y + )lb/
. Compare this
to the weight of the same lamina with density (x, y) = (
x + y + )lb/ .

states that the weight


S
A direct applica on of Deni on
of the lamina is R (x, y) dA. Since our lamina is in the shape of a circle, it
makes sense to approach the double integral using polar coordinates.
The density func on (x, y) = x + y + becomes (r, ) = (r cos ) +
(r sin ) + = r + . The circle is bounded by r and .
Thus the weight W is:
W=
=
=
=

(r + )r dr d
r + r

( ) d

)

d

lb.

Now compare this with the density func on (x,


x + y + . Con y) =
ver ng this to polar coordinates gives (r, ) =
(r cos ) + (r sin ) + =
Notes:

Figure . : Graphing the density funcons in Examples


and
.

Chapter

Mul ple Integra on


r + . Thus the weight W is:
W=

( r + )r dr d



( r + r ) d
( )
=
d
=

lb.

One would expect dierent density func ons to return dierent weights, as we
have here. The density func ons were chosen, though, to be similar: each gives
a density of at the origin and a density of at the outside edge of the circle,
as seen in Figure . .

(a)

(b)

Figure . : Graphing the density func ons in Example


. In (a) is the density

x +y + .
func on (x, y) = x + y + ; in (b) is (x, y) =

No ce how x + y +
weight.

x + y + over the circle; this results in less

Plo ng the density func ons can be useful as our understanding of mass
can be
related to our understanding of volume under a surface. We interpreted R f(x, y) dA as giving the volume under f over R; we can understand

(x, y) dA in the same way. The volume under over R is actually mass;
R
Notes:

.
by compressing the volume under onto the x-y plane, we get more mass
in some areas than others i.e., areas of greater density.
Knowing the mass of a lamina is one of several important measures. Another
is the center of mass, which we discuss next.

Center of Mass
Consider a disk of radius with uniform density. It is common knowledge
that the disk will balance on a point if the point is placed at the center of the
disk. What if the disk does not have a uniform density? Through trial-and-error,
we should s ll be able to nd a spot on the disk at which the disk will balance
on a point. This balance point is referred to as the center of mass, or center of
gravity. It is though all the mass is centered there. In fact, if the disk has a
mass of kg, the disk will behave physically as though it were a point-mass of
kg located at its center of mass. For instance, the disk will naturally spin with
an axis through its center of mass (which is why it is important to balance the
res of your car: if they are out of balance, their center of mass will be outside
of the axle and it will shake terribly).
We nd the center of mass based on the principle of a weighted average.
Consider a college class in which your homework average is %, your test average is %, and your nal exam grade is an %. Experience tells us that our
nal grade is not the average of these three grades: that is, it is not:
. + .

+ .

. %.

That is, you are probably not pulling a B in the course. Rather, your grades are
weighted. Lets say the homework is worth % of the grade, tests are % and
the exam is %. Then your nal grade is:
( . )( . ) + ( . )( .

) + ( . )( .

)= .

. %.

Each grade is mul plied by a weight.


In general, given values x , x , . . . , xn and weights w , w , . . . , wn , the weighted
average of the n values is
/ n
n

wi .
wi xi
i=

i=

In the grading example above, the sum of the weights . , . and . is ,


so we dont see the division by the sum of weights in that instance.
How this relates to center of mass is given in the following theorem.

Notes:

Center of Mass

Chapter

Mul ple Integra on

Theorem

Center of Mass of Discrete Linear System

Let point masses m , m , . . . , mn be distributed along the x-axis at locaons x , x , . . . , xn , respec vely. The center of mass x of the system is
located at
/ n
n

mi .
mi xi
x=
i=

Example

i=

Finding the center of mass of a discrete linear system

. Point masses of gm are located at x = , x = and x = are connected by a thin rod of negligible weight. Find the center of mass of the
system.
. Point masses of gm, gm and gm are located at x = , x = and
x = , respec vely, are connected by a thin rod of negligible weight. Find
the center of mass of the system.
S
x

. Following Theorem

x=

(a)

, we compute the center of mass as:


( ) + ( ) + ( )
=
+ +

= . .

So the system would balance on a point placed at x = / , as illustrated


in Figure . (a).

.
x

. Again following Theorem

, we nd:

x=

(b)

( ) + ( ) + ( )
=
.
+ +

Placing a large weight at the le hand side of the system moves the center
of mass le , as shown in Figure . (b).

Figure . : Illustra ng point masses


along a thin rod and the center of mass.

In a discrete system (i.e., mass is located at individual points, not along a


con nuum) we nd the center of mass by dividing the mass into a moment of
the system. In general, a moment is a weighted measure of distance from a parcular point or line. In the case described by Theorem
, we are nding a
weighted measure of distances from the y-axis, so we refer to this as the moment about the y-axis, represented by My . Le ng M be the total mass of the
system, we have x = My /M.

Notes:

Center of Mass

We can extend the concept of the center of mass of discrete points along a
line to the center of mass of discrete points in the plane rather easily. To do so,
we dene some terms then give a theorem.
Deni on

Moments about the x- and y- Axes.

Let point masses m , m , . . . , mn be located at points (x , y ),


(x , y ) . . . , (xn , yn ), respec vely, in the x-y plane.
. The moment about the y-axis, My , is My =

mi xi .

i=

. The moment about the x-axis, Mx , is Mx =

mi yi .

i=

One can think that these deni ons are backwards as My sums up x distances. But remember, x distances are measurements of distance from the
y-axis, hence dening the moment about the y-axis.
We now dene the center of mass of discrete points in the plane.
Theorem

Center of Mass of Discrete Planar System

Let point masses m , m , . . . , mn be located at points (x , y ),


n

mi .
(x , y ) . . . , (xn , yn ), respec vely, in the x-y plane, and let M =
The center of mass of the system is at (x, y), where
x=

My
M

and

y=

i=

Mx
.
M

Finding the center of mass of a discrete planar system


Example
Let point masses of kg, kg and kg be located at points ( , ), ( , ) and ( , ),
respec vely, and are connected by thin rods of negligible weight. Find the center
of mass of the system.
We follow Theorem

and My :
M=

Notes:

+ +

= kg.

and Deni on

(x, y)

to nd M, Mx

.
Figure . : Illustra ng the center of
mass of a discrete planar system in Example
.

Chapter

Mul ple Integra on

Mx =

mi yi

My =

mi xi

i=

i=

= ( )+ ( )+ ( )
= .

= ( )+ ( )+ ( )
= .
Thus the center of mass is (x, y) =
illustrated in Figure

My Mx
,
M M

=( .

, .

We nally arrive at our true goal of this sec on: nding the center of mass of
a lamina with variable density. While the above measurement of center of mass
is interes ng, it does not directly answer more realis c situa ons where we need
to nd the center of mass of a con guous region. However, understanding the
discrete case allows us to approximate the center of mass of a planar lamina;
using calculus, we can rene the approxima on to an exact value.
We begin by represen ng a planar lamina with a region R in the x-y plane
with density func on (x, y). Par on R into n subdivisions, each with area
Ai . As done before, we can approximate the mass of the i th subregion with
(xi , yi )Ai , where (xi , yi ) is a point inside the i th subregion. We can approximate the moment of this subregion about the y-axis with xi (xi , yi )Ai that is,
by mul plying the approximate mass of the region by its approximate distance
from the y-axis. Similarly, we can approximate the moment about the x-axis
with yi (xi , yi )Ai . By summing over all subregions, we have:

mass: M
moment about the x-axis: Mx
moment about the y-axis: My

i=
n

i=
n

(xi , yi )Ai

(as seen before)

yi (xi , yi )Ai
xi (xi , yi )Ai

i=

By taking limits, where size of each subregion shrinks to in both the x and
y direc ons, we arrive at the double integrals given in the following theorem.

Notes:

),

Theorem

Center of Mass

Center of Mass of a Planar Lamina, Moments

Let a planar lamina be represented by a region R in the x-y plane with


density func on (x, y).

(x, y) dA
. mass: M =
R

. moment about the x-axis: Mx =


. moment about the y-axis: My =

y(x, y) dA
x(x, y) dA

. The center of mass of the lamina is


)
(
My Mx
(x, y) =
.
,
M M

We start our prac ce of nding centers of mass by revisi ng some of the


lamina used previously in this sec on when nding mass. We will just set up the
integrals needed to compute M, Mx and My and leave the details of the integraon to the reader.
y

Finding the center of mass of a lamina


Example
Find the center mass of a square lamina, with side length , with a density of
= gm/cm . (Note: this is the lamina from Example
.)
We represent the lamina with a square region in the plane
as shown in Figure . as done previously.
Following Theorem
, we nd M, Mx and My :
S

M=
Mx =
My =

dA =

y dA =
x dA =

Thus the center of mass is (x, y) =

Notes:

dx dy = gm.
y dx dy = / = . .
x dx dy = / = . .

My Mx
,
M M

= ( . / , . / ) = ( . , . ).

0.5

0.5

Figure . : A region R represen ng a


lamina in Example
.

Chapter

Mul ple Integra on


This is what we should have expected: the center of mass of a square with constant density is the center of the square.
Example
Finding the center of mass of a lamina
Find the center of mass of a square lamina, represented by the unit square
with lower le hand corner at the origin (see Figure . ), with variable density (x, y) = (x + y + )gm/cm . (Note: this is the lamina from Example
.)
, to nd M, Mx and My :

We follow Theorem

M=
Mx =
My =

(x + y + ) dA =

y(x + y + ) dA =
x(x + y + ) dA =

(x + y + ) dx dy = gm.

y(x + y + ) dx dy =

x(x + y + ) dx dy =

) (
)
My Mx
=
( .
, .
).
,
,
Thus the center of mass is (x, y) =
M M
While the mass of this lamina is the same as the lamina in the previous example,
the greater density found with greater x and y values pulls the center of mass
from the center slightly towards the upper righthand corner.
(

Finding the center of mass of a lamina


Example
Find the center of mass of the lamina represented by the circle with radius ,
centered at the origin, with density func on (x, y) = (x + y + )lb/ . (Note:
this is one of the lamina used in Example
.)
As done in Example
, it is best to describe R using polar
S
coordinates. Thus when
we
compute
M
,
we
will(integrate
y
(
)
)( not x(x,
) y) = x(x +
y + ), but rather r cos (r cos , r sin ) = r cos r + . We compute
M, Mx and My :
M=
Mx =
My =

Notes:

(r + )r dr d =

. lb.

(r sin )(r + )r dr d = .
(r cos )(r + )r dr d = .

Center of Mass

Since R and the density of R are both symmetric about the x and y axes, it should
come as no big surprise that the moments about each axis is . Thus the center
of mass is (x, y) = ( , ).
Finding the center of mass of a lamina
Example
Find the center of mass of the lamina represented by the region R shown in Figure . , half an annulus with outer radius and inner radius , with constant
density lb/ .
Once again it will be useful to represent R in polar coorS
dinates. Using the descrip on of R and/or the illustra on, we see that R is
bounded by r and . As the lamina is symmetric about
the y-axis, we should expect My = . We compute M, Mx and My :
M=
Mx =
My =

( )r dr d =

(r cos )( )r dr d = .

(
)
Thus the center of mass is (x, y) = , ( , . ). The center of mass is
indicated in Figure . ; note how it lies outside of R!
This sec on has shown us another use for iterated integrals beyond nding
area or signed volume under the curve. While there are many uses for iterated
integrals, we give one more applica on in the following sec on: compu ng surface area.

Notes:

5
(x, y)

Figure .
Example

lb.

(r sin )( )r dr d =

: Illustra ng the region R in


.

Exercises

Terms and Concepts

. R is the annulus in the rst and second quadrants


bounded

by x + y = and x + y = ; (x, y) = x + y lb/

. Why is it easy to use mass and weight interchangeably,


even though they are dierent measures?
. Given a point (x, y), the value of x is a measure of distance
from the
-axis.
. We can think of

dm as meaning sum up lots of

. What is a discrete planar system?

. Why does Mx use R y(x, y) dA instead of R x(x, y) dA;


that is, why do we use y and not x?
. Describe a situa on where the center of mass of a lamina
does not lie within the region of the lamina itself.

Problems
In Exercises , point masses are given along a line or in
the plane. Find the center of mass x or (x, y), as appropriate.
(All masses are in grams and distances are in cm.)
. m =

at x = ;

m =

at x = ;

. m =
m =

at x = ; m = at x = ;
at x = ; m = at x =

. m =
m =

at ( , );
at ( , )

. m =
m =

at ( , ); m = at ( , );
at ( , ); m = at ( , )

m =

m =

at x =

In Exercises , nd the center of mass of the lamina described by the region R in the plane and its density func on
(x, y).
Note: these are the same lamina as in Exercises .
. R is the rectangle with corners ( , ), ( , ), ( , ) and
( , ); (x, y) = gm/cm
. R is the rectangle with corners ( , ), ( , ), ( , ) and
( , ); (x, y) = (x + y )gm/cm
. R is the triangle with corners ( , ), ( , ), and ( , );
(x, y) = lb/in
. R is the triangle with corners ( , ), ( , ), and ( , );
(x, y) = (x + y + )lb/in
. R is the circle centered at the origin with radius ; (x, y) =
(x + y + )kg/m
. R is the circle sector bounded
by x + y =

quadrant; (x, y) = ( x + y + )kg/m

in the rst

. R is the annulus in the rst and second quadrants bounded


by x + y = and x + y = ; (x, y) = lb/
. R is the annulus in the rst and second quadrants
bounded

by x + y = and x + y = ; (x, y) = x + y lb/

at ( , );

In Exercises
, nd the mass/weight of the lamina described by the region R in the plane and its density func on
(x, y).
. R is the rectangle with corners ( , ), ( , ), ( , ) and
( , ); (x, y) = gm/cm
. R is the rectangle with corners ( , ), ( , ), ( , ) and
( , ); (x, y) = (x + y )gm/cm
. R is the triangle with corners ( , ), ( , ), and ( , );
(x, y) = lb/in

The moment of iner a I is a measure of the tendency of a lamina to resist rota ng about an axis or con nue to rotate about
an axis. Ix is the moment of iner a about the x-axis, Ix is the
moment of iner a about the x-axis, and IO is the moment of
iner a about the origin. These are computed as follows:

y dm
Ix =
R

Iy =

IO =

x dm
)
(
x + y dm

In Exercises
, a lamina corresponding to a planar region R is given with a mass of units. For each, compute Ix ,
Iy and IO .

. R is the triangle with corners ( , ), ( , ), and ( , );


(x, y) = (x + y + )lb/in

. R is the square with corners at ( , ) and ( , )


with density (x, y) = .

. R is the circle centered at the origin with radius ; (x, y) =


(x + y + )kg/m

. R is the rectangle with corners at ( , ) and ( , )


with density (x, y) = .

. R is the circle sector bounded


by x + y =

quadrant; (x, y) = ( x + y + )kg/m

in the rst

. R is the rectangle with corners at ( , ) and ( , )


with density (x, y) = .

. R is the annulus in the rst and second quadrants bounded


by x + y = and x + y = ; (x, y) = lb/

. R is the circle with radius centered at the origin with density (x, y) = /.

Surface Area

In Sec on . we used denite integrals to compute the arc length of plane


curves of the form y = f(x). We later extended these ideas to compute the
arc length of plane curves dened by parametric or polar equa ons.
The natural extension of the concept of arc length over an interval to surfaces is surface area over a region.
Consider the surface z = f(x, y) over a region R in the x-y plane, shown in
Figure . (a). Because of the domed shape of the surface, the surface area will
be greater than that of the area of the region R. We can nd this area using the
same basic technique we have used over and over: well make an approxima on,
then using limits, well rene the approxima on to the exact value.
As done to nd the volume under a surface or the mass of a lamina, we
subdivide R into n subregions. Here we subdivide R into rectangles, as shown in
the gure. One such subregion is outlined in the gure, where the rectangle has
dimensions xi and yi , along with its corresponding region on the surface.
In part (b) of the gure, we zoom in on this por on of the surface. When xi
and yi are small, the func on is approximated well by the tangent plane at any
point (xi , yi ) in this subregion, which is graphed in part (b). In fact, the tangent
plane approximates the func on so well that in this gure, it is virtually indisnguishable from the surface itself! Therefore we can approximate the surface
area Si of this region of the surface with the area Ti of the corresponding por on
of the tangent plane.
This por on of the tangent plane is a parallelogram, dened by sides u and
v, as shown. One of the applica ons of the cross product from Sec on . is
that the area of this parallelogram is || u v ||. Once we can determine u and v,
we can determine the area.
u is tangent to the surface in the direc on of x, therefore, from Sec on . ,
u is parallel to , , fx (xi , yi ). The x-displacement of u is xi , so we know that
u = xi , , fx (xi , yi ). Similar logic shows that v = yi , , fy (xi , yi ). Thus:
surface area Si area of Ti
= || u v ||


= xi , , fx (xi , yi ) yi , , fy (xi , yi )

+ fx (xi , yi ) + fy (xi , yi ) xi yi .
=

Note that xi yi = Ai , the area of the i th subregion.


Summing up all n of the approxima ons to the surface area gives
n

+ fx (xi , yi ) + fy (xi , yi ) Ai .
surface area over R
i=

Notes:

Surface Area

(a)

(b)
Figure . : Developing a method of
compu ng surface area.

Chapter

Mul ple Integra on


Once again take a limit as all of the xi and yi shrink to ; this leads to a
double integral.

Note:
as done before, we think of

R dS as meaning sum up lots of


li le surface areas over R.
The concept of surface area is dened
here, for while we already have a no on
of the area of a region in the plane, we
did not yet have a solid grasp of what the
area of a surface in space means.

Deni on

Surface Area

Let z = f(x, y) where fx and fy are con nuous over a closed, bounded
region R. The surface area S over R is

dS
S=
R
=
+ fx (x, y) + fy (x, y) dA.
R

We test this deni on by using it to compute surface areas of known surfaces. We start with a triangle.
Finding the surface area of a plane over a triangle
Example
Let f(x, y) = x y, and let R be the region in the plane bounded by x = ,
y = and y = x/ , as shown in Figure . . Find the surface area of f over
R.
We follow Deni on
. We start by no ng that fx (x, y) =
S
and fy (x, y) = . To dene R, we use bounds y x/ and
x . Therefore

dS
S=
R

=
=
=
Figure . : Finding the area of a triangle in space in Example
.

x/

(
.

+ ( ) + ( ) dy dx

x)

dx

Because the surface is a triangle, we can gure out the area using geometry.
Considering the base of the
triangle to be the side in the x-y plane, we nd the
. We can nd the height using our knowledge of
length of the base to be
vectors: let u be the side in the x-z plane
and let v be the side in the x-y plane.
/ . Geometry states that the area is
The height is then || u proj v u || =
thus

/
=
.

We arm the validity of our formula.

Notes:

Surface Area

It is common knowledge that the surface area of a sphere of radius r is


r . We conrm this in the following example, which involves using our formula with polar coordinates.
The surface area of a sphere.
Example
Find the surface area of the sphere withradius a centered at the origin, whose
top hemisphere has equa on f(x, y) = a x y .
We start by compu ng par al deriva ves and nd

fx (x, y) =

a x y

and

y
fy (x, y) =
.
a x y

As our func on f only denes the top upper hemisphere of the sphere, we double our surface area result to get the total area:

+ fx (x, y) + fy (x, y) dA
S=
R

x +y
+
=
dA.
a
x y
R
The region R that we are integra ng over is the circle, centered at the origin,
with radius a: x + y = a . Because of this region, we are likely to have greater
success with our integra on by conver ng to polar coordinates. Using the subs tu ons x = r cos , y = r sin , dA = r dr d and bounds and
r a, we have:

a
r cos + r sin
r dr d
+
S=
a r cos r sin
a
r
=
r
dr d
+
a r

a
a
dr d.
( . )
r
=
a r
Apply subs tu on u = a r and integrate the inner integral, giving
=

a d

= a .
Our work conrms our previous formula.

