APEX 1calculus
APEX 1calculus
Edition: 5
Website: [Link]
©2021 Gregory Hartman
Licensed to the public under Creative Commons Attribution-Noncommercial 4.0
International Public License
Thanks
There are many people who deserve recognition 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 time and
attention at home. Many thanks to Troy Siemers, whose most important con-
tributions extend far beyond the sections he wrote or the 227 figures he coded
in Asymptote for 3D interaction. He provided incredible support, advice and
encouragement for which I am very grateful. My thanks to Brian Heinold and
Dimplekumar Chalishajar for their contributions and to Jennifer Bowen for read-
ing 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, Geoff Cox, Jessica Libertini and other faculty of VMI who have given me
numerous suggestions and corrections based on their experience with teaching
from the text. (Special thanks to Troy, Lee and Dan for their patience in teach-
ing Calc III while I was still writing the Calc III material.) Thanks to Randy Cone
for encouraging his tutors of VMI’s Open Math Lab to read through the text and
check the solutions, and thanks to the tutors for spending their time doing so.
A very special thanks to Kristi Brown and Paul Janiczek who took this opportu-
nity far above and beyond what I expected, meticulously checking every solution
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 time to work on this project. I am blessed to
have so many people give of their time to make this book better.
iv
Preface
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 better understand what
you will find beyond this page.
This text comprises a three—volume series on Calculus. The first part covers
material taught in many “Calc 1” courses: limits, derivatives, and the basics of
integration, found in Chapters 1 through 6.1. The second text covers material of-
ten taught in “Calc 2:” integration and its applications, including an introduction
to differential equations, along with an introduction to sequences, series and
Taylor Polynomials, found in Chapters 5 through 8. The third text covers topics
common in “Calc 3” or “multivariable calc:” parametric equations, polar coordi-
nates, vector-valued functions, and functions of more than one variable, found
in Chapters 10 through 15. All three are available separately for free at apexcal-
[Link]¹, and HTML versions of the book can be found at [Link]².
These three texts are intended to work together and make one cohesive text,
APEX Calculus, which can also be downloaded from the website.
Printing the entire 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 about
$15 at [Link]³.
For Students: How to Read this Text. Mathematics textbooks have a reputa-
tion for being hard to read. High—level mathematical writing often seeks to say
much with few words, and this style often seeps into texts of lower—level top-
ics. This book was written with the goal of being easier to read than many other
calculus textbooks, without becoming too verbose.
Each chapter and section starts with an introduction of the coming material,
hopefully setting the stage for “why you should care,” and ends with a look ahead
to see how the just—learned material helps address future problems.
v
vi
Interactive, 3D Graphics. Versions 3.0 and 4.0 of the textbook include inter-
active, 3D graphics in the pdf version. Nearly all graphs of objects in space can
be rotated, shifted, and zoomed in/out so the reader can better understand the
object illustrated. However, the only pdf viewers that support these 3D graphics
are Adobe Reader Acrobat (and only the versions for PC/Mac/Unix/Linux com-
puters, not tablets or smartphones).
The latest version of the book, which is authored in PreTeXt, is available in
html. In html, the 3D graphics are rendered using WebGL, and should work in
any modern web browser.
Interactive graphics are no longer supported within the pdf, but clicking on
any 3D graphic within the pdf will take you directly to the interactive version on
the web.
First PreTeXt Edition (Version 5.0). Key changes from Version 4.0 to 5.0:
• The underlying source code has been completely rewritten, to use the
PreTeXt⁶ language, instead of the original LATEX.
⁴[Link]/APEXCalculus
⁵[Link]/APEX
⁶[Link]
⁷[Link]/[Link]
⁸[Link]
A Brief History of Calculus
viii
Contents
Thanks iv
Preface v
1 Limits 2
1.1 An Introduction To Limits . . . . . . . . . . . . . . 2
1.2 Epsilon-Delta Definition of a Limit . . . . . . . . . . . 10
1.3 Finding Limits Analytically . . . . . . . . . . . . . 18
1.4 One-Sided Limits . . . . . . . . . . . . . . . . 30
1.5 Continuity . . . . . . . . . . . . . . . . . . . 39
1.6 Limits Involving Infinity . . . . . . . . . . . . . . 51
2 Derivatives 63
2.1 Instantaneous Rates of Change: The Derivative . . . . . . 63
2.2 Interpretations of the Derivative . . . . . . . . . . . 80
2.3 Basic Differentiation Rules . . . . . . . . . . . . . 87
2.4 The Product and Quotient Rules . . . . . . . . . . . 95
2.5 The Chain Rule . . . . . . . . . . . . . . . . . 105
2.6 Implicit Differentiation. . . . . . . . . . . . . . . 115
2.7 Derivatives of Inverse Functions . . . . . . . . . . . 126
ix
CONTENTS x
5 Integration 223
5.1 Antiderivatives and Indefinite Integration . . . . . . . . 223
5.2 The Definite Integral . . . . . . . . . . . . . . . 232
5.3 Riemann Sums . . . . . . . . . . . . . . . . . 244
5.4 The Fundamental Theorem of Calculus . . . . . . . . . 259
5.5 Numerical Integration . . . . . . . . . . . . . . . 273
5.6 Substitution . . . . . . . . . . . . . . . . . . 289
5.7 Hyperbolic Functions . . . . . . . . . . . . . . . 306
12 Vectors 636
12.1 Introduction to Cartesian Coordinates in Space . . . . . . 636
12.2 An Introduction to Vectors . . . . . . . . . . . . . 652
12.3 The Dot Product . . . . . . . . . . . . . . . . . 665
12.4 The Cross Product . . . . . . . . . . . . . . . . 678
12.5 Lines . . . . . . . . . . . . . . . . . . . . 688
12.6 Planes . . . . . . . . . . . . . . . . . . . . 697
Appendices
Back Matter
Index 1097
Part I
1
Chapter 1
Limits
0.6
x
−6 −4 −2 2 4 6 x
0.5 1 1.5
sin(0) “0”
y→ → .
0 0
2
y
CHAPTER 1. LIMITS 3
1
This is not a complete definition (that will come in the next section); this is a
pseudo-definition that will allow us to explore the idea of a limit.
Above, where f (x) = sin(x)/x, we approximated
sin(x) sin(x)
lim ≈ 0.84 and lim ≈ 1.
x→1 x x→0 x
(We approximated these limits, hence used the “≈” symbol, since we are work-
ing with the pseudo-definition of a limit, not the actual definition.)
Once we have the true definition of a limit, we will find limits analytically;
that is, exactly using a variety of mathematical tools. For now, we will approxi-
mate limits both graphically and numerically. Graphing a function can provide
a good approximation, though often not very precise. Numerical methods can [Link]/watch?v=__qzaSg4y1I
provide a more accurate approximation. We have already approximated limits
graphically, so we now turn our attention to numerical approximations. Figure 1.1.5 Investigating sin(x)/x
Consider again limx→1 sin(x)
x . To approximate this limit numerically, we can
create a table of x and f (x) values where x is “near” 1. This is done in Fig-
ure 1.1.6. x sin(x)/x
Notice that for values of x near 1, we have sin(x)/x near 0.841. The x = 1
0.9 0.870363
row is included, but we stress the fact that when considering limits, we are not
0.99 0.844471
concerned with the value of the function at that particular x value; we are only
concerned with the values of the function when x is near 1. 0.999 0.841772
Now approximate limx→0 sin(x) numerically. We already approximated the 1 0.841471
x
value of this limit as 1 graphically in Figure 1.1.4. Figure 1.1.7 shows the value 1.001 0.841170
of sin(x)/x for values of x near 0. Ten places after the decimal point are shown 1.01 0.838447
to highlight how close to 1 the value of sin(x)/x gets as x takes on values very 1.1 0.810189
near 0. We include the x = 0 row but again stress that we are not concerned
with the value of our function at x = 0, only on the behavior of the function Figure 1.1.6 Values of sin(x)/x with x
near 0. near 1
This numerical method gives confidence to say that 1 is a good approxima-
tion of limx→0 sin(x)
x ; that is,
sin(x) x sin(x)/x
lim ≈ 1. -0.1 0.9983341665
x→0 x
-0.01 0.9999833334
Later we will be able to prove that the limit is exactly 1.
-0.001 0.9999998333
We now consider several examples that allow us explore different aspects of
0 not defined
the limit concept.
0.001 0.9999998333
Example 1.1.8 Approximating the value of a limit. 0.01 0.9999833334
0.1 0.9983341665
Use graphical and numerical methods to approximate
Figure 1.1.7 Values of sin(x)/x with x
x2 − x − 6 near 0
lim .
x→3 6x2 − 19x + 3
CHAPTER 1. LIMITS 4
x2 − x − 6
y=
6x2 − 19x + 3
on a small interval that contains 3. To numerically approximate the limit,
create a table of values where the x values are near 3. This is done in Video Solution
Figure 1.1.9 and Figure 1.1.10, respectively.
0.35 y
x2 −x−6
x 6x2 −19x+3
2.9 0.29878
0.3
2.99 0.294569
2.999 0.294163
0.25 3 not defined
3.001 0.294073
3.01 0.293669 [Link]/watch?v=eHx3LmrQZXM
x
2.4 2.6 2.8 3 3.2 3.4 3.6 3.1 0.289773
x2 − x − 6
lim ≈ 0.294.
x→3 6x2 − 19x + 3
This example may bring up a few questions about approximating limits (and
the nature of limits themselves).
0, where (
x+1 x<0 Video Solution
f (x) = .
−x + 1 x > 0
2
Solution. Again we graph f (x) and create a table of its values near x =
0 to approximate the limit. Note that this is a piecewise defined function,
so it behaves differently on either side of 0. Figure 1.1.12 shows a graph
of f (x), and on either side of 0 it seems the y values approach 1. Note
that f (0) is not actually defined, as indicated in the graph with the open
circle.
y
1
[Link]/watch?v=7RAiKoLCpgU
0.8 x f (x)
0.6
−0.1 0.9
−0.01 0.99
0.4
−0.001 0.999
0.2 0.001 0.999999
x 0.01 0.9999
−1 −0.5 0.5 1 0.1 0.99
x f (x)
2
0.9 2.01
0.99 2.0001
1 0.999 2.000001
1.001 1.001
x
1.01 1.01
0.5 1 1.5 2 1.1 1.1
1
Figure 1.1.18 Video presentation for
Explore why limx→1 (x−1)2 does not exist. Examples 1.1.15–1.1.19
1
Solution. A graph and table of f (x) = (x−1) 2 are given in Figure 1.1.20
80 x f (x)
60
0.9 100.
0.99 10000.
40 0.999 1. × 106
20
1.001 1. × 106
x
1.01 10000.
0.5 1 1.5 2 1.1 100.
y 1 y
1
0.5
0.5
x sin(1/x)
x
x 0.1 −0.544021
−0.1 −5 · 10−2 5 · 10−2
−1 −0.5 0.5 1
0.1
0.01 −0.506366
−0.5 −0.5 0.001 0.82688
0.0001 −0.305614
−1 −1 1. × 10−5 0.0357488
1. × 10−6 −0.349994
(a) (b)
1. × 10−7 0.420548
Figure 1.1.23 Observing that f (x) = sin(1/x) has no limit as x → 0 in
Figure 1.1.24 Observing that f (x) =
Example 1.1.22
sin(1/x) has no limit as x → 0 in Ex-
It can be shown that in reality, as x approaches 0, sin(1/x) takes ample 1.1.22
on all values between −1 and 1 infinitely many times! Because of this
oscillation, limx→0 sin(1/x) does not exist.
Let f (x) represent the position function, in feet, of some particle that is
[Link]/watch?v=2NJmd0Jrt4U
moving in a straight line, where x is measured in seconds. Let’s say that when
x = 1, the particle is at position 10 ft., and when x = 5, the particle is at 20 ft. Figure 1.1.25 Video introduction to Sub-
Another way of expressing this is to say section 1.1.2
This difference quotient can be thought of as the familiar “rise over run” used
5
to compute the slopes of lines. In fact, that is essentially what we are doing:
given two points on the graph of f , we are finding the slope of the secant line x
through those two points. See Figure 1.1.26. 2 4 6
Now consider finding the average speed on another time interval. We again
start at x = 1, but consider the position of the particle h seconds later. That is,
consider the positions of the particle when x = 1 and when x = 1 + h. The
difference quotient (excluding units) is now
f (1 + h) − f (1) f (1 + h) − f (1)
= .
(1 + h) − 1 h
Let f (x) = −1.5x2 + 11.5x; note that f (1) = 10 and f (5) = 20, as in our
discussion. We can compute this difference quotient for all values of h (even
negative values!) except h = 0, for then we get “0/0,” the indeterminate form
introduced earlier. For all values h ̸= 0, the difference quotient computes the
average velocity of the particle over an interval of time of length h starting at
x = 1.
For small values of h, i.e., values of h close to 0, we get average velocities
over very short time periods and compute secant lines over small intervals. See
Figure 1.1.27. This leads us to wonder what the limit of the difference quotient
is as h approaches 0. That is,
f (1 + h) − f (1)
lim = ?
h→0 h
25 y 25 y 25 y
20 20 20
15 15 15
10 10 10
f (1+h)−f (1)
h h
5 5 5
−0.5 9.25
x x x
2 4 6 2 4 6 2 4 6 −0.1 8.65
−0.01 8.515
(a) h = 2 (b) h = 1 (c) h = 0.5
0.01 8.485
Figure 1.1.27 Secant lines of f (x) at x = 1 and x = 1 + h, for shrinking values 0.1 8.35
of h (i.e., h → 0) 0.5 7.75
As we do not yet have a true definition of a limit nor an exact method for
computing it, we settle for approximating the value. While we could graph the Figure 1.1.28 The difference quotient
difference quotient (where the x-axis would represent h values and the y-axis evaluated at values of h near 0
would represent values of the difference quotient) we settle for making a table.
See Figure 1.1.28. The table gives us reason to assume the value of the limit is
about 8.5.
Proper understanding of limits is key to understanding calculus. With limits,
we can accomplish seemingly impossible mathematical things, like adding up an
infinite number of numbers (and not get infinity) and finding the slope of a line
between two points, where the “two points” are actually the same point. These
are not just mathematical curiosities; they allow us to link position, velocity and
acceleration together, connect cross-sectional areas to volume, find the work
done by a variable force, and much more.
In the next section we give the formal definition of the limit and begin our [Link]/watch?v=YplEX5ohJk0
study of finding limits analytically. In the following exercises, we continue our
Figure 1.1.29 Video examples for dif-
introduction and approximate the value of limits.
ference quotients: once with direct com-
putation, and then by simplifying first
CHAPTER 1. LIMITS 9
1.1.3 Exercises
Terms and Concepts
1. In your own words, what does it mean to “find the limit of f (x) as x approaches 3”?
0
2. An expression of the form 0 is called .
3. (□ True □ False) The limit of f (x) as x approaches 5 is f (5).
4. Describe three situations where lim f (x) does not exist.
x→c
Problems
f (a+h)−f (a)
Exercise Group. Approximate the limit of the difference quotient, lim h , using h = ±0.1, ±0.01.
h→0
21. f (x) = 2 − 7x, a = 3 22. f (x) = 9x + 0.06, a = −1
Show your work. Show your work.
23. f (x) = x2 + 3x − 7, a = 1 24. 1
f (x) = x+1 ,a=2
Show your work. Show your work.
25. f (x) = 5x − 4x2 − 1, a = −3 26. f (x) = ln(x), a = 5
Show your work. Show your work.
27. f (x) = sin(x), a = π 28. f (x) = cos(x), a = π
Show your work. Show your work.
CHAPTER 1. LIMITS 10
The problem with these definitions 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, respectively?
The definition we describe in this section comes from formalizing “Near”. A
quick restatement gets us closer to what we want:
lim f (x) = L,
x→c
and means that given any ε > 0, there exists δ > 0 such that for all x in
CHAPTER 1. LIMITS 11
Mathematicians often enjoy writing ideas without using any words. Here is
the wordless definition of the limit:
lim f (x) = L
x→c
⇐⇒
∀ ε > 0, ∃ δ > 0 s.t. 0 < |x − c| < δ =⇒ |f (x) − L| < ε.
Note the order in which ε and δ are given. In the definition, the y-tolerance [Link]/watch?v=npoSY-AFvOY
ε is given first and then the limit will exist if we can find an x-tolerance δ that
works. Figure 1.2.3 Video presentation of De-
An example will help us understand this definition. Note that the explanation finition 1.2.2
is long, but it will take one through all steps necessary to understand the ideas.
|x − 4| < 1.75
=⇒ −1.75 < x − 4 < 1.75 < 2.25
y y
Choose ε > 0. Then …
ε = 0.5 ε = 0.5
2 2
ε = 0.5 ε = 0.5
…choose δ smaller
than each of these:
1 1
width width
1.75 2.25
x x
2 4 6 2 4 6
|y − 2| < ε
−ε < y − 2 < ε
√ √
−ε < x − 2 < ε (y = x)
√
2−ε< x<2+ε (Add 2)
(2 − ε) < x < (2 + ε)
2 2
(Square all)
4 − 4ε + ε < x < 4 + 4ε + ε
2 2
(Expand)
−4ε + ε < x − 4 < 4ε + ε
2 2
(Subtract 4)
The previous example was a little long in that we sampled a few specific
cases of ε before handling the general case. Normally
√ this is not done. The
previous example is also a bit unsatisfying in that 4 = 2; why work so hard
to prove something so obvious? Many ε-δ proofs are long and difficult to do.
In this section, we will focus on examples where the answer is, frankly, obvious,
because the non-obvious examples are even harder. In the next section we will
learn some theorems that allow us to evaluate limits analytically, that is, without
using the ε-δ definition.
CHAPTER 1. LIMITS 13
which implies
1 1
< ,
5 |x + 2|
which implies
ε ε
< . (1.2.1)
5 |x + 2|
This suggests that we set δ < 5ε . To see why, let consider what fol-
lows when we assume |x − 2| < δ:
|x − 2| < δ
ε
|x − 2| < (Our choice of δ)
5
ε
|x − 2| · |x + 2| < |x + 2| · (Multiply by |x + 2| )
5 y
ε
x2 − 4 < |x + 2| · (Simplify left side)
5
ε ε
x2 − 4 < |x + 2| · (Inequality (1.2.1), δ < 1)
|x + 2|
x2 − 4 < ε 4
Make note of the general pattern exhibited in these last two examples. In
some sense, each starts out “backwards.” That is, while we want to
1. start with |x − c| < δ and conclude that
2. |f (x) − L| < ε,
we actually start by doing what is essentially some “scratch-work” first:
1. assume |f (x) − L| < ε, then perform some algebraic manipulations to
give an inequality of the form
2. |x − c| < something.
When we have properly done this, the something on the “greater than” side
of the inequality becomes our δ. We can refer to this as the “scratch-work”
phase of our proof. Once we have δ, we can formally start the actual proof with
|x − c| < δ and use algebraic manipulations to conclude that |f (x) − L| < ε,
usually by using the same steps of our “scratch-work” in reverse order.
We highlight this process in the following example.
0<x<2
0 < x2 < 4 (Squared each term.)
Since 0 < x < 2, we can add 0, x and 2, respectively, to each part of the
inequality and maintain the inequality.
0 < x2 + x < 6
−1 < x2 + x − 1 < 5 (Subtracted 1 from each part.)
1 1
< 2 which implies that
5 x +x−1
ε ε
< 2 . (1.2.3)
5 x +x−1
So we set δ < ϵ/5. This ends our scratch-work, and we begin the
formal proof (which also helps us understand why this was a good choice
of δ).
Given ε, let δ < ε/5. We want to show that when |x − 1| < δ, then
(x3 − 2x) − (−1) < ε. We start with |x − 1| < δ:
|x − 1| < δ
ε
|x − 1| <
5
ε
|x − 1| < (Inequality (1.2.3), x near 1)
|x2 + x − 1|
|x − 1| · x2 + x − 1 < ε
x3 − 2x + 1 < ε
(x3 − 2x) − (−1) < ε,
We note that we could actually show that limx→c ex = ec for any con-
stant c. We do this by factoring out ec from both sides, leaving us to show
limx→c ex−c = 1 instead. By using the substitution u = x − c, this reduces
to showing limu→0 eu = 1 which we just did in the last example. As an added
benefit, this shows that in fact the function f (x) = ex is continuous at all values
of x, an important concept we will define in Section 1.5.
This formal definition of the limit is not an easy concept grasp. Our examples
are actually “easy” examples, using “simple” functions like polynomials, square
roots and exponentials. It is very difficult to prove, using the techniques given
above, that limx→0 sin(x)
x = 1, as we approximated in Section 1.1.
There is hope. Section 1.3 shows how one can evaluate complicated lim-
its using certain basic limits as building blocks. While limits are an incredibly
important part of calculus (and hence much of higher mathematics), rarely are
limits evaluated using the definition. Rather, the techniques of Section 1.3 are
employed.
CHAPTER 1. LIMITS 17
1.2.1 Exercises
Terms and Concepts
Problems
Solution.
(a) Using the Sums/Differences property, we know that
(b) Using the Scalar Multiples, Sums/Differences, and Powers proper- [Link]/watch?v=8x42kGfu9ts
ties, we find that
i.e., the limit at 2 could have been found just by plugging 2 into the function.
This holds true for all polynomials, and also for rational functions (which are
quotients of polynomials), as stated in the following theorem.
CHAPTER 1. LIMITS 20
p(x) p(c)
2. lim = q(c) , when q(c) ̸= 0.
x→c q(x)
It was likely frustrating in Section 1.2 to do a lot of work with ε and δ to prove
that
lim x2 = 4
x→2
as it seemed fairly obvious. The previous theorems state that many functions
behave in such an “obvious” fashion, as demonstrated by the rational function
in Example 1.3.6.
Polynomial and rational functions are not the only functions to behave in
such a predictable way. The following theorem gives a list of functions whose
behavior is particularly “nice” in terms of limits. In Section 1.5, we will give a
formal name to these functions that behave “nicely.”
Solution.
(a) This is a straightforward application of Theorem 1.3.7: lim cos(x) =
x→π
cos(π) = −1.
(b) We can approach this in at least two ways. First, by directly apply-
ing Theorem 1.3.7, we have:
lim sec2 (x) − tan2 (x) = sec2 (3) − tan2 (3).
x→3
(d) Again, we can approach this in two ways. First, we can use the ex-
ponential/logarithmic identity that eln(x) = x and evaluate lim eln(x) =
x→1
lim x = 1.
x→1
We can also use the Compositions rule. Using Theorem 1.3.7, we
have lim ln(x) = ln(1) = 0 and limx→0 ex = e0 = 1, satisfying
x→1
the conditions of the Compositions rule. Applying this rule,
then
lim g(x) = L. [Link]/watch?v=8Tv-GRQdAVA
x→c
Figure 1.3.11 Explaining the Squeeze
It can take some work to figure out appropriate functions by which to “squeeze” Theorem
a given function. However, that is generally the only place where work is neces-
sary; the theorem makes the “evaluating the limit part” very simple.
The Squeeze Theorem can be used to show that limits of sin(x) can be done
by direct substitution, as the videos in Figure 1.3.12 illustrate.
We use the Squeeze Theorem in the following example to finally prove that
lim sin(x)
x = 1.
x→0
tan(θ) θ sin(θ)
≥ ≥ .
2 2 2 [Link]/watch?v=pgjv3ojtXh4
(You may need to recall that the area of a sector of a circle is 12 r2 θ with
θ measured in radians.)
CHAPTER 1. LIMITS 23
tan(θ)
tan(θ)
tan(θ)
(1, tan(θ))
sin(θ)
(cos(θ), sin(θ))
θ θ θ
1 1 1
θ
Figure 1.3.14 Bounding the sector between two triangles
(1, 0)
2
Multiply all terms by sin(θ) , giving
1 θ
≥ ≥ 1.
cos(θ) sin(θ)
Taking reciprocals reverses the inequalities, giving Figure 1.3.15 The unit circle and re-
sin(θ) lated triangles
cos(θ) ≤ ≤ 1.