Notes:

Note: The inner integral in Equa on


( . ) is an improper
integral, as the
a
a
integrand of
r
dr is not dea r
ned at r = a. To properly evaluate this
integral, one must use the techniques of
Sec on . .
The reason this need
arises is that the
func on f(x, y) = a x y fails the
requirements of Deni on
, as fx and
fy are not con nuous on the boundary of
the circle x + y = a .
The computa on of the surface area is
s ll valid. The deni on makes stronger
requirements than necessary in part to
avoid the use of improper integra on, as
when fx and/or fy are not con nuous, the
resul ng improper integral may not converge. Since the improper integral does
converge in this example, the surface area
is accurately computed.

Chapter

Mul ple Integra on


Example
Finding the surface area of a cone
The general formula for a right cone with height h and base radius a is
h
f(x, y) = h
x +y ,
a
shown in Figure . . Find the surface area of this cone.
We begin by compu ng par al deriva ves.

xh
fx (x, y) =
a x +y

Figure . : Finding the surface area of


a cone in Example
.

and

yh

.
a x +y

Since we are integra ng over the circle x + y = a , we again use polar


coordinates. Using the standard subs tu ons, our integrand becomes

)
)
(
(
hr sin
hr cos

+
.
+
a r
a r
This may look in mida ng at rst, but there are lots of simple simplica ons to
be done. It amazingly reduces to just

h
=
+
a +h .
a
a

Our polar bounds are


Note: Note that once again fx and fy are
not con nuous on the domain of f, as
both are undened at ( , ). (A similar
problem occurred in the previous example.) Once again the resul ng improper
integral converges and the computa on
of the surface area is valid.

and r a. Thus
a
S=
a + h dr d
r
a
) a
(


a + h d
r
=
a

=
a a + h d

= a a + h .

This matches the formula found in the back of this text.

Finding surface area over a region


Example
Find the area of the surface f(x, y) = x y + over the region R bounded by
x y x, x , as pictured in Figure . .
It is straigh orward to compute fx (x, y) = x and fy (x, y) =
S
. Thus the surface area is described by the double integral


+ ( x) + ( ) dA =
+ x dA.
R

Figure
ample

: Graphing the surface in Ex-

Notes:

.
As with integrals describing arc length, double integrals describing surface area
are in general hard to evaluate directly because of the squareroot. This par cular integral can be easily evaluated, though, with judicious choice of our order
of integra on.

Integra ng with order dx dy requires us to evaluate


+ x dx. This can
be done, though it involves Integra
on By Parts and sinh x. Integra ng with
order dy dxhas as its rst integral
+ x dy, which is easy to evaluate: it
+ x + C. So we proceed with the order dy dx; the bounds are
is simply y
already given in the statement of the problem.

x
+ x dA =
+ x dy dx
R

=
=
Apply subs tu on with u =

(
y

(
x

) x
+ x dx
x

+ x

+ x :
=
=

(
(

+ x

dx.

)



)

u .

So while the region R over which we integrate has an area of u , the surface
has a much greater area as its z-values change drama cally over R.
In prac ce, technology helps greatly in the evalua on of such integrals. High
powered computer algebra systems can compute integrals that are dicult, or
at least me consuming, by hand, and can at the least produce very accurate approxima ons with numerical methods. In general, just knowing how to set up
the proper integrals brings one very close to being able to compute the needed
value. Most of the work is actually done in just describing the region R in terms
of polar or rectangular coordinates. Once this is done, technology can usually
provide a good answer.
We have learned how to integrate integrals; that is, we have learned to evaluate double integrals. In the next sec on, we learn how to integrate double integrals that is, we learn to evaluate triple integrals, along with learning some
uses for this opera on.

Notes:

Surface Area

Exercises

Terms and Concepts

. f(x, y) =

x +y +

R is the circle x + y = .

. Surface area is analogous to what previously studied concept?


. To approximate the area of a small por on of a surface, we
plane.
computed the area of its

. We interpret
.

dS as sum up lots of li le

. Why is it important to know how to set up a double integral to compute surface area, even if the resul ng integral
is hard to evaluate?

. f(x, y) = x y ; R is the rectangle with opposite corners


( , ) and ( , ).

. Why do z = f(x, y) and z = g(x, y) = f(x, y) + h, for some


real number h, have the same surface area over a region
R?
. Let z = f(x, y) and z = g(x, y) = f(x, y). Why is the surface area of g over a region R not twice the surface area of
f over R?

Problems

. f(x, y) =

In Exercises , set up the iterated integral that computes


the surface area of the given surface over the region R.

. f(x, y) = sin x cos y; R is the rectangle with bounds


x ,
y .

ex +
x and

R is the rectangle bounded by


y .

In Exercises
the region R.

, nd the area of the given surface over

. f(x, y) = x y + ; R is the rectangle with opposite corners ( , ) and ( , ).


. f(x, y) = x + y + ; R is the triangle with corners ( , ),
( , ) and ( , ).

. f(x, y) = x + y +

; R is the circle x + y =

. f(x, y) = x + y + over R, the triangle bounded by


y = x, y = x, y .
. f(x, y) = x + y over R, the triangle bounded by y = x,
y = and x = .
. f(x, y) = x / + y / over R, the rectangle with opposite
corners ( , ) and ( , ).

x + y over R, the circle x + y = .


. f(x, y) =

(This is the cone with height


and base radius ; be sure

to compare you result with the known formula.)


. Find the surface area of the spherewith radius by doubling the surface area of f(x, y) =
x y over R,
the circle x + y =
. (Be sure to compare you result
with the known formula.)
. Find the surface area of the ellipse formed by restric ng
the plane f(x, y) = cx + dy + h to the region R, the circle
x + y = , where c, d and h are some constants. Your
answer should be given in terms of c and d; why does the
value of h not ma er?

Chapter

Mul ple Integra on

Volume Between Surfaces and Triple Integra on

We learned in Sec on . how


to compute the signed volume V under a surface
z = f(x, y) over a region R: V = R f(x, y) dA. It follows naturally that if f(x, y)
g(x, y) on R, then the volume between f(x, y) and g(x, y) on R is

(
)
f(x, y) g(x, y) dA.
g(x, y) dA =
f(x, y) dA
V=
R

Theorem

Volume Between Surfaces

Let f and g be con nuous func ons on a closed, bounded region R, where
f(x, y) g(x, y) for all (x, y) in R. The volume V between f and g over R
is

(
)
V=
f(x, y) g(x, y) dA.
R

Finding volume between surfaces


Example
Find the volume of the space region bounded by the planes z = x + y and
z = x y in the st octant. In Figure . (a) the planes are drawn; in (b),
only the dened region is given.
We need to determine the region R over which we will inteS
grate. To do so, we need to determine where the planes intersect. They have
common z-values when x + y = x y. Applying a li le algebra, we
have:
(a)

x+y

x+ y=
x+y=

x y

The planes intersect along the line x+y = . Therefore the region R is bounded
by x = , y = , and y = x; we can convert these bounds to integra on
bounds of x , y x. Thus

(
)
x y ( x + y ) dA
V=
R

(b)
Figure
. : Finding the volume between the planes given in Example
.

u .

The volume between the surfaces is

Notes:

)
x y dy dx
cubic units.

Volume Between Surfaces and Triple Integra on

In the preceding example, we found the volume by evalua ng the integral


x
(
)
x y ( x + y ) dy dx.

Note how we can rewrite the integrand as an integral, much as we did in Sec on
. :

x y( x+y )=

x y

x+y

dz.

Thus we can rewrite the double integral that nds volume as


x (
x
(
)
x y( x+y ) dy dx =

x y

x+y

dz

dy dx.

This no longer looks like a double integral, but more like a triple integral.
Just as our rst introduc on to double integrals was in the context of nding the
area of a plane region, our introduc on into triple integrals will be in the context
of nding the volume of a space region.
To formally nd the volume of a closed, bounded region D in space, such as
the one shown in Figure . (a), we start with an approxima on. Break D into
n rectangular solids; the solids near the boundary of D may possibly not include
por ons of D and/or include extra space. In Figure . (b), we zoom in on a
por on of the boundary of D to show a rectangular solid that contains space not
in D; as this is an approxima on of the volume, this is acceptable and this error
will be reduced as we shrink the size of our solids.
The volume Vi of the i th solid Di is Vi = xi yi zi , where xi , yi
and zi give the dimensions of the rectangular solid in the x, y and z direc ons,
respec vely. By summing up the volumes of all n solids, we get an approxima on
of the volume V of D:
V

i=

Vi =

(a)

xi yi zi .

i=

Let ||D|| represent the length of the longest diagonal of rectangular solids
in the subdivision of D. As ||D|| , the volume of each solid goes to , as do
each of xi , yi and zi , for all i. Our calculus experience tells us that taking
a limit as ||D|| turns our approxima on of V into an exact calcula on of
V. Before we state this result in a theorem, we use a deni on to dene some
terms.
(b)
Figure . : Approxima ng the volume
of a region D in space.

Notes:

Chapter

Mul ple Integra on

Deni on

Triple Integrals, Iterated Integra on (Part I)

Let D be a closed, bounded region in space. Let a and b be real numbers, let g (x) and g (x) be
con nuous func ons of x, and let f (x, y) and f (x, y) be con nuous func ons of x and y.
. The volume V of D is denoted by a triple integral,

dV.
V=
D

. The iterated integral

b
a

b
a

g (x)
g (x)

g (x)

g (x)

f (x,y)
f (x,y)

f (x,y)
f (x,y)

dz dy dx is evaluated as

dz dy dx =

b
a

g (x)
g (x)

f (x,y)
f (x,y)

dz

dy dx.

Evalua ng the above iterated integral is triple integra on.

Our informal understanding of the nota on


dV is sum up
D
lots of li le
volumes over D, analogous to our understanding of R dA and R dm.
We now state the major theorem of this sec on.
Theorem

Triple Integra on (Part I)

Let D be a closed, bounded region in space and let D be any subdivision of D into n rectangular
solids, where the i th subregion Di has dimensions xi yi zi and volume Vi .
. The volume V of D is

V=
dV =
D

lim

||D||

Vi =

i=

lim

||D||

xi yi zi .

i=

. If D is dened as the region bounded by the planes x = a and x = b, the cylinders y = g( x)


and y = g (x), and the surfaces z = f (x, y) and z = f (x, y), where a < b, g (x) g (x)
and f (x, y) f (x, y) on D, then

dV =

b
a

g (x)
g (x)

f (x,y)
f (x,y)

dz dy dx.

. V can be determined using iterated integra on with other orders of integra on (there are
total), as long as D is dened by the region enclosed by a pair of planes, a pair of cylinders,
and a pair of surfaces.
Notes:

Volume Between Surfaces and Triple Integra on

We evaluated the area of a plane region R by iterated integra on, where


the bounds were from curve to curve, then from point to point. Theorem
allows us to nd the volume of a space region with an iterated integral with
bounds from surface to surface, then from curve to curve, then from point to
point. In the iterated integral

b
a

g (x)
g (x)

f (x,y)
f (x,y)

dz dy dx,

the bounds a x b and g (x) y g (x) dene a region R in the x-y plane
over which the region D exists in space. However, these bounds are also dening
surfaces in space; x = a is a plane and y = g (x) is a cylinder. The combina on
of these surfaces enclose, and dene, D.
Examples will help us understand triple integra on, including integra ng
with various orders of integra on.

Finding the volume of a space region with triple integra on


Example
Find the volume of the space region in the st octant bounded by the plane
z = y/ x/ , shown in Figure . (a), using the order of integra on
dz dy dx. Set up the triple integrals that give the volume in the other orders of
integra on.

(a)

Star ng with the order of integra on dz dy dx, we need to


S
rst nd bounds on z. The region D is bounded below by the plane z = (because we are restricted to the rst octant) and above by z = y/ x/ ;
z y/ x/ .
To nd the bounds on y and x, we collapse the region onto the x-y plane,
giving the triangle shown in Figure . (b). (We know the equa on of the line
y = x in two ways. First, by se ng z = , we have = y/ x/
y = x. Secondly, we know this is going to be a straight line between the
points ( , ) and ( , ) in the x-y plane.)
(b)

We dene that region R, in the integra on order of dy dx, with bounds

Notes:

Figure . : The region D used in Example


in (a); in (b), the region found by
collapsing D onto the x-y plane.

Chapter

Mul ple Integra on


y

x and

x . Thus the volume V of the region D is:

dV
V=
D

=
=
=
=

y x



z
x(

dz dy dz
)

y x

dz

dy dz

y x

dy dz
)
y x dy dz.

From this step on, we are evalua ng a double integral as done many mes before. We skip these steps and give the nal volume,
= u .
The order dz dx dy:
Now consider the volume using the order of integra on dz dx dy. The bounds
on z are the same as before, z y/ x/ . Collapsing the space region
on the x-y plane as shown in Figure . (b), we now describe this triangle with
the order of integra on dx dy. This gives bounds x y/ and y .
Thus the volume is given by the triple integral
V=

y x

dz dx dy.

The order dx dy dz:


Following our surface to surface. . . strategy, we need to determine the
x-surfaces that bound our space region. To do so, approach the region from
behind, in the direc on of increasing x. The rst surface we hit as we enter the
region is the y-z plane, dened by x = . We come out of the region at the plane
z = y/ x/ ; solving for x, we have x = y/ z/ . Thus the bounds
on x are: x y/ z/ .
Now collapse the space region onto the y-z plane, as shown in Figure . (a).
(Again, we nd the equa on of the line z = y/ by se ng x = in the equaon x = y/ z/ .) We need to nd bounds on this region with the order

Notes:

Volume Between Surfaces and Triple Integra on

dy dz. The curves that bound y are y = and y = z; the points that bound
z are and . Thus the triple integral giving volume is:
x y/ z/
y z
z

y/ z/

dx dy dz.

The order dx dz dy:


The x-bounds are the same as the order above. We now consider the triangle
in Figure . (a) and describe it with the order dz dy: z y/ and
y . Thus the volume is given by:
x y/ z/
z y/
y

y/

y/ z/

dx dz dy.
(a)

The order dy dz dx:


We now need to determine the y-surfaces that determine our region. Approaching the space region from behind and moving in the direc on of increasing y, we rst enter the region at y = , and exit along the plane z =
y/ x/ . Solving for y, this plane has equa on y = x z. Thus y
has bounds y x z.
Now collapse the region onto the x-z plane, as shown in Figure . (b). The
curves bounding this triangle are z = and z = x/ ; x is bounded by the
points x = to x = . Thus the triple integral giving volume is:
y x z
z x/
x

x/

x z

dy dz dx.
(b)

The order dy dx dz:


The y-bounds are the same as in the order above. We now determine the
bounds of the triangle in Figure . (b) using the order dy dx dz. x is bounded
by x = and x = z/ ; z is bounded between z = and z = . This leads
to the triple integral:
y x z
x z/
z

Notes:

z/

x z

dy dx dz.

Figure . : The region D in Example


is collapsed onto the y-z plane in (a);
in (b), the region is collapsed onto the x-z
plane.

Chapter

Mul ple Integra on


This problem was long, but hopefully useful, demonstra ng how to determine bounds with every order of integra on to describe the region D. In pracce, we only need , but being able to do them all gives us exibility to choose
the order that suits us best.
In the previous example, we collapsed the surface into the x-y, x-z, and y-z
planes as we determined the curve to curve, point to point bounds of integra on. Since the surface was a triangular por on of a plane, this collapsing, or
projec ng, was simple: the projec on of a straight line in space onto a coordinate plane is a line.
The following example shows us how to do this when dealing with more
complicated surfaces and curves.
Finding the projec on of a curve in space onto the coordiExample
nate planes
Consider the surfaces z = x y and z = y, as shown in Figure . (a).
The curve of their intersec on is shown, along with the projec on of this curve
into the coordinate planes, shown dashed. Find the equa ons of the projec ons
into the coordinate planes.

(a)

The two surfaces are z = x y and z = y. To nd


S
where they intersect, it is natural to set them equal to each other: x y =
y. This is an implicit func on of x and y that gives all points (x, y) in the x-y
plane where the z values of the two surfaces are equal.
We can rewrite this implicit func on by comple ng the square:
x y = y

y + y+x =

(y + ) + x = .

Thus in the x-y plane the projec on of the intersec on is a circle with radius ,
centered at ( , ).
To project onto the x-z plane, we do a similar procedure: nd the x and z
values where the y values on the surface are the same. We start by solving the
equa on of each surface for y. In this par cular case, it works well to actually
solve for y :
z= x y
y = x z
z= y y =z / .
Thus we have (a er again comple ng the square):
x z=z /

(b)
Figure . : Finding the projec ons of
the curve of intersec on in Example
.

(z + )

= ,

and ellipse centered at ( , ) in the x-z plane with a major axis of length and
a minor axis of length .

Notes:

Volume Between Surfaces and Triple Integra on

Finally, to project the curve of intersec on into the y-z plane, we solve equaon for x. Since z = y is a cylinder that lacks the variable x, it becomes our
equa on of the projec on in the y-z plane.
All three projec ons are shown in Figure . (b).
Finding the volume of a space region with triple integra on
Example
Set up the triple integrals that nd the volume of the space region D bounded
by the surfaces x + y = , z = and z = y, as shown in Figure . (a),
with the orders of integra on dz dy dx, dy dx dz and dx dz dy.
The order dz dy dx:

The region D is bounded below by the plane z = and above by the plane
z = y. The cylinder x + y = does not oer any bounds in the z-direc on,
as that surface is parallel to the z-axis. Thus z y.
Collapsing the region into the x-y plane, we get part of the circle with equaon x + y = as
shown in Figure . (b). As a func on
of x, this half circle
x . Thus y is bounded below by
x and above
has equa ony =
x y . The x bounds of the half circle are x .
by y = :
All together, the bounds of integra on and triple integral are as follows:

z y
x y
x

We evaluate this triple integral:

y

dz dy dx =

=
=

)
y
(

dz dy dx.

)
y dy dx

dx

( (
))


=
x x

Notes:

units .

(a)

dx

(b)
Figure . : The region D in Example
is shown in (a); in (b), it is collapsed onto
the x-y plane.

Chapter

Mul ple Integra on


With the order dy dx dz:
The region
is bounded below in the y-direc on by the surface x + y =
x and above by the surface y = z. Thus the y bounds are

y
=

x y z.

Collapsing the region onto the x-z plane gives the region shown in Figure
. (a); this half circle has equa on x + z = . (We nd this curve by solving
each surface for y , then se ng them equal to each other. We have y =
x
z
and y = z y =
z
.
Thus
x
+z
=
.)
It
is
bounded
below
by
x
=

z , where z is bounded by z . All together, we


and above by x =
have:

x y
z

z x
z
z

(a)

dy dx dz.
x

With the order dx dz dy:

y and above by
y .
D is bounded below by the surface x =
We then collapse the region onto the y-z plane and get the triangle shown in
Figure . (b). (The hypotenuse is the line z = y, just as the plane.) Thus z is
bounded by z y and y is bounded by y . This gives:
(b)

y x
y
z y
y

Figure . : The region D in Example


is shown collapsed onto the x-z plane
in (a); in (b), it is collapsed onto the y-z
plane.

dx dz dy.