θ
(These inequalities hold for all values of θ near 0, even negative values,
since cos(−θ) = cos(θ) and sin(−θ) = − sin(θ).)
Now take limits.
sin(θ)
lim cos(θ) ≤ lim ≤ lim 1
θ→0 θ→0θ θ→0
sin(θ)
cos(0) ≤ lim ≤1
θ→0 θ
sin(θ)
1 ≤ lim ≤1
θ→0 θ
sin(θ)
Clearly this means that lim θ = 1.
θ→0
sin(θ)
With the limit lim θ = 1 finally established, we can move on to other
θ→0
limits involving trigonometric functions, as the video in Figure 1.3.16 demon-
strates.
Two notes about the Example 1.3.13 are worth mentioning. First, one might
be discouraged by this application, thinking “I would never have come up with
that on my own. This is too hard!” Don’t be discouraged; within this text we
will guide you in your use of the Squeeze Theorem. As one gains mathematical [Link]/watch?v=Wd464IIls5Y
maturity, clever proofs like this are easier and easier to create. Figure 1.3.16 Finding limits involving
Second, this limit tells us more than just that as x approaches 0, sin(x)/x trigonometric functions
approaches 1. Both x and sin(x) are approaching 0, but the ratio of x and sin(x)
approaches 1, meaning that they are approaching 0 in essentially the same way.
Another way of viewing this is: for small x, the functions y = x and y = sin(x)
are essentially indistinguishable.
We include this special limit, along with three others, in the following theo-
rem.
CHAPTER 1. LIMITS 24
x2 − 1
lim .
x→1 x − 1
x2 − 1 12 − 1
lim =
x→1 x − 1 1−1
which is of the form 00 , an indeterminate form. We cannot apply the
theorem.
By graphing the function, as in Figure 1.3.19, we see that the function
seems to be linear, implying that the limit should be easy to evaluate.
Recognize that the numerator of our quotient can be factored:
y
x2 − 1 (x − 1)(x + 1) 3
= .
x−1 x−1
The function is not defined when x = 1, but for all other x, 2
x −1
2
(x − 1)(x + 1)
=
x−1 x−1
(x− 1
1)(x + 1)
=
(x
−1)
= x + 1, if x ̸= 1 x
0.5 1 1.5 2
x2 − 1
lim = lim (x + 1)
x→1 x − 1 x→1
=2
The key to Example 1.3.18 is that the functions y = (x2 − 1)/(x − 1) and
y = x+1 are identical except at x = 1. Since limits describe a value the function
is approaching, not the value the function actually attains, the limits of the two
functions are always equal.
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
x→c
lim f (x) = L.
x→c
The Fundamental Theorem of Algebra tells us that when dealing with a ra-
tional function of the form g(x)/f (x) and directly evaluating the limit lim fg(x)
(x)
x→c
returns “0/0”, then (x − c) is a factor of both g(x) and f (x). One can then
use algebra to factor this binomial out, cancel, then apply Theorem 1.3.20. We
demonstrate this once more.
Evaluate
x3 − 2x2 − 5x + 6 Video Solution
lim .
x→3 2x3 + 3x2 − 32x + 15
x2 + x − 2
= lim
x→3 2x2 + 9x − 5
10
=
40
CHAPTER 1. LIMITS 26
1
= .
4
Evaluate √
x−3
lim .
x→9 x−9 Video Solution
Solution. We begin by trying to apply the Quotients limit rule, but the
denominator evaluates to zero. In fact, this limit is of the indeterminate
form 0/0. We will do some algebra to resolve the indeterminate form. In
this case, we multiply the numerator and denominator by the conjugate
of the numerator.
√ √ √
x−3 x − 3 ( x + 3)
= · √
x−9 x − 9 ( x + 3)
[Link]/watch?v=vOW92eipOu4
x−9
= √
(x − 9)( x + 3)
We end this section by revisiting a limit first seen in Section 1.1, a limit of a
difference quotient. Let f (x) = −1.5x2 + 11.5x; we approximated the limit
lim f (1+h)−f
h
(1)
≈ 8.5. We formally evaluate this limit in the following exam-
h→0
ple.
This section contains several valuable tools for evaluating limits. One of the
main results of this section is Theorem 1.3.7; it states that many functions that
we use regularly behave in a very nice, predictable way. In Section 1.5 we give
a name to this nice behavior; we label such functions as continuous. Defining
that term will require us to look again at what a limit is and what causes limits
to not exist.
CHAPTER 1. LIMITS 28
1.3.1 Exercises
Terms and Concepts
1. Explain in your own words, without using ε-δ formality, why lim b = b.
x→c
2. Explain in your own words, without using ε-δ formality, why lim x = c.
x→c
3. What does the text mean when it says that certain functions’ “behavior is ‘nice’ in terms of limits”? What, in
particular, is “nice”?
4. Sketch a graph that visually demonstrates the Squeeze Theorem.
5. You are given the following information:
f (x)
lim f (x) = 0 lim g(x) = 0 lim =2
x→1 x→1 x→1 g(x)
What can be said about the relative sizes of f (x) and g(x) as x approaches 1?
6. (□ True □ False) lim ln x = 0.
x→1
Use a theorem to defend your answer.
Problems
Exercise Group. Use the following information to evaluate the given limit, when possible.
lim f (x) = 6 lim f (x) = 9 f (9) = 6
x→9 x→6
lim g(x) = 3 lim g(x) = 3 g(6) = 3
x→9 x→6
7. lim (f (x) + g(x)) 8. lim 3f (x)
x→9 g(x)
x→9
If it is not possible to determine the limit, state If it is not possible to determine the limit, state
why not. why not.
9. lim f (x)−2g(x)
g(x) 10. f (x)
lim 3−g(x)
x→9 x→6
If it is not possible to determine the limit, state If it is not possible to determine the limit, state
why not. why not.
11. lim g(f (x)) 12. lim f (g(x))
x→9 x→6
If it is not possible to determine the limit, state If it is not possible to determine the limit, state
why not. why not.
13. lim g(f (f (x))) 14. lim f (x)g(x) − f (x)2 + g(x)2
x→6 x→6
If it is not possible to determine the limit, state If it is not possible to determine the limit, state
why not. why not.
Exercise Group. Use the following information to evaluate the given limit, when possible. If it is not possible to
determine the limit, state why not.
lim f (x) = 2 lim f (x) = 1 f (1) = 1/5
x→1 x→10
lim g(x) = 0 lim g(x) = π g(10) = π
x→1 x→10
Exercise Group. The following exercises challenge your understanding of limits but can be evaluated using the
knowledge gained in Section 1.3.
sin(8x) sin(9x)
39. lim x 40. lim 8x
x→0 x→0
ln(1+x) sin(x)
41. lim x 42. lim x , where x is measured in degrees, not
x→0 x→0
radians.
43. Let f (x) = 0 and g(x) = xx .
(a) Explain why lim f (x) = 0.
x→2
(d) Explain why the previous statement does not violate the Composition Rule of Theorem 1.3.1.
CHAPTER 1. LIMITS 30
Left-Hand Limit
Let f be a function defined on (a, c) for some a < c and
let L be a real number. The statement that the limit of
f (x), as x approaches c from the left, is L, (alternatively,
that the left-hand limit of f at c is L) is denoted by
lim f (x) = L,
x→c−
and means that for any ε > 0, there exists δ > 0 such that
for all x ∈ (a, c), if |x − c| < δ, then |f (x) − L| < ε.
Right-Hand Limit
Let f be a function defined on (c, b) for some b > c and let
L be a real number. The statement that the limit of f (x),
as x approaches c from the right, is L, (alternatively, that
the right-hand limit of f at c is L) is denoted by
lim f (x) = L,
x→c+
and means that for any ε > 0, there exists δ > 0 such that
for all x ∈ (c, b), if |x − c| < δ, then |f (x) − L| < ε.
Solution. For these problems, the visual aid of the graph is likely more
Figure 1.4.4 A graph of f in Example 1.4.3
effective in evaluating the limits than using f itself. Therefore we will
refer often to the graph.
Video Solution
(a) As x goes to 1 from the left, we see that f (x) is approaching the
value of 1.
Therefore lim f (x) = 1.
x→1−
(b) As x goes to 1 from the right, we see that f (x) is approaching the
value of 2. Recall that it does not matter that there is an “open cir-
cle” there; we are evaluating a limit, not the value of the function.
Therefore lim f (x) = 2.
x→1+ [Link]/watch?v=NdBPwaP4Xkk
(c) The limit of f as x approaches 1 does not exist, as discussed in
Section 1.1. The function does not approach one particular value,
but two different values from the left and the right.
(d) Using the definition, and by looking at the graph, we see that
f (1) = 1.
(e) As x goes to 0 from the right, we see that f (x) is approaching
0. Therefore limx→0+ f (x) = 0. Note we cannot consider a left-
hand limit at 0 as f is not defined for values of x < 0.
(h) The graph and the definition of the function show that f (2) is not
defined.
Note how the left- and right-hand limits were different at x = 1. This, of
course, causes the limit to not exist. The following theorem states what is fairly
intuitive: the limit exists precisely when the left- and right-hand limits are equal.
CHAPTER 1. LIMITS 32
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 left and right
hand limits both equal L. If the limit is not equal to L, then at least one of the
left and right-hand limits is not equal to L (it may not even exist).
One thing to consider in Examples 1.4.3-Example 1.4.10 is that the value of
the function may/may not be equal to the value(s) of its left/right-hand limits,
even when these limits agree.
(a) lim f (x) (c) lim f (x) Figure 1.4.9 Graphing f in Example 1.4.8
x→1− x→1
[Link]/watch?v=HVFazve-Qxc
CHAPTER 1. LIMITS 34
Solution. It is clear from the definition of the function and its graph
Figure 1.4.11 Graphing f in Example 1.4.10
that all of the following are equal:
In Examples 1.4.3-Example 1.4.10 we were asked to find both limx→1 f (x) Video Solution
and f (1). Consider the following table:
1.4.1 Exercises
Terms and Concepts
1. What are the three ways in which a limit may fail to exist?
2. (□ True □ False) If lim f (x) = 5, then lim f (x) = 5.
x→1− x→1
Problems
4 4
2 2
x x
−1 1 2 3 4 5 6 −1 1 2 3 4 5 6
7. 8.
y y
8
4
4 2
2
x
x
−1 1 2 3
−1 1 2 3 4
2
2
x
1 −4 −2 2 4 6
−2
x
−1 1 2 3 4 5 6
−4
−1
−6
(a) lim− f (x)
x→2 (a) lim− f (x)
x→0
(b) lim f (x)
x→2+ (b) lim f (x)
x→0+
(c) lim f (x)
x→2 (c) lim f (x)
x→0
(d) f (2)
(d) f (0)
CHAPTER 1. LIMITS 37
11. 12.
4 y y
4
2 2
x
x
−4 −2 2 4
−4 −2 2 4
−2
−2
−4
−4
Let a be an integer with −3 ≤ a ≤ 3.
(a) lim f (x) (a) lim f (x)
x→−2− x→a−
(h) f (2)
Exercise Group. Evaluate the given limits of the piecewise defined function.
( (
x − 1 if x ≤ 3 2x − 2x2 − 5 if x < 3
13. f (x) = 14. f (x) =
x2 − 3 if x > 3 sin(x − 3) if x ≥ 3
(h) f (5)
(
1 − cos2 (x) x<a
if x < −1
17. f (x) = where a is a real x + 1
sin2 (x) x≥a 18. f (x) = x−1 if x = −1
number. x + 2 if x > −1
(a) lim− f (x)
x→ (a) lim f (x)
x→−1−
(b) lim+ f (x)
x→ (b) lim f (x)
x→−1+
(c) lim f (x)
x→ (c) lim f (x)
x→−1
(d) f ()
(d) f (−1)
(
a(x − b)2 + c x < b
x − 2x − 7
if x < −1
2
20. f (x) =
19. f (x) = x−1 if x = −1 a(x − b) + c x ≥ b
− x2 + x + 4
if x > −1 (a) lim f (x)
x→b−
(a) lim f (x)
x→−1− (b) lim+ f (x)
x→b
(b) lim f (x)
x→−1+ (c) lim f (x)
x→b
(c) lim f (x)
x→−1 (d) f (b)
(d) f (−1)
(
|x|
x x ̸= 0
21. f (x) =
0 x=0
(d) f (0)
CHAPTER 1. LIMITS 39
1.5 Continuity
As we have studied limits, we have gained the intuition that limits measure
“where a function is heading.” That is, if lim f (x) = 3, then as x is close to
x→1
1, f (x) is close to 3. We have seen, though, that this is not necessarily a good
indicator of what f (1) actually is. This can be problematic; functions can tend
to one value but attain another. This section focuses on functions that do not
exhibit such behavior.
y
Example 1.5.5 Finding intervals of continuity. 2
The floor function, f (x) = ⌊x⌋, returns the largest integer smaller than,
1
or equal to, the input x. (For example, f (π) = ⌊π⌋ = 3.) The graph of
f in Figure 1.5.6 demonstrates why this is often called a “step function.” x
Give the intervals on which f is continuous. −2 −1 1 2 3
Solution. We examine the three criteria for continuity.
−1
1. The limits lim f (x) do not exist at the jumps from one “step” to
x→c
the next, which occur at all integer values of c. Therefore the limits −2
5. The domain of f (x) = |x| is (−∞, ∞). We can define the ab-
solute value function as
(
−x x < 0
f (x) = .
x x≥0
Let f and g be continuous functions on an interval I, let c be a real num- Figure 1.5.10 Video presentation of The-
ber and let n be a positive integer. The following functions are continu- orem 1.5.11
CHAPTER 1. LIMITS 42
ous on I.
We have defined what it means
Sums/Difference f ±g for a function to be continuous
on an interval, but many functions,
Constant Multiple c·f such as f (x) = tan(x), have do-
Product f ·g mains that are the union of more
Quotient f /g (as long as g ̸= 0 on I) than one interval.
If the domain of a function is
Power fn a union of intervals, saying that
√
Root n
f (If n is even then require f (x) ≥ 0 on I.) a function is continuous on its do-
Compositions Adjust the definitions of f and g to: Let f be main means that the function is
continuous on I, where the range of f on I is continuous on each of those in-
J, and let g be continuous on J. Then g ◦ f , tervals. But be careful to note
i.e., g(f (x)), is continuous on I. that the converse is not true. As
we learned in Example 1.5.5, a
function can be continuous on a
collection of intervals, but not on
Theorem 1.5.12 Continuous Functions. their union.
As the video example in Figure 1.5.13 illustrates, the above theorems allow [Link]/watch?v=ewUiuE9bQlo
us to quickly construct new continuous functions from old ones.
We apply these theorems in the following Example. Figure 1.5.13 Continuity of composi-
tions
Example 1.5.14 Determining intervals on which a function is continu-
ous.
State the interval(s) on which each of the following functions is continu- Video Solution
ous.
√ √
1. f (x) = x−1+ 5−x 3. f (x) = tan(x)
p
2. f (x) = x sin(x) 4. f (x) = ln(x)
1. The square root terms are continuous on the intervals [1, ∞) and [Link]/watch?v=6Lm-0eBi-5E
(−∞, 5], respectively. As f is continuous only where each term
is continuous, f is continuous on [1, 5], the intersection of these
two intervals. A graph of f is given in Figure 1.5.15.
2. The functions y = x and y = sin(x) are each continuous every-
where, hence their product is, too.
CHAPTER 1. LIMITS 43
y y y
100
3 3
80
2 2 60
40
1 1
20
x x x
1 2 3 4 1 2 3 4 1 2 3 4
(a) The graph of a func- (b) The graph of a func- (c) The graph of a func-
tion with a removable tion with a jump discon- tion with an infinite dis-
discontinuity at x = 2 tinuity at x = 2 continuity at x = 2
Figure 1.5.17 Illustrating three common types of discontinuity
important concept as follows. Suppose f is defined on [1, 2], and f (1) = −10
5
and f (2) = 5. If f is continuous on [1, 2] (i.e., its graph can be sketched as a con-
tinuous curve from (1, −10) to (2, 5)) then we know intuitively that somewhere x
on the interval [1, 2] f must be equal to −9, and −8, and −7, −6, . . . , 0, 1/2, 1 1.5 2 2.5
etc. In short, f takes on all intermediate values between −10 and 5. It may take −5
on more values; f may actually equal 6 at some time, for instance, but we are
−10
guaranteed all values between −10 and 5.
While this notion seems intuitive, it is not trivial to prove and its importance
−15
is profound. Therefore the concept is stated in the form of a theorem.
Figure 1.5.18 Illustration of the Inter-
Theorem 1.5.19 Intermediate Value Theorem. mediate Value Theorem: the output
Let f be a continuous function on [a, b] and, without loss of generality, 3 is in between −10 and 5, and there-
let f (a) < f (b). Then for every value y, where f (a) < y < f (b), there fore any continuous function on [1, 2]
is at least one value c in (a, b) such that f (c) = y. with f (1) = −10 and f (2) = 5 will
achieve the output 3 somewhere in
One important application of the Intermediate Value Theorem is root find- [1, 2]
ing. Given a function f , we are often interested in finding values of x where
f (x) = 0. These roots may be very difficult to find exactly. Good approxima-
tions can be found through successive applications of this theorem. Suppose
through direct computation we find that f (a) < 0 and f (b) > 0, where a < b.
The Intermediate Value Theorem states that there is at least one c in (a, b) such
that f (c) = 0. The theorem does not give us any clue as to where to find such
a value in the interval (a, b), just that at least one such value exists.
There is a technique that produces a good approximation of c. Let d be the
midpoint of the interval [a, b], with f (a) < 0 and f (b) > 0 and consider f (d).
There are three possibilities:
1. f (d) = 0: We got lucky and stumbled on the actual value. We stop as we
found a root. [Link]/watch?v=Fx7Qu9tZlN4
Figure 1.5.20 Video presentation of The-
2. f (d) < 0: Then we know there is a root of f on the interval [d, b] — we
orem 1.5.19
have halved the size of our interval, hence are closer to a good approxima-
tion of the root.
CHAPTER 1. LIMITS 45
Approximate the root of f (x) = x − cos(x), accurate to three places Video Solution
after the decimal.
Solution. Consider the graph of f (x) = x − cos(x), shown in Fig-
ure 1.5.22. It is clear that the graph crosses the x-axis somewhere near
x = 0.8. To start the Bisection Method, pick an interval that contains
0.8. We choose [0.7, 0.9]. Note that all we care about are signs of f (x),
not their actual value, so this is all we display.
interval. While we do not know its exact value, we know it starts with
0.739.
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 end-
points differ by a preset small amount. One of the authors did write such
a program and found the zero of f to be 0.7390851332 , accurate to 10
places after the decimal. While it took a few minutes to write the pro-
gram, it took less than a thousandth of a second for the program to run
the necessary 35 iterations. In less than 8 hundredths of a second, the
zero was calculated to 100 decimal places (with less than 200 iterations).
1.5.1 Exercises
Terms and Concepts
9. (□ True □ False) If f is continuous on [0, 1) and [1, 2), then f is continuous on [0, 2).
10. (□ True □ False) The sum of continuous functions is also continuous.
Problems
Exercise Group. Use the graph to determine if the function is continuous at the given point.
11. Is f in the graph below continuous at 1? 12. Is f in the graph below continuous at 1?
y y
2 2
1 1
x x
−0.5 0.5 1 1.5 2 2.5 −0.5 0.5 1 1.5 2 2.5
13. Is f in the graph below continuous at 1? 14. Is f in the graph below continuous at 0?
y y
2 2
1 1
x x
−0.5 0.5 1 1.5 2 2.5 −0.5 0.5 1 1.5 2 2.5
2
1
x
−4 −2 2 4
x
−0.5 0.5 1 1.5 2 2.5
−2
(□ Yes. □ No.) −4
If not, state why it is not.
(□ Yes. □ No.)
If not, state why it is not.
17. Is f in the graph below continuous at −2, 0, and 18. Is f in the graph below continuous at 3π
2 ?
2? y
4 y
2
2
1
x
−4 −2 2 4
x
−2 − π2 π π 3π 2π
2 2
(□ Yes. □ No.)
−4 If not, state why it is not.
Exercise Group. Use the Bisection Method to approximate, accurate to two decimal places, the value of the root of
the given function in the given interval.
39. f (x) = x2 + 2x − 4 on the interval [1, 1.5]
Show the steps you used applying the Bisection Method.
40. f (x) = sin(x) − 12 on the interval [0.5, 0.55]
Show the steps you used applying the Bisection Method.
CHAPTER 1. LIMITS 50
1 60
lim = 0.
x→∞ x2
40
We explore both types of use of ∞ in turn.
x
Let I be an open interval containing c, and let f be a function defined
−1 −0.5 0.5 1
on I, except possibly at c.
Figure 1.6.1 Graphing f (x) = 1/x2
• The limit of f (x), as x approaches c, is infinity, denoted by
for values of x near 0
lim f (x) = ∞,
x→c
if given any N > 0, there exists δ > 0 such that for all x in I,
where x ̸= c, if |x − c| < δ, then f (x) > N .
if given any N < 0, there exists δ > 0 such that for all x in I,
where x ̸= c, if |x − c| < δ, then f (x) < N .
The first definition is similar to the ε-δ definition in Definition 1.2.2 from
Section 1.2. In that definition, given any (small) value ε, if we let x get close
enough to c (within δ units of c) then f (x) is guaranteed to be within ε of L.
Here, given any (large) value N , if we let x get close enough to c (within δ units [Link]/watch?v=UVhqWmKqHtw
of c), then f (x) will be at least as large as N . In other words, if we get close Figure 1.6.3 Video presentation of De-
enough to c, then we can make f (x) as large as we want. finition 1.6.2
It is important to note that by saying limx→c f (x) = ∞ we are implicitly
stating 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 infinity because it is helpful and descriptive. It is one specific
way in which a limit can fail to exist.
We define one-sided limits that approach infinity in a similar way.
lim f (x) = ∞,
x→c−
if given any N > 0, there exists δ > 0 such that for all a < x < c,
if |x − c| < δ, then f (x) > N .
y
• Let f be a function defined on (c, b) for some b > c. We say the 100
limit of f (x), as x approaches c from the right, is infinity, or, the
right-hand limit of f at c is infinity, denoted by 80
lim f (x) = ∞, 60
x→c+
40
if given any N > 0, there exists δ > 0 such that for all c < x < b,
if |x − c| < δ, then f (x) > N . 20
x
• The term left- (or, right-) hand limit of f at c is negative infinity is
0.5 1 1.5 2
defined in a manner similar to Definition 1.6.2.
Figure 1.6.6 Observing infinite limit as
x → 1 in Example 1.6.5
Example 1.6.5 Evaluating limits involving infinity.
Find lim 1
2 as shown in Figure 1.6.6. Video Solution
x→1 (x−1)
Solution. In Example 1.1.19 of Section 1.1, by inspecting values of x
close to 1 we concluded that this limit does not exist. That is, it cannot
equal any real number. But the limit could be infinite. And in fact, we
see that the function does appear to be growing larger and larger, as
f (0.99) = 104 , f (0.999) = 106 , f (0.9999) = 108 . A similar thing
happens on the other side of 1. From the graph and the numeric infor-
mation, we could state limx→1 1/(x − 1)2 = ∞. We can prove this by
using Definition 1.6.2 √
In general, let a “large” value√N be given. Let δ = 1/ N . If x is [Link]/watch?v=S3dUAUQiKFQ
within δ of 1, i.e., if |x − 1| < 1/ N , then:
1 y
|x − 1| < √ 40
N
1
(x − 1)2 < 20
N
1 x
> N, −1 −0.5
(x − 1)2 0.5 1
−20
which is what we wanted to show. So we may say limx→1 1/(x − 1)2 =
∞. −40
Example 1.6.7 Evaluating limits involving infinity. Figure 1.6.8 Evaluating lim 1
x→0 x
1
Find lim , as shown in Figure 1.6.8.
x→0 x
Solution. It is easy to see that the function grows without bound near Video Solution
0, but it does so in different ways on different sides of 0. Since its be-
havior is not consistent, we cannot say that limx→0 x1 = ∞. Instead,
we will say limx→0 x1 does not exist. However, we can make a state-
ment about one-sided limits. We can state that limx→0+ x1 = ∞ and
limx→0− x1 = −∞.