The following theorem states two things that should make common sense
to us. First, using the triple integral to nd volume of a region D should always
return a posi ve number; we are compu ng volume here, not signed volume.
Secondly, to compute the volume of a complicated region, we could break
it up into subregions and compute the volumes of each subregion separately,
summing them later to nd the total volume.

Notes:

Theorem

Volume Between Surfaces and Triple Integra on

Proper es of Triple Integrals

Let D be a closed, bounded region in


space, and let D and D be nonD .
overlapping regions such that D = D

dV
.
D

dV =

dV +

dV.

We use this la er property in the next example.


Finding the volume of a space region with triple integra on
Example
Find the volume of the space region D bounded by the coordinate planes, z =
x/ and z = y/ , as shown in Figure . (a). Set up the triple integrals
that nd the volume of D in all orders of integra on.
Following the boundsdetermining strategy of surface to
S
surface, curve to curve, and point to point, we can see that the most dicult
orders of integra on are the two in which we integrate with respect to z rst,
for there are two upper surfaces that bound D in the z-direc on. So we start
by no ng that we have
z

and

y.

We now collapse the region D onto the x-y axis, as shown in Figure . (b).
The boundary of D, the line from ( , , ) to ( , , ), is shown in part (b) of the
gure as a dashed line; it has equa on y = x. (We can recognize this in two
ways: one, in collapsing the line from ( , , ) to ( , , ) onto the x-y plane,
we simply ignore the z-values, meaning the line now goes from ( , ) to ( , ).
Secondly, the two surfaces meet where z = x/ is equal to z = y/ :
thus x/ = y/ y = x.)
We use the second property of Theorem
to state that

dV =
dV +
dV,
D

where D and D are the space regions above the plane regions R and R , respec vely. Thus we can say
)
)
( y/
( x/

dz dA.
dz dA +
dV =
D

Notes:

(a)

(b)
Figure . : The region D in Example
is shown in (a); in (b), it is collapsed onto
the x-y plane.

Chapter

Mul ple Integra on


All that is le is to determine bounds of R and R , depending on whether we
are integra ng with order dx dy or dy dx. We give the nal integrals here, leaving
it to the reader to conrm these results.
dz dy dx:
z x/
y x
x

dV =

x/

z y/
xy
x

dz dy dx

y/

dz dy dx

dz dx dy:
z x/
y/ x
y

(a)

dV

y/

x/

z y/
x y/
y

dz dx dy

y/

y/

dz dx dy

The remaining four orders of integra on do not require a sum of triple integrals. In Figure . we show D collapsed onto the other two coordinate
planes. Using these graphs, we give the nal orders of integra on here, again
leaving it to the reader to conrm these results.
dy dx dz:
y z
x z
z

y z
z x/
x

x/

dy dx dz

dy dz dx:
(b)
Figure . : The region D in Example
is shown collapsed onto the x-z plane
in (a); in (b), it is collapsed onto the y-z
plane.

Notes:

dy dx dz

Volume Between Surfaces and Triple Integra on

dx dy dz:
x z
y z
z

x z
z y/
y

y/

dx dy dz

dx dz dy:
z

dx dz dy

We give one more example of nding the volume of a space region.


Finding the volume of a space region
Example
Set up a triple integral that gives the volume of the space region D bounded by
z = x + and z = x y . These surfaces are plo ed in Figure . (a)
and (b), respec vely; the region D is shown in part (c) of the gure.

(a)

The main point of this example is this: integra ng with reS


spect to z rst is rather straigh orward; integra ng with respect to x rst is not.
The order dz dy dx:
The bounds on z are clearly x + z x y . Collapsing D onto the
x-y plane gives the ellipse shown in Figure . (c). The equa on of this ellipse
is found by se ng the two surfaces equal to each other:
x +

x y

x +y =

We can describe this ellipse with the bounds

x y
x and

x +

= .
(b)

x .

Thus we nd volume as

x + z x y

x y
x
x

x y
x +

dz dy dx .

The order dy dz dx:

(c)

Notes:

Figure . : The region D is bounded


by the surfaces shown in (a) and (b); D is
shown in (c).

Chapter

Mul ple Integra on


Integra ng with respect to y is not too dicult. Since the surface z = x +
is a cylinder whose directrix is the y-axis, it does not create a border for y. The
paraboloid z = x y does; solving for y, we get the bounds

x zy
x z.

Collapsing D onto the x-z axes gives the region shown in Figure . (a); the
lower curve is from the cylinder, with equa on z = x + . The upper curve is
from the paraboloid; with y = , the curve is z = x . Thus bounds on z are
x + z x ; the bounds on x are x . Thus we have:

x zy
x z

x x z
dy dz dx.

x + z x

x +
x z
x
The order dx dz dy:

(a)

This order takes more eort as D must be split into two subregions. The
two surfaces create two sets of upper/lower bounds in terms of x; the cylinder
creates bounds

z/ x z/
for region D and the paraboloid creates bounds

y / z / x
y / z /

(b)
Figure . : The region D in Example
is collapsed onto the x-z plane in (a); in
(b), it is collapsed onto the y-z plane.

for region D .
Collapsing D onto the y-z axes gives the regions shown in Figure . (b).
We nd the equa on of the curve z = y / by no ng that the equa on of
the ellipse seen in Figure . (c) has equa on

x +y / =
x=
y / .
Subs tute this expression for x in either surface equa on, z =
z = x + . In both cases, we nd
z=

y .

Region R , corresponding to D , has bounds


z

y / ,

and region R , corresponding to D , has bounds


y / z

Notes:

y ,

y .

x y or

Volume Between Surfaces and Triple Integra on

Thus the volume of D is given by:

z/

y /

dx dz dy +

z/

y
y /

y / z /

dx dz dy.

y / z /

If all one wanted to do in Example


was nd the volume of the region D,
one would have likely stopped at the rst integra on setup (with order dz dy dx)
and computed the volume from there. However, we included the other two
methods ) to show that it could be done, messy or not, and ) because somemes we have to use a less desirable order of integra on in order to actually
integrate.

Triple Integra on and Func ons of Three Variables


There are uses for triple integra on beyond merely nding volume, just as
there are uses for integra on beyond area under the curve. These uses start
with understanding how to integrate func ons of three variables, which is eecvely no dierent than integra ng func ons of two variables. This leads us to a
deni on, followed by an example.
Deni on

Iterated Integra on, (Part II)

Let D be a closed, bounded region in space, over which g (x), g (x),


f (x, y), f (x, y) and h(x, y, z) are all con nuous, and let a and b be real
numbers.
The iterated integral

b
a

g (x)
g (x)

Example

Evaluate
S

Notes:

f (x,y)
f (x,y)

b
a

g (x)
g (x)

f (x,y)
f (x,y)

h(x, y, z) dz dy dx =

b
a

h(x, y, z) dz dy dx is evaluated as

g (x)
g (x)

f (x,y)
f (x,y)

h(x, y, z) dz dy dx.

ng a triple integral of a func on of three variables


x Evalua
x+ y (
)
xy + xz dz dy dx.
x

x y

We evaluate this integral according to Deni on

Chapter

Mul ple Integra on


x

=
=
=
=

x+ y
x y

x (
x

x(

)
xy + xz dz dy dx

x+ y
x y

)
)
xy + xz dz dy dx

)
(
) x+ y
xyz + xz
dy dx
x y
x
)
(
x
(
)
dy dx
xy( x + y) + x( x + y) xy(x y) + x(x y)
x

x(
x

x +x y+ x +

x y+

xy

dy dx.

We con nue as we have in the past, showing fewer steps.


)
(
x x x +

dx

We now know how to evaluate a triple integral of a func on of three variables; we do not yet understand what it means. We build up this understanding
in a way very similar to how we have understood integra on and double integra on.
Let h(x, y, z) a con nuous func on of three variables, dened over some
space region D. We can par on D into n rectangularsolid subregions, each
with dimensions xi yi zi . Let (xi , yi , zi ) be some point in the i th subregion, and consider the product h(xi , yi , zi )xi yi zi . It is the product of a
func on value (thats the h(xi , yi , zi ) part) and a small volume Vi (thats the
xi yi zi part). One of the simplest understanding of this type of product is
when h describes the density of an object, for then h volume = mass.
We can sum up all n products over D. Again le ng ||D|| represent the
length of the longest diagonal of the n rectangular solids in the par on, we can
take the limit of the sums of products as ||D|| . That is, we can nd
S=

lim

||D||

i=

h(xi , yi , zi )Vi =

lim

||D||

h(xi , yi , zi )xi yi zi .

i=

While this limit has lots of interpreta ons depending on the func on h, in
the case where h describes density, S is the total mass of the object described
by the region D.

Notes:

Volume Between Surfaces and Triple Integra on

We now use the above limit to dene the triple integral, give a theorem that
relates triple integrals to iterated itera on, followed by the applica on of triple
integrals to nd the centers of mass of solid objects.
Deni on

Triple Integral

Let w = h(x, y, z) be a con nuous func on over a closed, bounded space


region D, and let D be any par on of D into n rectangular solids with
volume Vi . The triple integral of h over D is

h(x, y, z) dV =

lim

||D||

h(xi , yi , zi )Vi .

i=

Note: In the marginal note on page


,
we showed how the summa on of rectangles over a region R in the plane could
be viewed as a double sum, leading to
the double integral. Likewise, we can
n

h(xi , yi , zi )xi yi zi as
view the sum
i=

a triple sum,

The following theorem assures us that the above limit exists for con nuous
func ons h and gives us a method of evalua ng the limit.

p
m
n

k=

Theorem

Triple Integra on (Part II)

Let w = h(x, y, z) be a con nuous func on over a closed, bounded space


region D, and let D be any par on of D into n rectangular solids with
volume Vi .
. The limit

lim

||D||

h(xi , yi , zi )Vi exists.

i=

. If D is dened as the region bounded by the planes x = a and


x = b, the cylinders y = g (x) and y = g (x), and the surfaces
z = f (x, y) and z = f (x, y), where a < b, g (x) g (x) and
f (x, y) f (x, y) on D, then

h(x, y, z) dV =

b
a

g (x)
g (x)

f (x,y)
f (x,y)

h(x, y, z) dz dy dx.

We now apply triple integra on to nd the centers of mass of solid objects.

Mass and Center of Mass

One may wish to review Sec on


and concepts.

Notes:

. for a reminder of the relevant terms

j=

h(xi , yj , zk )xi yj zk ,

i=

which we evaluate as
)
)
( n ( m
p


h(xi , yj , zk )xi yj zk .
k=

j=

i=

Here we x a k value, which establishes


the z-height of the rectangular solids on
one level of all the rectangular solids
in the space region D. The inner double
summa on adds up all the volumes of the
rectangular solids on this level, while the
outer summa on adds up the volumes of
each level.
This triple summa
on understanding

leads to the
nota
on of the triple
D
integral, as well as the method of
evalua on shown in Theorem
.

Chapter

Mul ple Integra on

Deni on

Mass, Center of Mass of Solids

Let a solid be represented by a region D in space with variable density


func on (x, y, z).

(x, y, z) dV.
dm =
. The mass of the object is M =
D

. The moment about the x-y plane is Mxy =


. The moment about the x-z plane is Mxz =
. The moment about the y-z plane is Myz =

z(x, y, z) dV.
y(x, y, z) dV.
x(x, y, z) dV.

. The center of mass of the object is


(
)
(
)
Myz Mxz Mxy
x, y, z =
.
,
,
M M M

Finding the center of mass of a solid


Example
Find the mass and center of mass of the solid represented by the space region
bounded by the coordinate planes and z = y/ x/ , shown in Figure
. , with constant density (x, y, z) = gm/cm . (Note: this space region was
used in Example
.)
We apply Deni on
. In Example
, we found bounds
S
for the order of integra on dz dy dx to be z y/ x/ , y x
and x . We nd the mass of the object:

(x, y, z) dV
M=
D

=
=

Figure . : Finding the center of mass


of this solid in Example
.

= ( )=

y/ x/

( )

y/ x/

dz dy dx

dz dy dx

gm.

The evalua on of the triple integral is done in Example

Notes:

, so we skipped those

Volume Between Surfaces and Triple Integra on

steps above. Note how the mass of an object with constant density is simply
densityvolume.
We now nd the moments about the planes.

z dV
Mxy =
D

=
=
=

= .

y/ x/

( )
z dz dy dx

y/ x/

(
)
x
dx

dy dx

We omit the steps of integra ng to nd the other moments.

x dV
Myz =
D

=
Mxz =
=

y dV

The center of mass is


(
)
x, y, z =

)
, . , . .

Finding the center of mass of a solid


Example
Find the center of mass of the solid represented by the region bounded by the
planes z = and z = y and the cylinder x + y = , shown in Figure . ,
with density func on (x, y, z) =
+ x + y z. (Note: this space region
was used in Example
.)
As we start, consider the density func on. It is symmetric
S
about the y-z plane, and the farther one moves from this plane, the denser the
object is. The symmetry indicates that x should be .
As one moves away from the origin in the y or z direc ons, the object becomes less dense, though there is more volume in these regions.

Notes:

Figure . : Finding the center of mass


of this solid in Example
.

Chapter

Mul ple Integra on


Though none of the integrals needed to compute the center of mass are
par cularly hard, they do require a number of steps. We emphasize here the
importance of knowing how to set up the proper integrals; in complex situa ons
we can appeal to technology for a good approxima on, if not the exact answer.
We use the order of integra on dz dy dx, using the bounds found in Example
. (As these are the same for all four triple integrals, we explicitly show the
bounds only for M.)
M=
=
=
Myz =

)
+ x + y z dV

Mxy =
=

)
+ x + y z dV

)
+ x + y z dV

)
+ x + y z dV

(
x

= .

(
y
Mxz =
=

(
z

)
+ x + y z dV

Note how Myz = , as expected. The center of mass is


(

)
x, y, z =

.
,
.

.
,
.

, .

, .

)
.

As stated before, there are many uses for triple integra on beyond nding
volume. Whenh(x,
y, z) describes a rate of change func on over some space
region D, then

h(x, y, z) dV gives the total change over D. Our one specic

example of this was compu ng mass; a density func on is simply a rate of mass
change per volume func on. Integra ng density gives total mass.
While knowing how to integrate is important, it is arguably much more important to know how to set up integrals. It takes skill to create a formula that describes a desired quan ty; modern technology is very useful in evalua ng these

Notes:

Volume Between Surfaces and Triple Integra on

formulas quickly and accurately.


This chapter inves gated the natural followon to par al deriva ves: iterated integra on. We learned how to use the bounds of a double integral to
describe a region in the plane using both rectangular and polar coordinates,
then later expanded to use the bounds of a triple integral to describe a region in
space. We used double integrals to nd volumes under surfaces, surface area,
and the center of mass of lamina; we used triple integrals as an alternate method
of nding volumes of space regions and also to nd the center of mass of a region in space.
Integra on does not stop here. We could con nue to iterate our integrals,
next inves ga ng quadruple integrals whose bounds describe a region in
dimensional space (which are very hard to visualize). We can also look back to
regular integra on where we found the area under a curve in the plane. A
natural analogue to this is nding the area under a curve, where the curve is
in space, not in a plane. These are just two of many avenues to explore under
the heading of integra on.

Notes:

Exercises

Terms and Concepts


. The strategy for establishing bounds for triple integrals
to
,
to
and
is
to
.
. Give an informal interpreta on of what
means.

. D is bounded by the planes y = , y = , x = , z =


z = ( x)/ .

and

Evaluate the triple integral with order dx dy dz.

dV

. Give two uses of triple integra on.


. If an object has a constant density and a volume V, what
is its mass?

Problems
In Exercises , two surfaces f (x, y) and f (x, y) and a region R in the x, y plane are given. Set up and evaluate the
double integral that nds the volume between these surfaces
over R.
. f (x, y) = x y , f (x, y) = x + y;
R is the square with corners ( , ) and ( , ).

. D is bounded by the planes x = , x = , z = y and by


z=y / .
Evaluate the triple integral with the order dy dz dx.

. f (x, y) = x + y , f (x, y) = x y ;
R is the square with corners ( , ) and ( , ).
. f (x, y) = sin x cos y, f (x, y) = cos x sin y + ;
R is the triangle with corners ( , ), (, ) and (, ).
. f (x, y) = x + y + , f (x, y) =
R is the circle x + y = .

x y ;

In Exercises , a domain D is described by its bounding


surfaces, along with a graph. Set up the triple integrals that
give the volume of D in all orders of integra on, and nd
the volume of D by evalua ng the indicated triple integral.
. D is bounded by the coordinate planes and
z = x/ y.
Evaluate the triple integral with order dz dy dx.

. D is bounded by the planes z = , y = , x =

z= y x .
Do not evaluate any triple integral.

and by

. D is bounded by the planes x = , y = , z =


z= x+ y .

and

. D is bounded by the coordinate planes and by


z = y/ and z = x.

Evaluate the triple integral with order dx dy dz.

Evaluate the triple integral with the order dx dy dz.

In Exercises
. D is bounded by the plane z = y and by y =

x .

Evaluate the triple integral with the order dz dy dx.

. D is bounded by the coordinate planes and by


y = x and y = z .
Do not evaluate any triple integral. Which order is easier to
evaluate: dz dy dx or dy dz dx? Explain why.

/
/

x
x

, evaluate the triple integral.

x+y

(
z

y
y

)
cos x sin y sin z dz dy dx

)
x + y + z dz dy dx

)
sin(yz) dx dy dz
(

x y+y x
z x +y
e

dz dy dx

In Exercises , nd the center of mass of the solid represented by the indicated space region D with density func on
(x, y, z).
. D is bounded by the coordinate planes and
z = x/ y; (x, y, z) = gm/cm .
(Note: this is the same region as used in Exercise .)
. D is bounded by the planes y = , y = , x = , z = and
z = ( x)/ ; (x, y, z) = gm/cm .
(Note: this is the same region as used in Exercise .)
. D is bounded by the planes x = , y = , z = and
z = x + y ; (x, y, z) = x lb/in .
(Note: this is the same region as used in Exercise .)
. D is bounded by the plane z = y and by y = x .
(x, y, z) = y lb/in .
(Note: this is the same region as used in Exercise .)

A: S

T S

Chapter

. Let > be given. We wish to nd > such that when


|x | < , |f(x) | < .
Consider |f(x) | < , keeping in mind we want to make a
statement about |x |:

Sec on .
. Answers will vary.

|f(x)

. F
. Answers will vary.

|x | < /|x + |

. Limit does not exist.


.

Since x is near , we can safely assume that, for instance,


< x < . Thus

. Limit does not exist.


h
.
.
.
.

f(a+h)f(a)
h

h
.
.
.
.

f(a+h)f(a)
h

h
.
.
.
.

h
.
.
.
.

|<

|x | |x + | <

|<

|x + x

|<

|x + x

+
The limit seems to be exactly .

<x+

<

<x+

<

<

x+

<
x+

.
.
.
.

The limit is approx. .

Let =

. Then:

x+

|x | |x + | <
x+
|x | <

.
f(a+h)f(a)
h

.
.
.
.

<

|x | <

|x | <

The limit is approx. . .

.
.

<

f(a+h)f(a)
h

The limit is approx. .

Assuming x is near , x + is posi ve and we can drop the


absolute value signs on the right.

(x + )
|x | |x + | <
x+

Sec on .

|x + x

. should be given rst, and the restric on |x a| < implies


|f(x) K| < , not the other way around.

|(x + x )

. T

. Let > be given. We wish to nd >


|x | < , |f(x) ( )| < .
Consider |f(x) ( )| < :

such that when

|f(x) + | <

|( x) + | <
| x| <

<

x<

< x

< .