[Link]/watch?v=JP1k74FZE1I
CHAPTER 1. LIMITS 53
3x
Find the vertical asymptotes of f (x) = x2 −4 .
Solution. Vertical asymptotes occur where the function grows without Video Solution
bound; this can occur at values of c where the denominator is 0. When x
is near c, the denominator is small, which in turn can make the function
take on large values. In the case of the given function, the denominator
is 0 at x = ±2. Substituting in values of x close to 2 and −2 seems to
indicate that the function tends toward ∞ or −∞ at those points. We
can graphically confirm this by looking at Figure 1.6.12. Thus the vertical
asymptotes are at x = ±2.
−12
We have seen how the limits limx→0 sin(x)x and limx→1 xx−1 each return the
x
indeterminate form 0/0 when we blindly plug in x = 0 and x = 1, respectively.
−1 −0.5 0.5 1 1.5 2
However, 0/0 is not a valid arithmetical expression. It gives no indication that
the respective limits are 1 and 2. Figure 1.6.13 Graphically showing that
2
−1
f (x) = xx−1 does not have an as-
ymptote at x = 1
CHAPTER 1. LIMITS 54
With a little cleverness, one can come up with 0/0 expressions which have
a limit of ∞, 0, 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 0/0.
Rather, keep in mind that we are taking limits. What is really happening is that
the numerator is shrinking to 0 while the denominator is also shrinking to 0. The
respective rates at which they do this are very important and determine the ac-
tual value of the limit.
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 cancel-
ing), it may involve using trigonometric identities or logarithm rules, or it may
require a tool such as the Squeeze Theorem. In Section 4.6 we will learn yet
another technique called L’Hospital’s Rule that provides another way to handle
indeterminate forms.
Some other common indeterminate forms are ∞ − ∞, ∞ · 0, ∞/∞, 00 , ∞0
and 1∞ . Again, keep in mind that these are the “blind” results of directly sub-
stituting c into the expression, and each, in and of itself, has no meaning. The
expression ∞ − ∞ does not really mean “subtract infinity from infinity.” Rather,
it means “One quantity is subtracted from the other, but both are growing with-
out bound.” What is the result? It is possible to get every value between −∞
and ∞.
Note that 1/0 and ∞/0 are not indeterminate forms, though they are not
exactly valid mathematical expressions, either. In each, the function is growing
without bound, indicating that the limit will be ∞, −∞, or simply not exist if the
left- and right-hand limits do not match.
We can also define limits such as limx→∞ f (x) = ∞ by combining this defi-
nition with Definition 1.6.2.
[Link]/watch?v=7PwKJHgic7U
Figure 1.6.15 Video presentation of De-
finition 1.6.14
CHAPTER 1. LIMITS 55
x2
Approximate the horizontal asymptote(s) of f (x) = x2 +4 .
Solution. We will approximate the horizontal asymptotes by approxi-
2 2
mating the limits limx→−∞ x2x+4 and limx→∞ x2x+4 . (A rational function
can have at most one horizontal asymptote. So we could get away with
only taking x → ∞).
Figure 1.6.17(a) shows a sketch of f , and the table in Figure 1.6.17(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 asymp-
tote at y = 1.
y
1
0.8 x f (x)
0.6 10 0.9615
0.4
100 0.9996
10000 0.999996
0.2
−10 0.9615
x
−20 −10 10 20
−100 0.9996
−0.2 −10000 0.999996
(a) (b)
Figure 1.6.17 Using a graph and a table to approximate a horizontal as-
ymptote in Example 1.6.16
Later, we will show how to determine this analytically.
The video in Figure 1.6.18 shows how to prove the result from Example 1.6.16
using the limit definition.
Horizontal asymptotes can take on a variety of forms. Figure 1.6.19(a) shows
that f (x) = x/(x2 + 1) has a horizontal asymptote of y = 0, where 0 is ap-
proached from both above and below. √
Figure 1.6.19(b) shows that f (x) = x/ x2 + 1 has two horizontal asymp- [Link]/watch?v=kYmfeq-qKiI
totes; one at y = 1 and the other at y = −1. Figure 1.6.18 Using an ε-δ proof with
Figure 1.6.19(c) shows that f (x) = sin(x)/x has even more interesting be- Definition 1.6.14 in Example 1.6.16
havior than at just x = 0; as x approaches ±∞, f (x) approaches 0, but oscil-
lates as it does this.
1 y 1 y 1
y
0.5 0.5
x x 0.5
−0.5 −0.5 x
−20 −10 10 20
−1 −1
x3 + 2x + 1
lim .
x→∞ 4x3 − 2x2 + 9 [Link]/watch?v=v5SrtUsdMeU
A good way of approaching this is to divide through the numerator and de- Figure 1.6.20 Basic examples involv-
nominator by x3 (hence multiplying by 1), which is the largest power of x to ing limits at infinity
appear in the denominator. Doing this, we get
x3 + 2x + 1 1/x3 x3 + 2x + 1
lim = lim ·
x→∞ 4x − 2x + 9
3 2 x→∞ 1/x3 4x3 − 2x2 + 9
1 + 2/x2 + 1/x3
= lim .
x→∞ 4 − 2/x + 9/x3
Then using the rules for limits (which also hold for limits at infinity), as well
as the fact about limits of 1/xn , we see that the limit becomes
1+0+0 1
= .
4−0+0 4
This procedure works for any rational function. In fact, it gives us the follow-
ing theorem.
an xn + an−1 xn−1 + · · · + a1 x + a0
f (x) = ,
bm xm + bm−1 xm−1 + · · · + b1 x + b0
where m, n are positive integers and where any of the coefficients may
be 0 except for an and bm . Then:
1. If n = m, then
an
lim f (x) = lim f (x) = .
x→∞ x→−∞ bm
2. If n < m, then
3. If n > m, then limx→∞ f (x) and limx→−∞ f (x) are both infinite.
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 situation like the
example above, where we will divide by xn and in the limit all the terms will
approach 0 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 after dividing through by xm , all the
terms in the numerator will approach 0 in the limit, leaving us with 0/bm or 0.
CHAPTER 1. LIMITS 57
x2 x2 /x2
lim = lim
x→∞ x2 + 4 x→∞ x2 /x2 + 4/x2
1 [Link]/watch?v=cmZ39j1YI-o
= lim
x→∞ 1 + 4/x2
1
=
1+0
= 1.
x2 + 2x − 1 x2 − 1
1. lim 3. lim
x→−∞ x3 + 1 x→∞ 3 − x
x2 + 2x − 1
2. lim
x→∞ 1 − x − 3x2
Solution.
1. The highest power of x is in the denominator. Therefore, the limit
is 0; see Figure 1.6.24(a).
2. The highest power of x is x2 , which occurs in both the numerator
and denominator. The limit is therefore the ratio of the coeffi-
cients of x2 , which is −1/3. See Figure 1.6.24(b).
3. 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 x2 and −x. The expres-
sion in the limit will behave like x2 /(−x) = −x for large values of
x. Therefore, the limit is −∞. See Figure 1.6.24(c).
CHAPTER 1. LIMITS 58
0.6 y y x
0.4 y 10 20 30 40
0.4 −10
0.2
0.2
x −20
−40 −30 −20 −10 x
10 20 30 40 −30
−0.2
−0.2
−0.4 −40
−0.4
−0.6 −50
With care, we can quickly evaluate limits at infinity for a large number of
functions by considering the long run behavior using “dominant terms” of f (x).
For instance, consider again limx→±∞ √xx2 +1 , graphed in Figure 1.6.19(b). The
√
dominant terms are x in the numerator and x2 in the denominator. When x
is very large, x2 + 1 ≈ x2 . Thus
p √ x x
x2 + 1 ≈ x2 = |x| √ ≈ .
2
x +1 |x|
√1
x x x2
lim √ = lim √ ·
x→∞ x2 + 1 x→∞ x2 + 1 √1
x2
x
|x|
= lim q
x→∞ x2 +1
x2
1
= lim q for x > 0
x→∞ 1
1+ x2
1
=√
1+0
= 1.
The video in Figure 1.6.26 provides another example similar to Example 1.6.25.
[Link]/watch?v=vD1-zrRZQTI
Figure 1.6.26 Limits at infinity with a
radical function
CHAPTER 1. LIMITS 59
1.6.4 Exercises
Terms and Concepts
1. (□ True □ False) If lim f (x) = ∞, then we are implicitly stating that the limit exists.
x→5
2. (□ True □ False) If lim f (x) = 5, then we are implicitly stating that the limit exists.
x→5
Problems
Exercise Group. Evaluate the given limits using the graph of the function.
1 1
9. f (x) = (x+2)5
has the graph: 10. f (x) = (x−1)(x−2)2
has the graph:
y y
40 40
20 20
x x
−4 −3 −2 −1 1 −1 1 2 3
−20 −20
−40 −40
3
11. f (x) = e−x +1 has the graph: 12. f (x) = x3 sin(4πx) has the graph:
4 y y
1,000
3
500
2 x
−10 −5 5 10
1
−500
x
−10 −5 5 10 −1,000
−1
(a) lim f (x)
x→−∞
(a) lim f (x)
x→−∞
(b) lim f (x)
x→∞
(b) lim f (x)
x→∞
(c) lim− f (x)
x→0
(c) lim− f (x)
x→0
(d) lim f (x)
x→0+
(d) lim f (x)
x→0+
13. f (x) = sin(4x) has the graph: 14. f (x) = 2.4x − 9 has the graph:
y 20 y
1
0.5 10
x
x
−10 −5 5 10 −10 −5 5 10
−0.5
−10
−1
−20
(a) lim f (x)
x→−∞
(a) lim f (x)
x→−∞
(b) lim f (x)
x→∞
(b) lim f (x)
x→∞
x2 +13x+40 x2 −x−20
17. f (x) = x3 +7x2 −24x−180 18. f (x) = x2 +3x−4
Exercise Group. Identify the horizontal and vertical asymptotes, if any, of the given function.
2x2 +x−15 5x2 +x−4
19. f (x) = x2 −7x−18 20. f (x) = −2x2 −20x−18
4x2 −12x+8 2x2 −12x+16
21. f (x) = 6x3 −36x2 +48x 22. f (x) = −6x−18
x2 −10x+24 4x2 −44x+96
23. f (x) = 3x−18 24. f (x) = −x2 −4x−8
Derivatives
Chapter 1 introduced the most fundamental of calculus topics: the limit. This
chapter introduces the second most fundamental of calculus topics: the deriva-
tive. Limits describe where a function is going; derivatives describe how fast the
function is going.
63
CHAPTER 2. DERIVATIVES 64
Consider the interval from t = 2 to t = 3 (just before the riders hit the
ground). On that interval, the average velocity is
f (3) − f (2) 6 − 86
= = −80 ft/s,
3−2 1
where the minus sign indicates that the riders are moving down. By narrowing
the interval we consider, we will likely get a better approximation of the instan-
taneous velocity. On [2, 2.5] we have Units in Calculations. In the above
calculations, we left off the units
f (2.5) − f (2) 50 − 86 until the end of the problem. You
= = −72 ft/s. should always be sure that you
2.5 − 2 0.5
label your answer with the cor-
We can do this for smaller and smaller intervals of time. For instance, over
rect units. For example, if g(x)
a time span of one tenth of a second, i.e., on [2, 2.1], we have
gave you the cost (in $) of pro-
f (2.1) − f (2) 79.44 − 86 ducing x widgets, the units on
= = −65.6 ft/s. the difference quotient would be
2.1 − 2 0.1
$/widget.
Over a time span of one hundredth of a second, on [2, 2.01], the average
velocity is
f (2.01) − f (2) 85.3584 − 86
= = −64.16 ft/s.
2.01 − 2 0.01
What we are really computing is the average velocity on the interval [2, 2+h]
for small values of h. That is, we are computing
f (2 + h) − f (2)
h
where h is small.
We really want to use h = 0, but this, of course, returns the familiar “0/0” h Average Velocity ( fts )
indeterminate form. So we employ a limit, as we did in Section 1.1. 1 −80
We can approximate the value of this limit numerically with small values of 0.5 −72
h as seen in Figure 2.1.2. It looks as though the velocity is approaching −64 fts . 0.1 −65.6
Computing the limit directly gives 0.01 −64.16
0.001 −64.016
f (2 + h) − f (2) −16(2 + h)2 + 150 − (−16(2)2 + 150)
lim = lim
h→0 h h→0 h Figure 2.1.2 Approximating the instan-
−16(4 + 4h + h2 ) + 150 − 86 taneous velocity with average veloci-
= lim
h→0 h ties over a small time period h
−64 − 64h − 16h2 + 64
= lim
h→0 h
−64h − 16h2
= lim
h→0 h
= lim (−64 − 16h)
h→0
= −64.
y y
120
150
100
80
100
60
40
50
20
t
t
2 2.2 2.4 2.6 2.8 3 3.2 3.4
1 2 3
Figure 2.1.4 The function f (t) and a
Figure 2.1.3 The function f (t) and its secant line corresponding to t = 2
secant line corresponding to t = 2 and t = 3, zoomed in near t = 2
and t = 3
y y
120
100
100
80
60
50
40
20
t t
1.6 1.8 2 2.2 2.4 2.6 1.6 1.8 2 2.2 2.4 2.6
Figure 2.1.5 The function f (t) with Figure 2.1.6 The function f (t) with its
the same secant line, zoomed in fur- tangent line at t = 2
ther
As h → 0, these secant lines approach the tangent line, a line that goes
through the point (2, f (2)) with the special slope of −64. In Figure 2.1.5 and
Figure 2.1.6, we zoom in around the point (2, 86). We see the secant line, which
approximates f well, but not as well the tangent line shown in Figure 2.1.6.
We have just introduced a number of important concepts that we will flesh
out more within this section. First, we formally define two of them.
f (c + h) − f (c)
lim ,
h→0 h
provided the limit exists. If the limit exists, we say that f is differentiable
at c; if the limit does not exist, then f is not differentiable at c. If f is
differentiable at every point in I, then f is differentiable on I.
Solution.
(a) We compute this directly using Definition 2.1.7.
f (1 + h) − f (1)
f ′ (1) = lim
h→0 h
3(1 + h)2 + 5(1 + h) − 7 − (3(1)2 + 5(1) − 7)
= lim
h→0 h
3(1 + 2h + h2 ) + 5 + 5h − 7 − 1 [Link]/watch?v=4OMc0gJWcb0
= lim
h→0 h
3 + 6h + 3h2 + 5 + 5h − 8
= lim
h→0 h
3h2 + 11h
= lim
h→0 h
= lim (3h + 11)
h→0
= 11.
(b) The tangent line at x = 1 has slope f ′ (1) and goes through the
point (1, f (1)) = (1, 1). Thus the tangent line has equation, in
point-slope form, y = 11(x − 1) + 1. In slope-intercept form we
have y = 11x − 10.
(c) Again, using the definition,
f (3 + h) − f (3)
f ′ (3) = lim
h→0 h
3(3 + h)2 + 5(3 + h) − 7 − (3(3)2 + 5(3) − 7)
= lim
h→0 h
3(9 + 6h + h2 ) + 15 + 3h − 7 − 35
= lim
h→0 h
27 + 18h + 3h2 + 15 + 3h − 42 y
= lim 60
h→0 h
3h2 + 23h
= lim 40
h→0 h
= lim 3h + 23
h→0
20
= 23.
(d) The tangent line at x = 3 has slope 23 and goes through the x
−1 1 2 3 4
point (3, f (3)) = (3, 35). Thus the tangent line has equation
y = 23(x − 3) + 35 = 23x − 34.
Figure 2.1.11 A graph of f (x) = 3x2 +
A graph of f is given in Figure 2.1.11 along with the tangent lines at 5x − 7 and its tangent lines at x = 1
x = 1 and x = 3. and x = 3
CHAPTER 2. DERIVATIVES 67
Linear functions are easy to work with; many functions that arise in the
course of solving real problems are not easy to work with. A common practice
in mathematical problem solving is to approximate difficult functions with not-
so-difficult functions. Lines are a common choice. It turns out that at any given
point on the graph of a differentiable function f , the best linear approximation
to f is its tangent line. That is one reason we’ll spend considerable time finding
tangent lines to functions.
One type of function that does not benefit from a tangent line approximation
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.
CHAPTER 2. DERIVATIVES 68
f (1 + h) − f (1)
f ′ (1) = lim
h→0 h
3(1 + h) + 5 − (3 + 5)
= lim
h→0 h
3h
= lim
h→0 h
= lim 3
h→0
= 3.
We often desire to find the tangent line to the graph of a function without
knowing the actual derivative of the function. While we will eventually be able
to find derivatives of many common functions, the algebra and limit calculations
on some functions are complex. Until we develop further techniques, the best
we may be able to do is approximate the tangent line. We demonstrate this in
the next example.
sin(0 + h) − sin(0)
f ′ (0) ≈ 0.5
h
for a small value of h. We choose (somewhat arbitrarily) to let h = 0.1. x
Thus −π − π2 π π
sin(0.1) − sin(0) 2
f ′ (0) ≈ ≈ 0.9983.
0.1 −0.5
Thus our approximation of the equation of the tangent line is y =
0.9983(x − 0) + 0 = 0.9983x; it is graphed in Figure 2.1.18. The graph −1
seems to imply the approximation is rather good.
Figure 2.1.18 f (x) = sin(x) graphed
sin(x)
Recall from Section 1.3 that limx→0 x = 1, meaning for values of x near with an approximation to its tangent
line at x = 0
CHAPTER 2. DERIVATIVES 69
This process describes a function; 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 function occurs.
Instead of applying this function repeatedly for different 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:
input do something return
−→ −→
variable x to f and x function f ′ (x)
The output is the derivative function, f ′ (x). The f ′ (x) function will take a
number c as input and return the derivative of f at c. This calls for a definition.
f (x + h) − f (x)
f ′ (x) = lim
h→0 h
[Link]/watch?v=yPzNYlzA0Js
is the derivative of f .
Let y = f (x). The following notations all represent the derivative of Figure 2.1.20 Video presentation of De-
f: finition 2.1.19
dy df d d
f ′ (x) = y ′ = = = (f ) = (y).
dx dx dx dx
dy
Important: The notation dx is one symbol; it is not the fraction “dy/dx”. The
notation, while somewhat confusing at first, was chosen with care. A fraction-
looking symbol was chosen because the derivative has many fraction-like prop-
erties. Among other places, we see these properties at work when we talk about
the units of the derivative, when we discuss the Chain Rule, and when we learn
about integration (topics that appear in later sections and chapters).
Examples will help us understand this definition.
= lim (3h + 6x + 5)
h→0
= 6x + 5
f (x + h) − f (x)
f ′ (x) = lim
h→0 h
1
− 1
= lim x+h+1 x+1
h→0 h
Now find common denominator then subtract; pull 1/h out front to fa-
cilitate reading. [Link]/watch?v=JKHbXYanjDs
1 x+1 x+h+1
= lim · −
h→0 h (x + 1)(x + h + 1) (x + 1)(x + h + 1)
[Link]/watch?v=vsnDopbWHXQ
CHAPTER 2. DERIVATIVES 71
f (0 + h) − f (0)
f ′ (0) = lim .
h→0 h
Since x = 0 is the point where our function’s definition switches
from one piece to the other, we need to consider left and right-hand
limits. Consider the following, where we compute the left and right hand
limits side by side.
f (0 + h) − f (0) f (0 + h) − f (0)
lim lim
h→0− h h→0+ h
−h − 0 h−0
= lim− = lim+
h→0 h h→0 h
= lim− −1 = lim+ 1
h→0 h→0
y
= −1 =1
1
The last lines of each column tell the story: the left and right hand
limits are not equal. Therefore the limit does not exist at 0, and f is not 0.5
differentiable at 0. So we have
x
(
−1 −0.5
′ −1 x < 0 0.5 1
f (x) = . −0.5
1 x>0
−1
At x = 0, f ′ (x) does not exist; there is a jump discontinuity at 0; see
Figure 2.1.28. So f (x) = |x| is differentiable everywhere except at 0.
Figure 2.1.28 A graph of the deriva-
The point of non-differentiability came where the piecewise defined func- tive of f (x) = |x|
tion switched from one piece to the other. Our next example shows that this
does not always cause trouble.
f (π/2 + h) − f (π/2)
lim
h→0− h
sin(π/2 + h) − sin(π/2)
= lim−
h→0 h
sin( π2 ) cos(h) + sin(h) cos( π2 ) − sin( π2 )
= lim−
h→0 h
1 · cos(h) + sin(h) · 0 − 1
= lim−
h→0 h
cos(h) − 1 sin(h)
= lim− · lim−
h→0 h h→0 h
=1·0
= 0.
f (π/2 + h) − f (π/2) 1
lim+
h→0 h
1−1 0.5
= lim+
h→0 h
0
= lim x
h→0+ h π
2
= 0.
Since both the left and right hand limits are 0 at x = π/2, the limit Figure 2.1.31 A graph of f ′ (x) in Ex-
exists and f ′ (π/2) exists (and is 0). Therefore we can fully write f ′ as ample 2.1.29.
(
cos(x) x ≤ π/2
f ′ (x) = .
0 x > π/2
For all the functions f in this text, we can determine differentiability on [a, b]
by considering the limits limx→a+ f ′ (x) and limx→b− f ′ (x). This is often easier
to evaluate than the limit of the difference quotient.
[0, ∞). √
We state (without proof) that f ′ (x) = 1/ 2 x . Note that limx→0+ f ′ (x) =
∞; this limit was easier to evaluate than the limit of the difference quo-
tient, though it required us to already know the derivative of f .
Now consider g:
p √
g(a + h) − g(a) (0 + h)3 − 0
lim = lim+
h→0+ h h→0 h
h3/2 y
= lim+
h→0 h 1
= lim+ h1/2 = 0.
h→0
y = x1/2
As the one-sided limit exists at x = 0, we conclude g is differentiable
0.5
on its domain of [0, ∞). √
We state (without proof) that g ′ (x) = 3 x/2. Note that limx→0+ g ′ (x) = y = x3/2
0; again, this limit is easier to evaluate than the limit of the difference
quotient. x
−0.2
√ The two functions are graphed in Figure 2.1.35. Note how f (x) =
0.2 0.4 0.6 0.8 1 1.2
x seems to “go vertical” as x approaches 0, implying the slopes of its
tangent lines are growing toward
√ infinity. Also note how the slopes of Figure 2.1.35 A graph of y = x1/2 and
the tangent lines to g(x) = x3 approach 0 as x approaches 0. y = x3/2 in Example 2.1.34
Most calculus textbooks omit this topic and simply avoid specific cases where
it could be applied. We choose in this text to not make use of the topic unless
it is “needed.” Many theorems in later sections require a function f to be differ-
entiable on an open interval I; we could remove the word “open” and just use
“. . . on an interval I,” but choose to not do so in keeping with the current math-
ematical tradition. Our first use of differentiability on closed intervals comes in
Chapter 7, where we measure the lengths of curves.
This section defined the derivative; in some sense, it answers the question of
“What is the derivative?” The next section addresses the question “What does
the derivative mean?”
CHAPTER 2. DERIVATIVES 76
2.1.3 Exercises
Terms and Concepts
1. (□ True □ False) Let f be a position function. The average rate of change on [a, b] is the slope of the line
through the points (a, f (a)) and (b, f (b)).
2. (□ True □ False) The definition of the derivative of a function at a point involves taking a limit.