This implies we can let = . Then:

< x

| < ,

which is what we wanted to prove.


. Let > be given. We wish to nd > such that when
|x | < , |f(x) | < . However, since f(x) = , a constant
func on, the la er inequality is simply | | < , which is
always true. Thus we can choose any we like; we arbitrarily
choose = .
. Let > be given. We wish to nd > such that when
|x | < , |f(x) | < . In simpler terms, we want to show
that when |x| < , | sin x| < .
Set = . We start with assuming that |x| < . Using the hint,
we have that | sin x| < |x| < = . Hence if |x| < , we know
immediately that | sin x| < .

<

<

< (x )

<

< (x + ) ( ) <
| x ( )| < ,

which is what we wanted to prove.

|<

Sec on .

|x | <

< x

|x + |

. Answers will vary.


. Answers will vary.
. As x is near , both f and g are near , but f is approximately twice
the size of g. (I.e., f(x) g(x).)

. Limit does not exist.

(a)

. Not possible to know.

(b)

(d)

(c)

(e)

(f)

. .

(g)

. Limit does not exist


.
.

(h)
.

+ +

(a)

(c) sin a
(d) sin a

. /

(a)
(b)

(c)

(d)

(a)

(b)

(c) Does not exist

Sec on .
. The func on approaches dierent values from the le and right;
the func on grows without bound; the func on oscillates.
. F
.

(a)

(d)
.

Sec on .
. Answers will vary.

(c)

. A root of a func on f is a value c such that f(c) = .

(d)

. F

(e) As f is not dened for x < , this limit is not dened.

. T

(f)

. F

(a) Does not exist.

. No; lim f(x) = , while f( ) = .

(c) Does not exist.


(d) Not dened.
(e)

. No; f( ) does not exist.


. Yes
.

(a)
(b)
(c)
(d)
(a)
(b)
(c)
(d)
(e)
(f)
(g)
(h) Not dened

(a)
(b)

(c) Does not exist.

(d)

A.

(a) No; lim f(x) = f( )


x

(b) Yes

(f)

(b)

(b) Does not exist.

cos a = sin a

(b) sin a

(c) No; f( ) is not dened.


.

(a) Yes
(b) No; the le and right hand limits at are not equal.

(a) Yes
(b) No. limx f(x) =

. (, ] [ , )

. (, ] [ , )

/ = f( ) = .

. (, )

. ( , )

. (, ]

. Yes, by the Intermediate Value Theorem.


. We cannot say; the Intermediate Value Theorem only applies to
func on values between and ; as is outside this range,
we do not know.
. Approximate root is x = . . The intervals used are:
[ , . ] [ , . ] [ .
, . ]
[ .
, . ] [ .
, . ] [ .
, . ]
[ .
, .
] [ .
, .
]

. Approximate root is x = . . The intervals used are:


[ . , . ] [ .
, . ] [ .
, . ]
[ .
, .
] [ .
, .
]
.

(a)

. f (x) =
. r (x) =

(c) Limit does not exist


. Answers will vary.

(b) x =

(a) y = x +

Sec on .

(b) y = / (x )
(a) y = (x + ) +

. F

(b) y = / (x + ) +

. F

(a) y = (x ) +

. T

(b) y = (x ) +

. Answers will vary.

. y= .

(a)

. y= .

(b)

(a)

(x ) +

x+

(a) Approxima ons will vary; they should match (c) closely.

(b) f (x) = /(x + )

(b)

(a) y =

(d)

. Answers will vary.


. g (x) = x

(b)

. Answers will vary.

(c)

(d)

(c) At ( , ), slope is . At ( , . ), slope is / .


y

(a) Limit does not exist

(b) Limit does not exist


. Tables will vary.
(a)

x
.
.
.

(b)

x
.
.
.

f(x)
.

.
f(x)
.
.
.

It seems limx

f(x) = .

. .
y

.
It seems limx

f(x) = .

.5

. Tables will vary.

(b)

(c) It seems limx f(x) does not exist.

(a)

.5

x
.
.

f(x)
.

x
.
.

f(x)
.

It seems limx

It seems limx

f(x) = .
f(x) = .

. Approximately
.

. Horizontal asymptote at y = ; ver cal asymptotes at x = , .

(a) (, )

(b) (, ) ( , ) ( , )
(c) (, ]

(c) It seems limx f(x) = .

. Horizontal asymptote at y = ; ver cal asymptotes at x = , .

(d) [ , ]

Sec on .

. No horizontal or ver cal asymptotes.

. Velocity

. Linear func ons.

. Solu on omi ed.

. f( . ) is likely most accurate, as accuracy is lost the farther from


x=
we go.

. Yes. The only ques onable place is at x = , but the le and


right limits agree.

Chapter
Sec on .
. T

.
.
.

/s
(a) thousands of dollars per car
(b) It is likely that P( ) < . That is, nega ve prot for not
producing any cars.

. f(x) = g (x)

A.

. Either g(x) = f (x) or f(x) = g (x) is acceptable. The actual


answer is g(x) = f (x), but is very hard to show that f(x) = g (x)
given the level of detail given in the graph.

. f (x) = sin x + x cos x

. f (x) =

. h (x) = csc x ex

. f () .

. g (x) =

(b) f(x) = x when x = , so f (x) = x .


(c) They are equal.

. Power Rule.

. f (t) = t (sec t + et ) + t (sec t tan t + et )

ex .

. One answer is f(x) =

. g (x) =

. g(x) and h(x)


. One possible answer is f(x) =

. f (x) =

. f (x) =

. m (t) =

. f (r) = er

f (t)

g (x)

. f (x) =

x
et

(t cos t+ )( t sin t+t cos t)(t sin t+ )( t cos tt sin t)


(t cos t+ )

. g (x) = sin x sec x + x cos x sec x + x sin x sec x tan x =


tan x + x + x tan x = tan x + x sec x

. f (x) is a velocity func on, and f (x) is accelera on.

. h (t) =

(x+ )( x + x )(x + x )( )
(x+ )

(a) f (x) =

Sec on .

. f (x) =

(x )

. Tangent line: y = (x

t t +

Normal line: y = (x

. Tangent line: y = x
Normal line: y = / x

cos t + sin t

. x=

. f( ) (x) = cos x + x sin x

. f(

. f (x) = x f (x) =

x f (x) =

. h (t) = t et h (t) =

x f( ) (x) =

et h (t) = et h( ) (t) = et

. Tangent line: y = (x )
Normal line: y = / (x )

. Tangent line: y = x
Normal line: y = x +
Normal line: y =

= x

=x

. f () = cos + sin f () = sin + cos


f () = cos sin f( ) () = sin cos

. Tangent line: y =

. x= ,

x+

. .

(x )

(x )

6
y

10

. The tangent line to f(x) = ex at x =


e . y( . ) = . .

is y = x + ; thus

Sec on .
. F

. .5

. T
(a) f (x) = (x + x) + x( x + )
(b)

f (x)

(a)

h (s)

. T

= (s + ) + ( s )( )

(b) h (s) = s +

(c) They are equal.


.

(a) f (x) =
(b) f (x) =

x( x)(x + )
x

(a) h (s) =
(b) h (s) =

s ( ) (
s
/ s

(c) They are equal.

A.

. f (x) =

( x x) (

x )=(

. g () = (sin + cos ) (cos sin )


)
(
) (
. f (x) = x + x
x

)( x x)

. g (x) = sec ( x)

. p (t) = cos (t + t + ) sin(t + t + )( t + )

. f (x) = /x

(c) They are equal.


.

. T
. F

= x + x

(c) They are equal.


.

Sec on .

. F
.

s )

. g (r) = ln

. g (t) =

. f (x) =

( t + ) (ln ) t ( t + ) (ln ) t
( t+ )

x+ )( x +x)(ln x

(ln x x

. f (x) =

x)

. Compose f(g(x)) and g(f(x)) to conrm that each equals x.

. Tangent line: y = ( / ) +
Normal line: y = / ( / ) +

. Compose f(g(x)) and g(f(x)) to conrm that each equals x.


( )
. f ( ) = f ( ) = /
( )
. f ( / ) = f (/ ) =
( )
. f ( / ) = f ( ) =

. g (x) =

. g (t) = cos(t + t) cos( t )( t+ ) sin(t + t) sin( t )


. Tangent line: y =
Normal line: x =

. h (t) =

. In both cases the deriva ve is the same: /x.


(a)

F/mph

(b) The sign would be nega ve; when the wind is blowing at
mph, any increase in wind speed will make it feel colder,
i.e., a lower number on the Fahrenheit scale.

Sec on .
. Answers will vary.
. f (x) = x

. h (x) = . x
. g (x) =

dy
dx

= sin(x) sec(y)

dy
dx

x cos (x)
x

(a) f(x) = x, so f (x) =

(a) f(x) =
f (x)

= .

x , so f (x) = x

cos(cos

x)(

= x

/ ) + /

. y = / (x ) +

Sec on .

y
x

. Answers will vary.

sin(y) cos(y)
x

. Answers will vary.


. F
. A: abs. min B: none C: abs. max D: none E: none

x+y
y+x

(a) y =

(x ) +

d y
dx

d y
dx

y
x /

(x
+

)+

. f ( ) =

f ( . ) =

f ( ) is undened

) is not dened

max: (/ ,

/
e

. min: ( , )
max: (e, /e)
.
.

Tangent line: y = ( / )(x / ) + (/ )


)
(x+ )(x+ ) (
. y = (x+ )(x+ ) x+ + x+ x+ x+
Tangent line: y =

f (

. min: (, e )

yx /

(
)
. y = ( x)x x ln( x) + x
Tangent line: y = ( + ln )(x ) +
(
)
. y = xsin(x)+ cos x ln x + sin x+
x
x+ /

Sec on .

dy
dx

y(y x)
x(x y)

x +

Sec on .
. Answers will vary.
. Any c in [ , ] is valid.
. c= /

. Rolles Thm. does not apply.

. F
. The point (
inver ble).

f ( ) =

. min: ( . , . )
max: ( , )

. min: (/ ,
/ )
max: (/ , )

)
. min: ( ,
max: ( , / )

(a) y = (x ) +
(b) y =

. f ( ) =
.

(x . ) + .

(a) y =

(b) y = .
.

Chapter

x
y+

(b) y = .

f (x)

. y = (x

y+

(b)

cos(x)(x+cos(y))+sin(x)+y
sin(y)(sin(x)+y)+x+cos(y)

x( )(x+ )( / x / )
x

= .

dy
dx

sin (x)+cos (x)

h (x)

(b) f (x) = cos(sin x)

. f (t) = t

+ x

sin(t)
. g (t) = cos (t) cos(t)

. T

, ) lies on the graph of y =

(x) (assuming f is

. Rolles Thm. does not apply.


. c=

A.


. c= /

. Graph and verify.


. Possible points of inec on: none; concave down on (, )

. The Mean Value Theorem does not apply.

. c = sec ( / )
. c=

. Max value of
x= . .

at x = and x = ; min value of .

at

. They are the odd, integer valued mul ples of / (such as


, / , / , / , etc.)

Sec on .

. Answers will vary.


. Answers will vary.
. Increasing
. Graph and verify.
. Graph and verify.
. Graph and verify.
. Graph and verify.
. domain=(, )
c.p. at c = , ;
increasing on (, ) ( , );
decreasing on ( , );
rel. min at x = ;
rel. max at x = .

. domain=(, )
c.p. at c = ;
increasing on (, );

. domain=(, ) ( , ) ( , )
c.p. at c = ;
decreasing on (, ) ( , );
increasing on ( , ) ( , );
rel. min at x = ;

. domain=(, ) ( , );
c.p. at c = , ;
decreasing on (, ) ( , ) ( , );
increasing on ( , );
rel. min at x = ; rel. max at x = .
. domain = (, );
c.p. at c = , ;
decreasing on ( , );
increasing on (, ) ( , );
rel. min at x = ;
rel. max at x =

. c = cos ( /)

Sec on .

. Answers will vary.


. Yes; Answers will vary.
. Graph and verify.
. Graph and verify.
. Graph and verify.
. Graph and verify.
. Graph and verify.

A.

. Possible points of inec on: x = / ; concave down on


(, / ); concave up on ( / , )

. Possible points
); concave up on
of inec on: x= ( / )(
(( / )(
), (/ )( +
)); concave
down on
(, ( / )(
)) (( / )( +
), )

;
concave
downon
. Possible
points
of
inec
on:
x
=

( / , / ); concave up on (, / ) ( / , )
. Possible points of inec on: x = / , / ; concave down on
(/ , / ) concave up on (, / ) ( / , )
. Possible points of inec on: x = /e / ; concave down on
( , /e / ) concave up on ( /e / , )

. min: x =

. max: x = /

min: x = /

. min: x =
. min: x =

. cri cal values: x = , ; no max/min

. max: x = ; min: x =

. max: x =

. f has no maximal or minimal value


. f has a minimal value at x = /

. f has a rela ve
max at: x = ( / )( +
) rela ve min at:
x = ( / )(
)

. f has a rela ve max at x = / ; rela ve min at x = /


. f has a rela ve min at x = / ; rela ve max at x = /

. f has a rela ve min at x = / e = e /

Sec on .
. Answers will vary.
. T
. T
. A good sketch will include the x and y intercepts..
. Use technology to verify sketch.
. Use technology to verify sketch.
. Use technology to verify sketch.
. Use technology to verify sketch.
. Use technology to verify sketch.
. Use technology to verify sketch.
. Use technology to verify sketch.
. Use technology to verify sketch.
. Use technology to verify sketch.
n/ b

, where n is an odd integer Points of


. Cri cal points: x =
a
inec on: (n b)/a, where n is an integer.
.

dy
= x/y, so the func on is increasing in second and fourth
dx
quadrants, decreasing in the rst and third quadrants.

d y
= /y x /y , which is posi ve when y < and is
dx
nega ve when y > . Hence the func on is concave down in the
rst and second quadrants and concave up in the third and fourth
quadrants.

Chapter
Sec on .

. F
. x = . ,x = .
,x = .
x = .
,x = .

,x = .

. x = ,x = ,x = . ,x = .
x =
. x = ,x = .
x = .
. roots are: x = .
. x= .

,x = .

,x = .
,x = .

. roots are: x = .

,x = .

,x =

,x =

,x = ,x = .

(a)

,x =

/( ) .

and x = .

(c)
.

/(
mph

cm/s
cm/s

) .

. dy =

dx
x

xe

+ x e

dx

(tan x+ ) x sec x
dx
(tan x+ )

rad/s

. dy =

(b)

rad/s

. dy = (ex sin x + ex cos x)dx

(a)

(b)

(c)

(a)

(b)

rad/s

. dy =

/s

(a)

/s

.
.

, the rates

.
.

/min

(b)

/sec

(c)

/sec

. feet

(b)

feet

in , or /

(a)

. feet

(b)

(c) . %

/min

(a) The rope is

(d) About

dx

(x+ )

. dy = (ln x)dx

/s

(c) .
/min
As the tank holds about

, it will take about

minutes.

long.

. The isosceles triangle setup works the best with the smallest
percent error.

Chapter
Sec on .

feet.

. Answers will vary.

. The cone is rising at a rate of .

/s.

. Answers will vary.


. Answers will vary.

Sec on .

. velocity

. T
.

(d) Not dened; as the distance approaches


approaches .

. dy =

(a)

(c) In the limit, rate goes to .

. Use y = x ; dy = x dx with x = and dx = . . Thus


dy = . ; knowing =
, we have .
. .

. Use y = x; dy = /(
x)

dx
with
x
=
and
dx
=
. Thus

= , we have
. .
dy = . ; knowing

. Use y = x; dy = /( x ) dx withx = and dx = . . Thus


= , we have
dy
= / / = . ; knowing
. . .

. dy = ( x + )dx

. Due to the height of the plane, the gun does not have to rotate
very fast.

Sec on .

. Use y = cos x; dy = sin x dx with x = / . and


dx . . Thus dy = . ; knowing cos / = , we have
cos . . .

cm/s

/( ) .

. Use y = x ; dy = x dx with x = and dx = . . Thus


dy = . ; knowing = , we have .
. .

and x = .

. T
(b)

. Answers will vary.

Sec on .

. The largest area is formed by a square with sides of length

. F

and x = .

. The approxima ons alternate between x = , x =

feet along the shore before star ng

. T

,x = .

. x= .

. The dog should run about


to swim.

.
; the two numbers are each

. There is no maximum sum; the fundamental equa on has only


cri cal value that corresponds to a minimum.

. Area = / , with sides of length / .


. The radius should be about . cm and the height should be
r = . cm. No, this is not the size of the standard can.
. The height and width should be and the length should be ,
giving a volume of ,
in .

. miles should be run underground, giving a


. /
minimum cost of $
,
. .

/ x +C

. t+C
. /( t) + C

x+C

. cos + C
.

e + C

ln

+C

. t / +t / t +C

. e x + C
.

(a) x >

A.

(b)

. cos x sin x + tan x + C

/x

(c) x <
(d)
.

. ln |x| + csc x + C

Sec on .

/x

(e) ln |x| + C. Explana ons will vary.

. limits

ex +

. Rectangles.

. tan x +
.

/ x + x+

ex x

x ln ( )+ x +x ln )(ln

.
.
)+ln ( ) cos(x) ln ( )
ln ( )

. No answer provided.

(a)

+ + + + + =

i= (i )

. Answers may vary; i= ( )i ei

.
.

(c)

(d)

. / + /(

(e)

(f)

(a)

(c)
.

(d)
(e)

(d)

(a)

(a) Exact expressions will vary;


,

/
/

. T

(b)

. Answers will vary.

(d)

(a)

/s

. ( / )/ ln

(c)

Sec on .

(a)

(b)

. G(t) = sin t cos t

(d)

.
/s

(b)

seconds

(c)

seconds

/
/

. Answers can vary; one solu on is a = , b =

,b =

. Answers can vary; one solu on is a =

A.

, ,

. G(t) = / t / t + t +

(d) Never; the maximum height is

.
/

. F(x) = tan x +

(c)

(a)

(c)

(b) /

(b)

(b)

(b)

(a)

(a)

(c)

( +n)
n

(c)

(f)

) .

(a) Exact expressions will vary;


(b)

(c)

(b)

(b)

(c)

. + + +

. Answers will vary.

. Answers may vary;

Sec on .

+ + +

. Explana ons will vary. A sketch will help.

. c= /

/n.

. c=

/ .

x +

+C

/pi

ln | x + | + C

(x + )

/(e )

/s

/s

F (x)

= ( x + )x

(a)
(c)
(a)
(b)
(c)

/ .

(a)

.
.

(c)

(a)
(c)

+
.

+
) .
/ ( +
+

(b)

(b)

/ ( +
)

.
+

+C

+C

tan (x)

ex

+C

+C

(a) n =
(b) n =
(a) Area is

.
.

+ x + ln |x| + C

Chapter
Sec on .
. Chain Rule.
(x ) + C

cm

+ x ln |x + | + C

x x + ln |x + | + C
( )

tan x + C
.
( )
.
sin x + C
.

sec (|x|/ ) + C

tan

sin

+C
( x )
+C

+C
(x + )

.
x +C
.

ln | x + | + C


. ln x + x + + C


. x + ln x x + + x + C

(
)
(using max f (x) = )
(
)
(using max f ( ) (x) = )
(
)
(using max f (x) =
)
( () )
(using max f (x) = )

(b) Area is

. Trapezoidal Rule: .
Simpsons Rule: .
(b) n =

+C

ln (x) + C
( )
ln x + C

. cos (x) + C

. Trapezoidal Rule: .
Simpsons Rule: .

(a) n =

ln

.
.