3. In your own words, explain the difference between the average rate of change and instantaneous rate of change.
4. In your own words, explain the difference between Definitions 2.1.7 and Definition 2.1.19.
5. Let y = f (x). Give three different notations equivalent to “f ′ (x).”
6. If two lines are perpendicular, what is true of their slopes?
Problems
Exercise Group. Use the definition of the derivative to compute the derivative of the given function.
7. f (x) = 6 8. f (x) = 2x
Show your work. Show your work.
9. f (t) = 4 − 3t 10. g(x) = x2
Show your work. Show your work.
11. h(x) = x 3
12. f (x) = 3x2 − x + 4
Show your work. Show your work.
13. r(x) = x1 14. 1
r(s) = s−2
Show your work. Show your work.
Exercise Group. A function and an x-value are given. (Note: these functions are the same as those given in Exer-
cises 2.1.7 through Exercise 2.1.14.) Give the equations of the tangent line and the normal line at that x-value.
15. f (x) = 6 at x = −2 16. f (x) = 2x at x = 3
17. f (x) = 4 − 3x at x = 7 18. g(x) = x2 at x = 2
19. h(x) = x3 at x = 4 20. f (x) = 3x2 − x + 4 at x = −1
21. r(x) = 1
x at x = −2 22. r(x) = 1
x−2 at x = 3
Exercise Group. A function f and an x-value a are given. Approximate the equation of the tangent line to the graph
of f at x = a by numerically approximating f ′ (a), using h = 0.1.
23. f (x) = x2 − 2x + 5 and a = −2 24. f (x) = − x+8
10
and a = 4
25. f (x) = ex and a = −4 26. f (x) = cos(x) and a = 0
y
3
x
−2 −1 1 2
−1
(a) Use the graph to approximate the slope of the tangent line to f at (−1, 0), (0, −1), and (2, 3).
x
−1 1 2 3
(a) Use the graph to approximate the slope of the tangent line to f at (0, 1) and (1, 0.5).
(b) Using the definition of the derivative, find f ′ (x).
Show your work.
(c) Use the derivative to find the slope of the tangent line at the points (0, 1) and (1, 0.5).
Exercise Group. A graph of a function f (x) is given. Using the graph, sketch f ′ (x).
CHAPTER 2. DERIVATIVES 78
29. 30.
y y
3
2
2
1 x
−6 −4 −2 2
x
−2 −1 1 2 3 4 5
−2
−1
31. 32.
y 1 y
4
0.5
2
x x
−3 −2 −1 1 2 3 −2π −π π 2π
−2
−0.5
−4
−1
Exercise Group. Use the graph of the function to answer the following questions.
x 2
−2 −1 1 2
−2 x
−2 −1 1 2
−4
−2
Exercise Group. A function f (x) is given, along with its domain and derivative. Determine if f (x) is differentiable
on its domain.
CHAPTER 2. DERIVATIVES 79
p 3/2
35. f (x) = x5 (1 − x), domain is [0, 1], f ′ (x) = (5−6x)x
√
2 1−x
(□ yes □ no)
√
√ sin( x)
36. f (x) = cos ( x) , domain is [0, ∞), f ′ (x) = − 2√x
(□ yes □ no)
CHAPTER 2. DERIVATIVES 80
measured in units “P per Q”, or “P /Q.” Here we see the fraction-like behavior
dy units of y
of the derivative in the notation: the units of dx are units of x .
Let P (t) represent the world population t minutes after 12:00 a.m., Jan-
uary 1, 2012. It is fairly accurate to say that P (0) = 7,028,734,178
([Link]). It is also fairly accurate to state that P ′ (0) = 156; that is,
at midnight on January 1, 2012, the population of the world was growing
by about 156 people per minute (note the units). Twenty days later (or
28,800 minutes later) we could reasonably assume the population grew
by about 28,800 · 156 = 4,492,800 people.
f (c + h) ≈ f (c) + f ′ (c) · h.
is behaving by looking at the slopes of its tangent lines. We explore this idea in
10
the following example.
5
Example 2.2.7 Understanding the graph of the derivative.
Consider the graph of f (x) and its derivative, f ′ (x), in Figure 2.2.8. Use x
these graphs to find the slopes of the tangent lines to the graph of f at −1 1 2 3 4
x = 1, x = 2, and x = 3.
Figure 2.2.6 A graph of f (x) = x2
Solution. To find the appropriate slopes of tangent lines to the graph
and tangent lines at x = 1 and x = 3
of f , we need to look at the corresponding values of f ′ .
• The slope of the tangent line to f at x = 1 is f ′ (1); this looks to
be about −1. 5
y f (x)
Consider again the graph of f (x) and its derivative f ′ (x) in Example 2.2.7. f (x)
y
Use the tangent line to f at x = 3 to approximate the value of f (3.1). 5
Solution. Figure 2.2.11 shows the graph of f along with its tangent line,
f ′ (x)
zoomed in at x = 3. Notice that near x = 3, the tangent line makes an x
excellent approximation of f . Since lines are easy to deal with, often 1 2 3
it works well to approximate a function with its tangent line. (This is
especially true when you don’t actually know much about the function
at hand, as we don’t in this example.) −5
While the tangent line to f was drawn in Example 2.2.7, 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
Figure 2.2.9 Graphs of f and f ′ in Ex-
ample 2.2.7
CHAPTER 2. DERIVATIVES 84
= 3(3.1 − 3) + 4 4
= 0.1 · 3 + 4
= 4.3. 3
For instance, we can easily observe the location of an object and its instan-
taneous velocity at a particular point in time. We do not have a “function f ”
for the location, just an observation. This is enough to create an approximating
function for f .
This last example has a direct connection to our approximation method ex-
plained above after Example 2.2.2. 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 computing the tangent
line at (c, f (c)) is easy: y(x) = f ′ (c)(x − c) + f (c). In Example 2.2.10, we used
the tangent line to approximate a value of f . Let’s 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:
This is the exact same approximation method used above! Not only does
it make intuitive sense, as explained above, it makes analytical sense, as this
approximation method is simply using a tangent line to approximate a function’s
value.
The importance of understanding the derivative cannot be understated. When
f is a function 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.
CHAPTER 2. DERIVATIVES 85
2.2.5 Exercises
Terms and Concepts
Problems
Exercise Group. Graphs of functions f and g are given. Identify which function is the derivative of the other.
CHAPTER 2. DERIVATIVES 86
15. 16.
y y
f (x) f (x)
4 5
g(x) g(x)
x x
−2 −1 1 2 3 4
−4 −2 2 4
−2
−5
−4
• f is the derivative of g.
• f is the derivative of g.
• g is the derivative of f .
• g is the derivative of f .
17. 18.
y 2 y
f (x)
f (x)
4 g(x) g(x)
1
2
x x
−4 −2 2 4 −4 −2 2 4
−2
−1
−4
−2
Constant Rule d
= 0, where c is a constant.
dx (c)
Power Rule d n
= nxn−1 , where n is an integer, n >
dx (x )
0.
Other common d
dx (sin(x)) = cos(x)
functions
dx (cos(x)) = − sin(x)
d
d x x
dx (e ) = e
d 1
dx (ln(x)) = x , for x > 0.
This theorem starts by stating an intuitive fact: constant functions have zero
rate of change as they are constant. Therefore their derivative is 0 (they change
at the rate of 0). The theorem then states some fairly amazing things. The Power
Rule states that the derivatives of Power Functions (of the form y = xn ) are very [Link]/watch?v=wPJ8-zKc1n0
straightforward: multiply by the power, then subtract 1 from the power. We see Figure 2.3.2 Video explanation of The-
something incredible about the function y = ex : it is its own derivative. We orem 2.3.1
also see a new connection between the sine and cosine functions.
One special case of the Power Rule is when n = 1, i.e., when f (x) = x.
What is f ′ (x)? According to the Power Rule,
d d 1
f ′ (x) = (x) = x = 1 · x0 = 1.
dx dx
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?”
CHAPTER 2. DERIVATIVES 88
The answer is: 1. They change at the same rate. We can also interpret the
derivative as the slope of the tangent line to the function at a point (c, f (c)).
Since f (x) = x is a linear function with constant slope 1, we can say that the
derivative of f (x) = x is f ′ (x) = 1.
Theorem 2.3.1 states that the natural exponential function has a remarkable
propery: it is equal to its own derivative! The video in Figure 2.3.3 explains why
this is the case.
Let’s practice using this theorem.
Example 2.3.4 Using common derivative rules to find, and use, deriva-
tives. [Link]/watch?v=ipKTEdQFBjw
Solution.
1. The Power Rule states that if f (x) = x3 , then f ′ (x) = 3x2 .
2. To find the equation of the line tangent to the graph of f at x =
−1, we need a point and the slope. The point is (−1, f (−1)) =
(−1, −1). The slope is f ′ (−1) = 3. Thus the tangent line has [Link]/watch?v=lyAEJSmSr-A
equation y = 3(x − (−1)) + (−1) = 3x + 2.
3. We can use the tangent line to approximate (−1.1)3 since −1.1 is
close to −1. We have ℓ(x)
y
f ′ (x)
(−1.1)3 ≈ 3(−1.1) + 2 = −1.3. 4
f (x)
We can easily find the actual value: (−1.1)3 = −1.331. 2
Theorem 2.3.1 gives useful information, but we will need much more. For −2
instance, using the theorem, we can easily find the derivative of y = x3 , but it
does not tell how to compute the derivative of y = 2x3 , y = x3 + sin(x) nor −4
y = x3 sin(x). The following theorem helps with the first two of these examples
(the third is answered in the next section). Figure 2.3.5 A graph of f (x) = x3 ,
along with its derivative f ′ (x) = 3x2
and its tangent line at x = −1
CHAPTER 2. DERIVATIVES 89
Sum/Difference Rule
d d d
(f (x) ± g(x)) = (f (x)) ± (g(x))
dx dx dx
= f ′ (x) ± g ′ (x)
Constant Multiple
Rule d d
(c · f (x)) = c · (f (x))
dx dx
′
= c · f (x).
[Link]/watch?v=Hr0sQcVhQ9A
Figure 2.3.7 Video presentation of The-
orem 2.3.6
While we will be mainly focused on using these rules, it can also be interest-
ing to see where they come from. Fortunately, it is not too difficult to establish
these rules using the definition of the derivative. The video in Figure 2.3.8 shows
why the sum rule is true.
Theorem 2.3.6 allows us to find the derivatives of a wide variety of functions.
It can be used in conjunction with the Power Rule to find the derivatives of any
polynomial. Recall in Example 2.1.22 that we found, using the limit definition,
the derivative of f (x) = 3x2 + 5x − 7. We can now find its derivative without
expressly using limits:
d d 2 d d
3x2 + 5x − 7 = 3 x + 5 (x) − (7)
dx dx dx dx
= 3 · 2x + 5 · 1 − 0 [Link]/watch?v=nVVpyilxZTw
= 6x + 5.
Figure 2.3.8 Proving the sum rule
We were a bit pedantic here, showing every step. Normally we would do
d
all the arithmetic and steps in our head and readily find dx 3x2 + 5x + 7 =
6x + 5.
8
7.2
6
l(x)
4 7
2
f (x) 6.8
x
x
−1 1 2 3 4 5 2.6 2.8 3 3.2 3.4
In general, when finding the fourth derivative and on, we resort to the f (4) (x)
notation, not f ′′′′ (x); after a while, too many ticks is confusing.
Let’s practice using this new concept.
Note how we have come right back to f (x) again. (Can you quickly
figure what f (23) (x) is?)
3. Employing Theorem 2.3.1 and the Constant Multiple Rule, we can
see that
enthusiasts talk of how fast a car can go from 0 to 60 mph; they are bragging
about the acceleration of the car.
We started this chapter with amusement park riders free-falling with posi-
tion function f (t) = −16t2 + 150. It is easy to compute f ′ (t) = −32t ft/s and
f ′′ (t) = −32 (ft/s)/s. We may recognize this latter constant; it is the accelera-
tion due to gravity. In keeping with the unit notation introduced in the previous
section, we say the units are “feet per second per second.” This is usually short-
ened to “feet per second squared,” written as “ft/s2 .”
It can be difficult to consider the meaning of the third, and higher order,
derivatives. The third derivative is “the rate of change of the rate of change of
the rate of change of f .” That is essentially meaningless to the uninitiated. In
the context of our position/velocity/acceleration example, the third derivative
is the “rate of change of acceleration,” commonly referred to as “jerk.”
Make no mistake: higher order derivatives have great importance even if
their practical interpretations are hard (or “impossible”) to understand. The
mathematical topic of series makes extensive use of higher order derivatives.
CHAPTER 2. DERIVATIVES 93
2.3.3 Exercises
Terms and Concepts
d n
1. What is the name of the rule which states that dx (x ) = nxn−1 , where n > 0 is an integer?
d
2. What is dx (ln(x))?
• sin(x) cos(x)
√
• x
• 5 ln(x)
6. Explain in your own words how to find the third derivative of a function f (x).
7. Give an example of a function where f ′ (x) ̸= 0 and f ′′ (x) = 0.
8. Explain in your own words what the second derivative “means”.
9. If f (x) describes a position function, then f ′ (x) describes what kind of function? What kind of function is
f ′′ (x)?
10. Let f (x) be a function measured in pounds (lb), where x is measured in feet (ft). What are the units of f ′′ (x)?
Problems
(a) Rewrite this identity when b = e, i.e., using loge (x) = ln(x), with a = 10.
(b) Use part (a) to find the derivative of y = log10 (x).
Exercise Group. Compute the first four derivatives of the given function.
CHAPTER 2. DERIVATIVES 94
Exercise Group. Find the equations of the tangent and normal lines to the graph of the function at the given point.
33. f (x) = x3 + 8x at x = 2 34. f (t) = et − 2 at t = 0
35. g(x) = ln(x) at x = 1 36. f (x) = 4 sin(x) at x = π/6
37. f (x) = −2 cos(x) at x = π/6 38. f (x) = 9 − 9x at x = −9
CHAPTER 2. DERIVATIVES 95
d
Warning 2.4.3 dx (f (x)g(x)) ̸= f ′ (x)g ′ (x)! While this would be simpler than
the Product Rule, it is wrong.
We practice using this new rule in an example, followed by an example that Video Solution
demonstrates why this theorem is true.
Use the Product Rule to compute the derivative of y = 5x2 sin(x). Eval-
uate the derivative at x = π/2.
Solution. To make our use of the Product Rule explicit, let’s set f (x) =
5x2 and g(x) = sin(x). We easily compute/recall that f ′ (x) = 10x and
g ′ (x) = cos(x). Employing the rule, we have
[Link]/watch?v=37efDywkDyE
d d d
5x2 sin(x) = 5x2 sin(x) + 5x2 (sin(x))
dx dx dx
= 10x sin(x) + 5x2 cos(x).
20 y
At x = π/2, we have
π π π 2 π 15
′
y (π/2) = 10 · sin +5 cos = 5π.
2 2 2 2
10
We graph y and its tangent line at x = π/2, which has a slope of 5π,
in Figure 2.4.5. While this does not prove that the Product Rule is the 5
correct way to handle derivatives of products, it helps validate its truth.
x
We now investigate why the Product Rule is true. π
2
π
Proof of Product Rule. We can use the definition of the derivative to prove The-
Figure 2.4.5 A graph of y = 5x2 sin(x)
orem 2.4.2.
and its tangent line at x = π/2
By the limit definition, we have
d f (x + h)g(x + h) − f (x)g(x)
(f (x)g(x)) = lim .
dx h→0 h
We now do something a bit unexpected; add 0 to the numerator (so that
nothing is changed) in the form of − f (x)g(x + h) + f (x)g(x + h), then do
some regrouping as shown. Adding 0 in some clever form is
a common mathematical proof
technique.
CHAPTER 2. DERIVATIVES 96
d f (x + h)g(x + h) − f (x)g(x)
(f (x)g(x)) = lim
dx h→0 h
(now add 0 to the numerator)
f (x + h)g(x + h) − f (x)g(x + h) + f (x)g(x + h) − f (x)g(x)
= lim
h→0 h
(regroup)
[f (x + h)g(x + h) − f (x)g(x + h)] + [f (x)g(x + h) − f (x)g(x)]
= lim
h→0 h
(split fraction)
f (x + h)g(x + h) − f (x)g(x + h) f (x)g(x + h) − f (x)g(x)
= lim + lim
h→0 h h→0 h
(factor) Proving the product rule.
f (x + h) − f (x) g(x + h) − g(x)
= lim g(x + h) + lim f (x)
h→0 h h→0 h
(apply limit properties)
f (x + h) − f (x) g(x + h) − g(x)
= lim · lim g(x + h) + f (x) · lim
h→0 h h→0 h→0 h
(apply limits)
= f ′ (x)g(x) + f (x)g ′ (x)
(by definition of the derivative). [Link]/watch?v=i791Y97O5hI
We have proven the product rule as desired. (In the last step, we also relied
on the fact that since g is differentiable, it is also continuous, which guarantees
that limh→0 g(x + h) = g(x).) ■
It is often 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 derivative of a product of functions in two ways to verify that
the Product Rule is indeed “right.”
Instead, let’s apply the Product Rule to the original factored form:
d 2 d
y′ = x + 3x + 1 (2x2 − 3x + 1) + (x2 + 3x + 1) 2x2 − 3x + 1
dx dx [Link]/watch?v=-plvLFQ21Ig
= (2x + 3)(2x2 − 3x + 1) + (x2 + 3x + 1)(4x − 3)
= 4x3 − 7x + 3 + 4x3 + 9x2 − 5x − 3
= 8x3 + 9x2 − 12x.
Example 2.4.7 Using the Product Rule with a product of three func-
tions.
Video Solution
Let y = x3 ln(x) cos(x). Find y ′ .
Solution. We have a product of three functions while the Product Rule
only specifies how to handle a product of two functions. Our method
of handling this problem is to simply group the latter two functions to-
gether, and consider y = x3 · [ln(x) cos(x)]. Following the Product Rule,
we have
d 3 d
y′ = x ln(x) cos(x) + (x3 ) (ln(x) cos(x))
dx dx
[Link]/watch?v=PYK64WB4JUg
d
To evaluate dx (ln(x) cos(x)),
we apply the Product Rule again:
1
y ′ = 3x2 [ln(x) cos(x)] + (x3 ) cos(x) + ln(x)(− sin(x))
x
1
= 3x2 ln(x) cos(x) + x3 cos(x) + x3 ln(x)(− sin(x)).
x
Recognize the pattern in our answer above: when applying the Prod-
uct Rule to a product of three functions, there are three terms added
together in the final derivative. Each term contains only one derivative
of one of the original functions, and each function’s derivative shows up
in only one term. It is straightforward to extend this pattern to finding
the derivative of a product of four or more functions.
Ultimately though, we would simplify our final computation to:
If you check this answer with a cas, it may factor and give the answer:
Now that we have the hang of the product rule pattern, it’s not much more
difficult to move on to products of four or more functions, as the video in Fig-
ure 2.4.8 demonstrates.
We consider one more example before discussing another derivative rule.
Solution. Recalling that the derivative of ln(x) is 1/x, we use the Prod-
uct Rule to find our answers.
CHAPTER 2. DERIVATIVES 98
The Quotient Rule is not hard to use, although it might be a bit tricky to re-
member. A useful mnemonic works as follows. Consider a fraction’s numerator
and denominator as “HI” and “LO”, respectively. Then
[Link]/watch?v=IlsA8342GvQ
d HI LO · dHI − HI · dLO
= , Figure 2.4.11 Video presentation of The-
dx LO LOLO
orem 2.4.10
read “low dee high minus high dee low, over low low.” Said fast, that phrase can
roll off the tongue, making it easy to memorize. The “dee high” and “dee low”
parts refer to the derivatives of the numerator and denominator, respectively.
Let’s practice using the Quotient Rule.
Video Solution
Example 2.4.12 Using the Quotient Rule.
5x2
Let f (x) = sin(x) . Find f ′ (x).
Solution. Directly applying the Quotient Rule gives:
d 5x2 sin(x) · dx
d
5x2 − 5x2 · dxd
(sin(x))
= 2
dx sin(x) sin (x)
10x sin(x) − 5x2 cos(x)
= . [Link]/watch?v=Hr4bt6yFwPg
sin2 (x)
We include this result in the following theorem about the derivatives of the −5
trigonometric functions. Recall we found the derivative of y = sin(x) in Exam-
ple 2.1.24 and stated the derivative of the cosine function in Theorem 2.3.1. The −10
derivatives of the cotangent, cosecant and secant functions can all be computed
directly using Theorem 2.3.1 and the Quotient Rule. Figure 2.4.14 A graph of y = tan(x)
along with its tangent line at x = π/4
Theorem 2.4.15 Derivatives of Trigonometric Functions.
d d
1. (sin(x)) = cos(x) 4. (cot(x)) = − csc2 (x)
dx dx
d d
2. (cos(x)) = − sin(x) 5. (sec(x)) = sec(x) tan(x)
dx dx
d d
3. (tan(x)) = sec2 (x) 6. (csc(x)) = − csc(x) cot(x)
dx dx
To remember the above, it may be helpful to keep in mind that the deriv-
atives of the trigonometric functions that start with “c” have a minus sign in
them.
rem 2.4.15 and verify the two answers are the same.
10x sin(x)−5x2 cos(x)
Solution. We found in Example 2.4.12 that f ′ (x) = sin2 (x)
.
CHAPTER 2. DERIVATIVES 100
We now find f ′ using the Product Rule, considering f as f (x) = 5x2 csc(x).
d
f ′ (x) = 5x2 csc(x)
dx
d d
= 5x2 (csc(x)) + 5x2 csc(x)
dx dx
= 5x2 (− csc(x) cot(x)) + 10x csc(x) (now rewrite trig functions)
−1 cos(x) 10x
= 5x2 · · +
sin(x) sin(x) sin(x)
−5x2 cos(x) 10x
= + (get common denominator)
sin2 (x) sin(x)
10x sin(x) − 5x2 cos(x)
= .
sin2 (x)
Finding f ′ using either method returned the same result. At first, the
answers looked different, but some algebra verified they are the same.
In general, there is not one final form that we seek; the immediate result
from the Product Rule is fine. Work to “simplify” your results into a form
that is most readable and useful to you.
The Quotient Rule gives other useful results, as shown in the next example.
Example 2.4.17 Using the Quotient Rule to expand the Power Rule.
Video Solution
Find the derivatives of the following functions.
1
1. f (x) =
x
1
2. f (x) = , where n > 0 is an integer.
xn
Solution. We employ the Quotient Rule.
1.
[Link]/watch?v=jPqqK-ObPm4
x·0−1·1
f ′ (x) =
x2
1
=− 2
x
2.
xn · 0 − 1 · nxn−1
f ′ (x) =
(xn )2
n−1
nx
= − 2n
x
n
= − n+1 .
x
The derivative of y = x1n turned out to be rather nice. It gets better. Con-
sider:
d 1 d −n
n
= x (apply result from Example 2.4.17)
dx x dx
n
= − n+1 (rewrite algebraically)
x
CHAPTER 2. DERIVATIVES 101
= −nx−(n+1)
= −nx−n−1 .
This is reminiscent of the Power Rule: multiply by the power, then subtract
1 from the power. We now add to our previous Power Rule, which had the re-
striction of n > 0.
f ′ (x) = n · xn−1 .
x2 −3x+1
Let f (x) = x . Find f ′ (x) in each of the following ways:
2. By rewriting f , we can apply the Product Rule and Power Rule with
Integer Exponents as follows:
d −1 d 2
f ′ (x) = x2 − 3x + 1 x + x − 3x + 1 x−1
dx dx
= x2 − 3x + 1 · (−1)x−2 + (2x − 3) · x−1
x2 − 3x + 1 2x − 3
=− +
x2 x
x2 − 3x + 1 2x2 − 3x
=− +
x2 x2
CHAPTER 2. DERIVATIVES 102
x2 − 1 1
= = 1 − 2,
x2 x
the same result as above.
3. As x ̸= 0, we can divide through by x first, giving f (x) = x − 3 +
x−1 . Now apply the Power Rule with Integer Exponents.
1
f ′ (x) = 1 − ,
x2
the same result as before.
They are equal; they are all correct; only the first is “simple.” Work to make
answers simple.
In the next section we continue to learn rules that allow us to more easily
compute derivatives than using the limit definition directly. We have to memo-
rize the derivatives of a certain set of functions, such as “the derivative of sin(x)
is cos(x).” The Sum/Difference Rule, Constant Multiple Rule, Power Rule with
Integer Exponents, Product Rule and Quotient Rule show us how to find the de-
rivatives of certain combinations of these functions. The next section shows how
to find the derivatives when we compose these functions together.
CHAPTER 2. DERIVATIVES 103
2.4.1 Exercises
Terms and Concepts
1. (□ True □ False) The Product Rule states that d
x2 sin(x) = 2x cos(x).
dx
2
cos(x)
2. (□ True □ False) The Quotient Rule states that dx
d x
sin(x) = 2x .
3. (□ True □ False) The derivatives of the trigonometric functions that start with “c” have minus signs in
them.
4. What derivative rule is used to extend the Power Rule to include negative integer exponents?
5. (□ True □ False) Regardless of the function, there is always exactly one right way of computing its deriv-
ative.
6. In your own words, explain what it means to make your answers “clear.”
Problems
Exercise Group.
(a) Use the Product Rule to differentiate the function.
(b) Manipulate the function algebraically and differentiate without using the Product Rule.
Exercise Group.
csc(z) sec(θ)
29. g(z) = cos(z)+2 30. g(θ) = θ4 sec(θ) + θ4
Exercise Group. Find the equations of the tangent and normal lines to the graph of g at the indicated point.
37. g(x) = ex x2 − 7 at (0, −7) 38. g(x) = x cos(x) at 5π 5π
3 , 6
x2 sin(x)−2x
39. g(x) = x−(−4) at (−5, −25) 40. g(x) = x−8 at (0, 0)
Exercise Group. Find the x-values where the graph of the function has a horizontal tangent line.
41. f (x) = x2 − 17x − 29 42. f (x) = x sin(x) on [−1, 1]
2x 3x2
43. f (x) = −3x+3 44. f (x) = x−2
fore giving the corresponding differentiation rule, we note that the rule extends 1
to multiple compositions like f (g(h(x))) or f (g(h(j(x)))), etc.
To motivate the rule, let’s look at three derivatives we can already compute.
x
Example 2.5.3 Exploring similar derivatives. 0.5 1 1.5 2 2.5 3 3.5
Use the Chain Rule to find the derivatives of the functions F1 (x), F2 (x),
and F3 (x), as given in Example 2.5.3.
Solution. Example 2.5.3 ended with the recognition that each of the
given functions was actually a composition of functions. To avoid confu-
sion, we ignore most of the subscripts here.
y ′ = f ′ (g(x)) · g ′ (x)
= 2(1 − x) · (−1)
= −2(1 − x)
= 2x − 2.
CHAPTER 2. DERIVATIVES 107
y ′ = f ′ (g(x)) · g ′ (x)
= 3(1 − x)2 · (−1)
= −3(1 − x)2 .
F3 (x) = (1 − x)4 Finally, when y = (1−x)4 , we have f (x) = x4
and g(x) = (1 − x). Thus f ′ (x) = 4x3 and
f ′ (g(x)) = 4(1 − x)3 . Thus
y ′ = f ′ (g(x)) · g ′ (x)
= 4(1 − x)3 · (−1)
= −4(1 − x)3 .
d
· g ′ (x).
n−1
(g(x)n ) = n · (g(x))
dx
3. y = e−x .
2
1. y = sin(2x). 2. y = ln(4x3 −
2x2 ). Video Solution
Solution.
1. Consider y = sin(2x). Recognize that this is a composition of
functions, where f (x) = sin(x) and g(x) = 2x. Thus
y ′ = f ′ (g(x)) · g ′ (x)
d
= cos(2x) · (2x)
dx
[Link]/watch?v=yW1BbOeDFcM
CHAPTER 2. DERIVATIVES 108
= cos(2x) · 2
= 2 cos(2x).
2. Recognize that y = ln 4x3 − 2x2 is the composition of f (x) =
ln(x) and g(x) = 4x3 − 2x2 . Also, recall that
d 1
(ln(x)) = .
dx x
This leads us to:
1 d
y′ = · 4x3 − 2x2
4x3 − 2x2 dx
1
= 3 · 12x2 − 4x
4x − 2x 2
12x2 − 4x
= 3
4x − 2x2
4x(3x − 1)
=
2x(2x2 − x)
2(3x − 1)
= .
2x2 − x
Note that ln 4x3 − 2x2 = ln 4x2 (x − 1/2) was only defined
for x > 1/2, so the result of y ′ = 2(3x−1)
2x2 −x is only valid for x > 1/2
as well.
d
y ′ = e−x ·
2
−x2
dx
= e−x
2
· (−2x)
= −2xe−x .
2
Let f (x) = cos(x2 ). Find the equation of the line tangent to the graph
of f at x = 1.
Solution. The tangent line goes through the point (1, f (1)) ≈ (1, 0.54)
with slope f ′ (1). To find f ′ , we need the The Chain Rule.
f ′ (x) = − sin(x2 ) · (2x) = −2x sin(x2 ). Evaluated at x = 1, we
have f ′ (1) = −2 sin(1) ≈ −1.68. Thus the equation of the tangent line 1 y
is approximated by
0.5
y ≈ −1.68(x − 1) + 0.54.
x
The tangent line is sketched along with f in Figure 2.5.10. −3 −2 −1 1 2 3
The The Chain Rule is used often in taking derivatives. Because of this, one −0.5
can become familiar with the basic process and learn patterns that facilitate find-
−1
1. d n
dx (u ) = n · un−1 · u′ . 4. d
dx (cos(u)) = − sin(u) · u′ .
2. d u
dx (e ) = e u · u′ .
3. d
dx (sin(u)) = cos(u) · u′ . 5. d
dx (tan(u)) = sec2 (u) · u′ .
Of course, the The Chain Rule can be applied in conjunction with any of the
other rules we have already learned. We practice this next.
Solution.
1. We must use the Product Rule and The Chain Rule. Do not think
that you must be able to “see” the whole answer immediately;
rather, just proceed step-by-step.
d d 5
f ′ (x) = x5 · sin 2x3 + sin 2x3 · x
dx dx
d [Link]/watch?v=2QJLR-Y-Ht8
= x5 cos 2x3 · 2x3 + 5x4 sin 2x3
dx
5 2 3
= x 6x cos 2x + 5x4 sin 2x3
= 6x7 cos 2x3 + 5x4 sin 2x3 .
2. We must employ the Quotient Rule along with the The Chain Rule.
Again, proceed step-by-step.
e−x · dx e−x
2 2
d
5x3 − 5x3 · dx
d
f ′ (x) = 2
e−x2
CHAPTER 2. DERIVATIVES 110
e−x · 15x2 − 5x3 · e−x · dx
2 2
d
−x2
= 2
e−x 2
e−x 15x2 − 5x3 (−2x)e−x
2 2
= 2
e−x2
e−x 10x4 + 15x2
2
=
e−2x2
2
= ex 10x4 + 15x2 .
f ′ (x)
= ln x2 + 5x4 · x · − sin x−2 · −2x−3
+ 1 · cos x−2 − 2 sin e4x · cos e4x · 4e4x
!
−2
2x + 20x3
− x cos x − sin e
2 4x
· 2
x + 5x4
. 2
ln x2 + 5x4 .
The The Chain Rule also has theoretic value. That is, it can be used to find the
derivatives of functions that we have not yet learned as we do in the following
example.
formally as a theorem.
f ′ (x) = ln(a) · ax .
y ′ = f ′ (g(x)) · g ′ (x)
dy
= y ′ (u) · u′ (x) since y = f (u) and u = g(x)
dx
dy dy du
= · (using “fractional notation” for the derivative)
dx du dx
Here the “fractional” aspect of the derivative notation 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 derivative
dy
notation of du is one symbol. It is equally important to realize that this notation
was chosen precisely because of this behavior. It makes applying the The Chain
Rule easy with multiple variables. For instance,
dy dy d⃝ d△
= · · .
dt d⃝ d△ dt
where ⃝ and △ are any variables you’d like to use.
One of the most common ways of “visualizing” the The Chain Rule is to con-
sider a set of gears, as shown in Figure 2.5.17. The gears have 36, 18, and 6 teeth,
respectively. That means for every revolution of the x gear, the u gear revolves
twice. That is, the rate at which the u gear makes a revolution is twice as fast as
the rate at which the x gear makes a revolution.
CHAPTER 2. DERIVATIVES 113
dy
=3
du
2.5.1 Exercises
Terms and Concepts
1. (□ True □ False) The Chain Rule describes how to evaluate the derivative of a composition of functions.
n−1
2. (□ True □ False) The Generalized Power Rule states that d n
dx (g(x) ) = n (g(x)) .
3. (□ True □ False) dx
d
ln x2 = x12 .
4. (□ True □ False) d x
dx (3 ) ≈ 1.1 · 3x .
5. (□ True □ False) dx
dy = dx
dt · dt
dy .
6. (□ True □ False) Taking the derivative of f (x) = x2 sin(5x) requires the use of both the Product and
Chain Rules.
Problems
5
14. p(t) = cos(5t)
13. k(y) = y + y1
15. p(q) = tan(2q) 16. f (θ) = cot θ2 + 3
17. g(t) = sin t6 + 1
t3
18. g(q) = cos5 (7q)
19. h(y) = cos3 y 2 + 3y − 3 20. j(t) = ln(cos(t))
21. j(q) = ln q 8 22. k(y) = 3 ln(y)
23. p(t) = 6t 24. p(z) = 2csc(z)
25. f (x) = 810 4t
26. g(t) = 9t
6w +5 y
27. h(w) = 5w +6 28. h(y) = 7 5+8
y
2 3
29. j(r) = 5r −r 30. k(w) = w cot(5w)
6r 2
6 3
31. p(x) = x2 + 4x 7x4 + x 32. m(r) = sin(8 − 4r) cos 6r + r2
33. m(w) = cos(4w − 5) sin(9 + 7w) 34.
2
f (x) = e8x sin x1
cos(6r+4) (3z+5)2
35. g(r) = (3r+1)3 36. h(z) = sin(9z)
Exercise Group. Find the equations of tangent and normal lines to the graph of the function at the given point. Note:
the functions here are the same as in Exercises 2.5.7 through Exercise 2.5.10.
10
37. f (x) = 4x3 − x at x = 0 38. f (x) = (3x − 2)5 at x = 1
39. g(x) = (sin(x) + cos(x))3 at x = π/2. 40. h(x) = e3x
2
+x−1
at x = −1
d
41. Compute dx (ln(kx)) two ways. First by using the Chain Rule. Second, by using the logarithm rule ln(ab) =
ln(a) + ln(b) and then taking the derivative.
Show the work for both parts.
d
42. Compute dx ln xk two ways. First by using the Chain Rule. Second, by using the logarithm rule ln(ap ) =
p ln(a) (for positive a) and then taking the derivative.
Show the work for both parts.
CHAPTER 2. DERIVATIVES 115
y
2.6.1 The method of implicit differentiation
2
Implicit differentiation is a technique based on the The Chain Rule that is used to
find a derivative when the relationship between the variables is given implicitly
rather than explicitly (solved for one variable in terms of the other). x
We begin by reviewing the Chain Rule. Let f and g be functions of x. Then −3 −2 −1 1 2 3
d
(f (g(x))) = f ′ (g(x)) · g ′ (x).
dx −2
cos(y)y ′ + 3y 2 y ′ = −3x2 .
cos(y)y ′ + 3y 2 y ′ = −3x2
cos(y) + 3y 2 y ′ = −3x2
−3x2
y′ =
cos(y) + 3y 2
Implicit functions are generally harder to deal with than explicit functions.
With an explicit function, given an x value, we have an explicit formula for com-
puting the corresponding y value. With an implicit function, one often has to
find x and y values at the same time that satisfy the equation. It is much eas-
ier to demonstrate that a given point satisfies the equation than to actually find
such a point. √
For instance, we can affirm easily that the point 3 6, 0 lies on the graph of
the implicit function sin(y)√+ y 3 = 6 − x3 . Plugging in 0 for y, we see the left
hand side is 0. Setting x = 3 6, we see the right hand side is also 0; the equation
is satisfied. The following example finds the equation of the tangent line to this
function at this point.
−3x2
y′ = .
cos(y) + 3y 2
√
We find the slope of the tangent line at the point 3 6, 0 by substi-
√ √
tuting 3 6 for x and 0 for y. Thus at the point 3 6, 0 , we have the slope [Link]/watch?v=qYrcm4ObwOM
as √ 2 √
′ −3 3 6 −3 3 36
y = = ≈ −9.91.
cos(0) + 3 · 02 1
y
Therefore the equation of the tangent line
√ to the
implicitly defined
function sin(y) + y 3 = 6 − x3 at the point 3 6, 0 is 2
√ √
y = −3 36 x − 6 + 0 ≈ −9.91x + 18.
3 3
x
−3 −2 −1 1 2 3
The curve and this tangent line are shown in Figure 2.6.5.
This suggests a general method for implicit differentiation. For the steps be-
−2
low assume y is a function of x.
1. Take the derivative of each term in the equation. Treat the x terms like
normal. When taking the derivatives of y terms, the usual rules apply Figure 2.6.5 The function sin(y)+y 3 =
3
6−x
√ and its tangent line at the point
3
( 6, 0)
CHAPTER 2. DERIVATIVES 117
except that, because of the Chain Rule 2.5.4, we need to multiply each
term by y ′ .
2. Get all the y ′ terms on one side of the equal sign and put the remaining
terms on the other side.
(Practical Note: when working by hand, it may be beneficial to use the sym-
dy ′ 1
bol dx instead of y , as the latter can be easily confused for y or y .)
[Link]/watch?v=BMn-BU6VTQU
CHAPTER 2. DERIVATIVES 118
the first term. It requires both the The Chain Rule and Product Rule.
d d 2 2
sin x2 y 2 = cos x2 y 2 · x y
dx dx
= cos x2 y 2 · x2 (2yy ′ ) + 2xy 2
= 2 x2 yy ′ + xy 2 cos x2 y 2 .
We leave the derivatives of the other terms to the reader. After tak-
ing the derivatives of both sides, we have
2 x2 yy ′ + xy 2 cos x2 y 2 + 3y 2 y ′ = 1 + y ′ .
From here we can safely move around terms to get the following:
x
−1
2x2 y cos x2 y 2 y ′ + 3y 2 y ′ − y ′ = 1 − 2xy 2 cos x2 y 2 . 1
′ 1 − 2xy 2 cos x2 y 2
y = 2 .
2x y cos(x2 y 2 ) + 3y 2 − 1
Figure 2.6.9 A graph ofthe implicitly
A graph of this implicit function is given in Figure 2.6.9. defined curve sin x2 y 2 +y 3 = x+y
It is easy to verify that the points (0, 0), (0, 1) and (0, −1) all lie on
the graph. We can find the slopes of the tangent lines at each of these
points using our formula for y ′ .
• At (0, 0), the slope is −1. y
Quite a few “famous” curves have equations that are given implicitly. We can −1
use implicit differentiation to find the slope at various points on those curves.
We investigate two such curves in the next examples.
Example 2.6.11 Finding slopes of tangent lines to a circle. Figure 2.6.10 A graph of the implicitly
defined curve sin x2 y 2 +y 3 = x+y
2 2
Find the
√ slope
of the tangent line to the circle x + y = 1 at the point and certain tangent lines
1/2, 3/2 .
Solution. Taking derivatives, we get 2x+2yy ′ = 0. Solving for y ′ gives:
−x
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
CHAPTER 2. DERIVATIVES 119
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 per-
pendicular. √
At the point 1/2, 3/2 , we have the tangent line’s slope as
−1/2 −1
y′ = √ = √ ≈ −0.577. y √
3/2 3 1 1/2, 3/2
√
A graph of the circle and its tangent line at 1/2, 3/2 is given in
0.5
Figure 2.6.12, along with a thin dashed line from the origin that is per-
pendicular to the tangent line. (It turns out that all normal lines to a x
circle pass through the center of the circle.) −1 −0.5 0.5 1
This section has shown how to find the derivatives of implicitly defined func- −0.5
tions, whose graphs include a wide variety of interesting and unusual shapes.
Implicit differentiation can also be used to further our understanding of “regu- −1
lar” differentiation.
One hole in our current understanding of derivatives is this: what is the de- Figure 2.6.12 The unit
√ circle with its
rivative of the square root function? That is, tangent line at (1/2, 3/2)
d √ d 1/2
x = x =?
dx dx
We allude to a possible solution, as we can write the square root function as
a power function with a rational (or, fractional) power. We are then tempted to
apply the Power Rule with Integer Exponents and obtain
d 1/2 1 −1/2 1
x = x = √ .
dx 2 2 x
The trouble with this is that the Power Rule with Integer Exponents was ini-
tially defined only for positive integer powers, n > 0. While we did not justify
this at the time, generally the Power Rule with Integer Exponents is proved us-
ing something called the Binomial Theorem, which deals only with positive in-
tegers. The Quotient Rule allowed us to extend the Power Rule with Integer Ex-
ponents to negative integer powers. Implicit Differentiation allows us to extend
the Power Rule with Integer Exponents to rational powers, as shown below.
Let y = xm/n , where m and n are integers with no common factors (so
m = 2 and n = 5 is fine, but m = 2 and n = 4 is not). We can rewrite this
explicit function implicitly as y n = xm . Now apply implicit differentiation.
y = xm/n
y n = xm
d n d m
(y ) = (x )
dx dx
n · y n−1 · y ′ = m · xm−1
m xm−1
y′ = (now substitute xm/n for y)
n y n−1
m xm−1
= (apply lots of algebra)
n (xm/n )n−1
m (m−n)/n
= x
n
m
= xm/n−1 .
n
CHAPTER 2. DERIVATIVES 120
The above derivation is the key to the proof extending the Power Rule with
Integer Exponents to rational powers. Using limits, we can extend this once
more to include all powers, including irrational (even transcendental!) powers,
giving the following theorem.
x
−20 −10 10 20
2.6.2 Implicit Differentiation and the Second Derivative
−10
We can use implicit differentiation to find higher order derivatives. In theory,
dy
this is simple: first find dx , then take its derivative with respect to x. In practice,
−20
it is not hard, but it often requires a bit of algebra. We demonstrate this in an
example.
Figure 2.6.16 An astroid with a tangent
Example 2.6.17 Finding the second derivative. line
d2 y
Given x2 + y 2 = 1, find dx2 = y ′′ .
CHAPTER 2. DERIVATIVES 121 Video Solution
d ′
y ′′ = (y )
dx
d x
= − (Now use the Quotient Rule.)
dx y
y · 1 − x(y ′ ) [Link]/watch?v=V6piqsjn2mk
=− replace y ′ with − x/y:
y2
y − x(−x/y)
=−
y2
y + x2 /y
=− .
y2
While this is not a particularly simple expression, it is usable. We can
see that y ′′ > 0 when y < 0 and y ′′ < 0 when y > 0. In Section 3.4, we
will see how this relates to the shape of the graph. In calculus the expression 00 is
Also, if we remember that we are only considering points on the also considered well-defined and
curve x2 + y 2 = 1, then we know that x2 = 1 − y 2 . So we can re- equal to 1. This is easily confused
place the x2 in the expression for y ′′ to get with a limit of the form 00 , which
is indeterminate. We skirt the is-
′′ y + 1 − y 2 /y 1 sue here.
y =− =− 3
y2 y
which is a simpler expression. Recognizing when simplifications like this 4 y
are possible is not always easy.
3
y = xx
ln(y) = ln(xx ) (apply logarithm rule)
[Link]/watch?v=6eL6WBlmItk
ln(y) = x ln(x) (now use implicit differentiation)
Figure 2.6.20 Two approaches to solv-
d d
(ln(y)) = (x ln(x)) ing Example 2.6.19. First, taking the
dx dx log of both sides. Second, using the
inverse relationship eln(x) = x.
CHAPTER 2. DERIVATIVES 122
y′ 1
= ln(x) + x · 4 y
y x
y′
= ln(x) + 1 3
y
y ′ = y (ln(x) + 1) (substitute y = xx )
′ x 2
y = x (ln(x) + 1) .
1.5, 1.51.5
To “test” our answer, let’s use it to find the equation of the tangent 1
line at x = 1.5. Thepoint on the graph our tangent line must pass
through is 1.5, 1.51.5 ≈ (1.5, 1.837). Using the equation for y ′ , we x
find the slope as 0.5 1 1.5 2
We would not have been able to compute the derivative of the function in
Example 2.6.19 without logarithmic differentiation. But the method is also use-
ful in cases where the product and quotient rules could be used, but logarithmic
differentiation is simpler. The video in Figure 2.6.22 provides such an example.
Implicit differentiation proves to be useful as it allows us to find the instan-
taneous rates of change of a variety of functions. In particular, it extended the
Power Rule for Differentiation to rational exponents, which we then extended
to all real numbers. In Section 2.7, implicit differentiation will be used to find
the derivatives of inverse functions, such as y = sin−1 (x).
[Link]/watch?v=3Cv2EgjH9ZE
Figure 2.6.22 Using logarithmic differ-
entiation
CHAPTER 2. DERIVATIVES 123
2.6.4 Exercises
Terms and Concepts
1. In your own words, explain the difference between implicit functions and explicit functions.
2. Implicit differentiation is based on what other differentiation rule?
√
3. (□ True □ False) Implicit differentiation can be used to find the derivative of y = x.
4. (□ True □ False) Implicit differentiation can be used to find the derivative of y = x3/4 .
Problems
√ √
7. p(t) = 9 + t 2 8. m(w) = w tan(w)
9. m(y) = y 1.2 10. f (r) = rπ + r3.8 + π 3.8
√
11. g(w) = w+(−8)
√ 12. h(x) = 6 x(cos(x) + ex )
w
dy
Exercise Group. Find dx using implicit differentiation.
13. x4 + y 2 + y = 7 14. x2/5 + y 2/5 = 1
Show your work using implicit differentiation. Show your work using implicit differentiation.
15. cos(x) + sin(y) = 1 x
16. = 10
Show your work using implicit differentiation. y
Show your work using implicit differentiation.
y 18. x2 ex + 2 y = 5
17. = 10
x Show your work using implicit differentiation.
Show your work using implicit differentiation.
4
19. x2 tan(y) = 50 20. 3x2 + 2y 3 = 2
Show your work using implicit differentiation. Show your work using implicit differentiation.
2 x2 +y
21. y 2 + 2y − x = 200 22. x+y 2 = 17
Show your work using implicit differentiation. Show your work using implicit differentiation.
sin(x)+y
23. =1 24. ln x2 + y 2 = e
cos(y)+x
Show your work using implicit differentiation. Show your work using implicit differentiation.
25. ln x2 + xy + y 2 = 1
Show your work using implicit differentiation.
dy
26. Show that dx is the same for each of the following implicitly defined functions.
(a) xy = 1
(b) x2 y 2 = 1
(c) sin(xy) = 1
(d) ln(xy) = 1
Exercise Group. Find the equation of the tangent line to the graph of the implicitly defined function at the indicated
points. As a visual aid, the function is graphed.
CHAPTER 2. DERIVATIVES 124
1 y 1 y
√ √
( 0.6, 0.8)
0.5 0.5
(0.1, 0.281)
x x
−1 −0.5 0.5 1 −1 −0.5 0.5 1
−0.5 −0.5
−1 −1
x
x
−2.5 −2 −1.5 −1 −0.5 0.5
−4 −2 2 4
−2 −1
√
−4 (2, − 4 108)
(a) At (0, 1).