. Trapezoidal Rule: .
Simpsons Rule: .

+C

sin (x)

. Trapezoidal Rule: .
Simpsons Rule: .

.
.

(b)

+ C = (x ) x + + C

. x ex + C

. They are superseded by the Trapezoidal Rule; it takes an equal


amount of work and is generally more accurate.

. The key is to mul ply csc x by in the form


(csc x + cot x)/(csc x + cot x).

+x

. F

(x + )

. tan(x) x + C

Sec on .

. tan( x) + C

. F (x) = x(x + ) (x + )

yd

. tan ( x) + C
( x)
. sin
+C

(
)
x+
tan
ln x +
.


. x ln x + + C

x+


+C

. tan (cos(x)) + C

. ln | sec x + tan x| + C (integrand simplies to sec x)

.
x x+ +C

.
.

/
/

. /
. /

Sec on .
. T

A.

. Determining which func ons in the integrand to set equal to u


and which to set equal to dv.
. ex xex + C

. x cos x + x sin x + x cos x sin x + C

. x ex x ex + xex ex + C
.

(b) sec .
(
)

x x + + ln | x + + x| + C

e x ( sin( x) cos( x)) + C

. x tan(x) + ln | cos(x)| + C
(x )

. sec x + C

+ (x )

x
sec (x/
)+C

x + C (Trig. Subst. is not needed)


.

+C

x +

e
(
)
/ e + e
e

tan

( x+ )

Sec on .
. ra onal

. F
.

sin (x) + C

cos x

cos x + C
sin (x)

sin (x)

. sin (x) +

+
)
(
cos( x) cos( x) + C
.
(
)
.
sin( x) sin( x) + C
(
)
.
sin(x) + sin( x) + C
.

tan (x)

tan (x)

sec (x)

sin (x)

+C
+

tan (x)

+C

sec (x)

+C

+C

A
x

A
x

.
.

(sec x tan x ln | sec x + tan x|) + C

Sec on .

B
x+

ln |x | + ln |x + | + C

(ln |x + | ln |x |) + C

x+C


. ln x + x +

. x + ln |x | ln |x + | + C

B
x

. ln | x | + ln |x | + ln |x + | + C

tan x tan x + x + C

. x+ ln |x + | + C

. /

+ ln( +
)
.

Note: the new lower bound is


. sin ( / ) +
= sin ( / ) and the new upper bound is = sin ( / ).
The nal answer comes with recognizing that
sin( ( / ) = ) sin (( / ) and that )
/ .
cos sin ( / ) = cos sin ( / ) =

. F

(Trig. Subst. is not needed)

+C
(
)

x
x
sin (x/ ) + C

Sec on .

x+
x + x+

.
.

+C

sin x + x
x +C

. x csc x ln | csc x + cot x| + C


( )
( )

. sin
x
x cos
x +C

.
xe x e x + C

= sec

x x ln |x + x | + C

x + / + x| + C =
. x x + / + ln |

x x + + ln | x + + x| + C
(
)

.
x x / ln | x +
x / | +C=

x ln | x +
x |+C
x
( )
. sin x + C (Trig. Subst. is not needed)

x + x (ln |x|) x ln |x| + C

A.

(a) tan +

. x + x ln |x| + x x ln |x| + C
( )
. x ln x x + C

/ ex (sin x cos x) + C

. / cos x + C


. x tan ( x) ln x + + C

x + x sin x + C
.

. backwards

.
.


+x+

(
)

tan x+

+C

ln |x | + ln |x + x + | tan (x + ) + C


(
(
))
ln x + x + ln |x | + tan x+
+C
(
)

tan x+
ln x + x + + ln |x + |
+C

ln( / )

ln(

/ ) .

Sec on .
. Because cosh x is always posi ve.

coth x csch x =

ex + ex
ex ex

ex

+ + e x ) ( )
e x + e x
x
e + e x
= x
e + e x
=
( x
)
e + ex
cosh x =
=

(e

ex

e x

ex ex
x
x
e +e
ex + ex
= sech x tanh x

sinh x
dx
cosh x
Let u = cosh x; du = (sinh x)dx

tanh x dx =

du
u
= ln |u| + C

= ln(cosh x) + C.

sinh x

. a/b
.

.
.
.
.
.
.
.
.
.
.

Sec on .
. The interval of integra on is nite, and the integrand is
con nuous on that interval.

. e /

x +

(x+ )

/ ln

. sec x

. diverges

. y = / (x ln ) + /

.
. diverges

. y=x
.

/ ln(cosh( x)) + C

. diverges

/ sinh x + C or / cosh x + C

. diverges

. x cosh(x) sinh(x) + C

. cosh (x / ) + C = ln(x +
.

tan

(x/ ) +

. tan (ex ) + C

x )+C

ln |x | +

ln |x + | + C

.
. /
.

. diverges

. x tanh x + / ln |x | + C

. converges; Limit Comparison Test with /x

. converges; Direct Comparison Test with

Sec on .
.

. p>

. x cosh x

. /

. /

. converges; could also state <

. coth x
.

. Answers will vary.

. deriva ves; limits

.
]

(ex ex )
= x
(e + ex )(ex + ex )

. F

.
)

(ex ex )
=
(ex + ex )

.
ex

ex

+ e x ) +

cosh x +

d
d
[sech x] =
dx
dx

+ + e

(e

/ , /, , ,

xex .

. converges; Direct Comparison Test with xex .


,

. diverges; Direct Comparison Test with x/(x + cos x).

A.

. converges; Limit Comparison Test with /ex .

. F

Chapter

Sec on .

. / units

. T

. Answers will vary.


.

(a)

(b)

(c) /

(d)

(a)

(b) /

(c)

(d)

(
)

(
+ sinh ( ))

. On regions such as [/ , / ], the


area is
/ .
such as [/ , / ], the area is
.

/ . On regions

.
,

. Recall that dx does not just sit there; it is mul plied by A(x)
and represents the thickness of a small slice of the solid.
Therefore dx has units of in, giving A(x) dx the units of in .
/ units

. / units
.
.

/ units
/

units

. ln(
) .

+ x dx
.

.
+ x dx

x
+ x
.
dx

+ x dx
.
.

(a)

. Simpsons Rule fails, as it requires one to divide by . However,


recognize the answer should be the same as for y = x ; why?

(b)

. Simpsons Rule fails.

(c)

(d)

(a)
(b)

(c)

(a)

(b)

(c)
(d)

/
/

. In SI units, it is one joule, i.e., one Newtonmeter, or kgm/s m.


In Imperial Units, it is lb.
. Smaller.
.

. Orient the solid so that the x-axis is parallel to long side of the
base. All crosssec ons are trapezoids (at the far le , the
trapezoid is a square; at the far right, the trapezoid has a top
length of , making it a triangle). The area of the trapezoid at x is
A(x) = / ( / x + + )( ) = / x + . The volume is
. units .

Sec on .

x
dx =

+ x dx = / (
)

x
+ x/( x ) dx =

Sec on .

. The crosssec ons of this cone are the same as the cone in
Exercise . Thus they have the same volume of
/ units .

. T

. T

Sec on .

(a)
(b)

. T

.
.

Sec on .

A.

/ units

. units

.
/ units

(a)

(b)

lb

lb

.
.

lb

.
.

lb
,

,
lb. Note that the tank is oriented horizontally. Let the
origin be the center of one of the circular ends of the tank. Since
the radius is .
, the uid is being pumped to y = . ; thus
the distance the gas travels is h(y) = . y. A dieren al
element
of water is a rectangle, with length and width

.
y . Thusthe force required to move that slab of gas is
.

y dy. Total work is


F(y) =
.
.

(
.

y)
.
y dy. This can be
.
evaluated
without actual integra
on; split theintegral into

.
.
y dy + . .
.
. ( . )
.
(y)
.
y dy. The rst integral can be evaluated as
measuring half the area of a circle; the la er integral can be
shown to be without much diculty. (Use subs tu on and
realize the bounds are both .)

(a) approx.

(b) approx.

(c) approx
,
j (By volume, half of the water is between
the base of the cone and a height of .
m. If one
rounds this to m, the work is approx
,
j.)
.

. lb
. lb

. lb

. Let {an } be given such that lim |an | = . By the deni on of


n

the limit of a sequence, given any > , there is a m such that for
all n > m, | |an | | < . Since | |an | | = |an |, this
directly implies that for all n > m, |an | < , meaning that
lim an = .
n

. Le to reader

Sec on .
. Answers will vary.
. One sequence is the sequence of terms {a} . The other is the

sequence of nth par al sums, {Sn } = { ni= ai }.

lb

.
lb

(a)

lb

(b)

lb

(a)

(b)

lb

(a)

.
.
lb

lb
.

lb

(a)

, , ,

(a)

, ,

(a)

. , .

, .

, .

, .

lim an = ; by Theorem

the series diverges.

lim an = ; by Theorem

the series diverges.

lim an = e; by Theorem

the series diverges.

n
n

. Converges
. Converges

. Diverges
.

. Answers will vary.


, , ,

. an = n +
. an =

n(n+ )

(a) Sn =

(a) Sn =

/ n
/

(b) Converges to

,
,

(a) Sn =

(b) Diverges

. Answers will vary.

. Converges

Sec on .

, ,

. Converges

Chapter

, ,

(b) Plot omi ed

(b)

(a)

(b) Plot omi ed

(b) Plot omi ed

. never monotonic

. monotonically increasing

(b) Plot omi ed

. neither bounded above or below

. Answers will vary.

. bounded

. F

Sec on .

. bounded

( / )n
/

(b) Converges to / .
n

.
. diverges
. converges to
. diverges
. converges to e
. converges to
. converges to

.
.

(a) With par al frac ons, an =


(
)
Sn =
n+ n+ .

n+

. Thus

(b) Converges to /
(
)
(a) Sn = ln /(n + )

(b) Diverges (to ).

(a) an = n(n+ ) ; using par al frac ons, the resul ng


telescoping
( sum reduces to
)
+ + n+ n+ n+
Sn =

(b) Converges to

A.

(a) With par al frac ons, an =


(
)
/ n n+ .
Sn =

n+

. Thus

. Diverges; compare to

(a) The nth par al sum of the odd series is


+ + + + n . The nth par al sum of the even

. Diverges; compare to

(b) The nth par al sum of the odd series is


+ + + + n . The nth par al sum of plus the

even series is + + + + (n ) . Each term of the


even series is now greater than or equal to the
corresponding term of the odd series, with equality only on
the rst term. This gives us the result.

(c) If the odd series converges, the work done in (a) shows the
even series converges also. (The sequence of the nth
par al sum of the even series is bounded and
monotonically increasing.) Likewise, (b) shows that if the
even series converges, the odd series will, too. Thus if
either series converges, the other does.
Similarly, (a) and (b) can be used to show that if either
series diverges, the other does, too.
(d) If both the even and odd series converge, then their sum
would be a convergent series. This would imply that the
Harmonic Series, their sum, is convergent. It is not. Hence
each series diverges.

n=

(b) Converges to / .

series is + + + + n . Each term of the even


series is less than the corresponding term of the odd
series, giving us our result.

lim

sin( /n)
= .
/n

ln n
.
n
n=

n=

. Diverges; compare to

n=

. Converges; compare to

n=

. Just as lim
n

sin n
= ,
n

. Converges; Integral Test

. Diverges; the nth Term Test and Direct Comparison Test can be
used.
. Converges; the Direct Comparison Test can be used with sequence
/ n.
. Diverges; the nth Term Test can be used, along with the Integral
Test.
.

(a) Converges; use Direct Comparison Test as

an
n

(d) May converge; certainly nan > an but that does not mean
it does not converge.

. con nuous, posi ve and decreasing


. The Integral Test (we do not have a con nuous deni on of n!
yet) and the Limit Comparison Test (same as above, hence we
cannot take its deriva ve).

(e) Does not converge, using logic from (b) and nth Term Test.

Sec on .

. Converges

. algebraic, or polynomial.

. Diverges

. Integral Test, Limit Comparison Test, and Root Test

. Converges

. Converges

. Converges

. Converges

. Converges; compare to

n=

all n > .
. Diverges; compare to

n=

. Diverges; compare to

n=

, as /(n + n ) /n for

, as /n ln n/n for all n .

. Since n =

n >

/n / n for all n .
. Diverges; compare to

n=

. Diverges; compare to

n=

n
n
n
n

> , for all n .

n ,

<

n +n+
,
n

n
n

. Converges
. Converges; note the summa on can be rewri en as
from which the Ra o Test can be applied.

n=

n n!
n n!

. Converges
. Converges

. Diverges. The Root Test is inconclusive, but the nth -Term Test
shows divergence. (The terms of the sequence approach e , not
, as n .)

. Converges

. Diverges; Limit Comparison Test


. Converges; Ra o Test or Limit Comparison Test with / n .

. Note that

. The Ra o Test is inconclusive; the p-Series Test states it diverges.

. Diverges

n
n +n+
<
n
n

for all n .

as

< n.

(b) Converges; since original series converges, we know


limn an = . Thus for large n, an an+ < an .
(c) Converges; similar logic to part (b) so (an ) < an .

Sec on .

A.

. Converges; compare to

. Diverges; nth -Term Test or Limit Comparison Test with .


n

>

. Diverges; Direct Comparison Test with /n


. Converges; Root Test

Sec on .

. The signs of the terms do not alternate; in the given series, some
terms are nega ve and the others posi ve, but they do not
necessarily alternate.
. Many examples exist; one common example is an = ( )n /n.
.

(a) converges
(b) converges (p-Series)
(c) absolute

(a) diverges (limit of terms is not )


(b) diverges

(b) ( / , / )
.
.
.

(a) R =

(b) (, )
(a) R =

(b) [ , ]
.

(a) R =
(b) x =

(c) n/a; diverges


.

(a) R =
(b) ( , )

(a) converges

(a) f (x) =

(c) condi onal

(b)

(a) diverges (limit of terms is not )


(b) diverges

(c) n/a; diverges


.

(c) n/a; diverges


.

(a) converges
(b) converges (Geometric Series with r = / )
(c) absolute
(a) converges
(b) converges (Ra o Test)
(c) absolute

(a) converges
(b) diverges (p-Series Test with p = / )
(c) condi onal

. S = .
.
. S = .
.

;S = .
;

( )n
.

ln(n + )
n=

. n=

;S = .

( )n
n!

n=

f(x) dx = C +

. Using the theorem, we nd n =


guarantees the sum is within
.
of / . (Convergence is actually faster, as the sum is within
of / when n
.)

Sec on .

xn ;

f(x) dx = C +

n=

( , )

( , )

(n + )

xn+ ;

[ , )

+ x+ x + x +

+x+x +x +x

+x+ x x + x

Sec on .

. The Maclaurin polynomial is a special case of Taylor polynomials.


Taylor polynomials are centered at a specic x-value; when that
x-value is , it is a Maclauring polynomial.
. p (x) =
. p (x) =

+ x x .

x+ x x

. p (x) = x + x + x + x +
. p (x) =

+ x + x+

. p (x) = x x + x x +

. p (x) = + ( +x) ( +x) +

. p (x) =

+x

)
( +x

( +x)

x + (x )

( +x)
( +x)

(x ) +

( +x)

+x)
(

(x )

(x )

+ x+ x + x +

+x+

(a) R =

. p (x) = x x ; p ( . ) = .
.
! . .

(a) R =

. p (x) = + ( + x)
( + x) ; p ( ) = .
.

The third deriva ve of f(x) = x is bounded on ( , ) by .


.
.
Error is bounded by . ! = .

(b) (, )
(b) ( , ]

(a) R =
(b) ( , )

(a) R = /

n n+
x
;
n+

( )n x n
( )n+ x n+
=
;
( n )!
( n + )!
n=
n=
(, )

( )n x n+
(b)
f(x) dx = C +
; (, )
( n + )!
n=

)
( +x

( , )

(a) f (x) =

. p (x) =

n=

(a) f (x) =
(b)

(b) diverges

n=

(a) diverges (terms oscillate between )

n xn ;

n=

(b) diverges (Limit Comparison Test with /n)


.

. p (x) =

+x

+ ( + x)
. Error is bounded by

. The nth deriva ve of f(x) = ex is bounded by on intervals


containing and . Thus |Rn ( )| (n+ )! (n+ ) . When n = ,
this is less than .
.

A.

. The nth deriva ve of f(x) = cos x is bounded by on intervals


containing and / . Thus |Rn (/ )| (n+ )! (/ )(n+ ) .
When n = , this is less than .
. Since the Maclaurin
polynomial of cos x only uses even powers, we can actually just
use n = .
. The nth term is

n!

xn .

. The nth term is xn .


. The nth term is ( )n
.

x+

(x )n
.
n

x +

x +

. Given a value x, the magnitude of the error term Rn (x) is bounded


by





max f (n+ ) (z)
Rn (x)
(x )(n+ ) ,
(n + )!
where z is between and x.


n!
.
Note that f (n+ ) (x) = xn+
We consider the cases when x > and when x < separately.
n!
If x > , then < z < x and f (n+ ) (z) = zn+
< n!. Thus

For a xed x,

Sec on .

lim

. A Taylor polynomial is a polynomial, containing a nite number of


terms. A Taylor series is a series, the summa on of an innite
number of terms.

Since < x < , xn+ < and ( x)n+ < . We can then
extend the inequality from above to state

x
the Taylor series is
n!
n=

n+


Rn (x) x
( x)n+ <
.
n+
n+

. The nth deriva ve of /( x) is f (n) (x) = (n)!/( x)n+ ,


which evaluates to n! at x = .
The Taylor series starts + x + x + x + ;

xn
the Taylor series is

As n , /(n + ) . Thus by the Squeeze Theorem, we


conclude that lim Rn (x) = for all x, and hence
n

the Taylor series is

( )

n+

n=

(x / ) n+
( n + )!

(x / ) ;

the Taylor series is

( )n

n=

= (

)n+

n!
(x+ )n+

The Taylor series starts


+ (x ) (x ) +
the Taylor series is

( )n+

(x )n
n+

M
For any x, lim
n (n + )!
Theorem, we conclude that lim Rn (x) =
n

x
e =
n!
n=

A.

for all x.

for all x, and hence

( n)!

<x .

( )n

( )n

n=

M (n+ )
.
x
(n + )!
(n+ )
= . Thus by the Squeeze
x

for all

x n+
,
( n + )!

(
)
)
d(
x n+
d
=
( )n
sin x =
dx
dx n=
( n + )!

( n + )x n
xn
( )n
( )n
=
= cos x. (The
(
n
+
)!
(
n)!
n=
n=
summa on s ll starts at n = as there was no constant term in
the expansion of sin x).

n!



Rn (x)

( )n

n=

. Given sin x =

n+

. Given a value x, the magnitude of the error term Rn (x) is bounded


by




max f (n+ ) (z) (n+ )
,
x
Rn (x)
(n + )!
where z is between and x.
If x > , then z < x and f (n+ ) (z) = ez < ex . If x < , then
x < z < and f (n+ ) (z) = ez < . So given a xed x value, let
M = max{ex , }; f (n) (z) < M. This allows us to state

(x )n
n

(x) n
xn
=
= cos x, as all
( )n
( n)!
( n)!
n=
n=
powers in the series are even.

(x ) ;

n=

. Given cos x =

xn
.
n!

; at x = , f (n) ( ) = ( )n+

( )n+

n=

cos(x) =

. f (n) (x) = ( )n ex ; at x = , f (n) ( ) = when n is odd and


f (n) ( ) = when n is even.
The Taylor series starts x + x ! x + ;

ln x =

n=

. The Taylor series starts


(x / ) + x + (x / ) + x

(x )n+
= .
n+

n!
n!
If < x < , then x < z < and f (n+ ) (z) = zn+
< xn+
.
Thus
n+
n+



(x )(n+ ) = x
Rn (x) n!/x
( x)n+ .
(n + )!
n+

. All deriva ves of ex are ex which evaluate to at x = .