√
(a) At (0, 4).
(b) At − 34 , 3 4 3 .
√
(b) At 2, − 4 108 .
CHAPTER 2. DERIVATIVES 125
x
−2 −1 1 2
2 √
4+3 3
, 1.5 −1
2 √
−1, −1−2 5
x −2
2 4 6
√ (a) At (−1, 1).
7 6+3 3
(a) At 2, 2 . √
(b) At −1, 12 (−1 + 5) .
√
(b) At 4+3 3 3
,2 . √
2 (c) At −1, 12 (−1 − 5) .
d2 y
Exercise Group. An implicitly defined function is given. Find dx2 . Note: these are the same functions used in
Exercises 2.6.13 through 2.6.16.
33. x4 + y 2 + y = 7 34. x2/5 + y 2/5 = 1
35. cos(x) + sin(y) = 1 x
36. = 10
y
dy
Exercise Group. Use logarithmic differentiation to find dx , then find the equation of the tangent line at the indicated
x-value.
2
37. y = (1 + x)1/x at x = 1 38. y = (2x)x at x = 1
Show your work using logarithmic Show your work using logarithmic
differentiation. differentiation.
xx 40. y = xsin(x)+2 at x = π/2
39. y = at x = 1
x+1 Show your work using logarithmic
Show your work using logarithmic differentiation.
differentiation.
x+1 (x + 1)(x + 2)
41. y = at x = 1 42. y= at x = 0
x+2 (x + 3)(x + 4)
Show your work using logarithmic Show your work using logarithmic
differentiation. differentiation.
CHAPTER 2. DERIVATIVES 126
(0.375, −0.5)
d d
(f (y)) = (x) −1
dx dx
′ ′
f (y) · y = 1
Figure 2.7.4 Corresponding tangent lines
1 drawn to f and f −1
y′ = ′
f (y)
CHAPTER 2. DERIVATIVES 127
1
y′ = .
f ′ (g(x))
This leads us to the following theorem.
Let y = arcsin(x) = sin−1 (x). Find y ′ using Theorem 2.7.6. Video Solution
Solution. Adopting our previously defined notation, let g(x) = arcsin(x)
and f (x) = sin(x). Thus f ′ (x) = cos(x). Applying the theorem, we
have
1
g ′ (x) = ′
f (g(x))
1
= .
cos(arcsin(x))
This last expression is not immediately illuminating. Drawing a figure
will help, as shown in Figure 2.7.10. Recall that the sine function can [Link]/watch?v=xBZqkvQRSG4
be viewed as taking in an angle and returning a ratio of sides of a right
triangle, specifically, the ratio “opposite over hypotenuse.” This means
that the arcsine function takes as input a ratio of sides and returns an
angle. The equation y = arcsin(x) can be rewritten as y = arcsin(x/1);
1
that is, consider a right triangle where the hypotenuse has length 1 and x
the side opposite of the angle
√ with measure y has length x. This means
the final side has length 1 − x2 , using the Pythagorean Theorem.
Therefore
y
cos sin−1 (x) = cos(y) √
√ 1 − x2
1 − x2
=
Figure 2.7.10 A right triangle defined
p 1
= 1 − x2 , by y = sin−1 (x/1) with the length of
the third leg found using the Pythagorean
resulting in Theorem
d 1
(arcsin(x)) = √ .
dx 1 − x2
CHAPTER 2. DERIVATIVES 128
y π
2
y
√
π 3
1 3, 2 √
π 3 π
4 2 , 3
)
)
(x
(x
1
−
sin x x
sin
= −2 −1
=
− π2 − π4 y π π 1 2
y
4 2
− π4
−1
− π2
Figure 2.7.11 Graphs of sin(x) and sin−1 (x) along with corresponding tangent
lines
In Figure 2.7.11 we see f (x) = sin(x) and f −1 (x) = sin−1 (x) graphed
√ on
their respective domains. The line tangent to sin(x) at the point π/3, 3/2
has slope cos(π)/3
√ = 1/2. The slope of the corresponding point on sin−1 (x),
the point 3/2, π/3 , is
1 1
q √ 2 = p1 − 3/4
1− 3/2
1
=p
1/4
1
= = 2,
1/2
verifying yet again that at corresponding points, a function and its inverse have
reciprocal slopes.
Using similar techniques, we can find the derivatives of all the inverse trigono-
metric functions. In Table 2.7.12 we show the restrictions of the domains of the
standard trigonometric functions that allow them to be invertible.
CHAPTER 2. DERIVATIVES 129
Table 2.7.12 Domains and ranges of the trigonometric and inverse trigonomet-
ric functions
Function Domain Range
sin(x) [−π/2, π/2] [−1, 1]
sin−1 (x) [−1, 1] [−π/2, π/2]
cos(x) [0, π] [−1, 1]
cos−1 (x) [−1, 1] [0, π]
tan(x) (−π/2, π/2) (−∞, ∞)
tan−1 (x) (−∞, ∞) (−π/2, π/2)
csc(x) [−π/2, 0) ∪ (0, π/2] (−∞, −1] ∪ [1, ∞)
csc−1 (x) (−∞, −1] ∪ [1, ∞) [−π/2, 0) ∪ (0, π/2]
sec(x) [0, π/2) ∪ (π/2, π] (−∞, −1] ∪ [1, ∞)
sec−1 (x) (−∞, −1] ∪ [1, ∞) [0, π/2) ∪ (π/2, π]
cot(x) (0, π) (−∞, ∞)
cot−1 (x) (−∞, ∞) (0, π)
Note how each derivative is the negative of the derivative of its “co” function.
Because of this, derivatives of sin−1 (x), tan−1 (x), and sec−1 (x) are used almost
exclusively throughout this text.
d
In Section 2.3, we stated without proof or explanation that dx (ln(x)) = x1 .
We can justify that now using Theorem 2.7.6, as shown in the example.
In this chapter we have defined the derivative, given rules to facilitate its
computation, and given the derivatives of a number of standard functions. We
restate the most important of these in the following theorem, intended to be a
reference for further work.
d d
1. (c) = 0 14. (cos(x)) = − sin(x)
dx dx
d d
2. (x) = 1 15. (tan(x)) = sec2 (x)
dx dx
d n d
3. (x ) = nxn−1 16. (csc(x)) = − csc(x) cot(x)
dx dx
d
4. (f (x) ± g(x)) = d
dx 17. (sec(x)) = sec(x) tan(x)
f (x) ± g ′ (x)
′ dx
d d
5. (c · f (x)) = c · f ′ (x) 18. (cot(x)) = − csc2 (x)
dx dx
d d 1
6. (f (x) · g(x)) = f ′ (x) · 19. sin−1 (x) = √
dx dx 1 − x2
g(x) + f (x) · g ′ (x)
d
d ′ 20. cos−1 (x) =
7. (f (g(x))) = f (g(x)) · dx
dx 1
g ′ (x) −√
1 − x2
′ ′
d f (x) f (x) · g(x) − f (x) · g (x)
8. = d 1
dx g(x) (g(x))221. tan−1 (x) =
dx 1 + x2
d x
9. (e ) = ex d
dx 22. csc−1 (x) =
dx
d 1 1
10. (ln(x)) = − √
dx x |x| x2 − 1
d x d
11. (a ) = ln(a) · ax 23. sec−1 (x) =
dx dx
1
12.
d
(loga x) =
1
·
1 √
dx ln(a) x |x| x2 − 1
d d 1
13. (sin(x)) = cos(x) 24. cot−1 (x) = −
dx dx 1 + x2
CHAPTER 2. DERIVATIVES 131
2.7.1 Exercises
Terms and Concepts
Problems
Exercise Group. An invertible function f (x) is given along with a point that lies on its graph. Using Theorem 2.7.6,
′
evaluate f −1 (x) at the indicated value.
′
9. The point (9, 65) is on the graph of f (x) = 7x + 2. Find f −1 (65).
′
10. The point (−6, 51) is on the graph of f (x) = x2 − 2x + 3, x ≥ 1. Find f −1 (51).
√ ′ √3
11. The point 24 π
, 23 is on the graph of f (x) = cos(4x), 0 ≤ x ≤ π4 . Find f −1 2 .
′
12. The point (3, 576) is on the graph of f (x) = x3 − 27x2 + 267x − 9. Find f −1 (576).
−1 ′ 1
13. The point 2, 15 is on the graph of f (x) = 1+x 2 , x ≥ 0. Find f
1
5 .
′
14. The point (0, 3) is on the graph of f (x) = 3e4x . Find f −1 (3).
Exercise Group. Compute the derivative of the given function in two ways:
(a) By simplifying first, then taking the derivative, and
(b) by using the Chain Rule first then simplifying.
Verify that the two answers are the same.
25. f (x) = sin(sin−1 (x)) 26. f (x) = tan−1 (tan(x))
27. f (x) = sin(cos−1 (x)) 28. f (x) = sin(2 sin−1 (x))
Exercise Group. Find the equation of the line tangent to the graph of f at the indicated x value.
√ √
29. f (x) = sin−1 (x) at x = − 3
2 30. f (x) = cos−1 (2x) at x = 4
3
Chapter 3
Our study of limits led to continuous functions, a certain class of functions that
behave in a particularly nice way. Limits then gave us an even nicer class of
functions, functions that are differentiable.
This chapter explores many of the ways we can take advantage of the infor-
mation that continuous and differentiable functions provide.
132
CHAPTER 3. THE GRAPHICAL BEHAVIOR OF FUNCTIONS 133
y y y
4 4 4
2 2 2
x x x
−2 −1 1 2 −2 −1 1 2 −2 −1 1 2
This theorem states that f has extreme values, but it does not offer any ad- Figure 3.1.5 Video presentation of The-
vice about how/where to find these values. The process can seem to be fairly orem 3.1.4
easy, as the next example illustrates. After the example, we will draw on lessons
learned to form a more general and powerful method for finding extreme val-
y (5, 25)
ues.
20
Example 3.1.6 Approximating extreme values.
Consider f (x) = 2x3 − 9x2 on I = [−1, 5], as graphed in Figure 3.1.7. (0, 0) x
Approximate the extreme values of f . −2 2 4
Solution. The graph is drawn in such a way to draw attention to certain (−1, −11)
points. It certainly seems that the smallest y-value is −27, found when −20
x = 3. It also seems that the largest y-value is 25, found at the endpoint
of I, x = 5. We use the word seems, for by the graph alone we cannot (3, −27)
be sure the smallest value is not less than −27. Since the problem asks
Figure 3.1.7 A graph of f (x) = 2x3 −
for an approximation, we approximate the extreme values to be 25 and
9x2 as in Example 3.1.6
−27.
Notice how the minimum value came at “the bottom of a hill,” and the maxi- Alternative Vocabulary. The terms
mum value came at an endpoint. Also note that while 0 is not an extreme value, local minimum and local maximum
it would be if we narrowed our interval to [−1, 4]. The idea that the point (0, 0) are often used as synonyms for
is the location of an extreme value for some interval is important, leading us to relative minimum and relative max-
a definition of a relative maximum. In short, a “relative max” is a y-value that’s imum.
the largest y-value “nearby.” As it makes intuitive sense that
an absolute maximum is also a
Definition 3.1.8 Relative Minimum and Relative Maximum.
relative maximum, Definition 3.1.8
Let f be defined on an interval I containing c. allows a relative maximum to oc-
cur at an interval’s endpoint.
1. If there is a δ > 0 such that f (c) ≤ f (x) for all x in I where
|x − c| < δ, then f (c) is a relative minimum of f . We also say
that f has a relative minimum at (c, f (c)).
2. If there is a δ > 0 such that f (c) ≥ f (x) for all x in I where
|x − c| < δ, then f (c) is a relative maximum of f . We also say
CHAPTER 3. THE GRAPHICAL BEHAVIOR OF FUNCTIONS 134
What can we learn from the previous two examples? We were able to vi- Figure 3.1.13 A graph of f (x) = (x −
sually approximate relative extrema, and at each such point, the derivative was 1)2/3 + 2 as in Example 3.1.12
either 0 or it was not defined. This observation holds for all functions, leading
to a definition and a theorem.
In this text we use “critical num-
Definition 3.1.14 Critical Numbers and Critical Points. ber” and “critical value” interchange-
Let f be defined at c. The value c is a critical number (or critical value) ably. Other textbooks reserve the
of f if f ′ (c) = 0 or f ′ (c) is not defined. term critical value for the func-
If c is a critical number of f , then the point (c, f (c)) is a critical point tion value f (c), when c is a criti-
of f . cal number.
Note that all this was done without the aid of a graph; this work followed an
analytic algorithm and did not depend on any visualization. Figure 3.1.20 shows
f and we can confirm our answer, but it is important to understand that these
answers can be found without graphical assistance.
We practice again. [Link]/watch?v=LyhHlreZvhc
Example 3.1.22 Finding extreme values.
Solution. Here f is piecewise-defined, but we can still apply Key Idea 3.1.18
as it is continuous on [−4, 2] (one should check to verify that lim f (x) =
x→0
f (0)). Video Solution
Evaluating f at the endpoints gives:
f (−4) = 25 f (2) = 3.
20
15
x f (x) 10
−4 25
0 1 5
2 3 x
−4 −3 −2 −1 1 2
Figure 3.1.23 Finding the extreme
values of a piecewise-defined Figure 3.1.24 A graph of f (x) on
function in Example 3.1.22 [−4, 2] as in Example 3.1.22
√
this example we have five values to consider: x = 0, ±2, ± π. From
the table it is clear that the maximum value of f on [−2, 2] is 1; the
minimum value is −1. The graph in Figure 3.1.27 confirms our results.
1 y
0.5
x f (x)
−2 −0.65 x
√ −2 −1 1 2
− π −1
0 1 −0.5
√
π −1
2 −0.65 −1
0.5
x
−1 −0.5 0.5 1
−0.5 x f (x)
−1 0
−1 Circle Revisited. We implicitly found
0 1
1 0 the derivative of x2 + y 2 = 1,
Figure
√ 3.1.29 A graph of f (x) = the unit circle, in Section 2.6 Ex-
1 − x2 on [−1, 1] as in Exam- Figure 3.1.30 Finding the extrema ample 2.6.11 as dx dy
= −x/y. In
ple 3.1.28 of the half-circle in Example 3.1.28 Example 3.1.28, half of the unit
√ circle is given as y = f (x) =
Using the The Chain Rule, we find f ′ (x) = −x 1 − x2 . The critical √
points of f are found when f ′ (x) = 0 or when f ′ is undefined. It is 1 − x2 . √
straightforward to find that f ′ (x) = 0 when x = 0, and f ′ is undefined We found f ′ (x) = −x 1 − x2 .
when x = ±1, the endpoints of the interval (which are in the domain of Recognize that the denominator
f .) The table of important values is given in Figure 3.1.30. The maximum of this fraction is y; that is, we
value is 1, and the minimum value is 0. again found f ′ (x) = dx
dy
= −x/y.
3.1.1 Exercises
Terms and Concepts
Problems
Exercise Group. Identify each of the marked points as being an absolute maximum or minimum, a relative maximum
or minimum, or none of the above.
7. 8.
y B y C
2
G
D
2
1 B D
E
C x
x 1 2 3 4 5
1 2 3 4 5 6
−1
E
F
A
−2 A −2
1
2
x
x
−4 −2 2 4
−3 −2 −1 (0, 0)1 2 3
CHAPTER 3. THE GRAPHICAL BEHAVIOR OF FUNCTIONS 139
√
11. f (x) = sin(x) 12. f (x) = x2 4 − x
y y
16 512
5 , 25 5
√
(π/2, 1) 10
1
8
6
x
1 2 3 4 5 6 4
2
−1 (4, 0) x
(3π/2, −1)
−2 −1 (0, 0)1 2 3 4
√
x4 − 2x2 + 1
2 3
13. f (x) = 1 + (x−2) 3 14. f (x) =
x
y y
6 3
4 2
√
3
1
2
2 6, 1 + 3
(−1, 0) (1, 0) x
(2, 1)
x −2 −1 1 2
2 4 6 8 10
( (
x2 , x≤0 x2 , x≤0
15. f (x) = 16. f (x) =
5
x , x>0 x, x>0
y y
1 1
0.5 0.5
x x
−1 −0.5 (0, 0) 0.5 1 −1 −0.5 (0, 0) 0.5 1
−0.5 −0.5
Exercise Group. Find the extreme values of the function on the given interval.
17. f (x) = x2 + 2x − 1 on [−5, 1] 18. f (x) = x3 + 23 x2 − 18x − 6 on [0, 3]
√
19. f (x) = 4 cos(x) on 3π 7π
4 , 6 20. f (x) = x6 4 − x2 on [−2, 2]
2 x2
21. f (x) = x + x on [1, 4] 22. f (x) = on [−2, 2]
x2 +7
x x
23. f (x) = e cos(x) on [0, π] 24. f (x) = e sin(x) on [0, π]
ln(x)
f (x) = x( 4 ) − x3 on [0, 2]
3
25. f (x) = x2 on [1, 7] 26.
CHAPTER 3. THE GRAPHICAL BEHAVIOR OF FUNCTIONS 140
The slope at any point on the graph itself is given by the derivative f ′ (t). So,
since the answer to the question above is “yes,” this means that at some time
during the trip, the derivative takes on the value of 50 mph. Symbolically,
f (2) − f (0)
f ′ (c) = = 50
2−0
for some time 0 ≤ c ≤ 2.
How about more generally? Given any function y = f (x) and a range a ≤
x ≤ b does the value of the derivative at some point between a and b have to
match the slope of the secant line connecting the points (a, f (a)) and (b, f (b))?
Or equivalently, does the equation f ′ (c) = f (b)−f
b−a
(a)
have to hold for some
a < c < b?
Let’s look at two functions in an example.
Consider functions
1
f1 (x) = f2 (x) = |x|
x2
with a = −1 and b = 1 as shown in Figure 3.2.3. Both functions have a
value of 1 at a and b. Therefore the slope of the secant line connecting
the end points is 0 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 [−1, 1]
such that
f (1) − f (−1)
f ′ (c) = = 0.
1 − (−1)
CHAPTER 3. THE GRAPHICAL BEHAVIOR OF FUNCTIONS 141
y y
3 3
2 2
1 1
x x
−1 1 −1 1
So what went “wrong”? It may not be surprising to find that the discontinuity
of f1 and the corner of f2 play a role. If our functions had been continuous and
differentiable, would we have been able to find that special value c? This is our
motivation for the following theorem.
Note that the reasons that the functions in Example 3.2.2 fail are indeed that 6
f1 has a discontinuity on the interval [−1, 1] and f2 is not differentiable at the
4
origin.
We will give a proof of the Mean Value Theorem below. To do so, we use a 2
fact, called Rolle’s Theorem, stated here. x
−1 a c 1 b 2
Theorem 3.2.5 Rolle’s Theorem.
−2
Let f be continuous on [a, b] and differentiable on (a, b), where f (a) =
−4
f (b). There is some c in (a, b) such that f ′ (c) = 0.
Consider Figure 3.2.7 where the graph of a function f is given, where f (a) = Figure 3.2.7 A graph of f (x) = x3 −
f (b). It should make intuitive sense that if f is differentiable (and hence, con- 5x2 + 3x + 5, where f (a) = f (b).
tinuous) that there would be a value c in (a, b) where f ′ (c) = 0; that is, there Note the existence of c, where a <
would be a relative maximum or minimum of f in (a, b). Rolle’s Theorem guar- c < b, where f ′ (c) = 0.
antees at least one; there may be more.
Rolle’s Theorem is presented here as a stepping stone toward the Mean
Value Theorem, but it’s a useful result in its own right. It often turns up as a
tool in mathematical problem solving. The video in Figure 3.2.8 illustrates one
such use of Rolle’s Theorem.
Rolle’s 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 0, and the theorem
guarantees some c where f ′ (c) = 0. We will prove Rolle’s Theorem, then use it
to prove the Mean Value Theorem.
[Link]/watch?v=le-5zsb6O7o
Figure 3.2.8 Using Rolle’s Theorem to
show a polynomial has at most one
real root
CHAPTER 3. THE GRAPHICAL BEHAVIOR OF FUNCTIONS 142
[Link]/watch?v=ON7WYLJY2tE
CHAPTER 3. THE GRAPHICAL BEHAVIOR OF FUNCTIONS 143
f ′ (x) = 14 y
2 40
3x + 5 = 14
x2 = 3 20
√
x = ± 3 ≈ ±1.732
x
−3 −2 −1 1 2 3
We have found two values c in [−3, 3] where the instantaneous rate
of change is equal to the average rate of change; the Theorem 3.2.4 guar- −20
anteed at least one. In Figure 3.2.10, f is graphed with a line represent-
√
ing the average rate of change; the lines tangent to f at x = ± 3 are −40
also given. Note how these lines are parallel (i.e., have the same slope)
to the secant line. Figure 3.2.10 Demonstrating the Mean
Value Theorem in Example 3.2.9
While the Theorem 3.2.4 has practical use (for instance, the speed monitor-
ing application mentioned before), it is mostly used to advance other theory.
We will use it in the next section to relate the shape of a graph to its derivative.
Before ending this section, we give two important consequences of the Mean
Value Theorem. Each of these consequences has important applications to math-
ematical theory, and can be easily understood in the context of the position and
velocity of objects in motion.
First, we recall that the derivative of any constant function is zero. Is the
converse true? That is, are constant functions the only ones whose derivative is
zero? The Mean Value Theorem says yes. This officially establishes our intuition
about objects in (or, actually, not in) motion: if the velocity of an object is 0, then
the object’s position is unchanged; it is constant. Second, if two functions f and
g have the same derivative, what does this tell us about f and g? The Mean
Value Theorem implies that these functions must only differ by a constant; that
is, f (x) = g(x) + C, for some constant C.
This has an application to motion that is not intuitive to some. Suppose two
objects start moving while 5 ft apart, and always move with the same velocity.
Then the two objects will always be 5 ft apart. (If two pennies are dropped from
the 30th and 31st stories of a tall building at the same time, they will always be
1 story apart as they fall.)
Proof.
1. Choose any two points a and b in the interval I. By the Mean Value Theo-
rem, we must have
f (b) − f (a)
f ′ (c) =
b−a
for some c between a and b. But f ′ (c) = 0, so f (b) − f (a) = 0, or
f (a) = f (b). Since a and b were any two points, this tells us that f must
have the same value at every point; that is, f must be constant.
2. Suppose g ′ (x) = h′ (x) for each point x in I, and consider the function
f (x) = g(x) − h(x). By the difference rule for derivatives, we have
[Link]/watch?v=BzvxKeZMNtE
Figure 3.2.13 Showing that two func-
tions with the equal derivatives differ
by a constant
[Link]/watch?v=CGBTZtM9mFY
Figure 3.2.14 Demonstrating a prop-
erty of the sine function
CHAPTER 3. THE GRAPHICAL BEHAVIOR OF FUNCTIONS 145
3.2.1 Exercises
Terms and Concepts
1. Explain in your own words what the Mean Value Theorem states.
2. Explain in your own words what Rolle’s Theorem states.
Problems
Exercise Group. A function f (x) and interval [a, b] are given. Check if Rolle’s Theorem can be applied to f on [a, b];
if so, find c in (a, b) such that f ′ (c) = 0.
3. f (x) = 6 on [−1, 1] 4. f (x) = 6x on [−1, 1]
Explain why Rolle’s Theorem can or cannot be Explain why Rolle’s Theorem can or cannot be
applied. applied.
5. f (x) = x2 + x − 6 on [−3, 2] 6. f (x) = x2 + x − 2 on [−3, 2]
Explain why Rolle’s Theorem can or cannot be Explain why Rolle’s Theorem can or cannot be
applied. applied.
7. f (x) = x2 + x on [−2, 2] 8. f (x) = sin(x) on [π/6, 5π/6]
Explain why Rolle’s Theorem can or cannot be Explain why Rolle’s Theorem can or cannot be
applied. applied.