The Taylor series starts + x + x + ! x + ! x + ;

f (n) (x)


(x )n+
n!
.
(x )(n+ ) =
(n + )!
n+



Rn (x)

x
x

x
x

( )n

n=

( )n

n=

. x+x +
.

x
x

(x ) n
xn
=
.
( )n
( n)!
( n)!
n=
( x + ) n+
.
( n + )!

( )
sin x dx

Chapter
Sec on .

dx =

Sec on .

. When dening the conics as the intersec ons of a plane and a


double napped cone, degenerate conics are created when the
plane intersects the ps of the cones (usually taken as the origin).
Nondegenerate conics are formed when this plane does not
contain the origin.

. T
. rectangular

. Hyperbola

. With a horizontal transverse axis, the x term has a posi ve


coecient; with a ver cal transverse axis, the y term has a
posi ve coecient.
. y=

. x=y
. x=

10

(x + )

. x = (y ) +

. focus: ( , ); directrix: x = . The point P is

units from each.

.
x

. .

(x )
/

(x )

(y )

(x+ )

(y )

(x )

. x
.

(y )

); e =

/ .

=
+

= ; foci at ( ,

(y )

(x )

=
.

y
2
x

(x )

(y )

.
=

. (y )

(a) Solve for c in e = c/a: c = ae. Thus a e = a b , and


b = a a e . The result follows.

5
y

(b) Mercury: x /( .
) + y /( .
) =
Earth: x + y /( .
) =
Mars: x /( .
) + y /( .
) =

(c) Mercury: (x . ) /( .
) + y /( .
Earth: (x .
) + y /( .
) =
Mars: (x .
) /( .
) + y /( .

1
0.5

) =
) =

.
1

0.5

0.5
0.5

A.

10

.
5

10

(a) Traces a circle of radius counterclockwise once.


(b) Traces a circle of radius counterclockwise over

mes.

(c) Traces a circle of radius clockwise innite mes.


(d) Traces an arc of a circle of radius , from an angle of radians to radian, twice.
. x y =

dy
dx

= .

. t = ; limt

dy
dx

= .

d y
dx

d y
dx

.
.

. y=x

. y x =

. x + y = r ; circle centered at ( , ) with radius r.


(yk)

b
= ; hyperbola centered at (h, k) with
horizontal transverse axis and asymptotes with slope b/a. The
parametric equa ons only give half of the hyperbola. When
a > , the right half; when a < , the le half.

. x = ln t, y = t. At t = , x = , y = .
y = ex ; when x = , y = .
. x = /( t ), y = /( t). At t = , x = / , y = / .

y = /( x); when x = / , y = .
. t= ,

= (
( / , ).

A.

dy
dx

= t

; concave up on (, / ); concave down on

d y
= cot t; concave up on (, ); concave down on
dx
( , ).
d y
dx

( + cos( t))
(cos t+ cos( t))

, obtained with a computer algebra system;

(
)
concave up on tan ( / ), tan ( / ) , concave down

)
(
)
(
on / , tan ( / ) tan ( / ), /
=

. L .

(actual value: L = .

. L .

(actual value: L = .

. The answer is
dierent.

. SA .

(actual value SA = .

Sec on .

. Answers will vary.

)
)

for both (of course), but the integrals are

. x = cos t + , y = sin t + ; other answers possible

. x = sec t + , y =
tan t ; other answers possible

(a)

. L=

. L=

. x = cos t, y = sin t; other answers possible

t )

. t = . . . , , , . . .

. F

; normal line: y = x

= ; always concave up

. t=

. F

. T

. t = / , / , /

Sec on .

) + ; normal line:

cos t sin( t)+sin t cos( t)


sin t sin( t)+ cos t cos( t)

. t = ; limt

t +

. The solu on is non-trivial; use iden es sin( t) = sin t cos t and


cos( t) = cos t sin t to rewrite
g (t) = sin t( cos t sin t). On [ , ], sin t = when
t = , , ,
and cos t sin
t = when

t = tan ( ), tan ( ), tan ( ).

t, y =

(x

. The graph does not have a horizontal tangent line.

. x=

. t= /

(xh)
a

dy
dx

(a)

. t=

= csc t

(b) Tangent line: y = x

. x=

dy
dx

(a)

(b) t = / :
Tangent
line: y =
)+
y = / (x

. y=x

t+
t

(b) Tangent line: y = x + ; normal line: y = / x +

..

dy
dx

(a)

10

(b) Tangent line: y = (x ) + ; normal line:


y = / (x ) +

. A = P( . , / ) and P( . , / );
B = P( , / ) and P( , / );
C = P( , / ) and P( , / );
D = P( . , / ) and P( . , / );

)
. A=( ,

B = ( , )

C = P( , . )

D = P( , . )

2
2

x
8

.
y

.
2

1
4

.
.

2
4
y
2

.
5

2
4

. x + (y + ) =

. y= / x+ /
. y=

. x +y =
. = /

. r = sec

.
x

y
1

. r = cos / sin

. r=

. P( / , / ), P( , / ), P( / , / )

. P( , ) = P( , / ), P( , / )

. P( / , / ), P( / , / ), P( / , / )
. For all points, r = ; =
/ , / , / , /

.
1

. Answers will vary. If m and n do not have any common factors,


then an interval of n is needed to sketch the en re graph.

Sec on .
.

. Using x = r cos and y = r sin , we can write x = f() cos ,


y = f() sin .

(a)

dy
dx

= cot

(b) tangent line: y = (x


y=x

.
1

(a)

dy
dx

(a)

/ )+

/ ; normal line:

cos ( + sin )
cos sin ( +sin )

(b) tangent line: x =


dy
dx

/ ; normal line: y = /

cos +sin
cos sin

(b) tangent line: y = /x + / ; normal line:


y = / x + /

(a)

dy
dx

sin(t) cos( t)+sin( t) cos(t)


cos(t) cos( t)sin(t) sin( t)

(b) tangent line:


/ ) / ; normal line:
y = (x +
y= /
(x +
/ ) /

A.

. horizontal: = / , / ;
ver cal: = , ,

. horizontal: = tan ( / ), / , tan ( / ), +

tan ( / ), / , tan ( / );

), tan ( ), , +
ver cal: = , tan (
tan ( ), tan ( )
. In polar: =

= =

In rectangular: y =
. area =

. y +z =x

. z=( x +y ) =x +y

. area = /

. area =

. area = +
. area =

. area =

/
/
/

sin ( ) d

cos ( ) d =

. (a)

x=y +

. (b)

x +

( cos ) d +

(cos ) d =

. L .

; (actual value L = .

. SA =

. SA =

. SA =

Chapter
Sec on

. right hand
. curve (a parabola); surface (a cylinder)
. a hyperboloid of two sheets

. || AB || =
; || BC || =
; || AC || =
triangle as || AB || + || AC || = || BC || .

. Yes, it is a right

. Center at ( , , ); radius =

. Interior of a sphere with radius centered at the origin.


. The rst octant of space; all points (x, y, z) where each of x, y and
z are posi ve. (Analogous to the rst quadrant in the plane.)

Sec on

. Answers will vary.


. A vector with magnitude .
.

A.

. It stretches the vector by a factor of , and points it in the


opposite direc on.

#
. PQ = , = i + j
#
. PQ = , , = i + j
.

. Answers will vary; two possible answers are , , and


, , .

. projv u = / , / .

(a) u + v
= , , ; u v = , , ;
v =
u
,
,
.
(c) x = , , .

. projv u = / , / .

. projv u = , , .

. u = / , / + / , / .

. u = / , / + / , / .

. u = , , + , , .

u + v

.
.

x
v

lb

lb

lb

lb

Sec on

Sketch of u v shi ed for clarity.

. vector
. Perpendicular is one answer.

z
u

. Torque
. u v =

u + v

. u v =

. || u || =

, || v || =

, sin

, || u + v || =

, || u + v || =

, || u v || =

.
.

, .

. The force on each chain is

Sec on

; the

. , ,
. , ,
.

lb.
(about /

th of an inch).

lb

u (u v) = u , u , u (u v u v , (u v u v ), u v u v )
= u (u v u v ) u (u v u v ) + u (u v u v )

Sec on

. By considering the sign of the dot product of the two vectors. If


the dot product is posi ve, the angle is acute; if the dot product is
nega ve, the angle is obtuse.
.
.

. Answers will vary.


.

. Answers will vary; two possible answers are , and

. A point on the line and the direc on of the line.


. parallel, skew
. vector: (t) = , , + t , ,
parametric: x = + t, y = + t, z = + t
symmetric: (x )/ = (y + )/ = (z )/

. Answers can vary: vector: (t) = , , + t , ,


parametric: x = + t, y = t, z = t
symmetric: (x )/ = (y )/ = (z )

. not dened

= .

. Scalar

. = / =

. = .

lb

. With u = u , u , u and v = v , v , v , we have

lb.

weight is li ed

weight is li ed .

; the

= .

. The force on each chain is

sin cos + sin sin + cos

= sin (cos + sin ) + cos

= sin + cos

.
.

, || u v || =

|| u || =

. =

. i k = j

. u = / , / , /

. = .

. Answers will vary.

. || u || = , || v || =


. u = /
, /

. u = cos

. u v = , ,

u v
x

, ,

, .

. Answers can vary; here the direc on is given by d d : vector:


(t) = , , + t , ,
parametric: x = t, y = + t, z = + t
symmetric: x/ = (y )/ = (z )/

A.

. Answers can vary; here the direc on is given by d d : vector:


(t) = , , + t , ,
parametric: x = + t, y = t, z = + t
symmetric: x = y = (z + )/

. vector: (t) = , + t ,
parametric: x = + t, y = + t
symmetric: (x )/ = (y )/

. Answers
may vary:

x = t
=

y=
z=

. ( , , )

. parallel

. intersec ng;
( )=
( ) = , ,

. skew

. No point of intersec on; the plane and line are parallel.

. same

.
/

.
/

. /

(a) The distance formula cannot be used because since d and


d are parallel,c is and we cannot divide by || ||.
#
(b) Since d and d are parallel, P P lies in the plane formed
#
by the two lines. Thus P P d is orthogonal to this
#
plane, andc = (P P d ) d is parallel to the plane,
but s ll orthogonal to both d and d . We desire the length
#
of the projec on of P P ontoc, which is what the formula
provides.
(c) Since the lines are parallel, one can measure the distance
between the lines at any loca on on either line (just as to
nd the distance between straight railroad tracks, one can
use a measuring tape anywhere along the track, not just at
one specic place.) Let P = P and Q = P as given by the
equa ons of the lines, and apply the formula for distance
between a point and a line.

Sec on

#
. Since both P and Q are on the line, PQ is parallel to d. Thus
#
PQ d = , giving a distance of .
.

#
. If P is any point in the plane, and Q is also in the plane, then PQ
lies parallel to the plane and is orthogonal to n, the normal vector.
#
Thus n PQ = , giving the distance as .

Chapter
Sec on

. parametric equa ons


. displacement
y

. A point in the plane and a normal vector (i.e., a direc on


orthogonal to the plane).

. Answers will vary.

. Answers will vary.

. Standard form: (x ) (y ) + (z ) =
general form: x y + z =

. Answers may vary;


Standard form: (x ) + (y ) (z ) =
general form: x + y z =

. Answers may vary;


Standard form: (x ) + (y ) + (z ) =
general form: x + y + z =

.5

.5

. Answers may vary;


Standard form: (x ) (y + ) (z ) =
general form: x y z =

.
y

. Answers may vary;


Standard form: (x ) + (y ) + (z ) =
general form: x + y + z =

. Answers may vary;


Standard form: (x + ) + (y )
general form: x + y z =

. Answers may vary;


Standard form: (x ) (y ) =
general form: x y =

(z ) =

A.

. r (t) = /t , /( t + ) , sec t

. r (t) = t, sin t, t + + t + , t cos t, =


(t + ) cos t + t sin t + t +

.
5

10

10

2
r (1)

r (t) = t + , t

y
r (1)

1
2
2

1
x 1

.y

. ||r(t) || =

cos t + sin t.

. ||r(t) || = cos t + t + t .

. Answers may vary; three solu ons are


r(t) = sin t + , cos t + ,
r(t) = cos t + , sin t + and
r(t) = cos t + , sin t + .

. Answers may vary, though most direct solu ons are


r(t) = cos t + , sin t ,
r(t) = cos t + , sin t and
r(t) = sin t + , cos t .

. Answers may vary, though most direct solu ons are


r(t) = t, / (t ) + ,
r(t) = t + , / t + ,
r(t) = t + , t + and
r(t) = t + , t + .

. Answers may vary, though most direct solu on is


r(t) = cos( t), sin( t), t.

. ,

Sec on

r (t) =

t, t

. (t) = , + t ,

. t = n, where n is an integer; so
t = . . . , , , , , . . .

. r(t) is not smooth at t = / + n, where n is an integer

. Both deriva ves return t , t t , t .

.
Both deriva ves return

t et , cos t t , (t + t)et (t ) cos t sin t .

. tan t, tan t + C
. ,

. r(t) = ln |t + | + , ln | cos t| +

. r(t) = cos t + , t sin t, et t


.
.

( e )

Sec on

. Velocity is a vector, indica ng an objects direc on of travel and its


rate of distance change (i.e., its speed). Speed is a scalar.

. component
. It is dicult to iden fy the points on the graphs ofr(t) andr (t)
that correspond to each other.

. e ,
. t, ,
. ( , )

. (t) = , , + t , ,

. , ,

. The average velocity is found by dividing the displacement by the


me traveled it is a vector. The average speed is found by
dividing the distance traveled by the me traveled it is a scalar.
. One example is traveling at a constant speed s in a circle, ending
at the star ng posi on. Since the displacement is , the average
velocity is , hence || || = . But traveling at constant speed s
means the average speed is also s > .
. v(t) = , , , a(t) = , ,

. v(t) = sin t, cos t, a(t) = cos t, sin t

A.

. The posi on func on isr(t) =


t, t +
. The
y-component is when t = . ;r( . ) =
. , , meaning
the box will travel about
horizontally before it lands.

. v(t) = , cos t, a(t) = , sin t


y

Sec on

v(/4)

1.5

.
. T(t) and
N(t).

. T(t) =

t
t t+

0.5

0.5

1.5

. v(t) = t + , t + , a(t) = ,
y

v(/4)
2

4
a(/4)

cos t+sin t

. || v(t) || =
t + .
Min at t = ; Max at t = .

. Displacement: , , ; distance traveled:


average velocity: , , ; average speed:

.
.
/s

. Displacement: , ; distance traveled: .


velocity: , ; average speed: /s

, /

)/( ) + n/ .

a +b

a sin t, a cos t, b;
N(t) = cos t, sin t,

and aN =

t
+ t

At t = , aT = and

aN = ;
and aN = / .
At t = , aT = /
At t = , all accelera on comes in the form of changing the
direc on of velocity and not the speed; at t = , more
accelera on comes in changing the speed than in changing
direc on.

. aT = and aN =
At t = , aT = and aN = ;
At t = / , aT = and aN = .
The object moves at constant speed, so all accelera on comes
from changing direc on, hence aT = . a(t) is always parallel to

N(t), but twice as long, hence aN = .

; average

+ n/ seconds,

(a) Holding the crossbow at an angle of .


radians,
will hit the target . s later. (Another solu on
.
exists, with an angle of , landing . s later, but this is
imprac cal.)
(b) In the . seconds the arrow travels, a deer, traveling at
mph or .
/s, can travel . . So she needs to lead
the deer by . .

A.

, cos t, sin t;
N(t) = , sin t, cos t

t
+ t

. aT =

(a) r ( ) = , ;r ( ) = ,

(b) v ( ) = , ; || v ( ) || = ; a ( ) = ,
v ( ) = , ; || v ( ) || =
; a ( ) = ,

. v(t) = t + , t + ,r(t) = t + t + , t / + t

. T(t) =

(a) r ( ) = , ;r ( ) = ,

(b) v ( ) = , ; || v ( ) || = ; a ( ) = ,
v ( ) = , ; || v ( ) || =
; a ( ) = ,

. v(t) = sin t, cos t,r(t) = cos t, sin t

(a) Be sure to show work

(b)
N( ) = ,

. T(t) =

. || v(t) || = .
speed is constant, so there is no dierence between min/max

t + + t /( t ).
. || v(t) || =
min: t = ; max: there is no max; speed approaches as
t

. At t-values of sin ( /
where n is an integer.

; T( ) =

cos t+sin t

(a) Be sure to show work

, /
(b)
N(/ ) = /

. || v(t) || = .
Speed is constant, so there is no dierence between min/max

. || v(t) || = | sec t| tan t + sec t.


min: t = ; max: t = /

. T(t) = sin t, cos t;


N(t) = cos t, sin t

. T(t) = sin t
;
, cos t
cos t+sin t
cos t+sin t

, sin t
N(t) = cos t

cos t sin t
cos t sin t

t
t t+

cos t, sin t. (Be careful; this cannot be

t sin t = cos t sin t, but


simplied as just cos t, sin t as cos
rather | cos t sin t|.) T(/ ) = / ,
/


, /
; in parametric form,
. (t) = , + t /

{
x =
+
t/
(t) =
y =
t/

/ ,
/ +t / ,
/ ; in parametric form,
. (t) =

{
x = / t/
(t) =
y =
/ +
t/

. T(t) =

a(/4)

. aT = and aN = a
At t = , aT = and aN = a;
At t = / , aT = and aN = a.
The object moves at constant speed, meaning that aT is always .
The object rises along the z-axis at a constant rate, so all
accelera on comes in the form of changing direc on circling the
z-axis. The greater the radius of this circle the greater the
accelera on, hence aN = a.

Sec on
.

me and/or distance

. Answers may include lines, circles, helixes


.
. s = t, sor(s) = s/ , s/ , s/

t, sor(s) = cos(s/
), sin(s/

. s=

), s/

. =

| x|

+( x )

( ) = , ( / ) =
| cos x|

. =

.
2

c = 2

( +sin x)
( ) = , (/ ) =

. =

| cos t cos( t)+ sin t sin( t)|

cos ( t)+sin t)

c=0

c=2
4

;
2

( ) = / , (/ ) =
. =

| t + |

t +( t )

.
.

( ) = , ( ) =

. Level curves are circles, centered at ( /c, /c) with radius


/c . When c = , the level curve is the line y = x.

. = ;

( ) = , ( ) =
. =

c=

( ) = /

c=

, (/ ) = /

. maximized at x =

. maximized at t = /

. radius of curvature is

/ .

. radius of curvature is .
. x + (y / ) = / , orc(t) = / cos t, / sin t + /
. x + (y + ) =

, orc(t) = cos t, sin t

Chapter
Sec on

c=

. Level curves are ellipses of the form


and b = c/ .
4

x
c

y
c /

= , i.e., a = c

. Answers will vary.


4

. topographical

. surface

. domain: R
range: z

. domain: x + y z = ; the set of points in R NOT in the


domain form a plane through the origin.
range: R

. domain: R
range: R
. domain: R
range: < z
. domain: {(x, y) | x + y }, i.e., the domain is the circle and
interior of a circle centered at the origin with radius .
range: z
. Level curves are lines y = ( / )x c/ .