9. f (x) = cos(x) on [0, π] 1
10. f (x) = x2 −2x+1 on [0, 2]
Explain why Rolle’s Theorem can or cannot be Explain why Rolle’s Theorem can or cannot be
applied. applied.
Exercise Group. A function f (x) and interval [a, b] are given. Check if The Mean Value Theorem of Differentiation
can be applied to f on [a, b]; if so, find c in (a, b) guaranteed by the Mean Value Theorem.
11. f (x) = x2 + 3x − 1 on [−2, 2] 12. f (x) = 5x2 − 6x + 8 on [0, 5]
Explain why the Mean Value Theorem can or Explain why the Mean Value Theorem can or
cannot be applied. cannot be applied.
√ √
13. f (x) = 9 − x2 on [0, 3] 14. f (x) = 25 − x on [0, 9]
Explain why the Mean Value Theorem can or Explain why the Mean Value Theorem can or
cannot be applied. cannot be applied.
f (x) = xx2 −9
2
15. 16. f (x) = ln(x) on [1, 5]
−1 on [0, 2]
Explain why the Mean Value Theorem can or Explain why the Mean Value Theorem can or
cannot be applied. cannot be applied.
17. f (x) = tan(x) on [−π/4, π/4] 18. f (x) = x3 − 2x2 + x + 1 on [−2, 2]
Explain why the Mean Value Theorem can or Explain why the Mean Value Theorem can or
cannot be applied. cannot be applied.
19. f (x) = 2x3 − 5x2 + 6x + 1 on [−5, 2] 20. f (x) = sin−1 (x) on [−1, 1]
Explain why the Mean Value Theorem can or Explain why the Mean Value Theorem can or
cannot be applied. cannot be applied.
CHAPTER 3. THE GRAPHICAL BEHAVIOR OF FUNCTIONS 146
5 y
Our study of “nice” functions f in this chapter has so far focused on individual 3
points: points where f is maximal/minimal, points where f ′ (x) = 0 or f ′ does
not exist, and points c where f ′ (c) is the average rate of change of f on some 2
interval.
1
In this section we begin to study how functions behave between special
points; we begin studying in more detail the shape of their graphs. x
We start with an intuitive concept. Given the graph in Figure 3.3.1, where 1 2 3
would you say the function is increasing? Decreasing? Even though we have
Figure 3.3.1 A graph of a function f
not defined these terms mathematically, one likely answered that f is increasing
used to illustrate the concepts of in-
when x > 1 and decreasing when x < 1. We formally define these terms here.
creasing and decreasing
Definition 3.3.2 Increasing and Decreasing Functions.
Let f be a function defined on an interval I.
1. f is increasing on I if for every a < b in I, f (a) < f (b).
1. If f ′ (c) > 0 for all c in (a, b), then f is increasing on [a, b].
2. If f ′ (c) < 0 for all c in (a, b), then f is decreasing on [a, b].
3. If f ′ (c) = 0 for all c in (a, b), then f is constant on [a, b].
The conclusions of Item 1 and Item 2 also hold if f ′ (c) = 0 for a finite
number of nonadjacent values of c in I.
Figure 3.3.10 summarizes our work. Figure 3.3.10 Completed number line
We can verify our calculations by considering Figure 3.3.11, where for f in Example 3.3.8
f is graphed. The graph also presents f ′ ; note how f ′ > 0 when f is
increasing and f ′ < 0 when f is decreasing.
y
One is justified in wondering why so much work is done when the graph 10
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 significant hap- 5 f (x)
f ′ (x)
pens near x = −1 and x = 0.3, but we cannot determine exactly where from
the graph.
One could argue that just finding critical values is important; once we know x
the significant points are x = −1 and x = 1/3, the graph shows the increasing/ −2 −1 1/3 1 2
decreasing traits just fine. That is true. However, the technique prescribed here
relative extrema of f .
2
Remark 3.3.14 Importance of Continuity. The continuity of f when using the
first derivative test is very important. Without continuity, almost anything can x
happen at a critical number. For example, we can construct a piecewise function −1 1 2 3
where the sign of f ′ switches to positive to negative at c and f (c) is not a local
maximum. This is shown in Figure 3.3.15.
−2
x2 − 2x − 3
f ′ (x) = .
(x − 1)2
We need to find the critical values of f ; we want to know when Video Solution
f ′ (x) = 0 and when f ′ is not defined. That latter is straightforward:
when the denominator of f ′ (x) is 0, f ′ is undefined. That occurs when
x = 1, which we’ve already recognized as an important value, but not a
critical number.
f ′ (x) = 0 when the numerator of f ′ (x) is 0. That occurs when
x − 2x − 3 = (x − 3)(x + 1) = 0; i.e., when x = −1, 3.
2
One is often tempted to think that functions always alternate “increasing, de-
creasing, increasing, decreasing,…” around critical values. Our previous example
demonstrated that this is not always the case. While x = 1 was not technically
a critical value, it was an important value we needed to consider. We found that
f was decreasing on “both sides of x = 1.”
We examine one more example.
Find the intervals on which f (x) = x8/3 − 4x2/3 is increasing and de- Video Solution
creasing and identify the relative extrema.
Solution. The domain of f is R (you can take the odd root of both pos-
itive and negative nubmers). Next, we take the first derivative. Since we
know we want to solve f ′ (x) = 0, we will do some algebra after taking
the derivative.
8 2
f (x) = x 3 − 4x 3
8 5 8 1
f ′ (x) = x 3 − x− 3
3 3 [Link]/watch?v=T4RxcQnNotc
8 −1 6
= x 3 x3 − 1
3
8 1
= x− 3 x2 − 1
3
8 1
= x− 3 (x − 1)(x + 1).
3
This derivation of f ′ shows that f ′ (x) = 0 when x = ±1 and f ′ is
not defined when x = 0. Thus we have three critical values, breaking
the number line into four subintervals as shown in Figure 3.3.20.
Interval 4: (1, ∞) Similar work to that done for the other three rel rel rel
intervals shows that f ′ (x) > 0 on (1, ∞), so min max min
f is increasing on this interval. f′ < 0 f′ > 0 f′ < 0 f′ > 0
f decr f incr f decr f incr
We conclude by stating that f is increasing on the intervals (−1, 0)
−1 0 1
and (1, ∞) and decreasing on the intervals (−∞, −1) and (0, 1). The
sign of f ′ changes from negative to positive around x = −1 and x = 1, Figure 3.3.20 Number line for f in Ex-
meaning by Theorem 3.3.12 that f (−1) and f (1) are relative minima of ample 3.3.19
f . As the sign of f ′ changes from positive to negative at x = 0, we have a
relative maximum at f (0). Figure 3.3.21 shows a graph of f , confirming
our result. We also graph f ′ , highlighting once more that f is increasing
when f ′ > 0 and is decreasing when f ′ < 0.
y
We have seen how the first derivative of a function helps determine when f (x) 10
the graph of a function is going “up” or “down.” In the next section, we will see
how the second derivative helps determine how the graph of a function curves.
5
f ′ (x)
x
−3 −2 −1 1 2 3
−5
3.3.1 Exercises
Terms and Concepts
Problems
Exercise Group. A function f (x) is given. Graph f and f ′ on the same axes (using technology is permitted) and verify
Theorem 3.3.5.
7. f (x) = 2x + 3 8. f (x) = x2 − 3x + 5
9. f (x) = cos(x) 10. f (x) = tan(x)
11. f (x) = x − 5x + 7x − 1
3 2
12. f (x) = 2x3 − x2 + x − 1
13. f (x) = x4 − 5x2 + 4 14. f (x) = 1
x2 +1
(f) Use the First Derivative Test to determine which critical points are a relative minimum.
15. f (x) = x2 + 4x 16. f (x) = x3 + 2x2 + 9
17. f (x) = 7x3 − 17x2 − 35x + 1 18. f (x) = x3 − 9x2 + 27x − 27
19. f (x) = 1 x2 −1
x2 −10x+34 20. f (x) = x2 −36
x 2
21. f (x) = x2 +12x+35 (x−(−5)) 3
22. f (x) = x
23. f (x) = sin(x) cos(x) on (−π, π) 6
24. f (x) = x + 192x
CHAPTER 3. THE GRAPHICAL BEHAVIOR OF FUNCTIONS 154
y y
6 6
4 4
2 2
x x
−1 1 2 −1 1 2
(a) A graph that is concave up. No- (b) A graph that is concave down. No-
tice how the secant line lies above the tice how the secant line lies below the Loose Language. We often state
graph. graph. that “f is concave up” instead of
“the graph of f is concave up” for
Figure 3.4.3 Illustrating the nature of concave up and concave down simplicity.
CHAPTER 3. THE GRAPHICAL BEHAVIOR OF FUNCTIONS 155
Theorem 3.4.4
y
Let f be a continuous function on [a, b] and differentiable on (a, b). 30
1. If f ′′ (c) > 0 for all c in (a, b), then f is concave up on [a, b].
20
2. If f ′′ (c) < 0 for all c in (a, b), then f is concave down on [a, b].
3. If f ′′ (c) = 0 for all c in (a, b), then f is linear on [a, b].
10
′
The graph of a function f is concave up when f is increasing. That means as
one looks at a concave up graph from left to right, the slopes of the tangent lines x
will be increasing. Consider Figure 3.4.5, where a concave up graph is shown −3 −2 −1 1 2 3
along with some tangent lines. Notice how the tangent line on the left is steep,
downward, corresponding to a lesser (large negative) value of f ′ . On the right, Figure 3.4.5 A function f with a con-
the tangent line is steep, upward, corresponding to a greater (large positive) cave up graph. Notice how the slopes
value of f ′ . of the tangent lines, when looking from
If a function is decreasing and concave up, then its rate of decrease is slowing; left to right, are increasing. (The slope
it is “leveling off.” You can see this in the left side of Figure 3.4.5. If the function is values pictured are −12, −6, 6 and 12).
increasing and concave up, then the rate of increase is increasing. The function
is increasing at a faster and faster rate. You can see this in the right side of
Figure 3.4.5.
Now consider a function which is concave down. We essentially repeat the
above paragraphs with slight variation.
The graph of a function f is concave down when f ′ is decreasing. That means
as one looks at a concave down graph from left to right, the slopes of the tangent
CHAPTER 3. THE GRAPHICAL BEHAVIOR OF FUNCTIONS 156 y
30
lines will be decreasing. Consider Figure 3.4.6, where a concave down graph is
shown along with some tangent lines. Notice how the tangent line on the left
20
is steep, upward, corresponding to a greater (large positive) value of f ′ . On
the right, the tangent line is steep, downward, corresponding to a lesser (large
negative) value of f ′ . 10
If a function is increasing and concave down, then its rate of increase is slow-
ing; it is “leveling off.” If the function is decreasing and concave down, then the
rate of decrease is decreasing. The function is decreasing at a faster and faster x
rate. −3 −2 −1 1 2 3
Our definition of concave up and concave down is given in terms of when Figure 3.4.6 A function f with a con-
the first derivative is increasing or decreasing. We can apply the results of the cave down graph. Notice how the slopes
previous section to find intervals on which a graph is concave up or down. That of the tangent lines, when looking from
is, we recognize that f ′ is increasing when f ′′ > 0, etc. left to right, are decreasing.
Theorem 3.4.7 Test for Concavity.
Let f be twice differentiable on an interval I. The graph of f is concave
up if f ′′ > 0 on I, and is concave down if f ′′ < 0 on I.
that the sign of f ′′ is changing from positive to negative (or, negative to positive)
at x = c. A sign change may occur when f ′′ = 0 or f ′′ is undefined. This leads
to the following theorem.
We have identified the concepts of concavity and points of inflection. It is Video Solution
now time to practice using these concepts; given a function, we should be able
to find its points of inflection and identify intervals on which it is concave up or
down. We do so in the following examples.
Let f (x) = x/(x2 − 1). Find the inflection points of f and the intervals
−2
on which it is concave up/down.
Solution. We need to find f ′ and f ′′ . Using the Quotient Rule 2.4.10
and simplifying, we find
Figure 3.4.14 A graph of f (x) used in
−(1 + x2 ) 2x(x2 + 3) Example 3.4.12
f ′ (x) = f ′′ (x) = .
(x2 − 1)2 (x2 − 1)3
Video Solution
To find the possible points of inflection, we seek to find where f ′′ (x) =
0 and where f ′′ is not defined. Solving f ′′ (x) = 0 reduces to solving
2x(x2 + 3) = 0; we find x = 0. We find that f ′′ is not defined when
x = ±1, for then the denominator of f ′′ is 0. We also note that f itself is
not defined at x = ±1, having a domain of (−∞, −1)∪(−1, 1)∪(1, ∞).
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 inflection at x = ±1 as they are not part of the
domain, but we must still consider these x-values to be important and
will include them in our number line. [Link]/watch?v=eC6QLbsuRVs
CHAPTER 3. THE GRAPHICAL BEHAVIOR OF FUNCTIONS 158
Interval 1: Select a number c in this interval with a large Figure 3.4.16 Number line for f in Ex-
(−∞, −1) magnitude (for instance, c = −100). The de- ample 3.4.15
nominator of f ′′ (x) will
be positive. In the nu-
merator, the c2 + 3 factor will be positive
and the 2c factor will be negative. Thus the nu-
merator is negative and f ′′ (c) is negative. We
conclude f is concave down on (−∞, −1). 10 y
f (x) f ′′ (x)
Interval 2: (−1, 0) For any number c in this interval, the factor
2c in the numerator will be negative, the fac- 5
tor c2 + 3 in the numerator will be positive,
3 x
and the factor c2 − 1 in the denominator −3 −2 −1 1 2 3
will be negative. Thus f ′′ (c) > 0 and f is con-
cave up on this interval. −5
Interval 3: (0, 1) Any number c in this interval will be posi-
tive and “small.” Thus the numerator is pos- −10
itive while the denominator is negative. Thus
f ′′ (c) < 0 and f is concave down on this in- Figure 3.4.17 A graph of f (x) and f ′′ (x)
terval. in Example 3.4.15
Interval 4: (1, ∞) Choose a large value for c. It is evident that
f ′′ (c) > 0, so we conclude that f is concave
up on (1, ∞).
We conclude that f is concave up on (−1, 0) and (1, ∞) and con-
cave down on (−∞, −1) and (0, 1). There is only one point of inflection,
(0, 0), as f is not defined at x = ±1. Our work is confirmed by the graph
20 y
of f in Figure 3.4.17. Notice how f is concave up whenever f ′′ is positive,
and concave down when f ′′ is negative. The inflection in f occurs where
f ′′ changes sign. 15
S(t)
Recall that relative maxima and minima of f are found at critical points of 10
f ; that is, they are found when f ′ (x) = 0 or when f ′ is undefined. Likewise,
the relative maxima and minima of f ′ are found when f ′′ (x) = 0 or when f ′′ is
5
undefined; note that these are the inflection points of f .
What does a “relative maximum of f ′ ” mean? The derivative measures the t
rate of change of f ; maximizing f ′ means finding where f is increasing the most 0.5 1 1.5 2 2.5 3
— where f has the steepest tangent line. A similar statement can be made for
minimizing f ′ ; it corresponds to where f has the steepest negatively-sloped tan- Figure 3.4.19 A graph of S(t) in Ex-
gent line. ample 3.4.18, modeling the sale of a
We utilize this concept in the next example. product over time
The sales of a certain product over a three-year span are modeled by Video Solution
S(t) = t4 −8t2 +20, where t is the time in years, shown in Figure 3.4.19.
Over the first two years, sales are decreasing. Find the point at which
sales are decreasing at their greatest rate.
Solution. We want to maximize the rate of decrease, which is to say,
we want to find where S ′ has a minimum. To do this, we find where S ′′
[Link]/watch?v=yjcSXaGkL5Y
20 y
CHAPTER 3. THE GRAPHICAL BEHAVIOR OF FUNCTIONS 159
S(t)
10
is 0 and S ′′ changes from negative to positive. We find S ′ (t) = 4t3 −16t
′′ ′′
p S (t) = 12t − 16. Setting S (t) = 0 and solving, we get t =
2
and t
4/3 ≈ 1.16 (we ignore the negative solution for t since it does not lie 0.5 1 1.5 2 2.5 3
in the domain of our function S). p
Since S ′′ (1) = −4 < 0 and S ′′ (2) = 32 > 0, we can say S ′ ( 4/3) −10 S ′ (t)
is a local minimum of S ′ . This is both the inflection point and the point
of maximum decrease. This is the point at which things first start looking
up for the company. After the inflection point, sales are still decreasing, Figure 3.4.20 A graph of S(t) in Exam-
but not decreasing quite as quickly as they had been. ple 3.4.18, along with S ′ (t)
A graph of S(t) and S ′ (t) is given in Figure 3.4.20. When S ′ (t) < 0,
sales are decreasing; note how at t ≈ 1.16, S ′ (t) is minimized. That y
1
is, sales are decreasing at the fastest rate at t ≈ 1.16. On the interval
of (1.16, 2), S is decreasing but concave up, so the decline in sales is 0.8
“leveling off.”
0.6
Not every critical point corresponds to a relative extrema; f (x) = x3 has a
critical point at (0, 0) but no relative maximum or minimum. Likewise, just be- 0.4
cause f ′′ (x) = 0 we cannot conclude concavity changes at that point. We were
0.2
careful before to use terminology “possible point of inflection” since we needed
to check to see if the concavity changed. The canonical example of f ′′ (x) = 0 x
without concavity changing is f (x) = x4 . At x = 0, f ′′ (x) = 0 but f is always −1 −0.5 0.5 1
concave up, as shown in Figure 3.4.21.
Figure 3.4.21 A graph of f (x) = x4 .
Clearly f is always concave up, despite
3.4.2 The Second Derivative Test the fact that f ′′ (x) = 0 when x = 0.
The first derivative of a function gave us a test to find if a critical value corre- It this example, the possible point of
sponded to a relative maximum, minimum, or neither. The second derivative inflection (0, 0) is not a point of inflec-
gives us another way to test if a critical point is a local maximum or minimum. tion.
The following theorem officially states something that is intuitive: if a critical 10 y
value occurs in a region where a function f is concave up, then that critical value concave down
must correspond to a relative minimum of f , etc. See Figure 3.4.22 for a visual- =⇒ rel max
5
ization of this.
x
Theorem 3.4.23 The Second Derivative Test.
−2 −1 1 2
Let c be a critical value of f where f ′′ (c) is defined.
−5
1. If f ′′ (c) > 0, then f has a local minimum at (c, f (c)). concave up
=⇒ rel min
−10
2. If f ′′ (c) < 0, then f has a local maximum at (c, f (c)).
Figure 3.4.22 Demonstrating the fact
The Second Derivative Test relates to the First Derivative Test in the following that relative maxima occur when the
way. If f ′′ (c) > 0, then the graph is concave up at a critical point c and f ′ itself graph is concave down and relative min-
is growing. Since f ′ (c) = 0 and f ′ is growing at c, then it must go from negative ima occur when the graph is concave
to positive at c. This means the function goes from decreasing to increasing, up
indicating a local minimum at c.
Let f (x) = 100/x + x. Find the critical points of f and use the The
Second Derivative Test to label them as relative maxima or minima.
Solution. We find f ′ (x) = −100/x2 + 1 and f ′′ (x) = 200/x3 . We
set f ′ (x) = 0 and solve for x to find the critical values (note that f ′ is
not defined at x = 0, but neither is f so this is not a critical value.) We
[Link]/watch?v=4hWldEUoG2U
Figure 3.4.24 Video presentation of The-
orem 3.4.23
Video Solution
CHAPTER 3. THE GRAPHICAL BEHAVIOR OF FUNCTIONS 160
find the critical values are x = ±10. We now evaluate the second deriv-
ative at these critical numbers. Evaluating f ′′ (10) = 0.1 > 0, so there
is a local minimum at x = 10. Evaluating f ′′ (−10) = −0.1 < 0, deter-
mining a relative maximum at x = −10. These results are confirmed in
Figure 3.4.26.
[Link]/watch?v=_DhmPXRZfi8
We have been learning how the first and second derivatives of a function
relate information about the graph of that function. We have found intervals of
increasing and decreasing, intervals where the graph is concave up and down,
y
along with the locations of relative extrema and inflection points. In Chapter 1 40
we saw how limits explained asymptotic behavior. In the next section we com-
bine all of this information to produce accurate sketches of functions. 20
f ′′ (10) > 0
x
−20 −10 10 20
f ′′ (−10) < 0
−20
−40
3.4.3 Exercises
Terms and Concepts
1. Sketch a graph of a function f (x) that is concave up on (0, 1) and is concave down on (1, 2).
2. Sketch a graph of a function f (x) that is:
Problems
Exercise Group. A function f (x) is given. Graph f and f ′′ on the same axes (using technology is permitted) and
verify Theorem 3.4.7.
5. f (x) = −7x + 3 6. f (x) = −4x2 + 3x − 8
7. f (x) = 4x2 + 3x − 8 8. f (x) = x3 − 3x2 + x − 1
9. f (x) = −x3 + x2 − 2x + 5 10. f (x) = sin(x)
11. f (x) = tan(x) 1
12. f (x) = 2
x +1
1
13. f (x) = x 14. f (x) = x12
Exercise Group. A function f (x) is given. Find the critical points of f and use the Second Derivative Test, when
possible, to determine the relative extrema. (Note: these are the same functions as in Exercise Group 15–28.)
29. f (x) = x2 + 14x + 49 30. f (x) = −x2 − 5x + 3
31. f (x) = x3 − 4x − 4 32. f (x) = −x3 + 8x2 − 25x − 3
33. f (x) = x 4
+ 64x − 9 34. f (x) = 2x4 − 8x3 − 16x2 + 96x + 9
4
CHAPTER 3. THE GRAPHICAL BEHAVIOR OF FUNCTIONS 162
35. f (x) = x4 − 12x3 + 54x2 − 108x + 81 36. f (x) = sec(x) on (−3π/2, 3π/2)
1 1
37. f (x) = x2 +18x+83 38. f (x) = x2 −49
39. f (x) = sin(x) + cos(x) on (−π, π) 2 x
40. f (x) = x e
f (x) = x2 ln(x) f (x) = e−x
2
41. 42.
Exercise Group. A function f (x) is given. Find the x values where f ′ (x) has a relative maximum or minimum. (Note:
these are the same functions as in Exercise Group 15–28.)
43. f (x) = x2 − 8x + 16 44. f (x) = −x2 + 6x + 4
45. f (x) = x − 9x − 2
3
46. f (x) = −9x3 − 8x2 − 7x − 1
47. f (x) = x4 3
+ 14 x3 + 7 48. f (x) = 3x4 − 24x3 + 66x2 − 72x − 6
4
49. f (x) = x + 4x3 + 6x2 + 4x + 1
4 50. f (x) = sec(x) on (−3π/2, 3π/2)
1 1
51. f (x) = x2 −2x+4 52. f (x) = x2 −13x+36
53. f (x) = sin(x) + cos(x) on (−π, π) 2 x
54. f (x) = x e
f (x) = e−x
2 2
55. f (x) = x ln(x) 56.
CHAPTER 3. THE GRAPHICAL BEHAVIOR OF FUNCTIONS 163
√ √
1
10 − 37
10
9 ≈ 1
10 + 37
9 9
1.111
≈ 0.435 ≈ 1.787
10 y 10 y
10
8
5 5 6
4
x x
2
−1 1 2 3 −1 1 2 3 1.0 0.5 0.5 1.0 1.5 2.0 2.5 3.0
2
4
−5 −5
x2 − x − 2 Video Solution
Sketch f (x) = .
x2 − x − 6
Solution. We again follow the steps outlined in Key Idea 3.5.1.
24x2 − 24x + 56
f ′′ (x) = .
(x − 3)3 (x + 2)3
In Figure 3.5.7(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 function looks like (these lines effectively only
convey increasing/decreasing information). In Figure 3.5.7(b), we
adjust the graph with the appropriate concavity. We also show f
crossing the x-axis at x = −1 and x = 2 and crossing the y-axis
at y = 1/3. Finally, Figure 3.5.7(c) shows a computer generated
graph of f , which verifies the accuracy of our sketch.