.
The level surfaces are spheres, centered at the origin, with radius
c.
x
c

. The level curves for each surface are similar; for z =

the level curves are ellipses of the form

. domain: z x y ; the set of points in R above (and


including) the hyperbolic paraboloid z = x y .
range: [ , )

. The level surfaces are paraboloids of the form z =


larger c, the wider the paraboloid.

.
. Level curves are parabolas x = y + c.

x
c

y
c /

y
c

; the

x + y

= , i.e., a = c

and b = c/ ; whereas for z = x + y the level curves are

y
ellipses of the form xc + c/
= , i.e., a = c and b = c/ .
The rst set of ellipses are spaced evenly apart, meaning the
func on grows at a constant rate; the second set of ellipses are
more closely spaced together as c grows, meaning the func on
grows faster and faster as c increases.

The func on z = x + y can be rewri en as z = x + y ,


an ellip c cone; the func on z = x + y is a paraboloid, each
matching the descrip on above.

A.

Sec on

. fx = xy + x, fy = x +
fxx = y + , fyy =
fxy = x, fyx = x

. Answers will vary.


. Answers will vary.
One possible answer: {(x, y)|x + y }

. fx = /y, fy = x/y
fxx = , fyy = x/y
fxy = /y , fyx = /y

. fx = xex +y , fy = yex +y
fxx = ex +y + x ex +y , fyy = ex
fxy = xyex +y , fyx = xyex +y

. Answers will vary.


One possible answer: {(x, y)|x + y < }
(a) Answers will vary.
interior point: ( , )
boundary point: ( , )
(c) S is bounded
(a) Answers will vary.
interior point: none
boundary point: ( , )

(b) S is a closed set, consis ng only of boundary points


(c) S is bounded
{
}
(a) D = (x, y) | x y .

fxx =

fxy =

(a) Along y = , the limit is .


(b) Along x = , the limit is .
Since the above limits are not equal, the limit does not exist.
mx( m)
.
m x+
(b) Along x = , the limit is .
Since the above limits are not equal, the limit does not exist.
(a) Along y = mx, the limit is

lim

x+y
x

x
= lim
x x
= lim

x+

= / .
(b) Along y = x + , the limit is:
lim

(x,y)( , )

x+y
x

x+

= .
Since the limits along the lines y =
overall limit does not exist.

and y = x + dier, the

. A constant is a number that is added or subtracted in an


expression; a coecient is a number that is being mul plied by a
nonconstant func on.
. fx
. fx = xy , fy = x +
fx ( , ) = , fy ( , ) =
. fx = sin x sin y, fy = cos x cos y
fx (/ , / ) = / , fy (/ , / ) = /

A.

xy +

, fyy =
+

y
xy +

x y
xy +

x
xy +

, fyx =

xy

xy +

x
, fy = (x +y y+ )
+y + )
x
(x +y + ) , fyy = (x +yy + )
(x +y + )
xy
, fyx = (x +yxy+ )
(x +y + )

y
xy +

(x +y + )

. fx = x, fy =
fxx = , fyy =
fxy = , fyx =
. fx =

xy

fxx =

, fy = lny x
x y

fxy =

xy

, fyy =

ln x
y

, fyx =

xy

. f(x, y) = x y xy + y + C, where C is any constant.

. fx = xe y z , fy = x e y z , fz = x e
fyz = x e y z , fzy = x e y z
. fx =

fyz =

Sec on
(x )
= lim
x
x
= lim

Sec on

xy +
xy

xy
xy +

. f(x, y) = x sin y + x + C, where C is any constant.

(a) Along y = , the limit is:


(x,y)( , )

, fy =

. fx = (x

(c) D is unbounded.

y
xy +
y

fxy =

(b) D is an open set.

+y

sin( xy ), fy = xy sin( xy )
y cos( xy ),
x y cos( xy ) xy sin( xy )
xy cos( xy ) y sin( xy ),
xy cos( xy ) y sin( xy )

fxx =

(c) D is bounded.
}
{
(a) D = (x, y) | y > x .

. fx = y
fxx =
fyy =
fxy =
fyx =
. fx =

(b) D is a closed set.


.

+ y ex

. fx = cos x cos y, fy = sin x sin y


fxx = sin x cos y, fyy = sin x cos y
fxy = sin y cos x, fyx = sin y cos x

(b) S is a closed set


.

+y

, f = yx z , fz
y z y
x
, fzy = y xz
y z

y z

x
y z

. T
. T
. dz = (sin y + x)dx + (x cos y)dy
. dz = dx dy

. dz = x

x +y

dx +

x +y

dy, with dx = .

( , ), dz = / ( . ) + / (. ) = .
f( . , . ) .
+ = .
.

and dy = . . At
, so

. dz = ( xy y )dx + (x xy)dy, with dx = . and


dy = . . At ( , ), dz = ( . ) + ( )( . ) = .
f( . , . ) . = . .

, so

. The total dieren al of volume is dV = dr + dh. The


coecient of dr is greater than the coecient of dh, so the
volume is more sensi ve to changes in the radius.
. Using trigonometry, = x tan , so d = tan dx + x sec d.
With = and x = , we have d = . dx +
. d.
The measured length of the wall is much more sensi ve to errors

in than in x. While it can be dicult to compare sensi vi es


between measuring feet and measuring degrees (it is somewhat
like comparing apples to oranges), here the coecients are so
dierent that the result is clear: a small error in degree has a
much greater impact than a small error in distance.
. dw = xyz dx + x z dy + x yz dz
. dx = . , dy = . . dz = (.
f( . , . ) + . = . .

) + ( )( . ) = .

. dx = . , dy = . , dz = . .
dw = ( . ) + ( )( . ) + . ( . ) = .
f( . , . , . ) . = . .

Sec on

. So

, so

. Because the parametric equa ons describe a level curve, z is


constant for all t. Therefore dz
= .
dt
.

dx
,
dt

and

f
y

. F
.

(a)

dz
dt

= ( t) + ( ) = t + .

(b) At t = ,

dz
dt

(a)

(b)

. f = xy + y + y, x + xy + x
(a) f( , ) = ,

dz
dt

. t = / ; this corresponds to a minimum

(a)

dz
dt

z
s
z
t

dz
= cos t sin t.
dt
when t = n or n + / , where n is any integer.

= xy( ) + x ( ) = xy + x ;
= xy( ) + x ( ) = xy + x

(b) With s = , t = , x =
z
=
t

and y = . Thus

(a)

(c) ,


/ , /

x
, (x
. f = (x +y
+ )

z
s

and

(x + y )( x) (x + y)( )
,
(x + y )
(x + y )( ) (x + y)( y)
;
fy =
(x + y )
x(x + y ) (x + y)
dy
=
dx
x + y y(x + y)

= ( ) + ( ) = .

z
s
z
t

= ( )+ ( )=

= ( ) + ( ) =

Sec on

(a) f( , ) = / , / .

(c) / , /


/ , /

. f = x y , y x
(a) f( , ) = ,

(b)

(c) ,

.
.

. fx =

dz
dt

y
+y + )

(b) || f( , ) || = || / , / || =

z
s
z
t

. fx = x tan y, fy = x sec y;
dy
tan y
=
dx
x sec y

(d)

= x(cos t) + y(sin t) = x cos t + y sin t;


= x(s sin t) + y(s cos t) = xs sin t + ys cos t

=
and y =
. Thus z
(b) With s = , t = / , x =
s
and z
=
t

(b) || f( , ) || = || , || =

(d)

. t = tan ( / ) + n, where n is an integer

Thus

(a) / (u = / , / )

(u = / ,
)
(b)

y
x
, (x +y
; f( , ) = / , / . Be
. f = (x +y
+ )
+ )
sure to change all direc ons to unit vectors.


)
(a) (u = / , /

/ (u = / , /
)
(b)

sin t + cos t

(a)

. We nd that

. f = x y , y x

. f = xy + y + y, x + xy + x ; f( , ) = , . Be
sure to change all direc ons to unit vectors.

(a)

dz
dt

dz
= x(cos t) + y( cos t).
dt

/ ,y =
/ , and
(b) At t = / , x =

. maximal, or greatest

. f = xy + y + y, x + xy + x

y
x
. f = (x +y
, (x +y
+ )
+ )

. f = x y , y x; f( , ) = , .

= ( sin t) + (cos t) =

=
.
(b) At t = / , dz
dt

. u = ,

. A par al deriva ve is essen ally a special case of a direc onal


deriva ve; it is the direc onal deriva ve in the direc on of x or y,
i.e., , or , .

(d) All direc ons give a direc onal deriva ve of .

(a) F(x, y, z) = xz + y, x, x z z

(b)
/

(a) F(x, y, z) = xy , y(x z ), y z


(b)

Sec on

. Answers will vary. The displacement of the vector is one unit in


the x-direc on and units in the z-direc on, with no change in y.
Thus along a line parallel to v, the change in z is
mes the
change in x i.e., a slope of . Specically, the line in the x-z
plane parallel to z has a slope of .
. T
.

x= +t
y=
(a) x (t) =

z=

x=
y= +t
(b) y (t) =

z=

A.

x = + t/
(c) u (t) =
y = + t/

/ t
z=

x= +t
y=
.
(a) x (t) =

z= + t

x=
y= +t
(b) y (t) =

z= t

x = + t/
(c) u (t) =
y = +
t/

z=
t

x= t
y= t
. n (t) =

z= t

x= + t
y= t
. n (t) =

z= t
. ( .

, .

. ( .

, .

.
.

, .

(x )

), ( .

) and ( .

, .

, .

, .

(y ) (z +

. The triangle is bound by the lines y = , y = x + and


y= x+ .
Along y = , there is a cri cal point at ( , ).
Along y = x + , there is a cri cal point at ( / , / ).
Along y = x + , there is a cri cal point at ( / , / ).
The func on f has one cri cal point, irrespec ve of the constraint,
at ( , / ).
Checking the value of f at these four points, along with the three
ver ces of the triangle, we nd the absolute maximum is at
( , , ) and the absolute minimum is at ( , / , / ).
. The region has no corners or ver ces, just a smooth edge.
To nd cri cal points along the circle x + y =
, we solve for y :
y = x . We can go further and state y =
x .
We can rewrite f as

f(x) = x + x + (
x = x+ +
x . (We
x ) +

will return
and use

x later.) Solving f (x) = , we get


x=
y=
. f (x) is also undened at x = , where
y= .

Using y =

x , we rewrite f(x, y) as
f(x) =x + x . Solving f (x) = , we get
x= , y= .

)=

(x ) (y ) (z ) = (Note that this tangent plane


is the same as the original func on, a plane.)

/ ,
/
. F = x/ , y/ , z/ ; at P, F = / ,

x = + t/
y = + t/
(a) n (t) =

z=
+
t/

(y
(z
)+
)= .
(b) (x ) +

. F = y z , xy, xz ; at P, F = , ,

x=
y= + t
(a) n (t) =

z= + t
(b)

(y ) + (z + ) = .

Sec on

The func on f itself has a cri cal point


at (
, ).

Checking the value of f at ( , ), ( ,


), ( , ),
( , ) and ( , ), we nd the absolute maximum is at ( , , )
and the absolute minimum is at ( , , ).

Chapter
Sec on

. C(y), meaning that instead of being just a constant, like the


number , it is a func on of y, which acts like a constant when
taking deriva ves with respect to x.
. curve to curve, then from point to point
.

x +

(a)
(b)

(a) x / x + x /

(b)

(a) sin y

(b) /


dy dx and
.

. F; it is the other way around.


. T

. One cri cal point at ( , ); fxx = and D = , so this point


corresponds to a rela ve minimum.
. One cri cal point at ( , ); D = , so this point corresponds
to a saddle point.
. Two cri cal points: at ( , ); fxx = and D = , so this point
corresponds to a saddle point;
at ( , ), fxx = and D = , so this corresponds to a rela ve
minimum.
. One cri cal point at ( , ). D = x y , so at ( , ), D = and
the test is inconclusive. (Some elementary thought shows that it
is the absolute minimum.)

y
(x ) y ) /

and D =

(x ) y )

dx dy.

x is bounded above by two dierent func ons. This gives:


y
y+
dx dy.
dx dy +

area of R = u

x
dy dx and
.
x

area of R = /

dx dy

y
4

Both fx and fy are undened along the circle (x ) + y = ;


at any point along this curve, f(x, y) = , the absolute minimum
of the func on.

area of R = u
x
dy dx. The order dx dy needs two iterated integrals as
.

. One cri cal point: fx = when x = ; fy = when y = , so one


cri cal point at ( , ), which is a rela ve maximum, where
fxx =

A.

y = 4 x2

R
2

area of R =

dx dy

(a)

R
.

x2 /16 + y2 /4 = 1

(b)
2

(c)

x /

area of R =

x y dy dx =

x y dx dy.

dy dx

x /

y
x+

.
y=

(a)

(b)

y=x

(c)

area of R =

Sec on

x+

dy dx

x+

y=

(d)

.
/ x

/ y

)
x y dy dx =

)
x y dx dy.

y
3

. volume
. The double integral gives the signed volume under the surface.
Since the surface is always posi ve, it is always above the x-y
plane and hence produces only posi ve volume.

R
.

(a)
3

/ ;

/ ;

x
+
y
y

)
(

dy dx

x y+

dx dy
(b)

(x + y + ) dy dx

(c)

y= x

y = x2

x
y
y

)
x y x dy dx =

)
x y x dx dy.
x

x y dy dx =

. average value of f = / =
. average value of f =

x y dx dy.

ex dy dx = e .

y
dx gives tan ( /y) / ; integra ng
x
+
y
y
tan ( /y) is hard.
x
y
dy dx = ln .
x +y

. Integra ng

(c)

. Integra ng ex with respect


to x is not possible in terms of

(a)

(b)

elementary func ons.

Sec on

(
)
. f r cos , r sin , r dr d

A.

.
.

/
/

r dr d =

)
r sin r dr d =

)
(
ln(r ) r dr d = ln

cos

cos

/
/

r cos r sin +

. SA =

SA =

=
=

)
r cos( ) r dr d =

volume under the surface


plane.

e(x +y )

SA =

. (x, y) = ( , )
gm;

Sec on

kg

lb

gm; My =

; Mx =

kg; My = ; Mx = ; (x, y) = ( / , / )

lb; My = ; Mx =

. Ix =

/ ; Iy =

/ ; IO =

. Ix =

/ ; Iy =

/ ; IO =

; (x, y) = ( , .

. SA =

. V=
. V=

+ x + y dx dy

x(

dz dx dy:

. Intui vely, adding h to f only shi s f up (i.e., parallel to the z-axis)


and does not change its shape. Therefore it will not change the
surface area over R.
Analy cally, fx = gx and fy = gy ; therefore, the surface area of
each is computed with iden cal double integrals.

+ cos x cos y + sin x sin y dx dy
. SA =

r
+ c + d dr d

+c +d

+c +d .

dy

. dz dy dx:

. surface areas

. Answers can vary. From this sec on we used triple integra on to


nd the volume of a solid region, the mass of a solid, and the
center of mass of a solid.

. arc length

A.

r cos t + r sin t
dr d
r sin t + r cos t

. surface to surface, curve to curve and point to point

; (x, y) = ( , . )

. M = lb; My = ; Mx = / ; (x, y) = ( , / )

Sec on

The value of h does not ma er as it only shi s the plane ver cally
(i.e., parallel to the z-axis). Dierent values of h do not create
dierent ellipses in the plane.

. M = lb

. M=

+ + x dy dx

dr d
r

SA =

. x= .

. M=

. Integra ng in polar is easiest considering R:

. Mx measures the moment about the x-axis, meaning we need to


measure distance from the x-axis. Such measurements are
measures in the y-direc on.

. M=

)
(
+ x dx
x

. li le masses

+ r dr d

. This is easier in polar:

over the en re x-y

. Because they are scalar mul ples of each other.

. M=

. M=

+ r cos t + r sin t dr d

. M=

lim ( ea ) = . This implies that there is a nite

Sec on

SA =

(a) This is impossible to integrate with rectangular coordinates


as e(x +y ) does not have an an deriva ve in terms of
elementary func ons.
a
rer dr d = ( ea ).
(b)
(c)

)
r cos + r sin r dr d =

. This is easier in polar:

r cos r sin r dr d =

dx dy =

+ +

( )
r dr d =

dy dz dx:
dy dx dz:
dx dz dy:

)
x y ( x + y) dx dy =

x/

)
cos x sin y + sin x cos y dy dx = .

dx dy dz:

V=

x/

x/
z/
y
z/

x/ y

x/ y

dz dx dy

x/ z/
x/ z/
y z/
y z/

x/ y

dz dy dx

dy dz dx
dy dx dz

dx dz dy
dx dy dz

dz dy dx = .

. dz dy dx:
dz dx dy:
dy dz dx:
dy dx dz:
dx dz dy:
dx dy dz:
V=

. dz dy dx:
dz dx dy:
dy dz dx:
dy dx dz:
dx dz dy:
dx dy dz:

/
yz

z
z

x/

y
x

z/

dz dx dy

z
y

yz

z/

z/

dy dz dx

dz dx dy:

dy dx dz

dy dz dx:

dx dz dy
dy dx dz:
dx dy dz
dx dz dy:

x+ y
x+ y

y+

z/ y+

dx dy dz:

dz dy dx

dy dz dx

dy dx dz

dx dz dy
dx dy dz

z/

yz/
x

dz dy dx

dz dx dy

dy dz dx +
x

dy dx dz +

dx dy dz = / .

z
x

dy dz dx
dy dx dz

dx dz dy
dx dy dz

Answers will vary. Neither order is par cularly hard. The order
dz dy dx requires integra ng a square root, so powers can be
messy; the order dy dz dx requires two triple integrals, but each
uses only polynomials.

dz dx dy

x/ +

z/ x/ +

z/

. dz dy dx:

dy dz dx = / .

V=

dz dy dx

/
y y

. M = , Myz = / , Mxz = / , Mxy = ;


(x, y, z) = ( / , / , / )
. M = / , Myz = / , Mxz =
/ , Mxy =
(x, y, z) = ( / , / , / ) ( . , . , .

/ ;
)

A.

Index
!,
Absolute Convergence Theorem,
absolute maximum,
absolute minimum,
Absolute Value Theorem,
accelera on, ,
Alterna ng Harmonic Series,
,
Alterna ng Series Test
for series,
aN ,
,
analy c func on,
angle of eleva on,
an deriva ve,
arc length,
,
,
,
,
arc length parameter,
,
asymptote
horizontal,
ver cal,
aT ,
,
average rate of change,
average value of a func on,
average value of func on,
Binomial Series,
Bisec on Method,
boundary point,
bounded sequence,
convergence,
bounded set,
center of mass,

,
Chain Rule,
mul variable,
,
nota on,
circle of curvature,
closed,
closed disk,
concave down,
concave up,
concavity,
,
inec on point,
test for,
conic sec ons,
degenerate,
ellipse,
hyperbola,
parabola,
Constant Mul ple Rule
of deriva ves,
of integra on,
of series,

constrained op miza on,


con nuous func on, ,
proper es, ,
vectorvalued,
contour lines,
convergence
absolute,
,
Alterna ng Series Test,
condi onal,
Direct Comparison Test,
for integra on,
Integral Test,
interval of,
Limit Comparison Test,
for integra on,
nth term test,
of geometric series,
of improper int.,
,
,
of monotonic sequences,
of p-series,
of power series,
of sequence,
,
of series,
radius of,
Ra o Comparison Test,
Root Comparison Test,
cri cal number,
cri cal point,
,

cross product
and deriva ves,
applica ons,
area of parallelogram,
torque,
volume of parallelepiped,
deni on,
proper es,
,
curvature,
and mo on,
equa ons for,
of circle,
,
radius of,
curve
parametrically dened,
rectangular equa on,
smooth,
curve sketching,
cusp,
cycloid,
cylinder,
decreasing func on,

nding intervals,
strictly,
denite integral,
and subs tu on,
proper es,
deriva ve
accelera on,
as a func on,
at a point,
basic rules,
Chain Rule, ,
,
,
Constant Mul ple Rule,
Constant Rule,
dieren al,
direc onal,
,
,
,
,
exponen al func ons,
First Deriv. Test,
Generalized Power Rule,
higher order,
interpreta on,
hyperbolic funct.,
implicit,
,
interpreta on,
inverse func on,
inverse hyper.,
inverse trig.,
Mean Value Theorem,
mixed par al,
mo on,
mul variable dieren ability,
,
normal line,
nota on, ,
parametric equa ons,
par al,
,
Power Rule, , ,
power series,
Product Rule,
Quo ent Rule,
Second Deriv. Test,
Sum/Dierence Rule,
tangent line,
trigonometric func ons,
vectorvalued func ons,
,
,
velocity,
dieren able, ,
,
dieren al,
nota on,
Direct Comparison Test
for integra on,
for series,
direc onal deriva ve,
,
,
,
,
directrix,
,
Disk Method,
displacement,
,
,
distance
between lines,
between point and line,
between point and plane,
between points in space,
traveled,

divergence
Alterna ng Series Test,
Direct Comparison Test,
for integra on,
Integral Test,
Limit Comparison Test,
for integra on,
nth term test,
of geometric series,
of improper int.,
,
,
of p-series,
of sequence,
of series,
Ra o Comparison Test,
Root Comparison Test,
dot product
and deriva ves,
deni on,
proper es,
,
double integral,
,
in polar,
proper es,
eccentricity,
,
elementary func on,
ellipse
deni on,
eccentricity,
parametric equa ons,
reec ve property,
standard equa on,
extrema
absolute,
,
and First Deriv. Test,
and Second Deriv. Test,
nding,
rela ve,
,
,
Extreme Value Theorem,
,
extreme values,
factorial,
First Deriva ve Test,
oor func on,
uid pressure/force,
,
focus,
,
,
Fubinis Theorem,
func on
of three variables,
of two variables,
vectorvalued,
Fundamental Theorem of Calculus,
and Chain Rule,
Gabriels Horn,
Generalized Power Rule,
geometric series,
,
gradient,
,
,
,
and level curves,
and level surfaces,
Harmonic Series,

Head To Tail Rule,


Hookes Law,
hyperbola
deni on,
eccentricity,
parametric equa ons,
reec ve property,
standard equa on,
hyperbolic func on
deni on,
deriva ves,
iden es,
integrals,
inverse,
deriva ve,
integra on,
logarithmic def.,
implicit dieren a on,
,
improper integra on,
,
increasing func on,
nding intervals,
strictly,
indenite integral,
indeterminate form, , ,
,
inec on point,
ini al point,
ini al value problem,
Integral Test,
integra on
arc length,
area,
,
,
area between curves,
,
average value,
by parts,
by subs tu on,
denite,
and subs tu on,
proper es,
Riemann Sums,
displacement,
distance traveled,
double,
uid force,
,
Fun. Thm. of Calc.,
,
general applica on technique,
hyperbolic funct.,
improper,
,
,
,
indenite,
inverse hyper.,
iterated,
Mean Value Theorem,
mul ple,
nota on,
,
,
,
numerical,
Le /Right Hand Rule,
,
Simpsons Rule,
,
,
Trapezoidal Rule,
,
,
of mul variable func ons,
of power series,

of trig. func ons,


of trig. powers,
,
of vectorvalued func ons,
par al frac on decomp.,
Power Rule,
Sum/Dierence Rule,
surface area,
,
,
trig. subst.,
triple,
,
,
volume
cross-sec onal area,
Disk Method,
Shell Method,
,
Washer Method,
,
work,
interior point,
Intermediate Value Theorem,
interval of convergence,
iterated integra on,
,
,
,
changing order,
proper es,
,
LHpitals Rule,
,
lamina,
Le Hand Rule,
,
,
Le /Right Hand Rule,
level curves,
,
level surface,
,
limit
Absolute Value Theorem,
at innity,
deni on,
dierence quo ent,
does not exist, ,
indeterminate form, , ,
,
LHpitals Rule,
,
le handed,
of innity,
of mul variable func on,
,
of sequence,
of vectorvalued func ons,
one sided,
proper es, ,
pseudo-deni on,
right handed,
Squeeze Theorem,
Limit Comparison Test
for integra on,
for series,
lines,
distances between,
equa ons for,
intersec ng,
parallel,
skew,
logarithmic dieren a on,

Maclaurin Polynomial, see Taylor Polynomial


deni on,
Maclaurin Series, see Taylor Series

deni on,
magnitude of vector,
mass,
,
,
center of,
maximum
absolute,
,
and First Deriv. Test,
and Second Deriv. Test,
rela ve/local,
,
,
Mean Value Theorem
of dieren a on,
of integra on,
Midpoint Rule,
,
minimum
,
absolute,
and First Deriv. Test,
,
rela ve/local,
,
,
moment,
,
,
monotonic sequence,
mul ple integra on, see iterated integra on
mul variable func on,
,
con nuity,

,
,
dieren ability,
,
,
,
domain,
,
level curves,
level surface,
limit,
,
,
range,
,
Newtons Method,
norm,
normal line, ,
,
normal vector,
nth term test,
numerical integra on,
Le /Right Hand Rule,
Simpsons Rule,
,
error bounds,
Trapezoidal Rule,
,
error bounds,

open,
open ball,
open disk,
op miza on,
constrained,
orthogonal,
,
decomposi on,
orthogonal decomposi on of vectors,
orthogonal projec on,
oscula ng circle,
p-series,
parabola
deni on,
general equa on,
reec ve property,
parallel vectors,
Parallelogram Law,
parametric equa ons
arc length,

concavity,
deni on,
nding ddxy ,
nding dy
dx ,
normal line,
surface area,
tangent line,
par al deriva ve,
,
high order,
meaning,
mixed,
second deriva ve,
total dieren al,
,
perpendicular, see orthogonal
planes
coordinate plane,
distance between point and plane,
equa ons of,
introduc on,
normal vector,
tangent,
point of inec on,
polar
coordinates,
func on
arc length,
gallery of graphs,
surface area,
func ons,
area,
area between curves,
nding dy
dx ,
graphing,
polar coordinates,
plo ng points,
Power Rule
dieren a on, , , ,
integra on,
power series,
algebra of,
convergence,
deriva ves and integrals,
projec le mo on,
,
,
quadric surface
deni on,
ellipsoid,
ellip c cone,
ellip c paraboloid,
gallery,

hyperbolic paraboloid,
hyperboloid of one sheet,
hyperboloid of two sheets,
sphere,
trace,
Quo ent Rule,
R,
radius of convergence,
radius of curvature,

Ra o Comparison Test
for series,
rearrangements of series,
,
related rates,
Riemann Sum,
,
,
and denite integral,
Right Hand Rule,
,
,
right hand rule
of Cartesian coordinates,
Rolles Theorem,
Root Comparison Test
for series,
saddle point,
,
Second Deriva ve Test,
,
sensi vity analysis,
sequence
Absolute Value Theorem,
posi ve,
sequences
boundedness,
convergent,
,
,
deni on,
divergent,
limit,
limit proper es,
monotonic,
series
absolute convergence,
Absolute Convergence Theorem,
alterna ng,
Approxima on Theorem,
Alterna ng Series Test,
Binomial,
condi onal convergence,
convergent,
deni on,
Direct Comparison Test,
divergent,
geometric,
,
Integral Test,
interval of convergence,
Limit Comparison Test,
Maclaurin,
nth term test,
p-series,
par al sums,
power,
,
deriva ves and integrals,
proper es,
radius of convergence,
Ra o Comparison Test,
rearrangements,
,
Root Comparison Test,
Taylor,
telescoping,
,
Shell Method,
,
signed area,
signed volume,
,
Simpsons Rule,
,

error bounds,
smooth,
smooth curve,
speed,
sphere,
Squeeze Theorem,
Sum/Dierence Rule
of deriva ves,
of integra on,
of series,
summa on
nota on,
proper es,
surface area,
solid of revolu on,
,
surface of revolu on,
,

tangent line, ,
,
,
direc onal,
tangent plane,
Taylor Polynomial
deni on,
Taylors Theorem,
Taylor Series
common series,
deni on,
equality with genera ng func on,
Taylors Theorem,
telescoping series,
,
terminal point,
total dieren al,
,
sensi vity analysis,
total signed area,
trace,
Trapezoidal Rule,
,
error bounds,
triple integral,
,
,
proper es,
unbounded sequence,
unbounded set,
unit normal vector
aN ,
and accelera on,
,
and curvature,
deni on,
in R ,
unit tangent vector
and accelera on,
,
and curvature,
,
aT ,
deni on,
in R ,
unit vector,
proper es,
standard unit vector,
unit normal vector,
unit tangent vector,
vectorvalued func on
algebra of,

arc length,
average rate of change,
con nuity,
deni on,
deriva ves,
,
,
describing mo on,
displacement,
distance traveled,
graphing,
integra on,
limits,
of constant length,
,
,
projec le mo on,
,
smooth,
tangent line,
vectors,
algebra of,
algebraic proper es,
component form,
cross product,
,
,
deni on,
dot product,

Head To Tail Rule,


magnitude,
norm,
normal vector,
orthogonal,
orthogonal decomposi on,
orthogonal projec on,
parallel,
Parallelogram Law,
resultant,
standard unit vector,
unit vector,
,
zero vector,
velocity, ,
volume,
,
,
Washer Method,
work,
,

Dieren a on Rules
.

d
(cx) = c
dx

d x
(a ) = ln a ax
dx

d
(u v) = u v
dx

d
(ln x) =
dx
x

d
(loga x) =

dx
ln a x

d
(sin x) = cos x
dx

d
(u v) = uv + u v
dx
( )
vu uv
d u
=
.
dx v
v

d ( )
sin x =
dx
x

d ( )

cos x =
dx
x

d ( )

csc x =
dx
|x| x

d ( )
sec x =
dx
|x| x
d ( )
tan x =
dx

d
(sech x) = sech x tanh x
dx

d
(csch x) = csch x coth x
dx

d
(coth x) = csch x
dx

d
(u(v)) = u (v)v
dx

d
(cos x) = sin x
dx

d
(c) =
dx

d
(csc x) = csc x cot x
dx

d
(x) =
dx

d
(sec x) = sec x tan x
dx

d
(cosh x) = sinh x
dx

d n
(x ) = nxn
dx

d
(tan x) = sec x
dx

d
(sinh x) = cosh x
dx

d
.
(cot x) = csc x
dx

d x
(e ) = ex
.
dx

d ( )
cot x =
dx

+x

+x

d
.
(tanh x) = sech x
dx

)
d (
cosh x =
dx
x
)
d (
sinh x =
dx
x +

)
d (

sech x =
dx
x
x

)
d (

csch x =
dx
|x|
+x
)
d (
tanh x =
dx

)
d (
.
coth x =
dx

x
x

Integra on Rules
.

c f(x) dx = c

f(x) dx

f(x) g(x) dx =

f(x) dx g(x) dx

.
dx = C
.

dx = x + C
xn dx =

n+

xn+ + C, n =

n =

.
ex dx = ex + C

tan x dx = ln | cos x| + C

sec x dx = ln | sec x + tan x| + C

csc x dx = ln | csc x + cot x| + C

cot x dx = ln | sin x| + C

sinh x dx = cosh x + C

sec x dx = tan x + C

tanh x dx = ln(cosh x) + C

csc x dx = cot x + C

coth x dx = ln | sinh x| + C

sec x tan x dx = sec x + C

csc x cot x dx = csc x + C



a + x
+C
ln
a x
a x
)
(

+C
dx = ln
.
a
x a x
a+ a x



x
+C

.
dx = ln
a
x x +a
a+ x +a

dx = ln |x| + C

cos x dx =

x+

sin

cos x dx = sin x + C

sin x dx =

sin

sin x dx = cos x + C

ax dx =

ln a

ax + C

( )
x
+C
a
a x
( )

|x|

.
+C
dx = sec
a
a
x x a

.
cosh x dx = sinh x + C
.

x +a

dx =

tan

( )
x +C

( )
x +C

( )
x
+C
a

dx = sin

x a
x +a



dx = ln x + x a + C



dx = ln x + x + a + C

dx =

The Unit Circle

Deni ons of the Trigonometric Func ons

Unit Circle Deni on

,
)

,
)

( , )

/
/

,
(

)
,

(x, y)
)

sin = y

cos = x

csc =

sec =

tan =

( , )

,
(
)
,

/
(

( , )

(
,
)

= sec x

+ cot x = csc x

,
)

e
us

en
ot

p
Hy

x = sec x
)
sec
x = csc x
)
(
x = tan x
cot
csc

PowerReducing Formulas
cos x
sin x =

sin x sin y = sin

cos x =

cos x + cos y = cos

xy

cos x cos y = sin

x+y

cos

x+y

cos

x+y

sin

sin x cos y =

xy

xy

Product to Sum Formulas


(
)
cos(x y) cos(x + y)
sin x sin y =

cos x cos y =

O
H

csc =

H
O

cos =

A
H

sec =

H
A

tan =

O
A

cot =

A
O

es

Sum to Product Formulas


(
)
(
)
x+y
xy
sin x + sin y = sin
cos
(

cot =

sin =

Adjacent

Cofunc on Iden es
(
)
sin
x = cos x
(
)
cos
x = sin x
)
(
x = cot x
tan

es

sin x + cos x =
tan x +

Common Trigonometric Iden


Pythagorean Iden

Opposite

Right Triangle Deni on

/
(

( , )

y
x

cos(x y) + cos(x + y)

sin(x + y) + sin(x y)

tan x =

Double Angle Formulas


sin x = sin x cos x
cos x = cos x sin x
= cos x
=

tan x =

+ cos x
cos x
+ cos x

sin x
tan x
tan x

Even/Odd Iden

es

sin(x) = sin x
cos(x) = cos x

tan(x) = tan x

csc(x) = csc x
sec(x) = sec x

cot(x) = cot x

)
Angle Sum/Dierence Formulas
sin(x y) = sin x cos y cos x sin y

cos(x y) = cos x cos y sin x sin y


tan x tan y
tan(x y) =
tan x tan y

x
x
y

Areas and Volumes


Triangles
h = a sin

Right Circular Cone


c

Area = bh

Law of Cosines:
c = a + b ab cos

Parallelograms

Volume = r h
Surface Area =

r r + h + r

h
r

Right Circular Cylinder

Area = bh

Volume = r h

Surface Area =
rh + r

Trapezoids

Sphere

Area = (a + b)h

Volume = r

Surface Area = r
b

Circles

General Cone

Area = r

Area of Base = A

Circumference = r

Volume = Ah
A

General Right Cylinder

Sectors of Circles
in radians
Area = r
s = r

Area of Base = A
Volume = Ah

h
A

Algebra
Factors and Zeros of Polynomials

Let p(x) = an xn + an xn + + a x + a be a polynomial. If p(a) = , then a is a zero of the polynomial and a solu on of
the equa on p(x) = . Furthermore, (x a) is a factor of the polynomial.

Fundamental Theorem of Algebra

An nth degree polynomial has n (not necessarily dis nct) zeros. Although all of these zeros may be imaginary, a real
polynomial of odd degree must have at least one real zero.

Quadra c Formula

If p(x) = ax + bx + c, and

b ac, then the real zeros of p are x = (b

b ac)/ a

Special Factors

x a = (x a)(x + a)
x a = (x a)(x + ax + a )
x + a = (x + a)(x ax + a )
x a = (x a )(x + a )
(x + y)n = xn + nxn y + n(n! ) xn y + + nxyn + yn
(x y)n = xn nxn y + n(n! ) xn y nxyn yn

Binomial Theorem

(x + y) = x + xy + y
(x + y) = x + x y + xy + y
(x + y) = x + x y + x y + xy + y

(x y) = x xy + y
(x y) = x x y + xy y
(x y) = x x y + x y xy + y

Ra onal Zero Theorem

If p(x) = an xn + an xn + + a x + a has integer coecients, then every rational zero of p is of the form x = r/s,
where r is a factor of a and s is a factor of an .

Factoring by Grouping

acx + adx + bcx + bd = ax (cs + d) + b(cx + d) = (ax + b)(cx + d)

Arithme c Opera ons


ab + ac = a(b + c)
(a)

( bc ) =
d

(a) (d)
b

ad
bc

( )
b
ab
=
c
c

c
ad + bc
a
+ =
b d
bd
(a)

a+b
a b
= +
c
c
c

ba
ab
=
cd
dc

ab + ac
=b+c
a

b
c

a
ac
( )=
b
b
c

a
bc

Exponents and Radicals


a = , a =
( a )x
b

ax
= x
b

(ab)x = ax bx

ax ay = ax+y

am

=a

m/n

ax

a=a


ab = n a n b

ax
= axy
ay
x y

(a ) = a

xy

a=a

/n

n
a
a
=
n
b
b

Addi onal Formulas


Summa
on Formulas:
n

i=

c = cn
i =

i=

n(n + )

i=

n(n + )( n + )

i=

i=

i =

n(n + )

Trapezoidal
Rule:

x [

f(x) dx

with Error

f(x ) + f(x ) + f(x ) + ... + f(xn ) + f(xn+ )


]
(b a) [
max f (x)
n

Simpsons
Rule:

f(x) dx

with Error

x [

f(x ) + f(x ) + f(x ) + f(x ) + ... + f(xn ) + f(xn ) + f(xn+ )

]

(b a) [
max f ( ) (x)
n

Arc Length:
L=

b
a

Surface of Revolu on:


f (x)

S=

dx

b
a

f(x)

(where f(x) )

S=

b
a

+ f (x) dx

+ f (x) dx

(where a, b )

Work Done by a Variable Force:


W=

b
a

F(x) dx

Force Exerted by a Fluid:


F=

b
a

w d(y) (y) dy

Taylor Series Expansion for f(x):


pn (x) = f(c) + f (c)(x c) +

f (c)
f (c)
f (n) (c)
(x c) +
(x c) + ... +
(x c)n
!
!
n!

Maclaurin Series Expansion for f(x), where c = :


pn (x) = f( ) + f ( )x +

f ( )
f ( )
f (n) ( ) n
x +
x + ... +
x
!
!
n!

Summary of Tests for Series:


Test

Series

nth-Term

n=

Geometric Series

Condi on(s) of
Convergence

an

p-Series

rn

n=

Integral Test

(bn bn+a )

n=

an

Direct Comparison

an

n=

Limit Comparison

an

n=

Ra o Test

an

n=

Root Test

n=

This test cannot be used to


show convergence.
Sum =

|r|

lim bn = L

Sum =

p>

r
)

bn

a(n) dn

( a

n=

is convergent

n=

lim an =

|r| <

(an + b)p

Comment

n=

Telescoping Series

Condi on(s) of
Divergence

a(n) dn

is divergent

bn

n=

an = a(n) must be
con nuous

bn

n=

converges and
an bn

bn
n=

diverges and
bn an

bn
n=

converges and
lim an /bn

diverges and
lim an /bn >

an+
lim
<
n an

an+
lim
>
n an

Also diverges if
lim an /bn =

{an } must be posi ve


Also diverges if
lim an+ /an =

an

( )
lim an

/n

<

( )
lim an

/n

{an } must be posi ve


>

Also diverges if
( ) /n
=
lim an

Vous aimerez peut-être aussi