−2 1 3
2
y y
4 4
4
2 2
2
x x
−4 −2 −4 −2
2 4 2 4
4 2 2 4
−2 −2 2
−4 −4 4
5(x − 2)(x + 1)
Sketch f (x) = .
x2 + 2x + 4
Solution. We again follow Key Idea 3.5.1.
8 y 8 y
8
6 6
6
4 4
4
2 2
2
x x
−6 −4 −2 2 4 6 −6 −4 −2 2 4 6 6 4 2 2 4 6
−2 −2 2
To get some more practice with curve sketching, we include a few more
video examples to illustrate the process. (The last of these could be considered
“archival footage”: it was from a first run at using our new lightboard.)
In each of our examples, we found a few significant points on the graph of
f that corresponded to changes in increasing/decreasing or concavity. We con-
nected these points with straight lines, then adjusted for concavity, and finished
by showing a very accurate, computer generated graph. [Link]/watch?v=S3j-vuUZPjE
Why are computer graphics so good? It is not because computers are “smarter” Figure 3.5.11 Sketching the polynomial
than we are. Rather, it is largely because computers are much faster at comput- f (x) = x2 (5 − x)3
ing than we are. In general, computers graph functions much like most students
do when first learning to draw graphs: they plot equally spaced points, then con-
nect the dots using lines. By using lots of points, the connecting lines are short
and the graph looks smooth.
This does a fine 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 noticeable sharp edges on the graph unless a large number of
points are used. High quality computer algebra systems, such as Mathematica
and Sage, use special algorithms to plot lots of points only where the graph is
“curvy.”
In Figure 3.5.14, two graph of y = sin(x) is given, generated by Sage and
Mathematica. The small points represent each of the places where each cas
sampled the function. Notice how at the “bends” of sin(x), lots of points are
used; where sin(x) is relatively straight, fewer points are used. (In the Math- [Link]/watch?v=fdnoec_9Yw4
ematica plot, many points are also used at the endpoints to ensure the “end Figure 3.5.12 Sketching the graph of
behavior” is accurate.) the trigonometric function f (x) = sin(2x)−
2 sin(x)
1.0
0.5
1 2 3 4 5 6
0.5
1.0
(a) Sage output (b) Mathematica output
Figure 3.5.14 CAS plots of y = sin(x) illustrating the sample points
How does Sage know where the graph is “curvy”? Calculus. When we study [Link]/watch?v=JR31YX5N3M8
curvature in a later chapter, we will see how the first and second derivatives of a Figure 3.5.13 Sketching the graph of
f (x) = x4/3 − 4x1/3
CHAPTER 3. THE GRAPHICAL BEHAVIOR OF FUNCTIONS 169
3.5.1 Exercises
Terms and Concepts
Problems
Exercise Group. In the following exercises, practice using Key Idea 3.5.1 by applying the principles to the given
functions with familiar graphs.
6. Use Key Idea 3.5.1 to sketch a graph of f (x) = 2x + 4
7. Use Key Idea 3.5.1 to sketch a graph of f (x) = −x2 + 1
8. Use Key Idea 3.5.1 to sketch a graph of f (x) = sin(x)
9. Use Key Idea 3.5.1 to sketch a graph of f (x) = ex
1
10. Use Key Idea 3.5.1 to sketch a graph of f (x) =
x
1
11. Use Key Idea 3.5.1 to sketch a graph of f (x) = 2
x
Exercise Group. In the following exercises, sketch a graph of the given function using Key Idea 3.5.1. Show all work;
check your answer with technology.
12. Use Key Idea 3.5.1 to sketch a graph of f (x) = x3 − 2x2 + 4x + 1
13. Use Key Idea 3.5.1 to sketch a graph of f (x) = −x3 + 5x2 − 3x + 2
14. Use Key Idea 3.5.1 to sketch a graph of f (x) = x3 + 3x2 + 3x + 1
15. Use Key Idea 3.5.1 to sketch a graph of f (x) = x3 − x2 − x + 1
16. Use Key Idea 3.5.1 to sketch a graph of f (x) = (x − 2) ln(x − 2)
17. Use Key Idea 3.5.1 to sketch a graph of f (x) = (x − 2)2 ln(x − 2)
x2 − 4
18. Use Key Idea 3.5.1 to sketch a graph of f (x) =
x2
x2 − 4x + 3
19. Use Key Idea 3.5.1 to sketch a graph of f (x) = 2
x − 6x + 8
x2 − 2x + 1
20. Use Key Idea 3.5.1 to sketch a graph of f (x) =
x2 − 6x + 8
√
21. Use Key Idea 3.5.1 to sketch a graph of f (x) = x x + 1
22. Use Key Idea 3.5.1 to sketch a graph of f (x) = x2 ex
23. Use Key Idea 3.5.1 to sketch a graph of f (x) = sin(x) cos(x) on [−π, π]
24. Use Key Idea 3.5.1 to sketch a graph of f (x) = (x − 3)2/3 + 2
(x − 1)2/3
25. Use Key Idea 3.5.1 to sketch a graph of f (x) =
x
CHAPTER 3. THE GRAPHICAL BEHAVIOR OF FUNCTIONS 171
Exercise Group. In the following exercises, a function with the parameters a and b are given. Describe the critical
points and possible points of inflection of f in terms of a and b.
a
26. f (x) = 2
x + b2
(a) Find the critical points of f .
(b) Find the inflection points of f .
27. f (x) = sin(ax + b)
In Chapter 3, we learned how the first and second derivatives of a function influ-
ence its graph. In this chapter we explore other applications of the derivative.
1 y 1 y 1 y
x x x
x0 x1 x0 x2 x1 x0 x2 x3 x1
−1 −1 −1
172
CHAPTER 4. APPLICATIONS OF THE DERIVATIVE 173
This line crosses the x-axis when y = 0, and the x-value where it crosses is
what we called x1 . So let y = 0 and replace x with x1 , giving the equation:
f (x1 )
x2 = x1 − .
f ′ (x1 )
f (xn )
xn+1 = xn − .
f ′ (xn )
Newton’s Method is not Infalli-
3. Stop the iterations when successive approximations do not differ
ble. The sequence of approximate
in the first d places after the decimal point.
values may not converge, or it may
converge so slowly that one is “tricked”
Let’s practice Newton’s Method with a concrete example.
into thinking a certain approxima-
Example 4.1.3 Using Newton’s Method. tion is better than it actually is.
These issues will be discussed at
Approximate the real root of x3 − x2 − 1 = 0, accurate to the first the end of the section.
three places after the decimal, using Newton’s Method and an initial
approximation of x0 = 1.
Solution. To begin, we compute f ′ (x) = 3x2 − 2x. Then we apply the
Newton’s Method algorithm, outlined in Key Idea 4.1.2.
CHAPTER 4. APPLICATIONS OF THE DERIVATIVE 174
f (1) f (1.625)
x1 = 1 − x3 = 1.625 −
f ′ (1) f ′ (1.625)
13 − 12 − 1 1.6253 − 1.6252 − 1
=1− = 1.625 −
3 · 12 − 2 · 1 3 · 1.6252 − 2 · 1.625
=2 ≈ 1.48579
f (1.48579)
x4 = 1.48579 −
f ′ (1.48579)
f (2)
x2 = 2 − ≈ 1.46596
f ′ (2)
23 − 22 − 1 f (1.46596)
=2− x5 = 1.46596 −
3 · 22 − 2 · 2 f ′ (1.46596)
= 1.625 ≈ 1.46557
We can automate this process on a calculator that has an ANS key that returns −1
the result of the previous calculation. Start by pressing 1 and then Enter. (We
have just entered our initial guess, x0 = 1.) Now compute −1.5
We can continue this way, but it is really best to automate this process.
On a calculator with an ANS key, we would start by entering 0.75, then
Enter, inputting our initial approximation. We then enter:
ANS - (cos(ANS)-ANS)/(-sin(ANS)-1)
Repeatedly pressing the Enter key gives successive approximations.
We quickly find:
x3 = 0.7390851332
x4 = 0.7390851332.
Our approximations x2 and x3 did not differ for at least the first five
places after the decimal, so we could have stopped. However, using our
calculator in the manner described is easy, so finding x4 was not hard.
It is interesting to see how we found an approximation, accurate to as
many decimal places as our calculator displays, in just four iterations.
If you know how to program, you can translate the following pseudocode
into your favorite language to perform the computation in this problem.
x = 0.75
while true
oldx = x
x = x - (cos(x)-x)/(-sin(x)-1)
print x
if abs(x-oldx) < 0.0000000001
break
This code calculates x1 , x2 , etc., storing each result in the variable x. The previ-
ous approximation is stored in the variable oldx. We continue looping until the
difference between two successive approximations, abs(x-oldx), is less than
some small tolerance, in this case, 0.0000000001.
Convergence of Newton’s Method. What should one use for the initial guess,
x0 ? Generally, the closer to the actual root the initial guess is, the better. How-
CHAPTER 4. APPLICATIONS OF THE DERIVATIVE 176
0.5 y
ever, some initial guesses should be avoided. For instance, consider Example 4.1.3
where we sought the root to f (x) = x3 − x2 − 1. Choosing x0 = 0 would have x
been a particularly poor choice. Consider Figure 4.1.7, where f (x) is graphed −0.5 0.5 1 1.5
along with its tangent line at x = 0. Since f ′ (0) = 0, the tangent line is horizon-
tal and does not intersect the x-axis. Graphically, we see that Newton’s Method −0.5
fails.
We can also see analytically that it fails. Since −1
f (0)
x1 = 0 − −1.5
f ′ (0)
and f ′ (0) = 0, we see that x1 is not well defined. Figure 4.1.7 A graph of f (x) = x3 −
This problem can also occur if, for instance, it turns out that f ′ (x5 ) = 0. x2 −1, showing why an initial approxi-
Adjusting the initial approximation x0 by a very small amount will likely fix the mation of x0 = 0 with Newton’s Method
problem. fails
It is also possible for Newton’s Method to not converge while each successive
approximation is well defined. Consider f (x) = x1/3 , as shown in Figure 4.1.8.
It is clear that the root is x = 0, but let’s approximate this with x0 = 0.1. Fig-
ure 4.1.8(a) shows graphically the calculation of x1 ; notice how it is farther from
the root than x0 . Figure 4.1.8(b) and Figure 4.1.8(c) show the calculation of x2
and x3 , which are even farther away; our successive approximations are getting
worse. (It turns out that in this particular example, each successive approxima-
tion is twice as far from the true answer as the previous approximation.)
y y y
1 1 1
x x x
−1 x1 x0 1 −1 x1 x0 x2 1 −1 x3 x1 x0 x2 1
−1 −1 −1
4.1.1 Exercises
Terms and Concepts
1. (□ True □ False) Given a function f (x), Newton’s Method produces an exact solution to f (x) = 0.
2. (□ True □ False) In order to get a solution to f (x) = 0 accurate to d places after the decimal, at least
d + 1 iterations of Newton’s Method must be used.
Problems
Exercise Group. The roots of the function f (x) are known or are easily found. Use five iterations of Newton’s Method
with the given initial approximation to approximate the root. Compare it to the known value of the root.
3. f (x) = cos(x), x0 = 1.5 4. f (x) = sin(x), x0 = 1
Compare x5 to the known value of the root. Compare x5 to the known value of the root.
5. f (x) = x2 + x − 2, x0 = 0 6. f (x) = x2 − 2, x0 = 1.5
Compare x5 to the known value of the root. Compare x5 to the known value of the root.
7. f (x) = ln(x), x0 = 2 8. f (x) = x3 − x2 + x − 1, x0 = 2
Compare x5 to the known value of the root. Compare x5 to the known value of the root.
Exercise Group. Use Newton’s Method to approximate all roots of the given function accurate to three places after
the decimal. If an interval is given, find only the roots that lie within that interval. Use technology to obtain good
initial approximations.
9. f (x) = x3 + 5x2 − x − 1
Show the steps you took using Newton’s Method.
10. f (x) = x4 + 2x3 − 7x2 − x + 5
Show the steps you took using Newton’s Method.
11. f (x) = x17 − 2x13 − 10x8 + 10 on (−2, 2)
Show the steps you took using Newton’s Method.
12. f (x) = x2 cos(x) + (x − 1) sin(x) on (−3, 3)
Show the steps you took using Newton’s Method.
Exercise Group. Use Newton’s Method to approximate when the given functions are equal, accurate to 3 places after
the decimal. Use technology to obtain good initial approximations.
13. f (x) = x2 , g(x) = cos(x)
Show the steps you took using Newton’s Method.
14. f (x) = x2 − 1, g(x) = sin(x)
Show the steps you took using Newton’s Method.
2
15. f (x) = ex , g(x) = cos(x)
Show the steps you took using Newton’s Method.
16. f (x) = x, g(x) = tan(x) on [−6, 6]
Show the steps you took using Newton’s Method.
17. Why does Newton’s Method fail in finding a root of f (x) = x3 − 3x2 + x + 3 when x0 = 1?
18. Why does Newton’s Method fail in finding a root of f (x) = −17x4 + 130x3 − 301x2 + 156x + 156 when
x0 = 1?
CHAPTER 4. APPLICATIONS OF THE DERIVATIVE 178
The radius of a circle is growing at a rate of 5 inh . At what rate is the Video Solution
circumference growing?
Solution. The circumference and radius of a circle are related by C =
2πr. We are given information about how the length of r changes with
in
respect to time; that is, we are told dr
dt is 5 h . We want to know how the
length of C changes with respect to time, i.e., we want to know dC dt .
Implicitly differentiate both sides of C = 2πr with respect to t:
C = 2πr
d d [Link]/watch?v=Qg3GStrQ8pY
(C) = (2πr)
dt dt
dC dr
= 2π .
dt dt
As we know dr
dt is 5 inh , we know
dC
= 2π5 = 10π ≈ 31.4 in/hr .
dt
This problem was relatively straightforward, owing to the linear re-
lationship between radius and circumference. The video in Figure 4.2.4
explores what would happen if we had instead been asked for the rate
at which the area is changing.
1. Read the problem carefully and identify the quantities that are
changing with time. (There may be many quantities that change
with time, try to identify which variables are important to your
goal and only focus on these quantities.)
CHAPTER 4. APPLICATIONS OF THE DERIVATIVE 179
Solution.
1. We can answer this question two ways: using “common sense” or
related rates. The common sense method states that the volume
3
of the puddle is growing by 2 ins , where
Since the depth is constant at 1/8 in, the area must be growing [Link]/watch?v=8ctKxMoFWkU
2
by 16 ins since 16 · 18 = 2. This approach reveals the underlying
related rates principle.
Now let’s solve the problem using Key Idea 4.2.5. Based on the
problem description, the quantities that change with time are the
volume of water (the volume of the puddle), the area of the circu-
lar puddle and the radius of the circle. We don’t need a diagram
for this problem. The important variables for this part of the prob-
lem are the volume and area.
Let V and A represent the Volume and Area of the puddle. We
know V = A × 18 . Take the derivative of both sides with respect
to t, employing implicit differentiation.
1
V = A
8
CHAPTER 4. APPLICATIONS OF THE DERIVATIVE 180
d d 1
(V ) = A
dt dt 8
dV 1 dA
=
dt 8 dt
We know the change in volume, dV dt = 2, so we substitute this
value into our related rates equation: 2 = 18 dA dA
dt , and hence dt =
2
16. Thus the area is growing by 16 ins .
2. We already identified the quantities that are changing in Part 1.
The variables of interest in this problem are the radius and the
volume. We need an equation that relates the volume of the circle
to the radius. Since the puddle is a right circular cylinder, we will
use a known volume formula, V = πr2 h where V is the volume
of the puddle (in in3 , r is the radius (in inches) and h is the height
(i.e. depth) of the puddle in inches. (Notice that this formula is
equivalent to V = area×depth.) We know that the height (depth)
is a constant 1/8 inch. Since this quantity does not change in the
problem, we can safely substitute this value now.
Implicitly derive both sides of V = πr2 18 with respect to t:
1 2
V = πr
8
d d 1 2
V = πr
dt dt 8
dV 1 dr
= 2πr
dt 8 dt
dV 1 dr
= πr
dt 4 dt
3
We know that dV
dt is 2 ins . So we have:
1 dr
2= πr
4 dt
dr
Solving for dt , we have
dr 8
= .
dt πr
Note how our answer is not a number, but rather a function of r.
In other words, the rate at which the radius is growing depends on
3
how big the circle already is. If the circle is very large, adding 2 ins
of water will not make the circle much bigger at all. If the circle is
dime-sized, adding the same amount of water will make a radical
change in the radius of the circle.
In some ways, our problem was (intentionally) ill-posed. We need
to specify a current (instantaneous) value of the radius in order to
know a rate of change. When the puddle has a radius of 10 in, the
radius is growing at a rate of
dr 8 4
= = ≈ 0.25 in/s .
dt 10π 5π
CHAPTER 4. APPLICATIONS OF THE DERIVATIVE 181
Radar guns measure the rate of distance change between the gun and
the object it is measuring. For instance, a reading of “55 mph” means
the object is moving away from the gun at a rate of 55 miles per hour,
whereas a measurement of “−25 mph” would mean that the object is
approaching the gun at a rate of 25 miles per hour.
If the radar gun is moving (say, attached to a police car) then radar N
readouts are only immediately understandable if the gun and the object B = 1/2
are moving along the same line. If a police officer is traveling 60 mph and E
gets a readout of 15 mph, he knows that the car ahead of him is moving Car
away at a rate of 15 miles an hour, meaning the car is traveling 75 mph.
A = 1/2
(This straight-line principle is one reason officers park on the side of the
highway and try to shoot straight back down the road. It gives the most C
accurate reading.)
Suppose an officer is driving due north at 30 mph and sees a car mov-
ing due east, as shown in Figure 4.2.8. Using his radar gun, he measures
a reading of 20 mph. By using landmarks, he believes both he and the Officer
other car are about 1/2 mile from the intersection of their two roads.
If the speed limit on the other road is 55 mph, is the other driver
speeding? Figure 4.2.8 A sketch of a police car
(at bottom) attempting to measure the
Solution. The important quantities that are changing are: the distance
speed of a car (at right) in Example 4.2.7
of the officer to the intersection, the distance of the car to the intersec-
tion, and the distance of the officer to the car. (There are other quanti-
ties that are changing as well such as the angles and area of the triangle, Video Solution
but these are not important to this problem.)
Using the diagram in Figure 4.2.8, let’s label what we know about
the situation. As both the police officer and other driver are 1/2 mile
from the intersection, we have√A = 1/2, B = 1/2, and through the
Pythagorean Theorem, C = 1/ 2 ≈ 0.707. These values are “instanta-
neous” values for our variables, so we won’t use them until the end of
the problem. Instead, we will use the variables A, B, and C.
We need an equation that relates A, B, and C. The Pythagorean
Theorem is a good choice: A2 + B 2 = C 2 . Differentiate both sides with [Link]/watch?v=8fFao8XCQC0
respect to t:
A2 + B 2 = C 2
d d
A2 + B 2 = C2
dt dt
dA dB dC
2A + 2B = 2C
dt dt dt
dt = −30.
We know the police officer is traveling at 30 mph; that is, dA
The reason this rate of change is negative is that A is getting smaller; the
distance between the officer and the intersection is shrinking. The radar
measurement is dC dt = 20. We want to find dt .
dB
dB
We have values for everything except dt . Solving for this we have:
dB C dC − A dA
= dt dt
.
dt B
Now we substitue in our known rates and instantaneous values of our
variables:
dB 0.707(20) − 0.5(−30)
≈
dt (0.5)
CHAPTER 4. APPLICATIONS OF THE DERIVATIVE 182 Practicality. Example 4.2.7 is both
interesting and impractical. It high-
lights the difficulty in using radar
= 58.28 mph . in a nonlinear fashion, and explains
why “in real life” the police offi-
The other driver appears to be speeding slightly. cer would follow the other dri-
ver to determine their speed, and
not pull out pencil and paper.
Example 4.2.9 Studying related rates. The principles here are impor-
tant, though. Many automated
A camera is placed on a tripod 10 ft from the side of a road. The camera
vehicles make judgments about
is to turn to track a car that is to drive by at 100 mph for a promotional
other moving objects based on
video. The video’s planners want to know what kind of motor the tripod
perceived distances, radar-like mea-
should be equipped with in order to properly track the car as it passes
surements and the concepts of
by. Figure 4.2.10 shows the proposed setup.
related rates.
How fast must the camera be able to turn to track the car?
Solution. The quantities that changing are x and θ as drawn on Fig-
ure 4.2.10. (The hypotenuse of the triangle is also changing, but this
isn’t important to the problem). We seek information about how fast
the camera is to turn; therefore, we need an equation that will relate an
angle θ to the position of the camera and the speed and position of the 100mph
car.
Figure 4.2.10 suggests we use a trigonometric equation. Letting x x
represent the distance the car is from the point on the road directly in
front of the camera, we have 10ft
x
tan(θ) = . (4.2.1) θ
10
Now take the derivative of both sides of Equation (4.2.1) using im- Figure 4.2.10 Tracking a speeding car
plicit differentiation: (at left) with a rotating camera
x
tan(θ) =
10
d dx Video Solution
(tan(θ)) =
dt dt 10
dθ 1 dx
sec2 (θ) =
dt 10 dt
dθ
Now we solve for dt :
dθ cos2 (θ) dx
= (4.2.2)
dt 10 dt
As the car is moving at 100 mph, we have that dx dt is −100 mph (as
in the last example, since x is getting smaller as the car travels, dx
dt is [Link]/watch?v=B85LlGHgVQo
negative). We need to convert the measurements so they use the same
units (we chose ft); rewrite −100 mph in terms of fts :
dx mi
= −100
dt hr
mi ft 1 hr
= −100 · 5280 ·
hr mi 3600 s
= −146.6 ft/s .
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
θ = 0). Our mathematics bears this out. In Equation (4.2.2) we see this
is when cos2 (θ) is largest; this is when cos(θ) = 1, or when θ = 0. We
CHAPTER 4. APPLICATIONS OF THE DERIVATIVE 183
dθ 1
≈− 146.67 ft/s
dt 10 ft
= −14.667 radians/s
We find that dθ
dt is negative; this matches our diagram in Figure 4.2.10
for θ is getting smaller as the car approaches the camera.
What is the practical meaning of −14.667 rad s ? Recall that 1 circular
revolution goes through 2π radians, thus 14.667 rad s means 14.667/(2π) ≈
2.33 revolutions per second. The negative sign indicates the camera is
rotating in a clockwise fashion.
4.2.1 Exercises
Terms and Concepts
1. (□ True □ False) Implicit differentiation is often used when solving “related rates” type problems.
2. (□ True □ False) A study of related rates is part of the standard police officer training.
Problems
cm3
3. Water flows onto a flat surface at a rate of 4 s forming a circular puddle 8 mm deep. How fast is the radius
growing when the radius is:
(a) 2 cm
(b) 20 cm
(c) 200 cm
cm3
4. A spherical balloon is inflated with air flowing at a rate of 5 s . How fast is the radius of the balloon increasing
when the radius is:
(a) 1 cm
(b) 10 cm
(c) 100 cm
5. Consider the traffic situation introduced in Example 4.2.7. How fast is the “other car” traveling if the officer and
the other car are each 34 mile from the intersection, the other car is traveling due west, the officer is traveling
north at 55 mph, and the radar reading is −75 mph?
6. Consider the traffic situation introduced in Example 4.2.7. Calculate how fast the “other car” is traveling in each
of the following situations.
(a) The officer is traveling due north at 50 mph and is 34 mile from the intersection, while the other car is 1
mile from the intersection traveling west and the radar reading is −85 mph?
3
(b) The officer is traveling due north at 50 mph and is 1 mile from the intersection, while the other car is 4
mile from the intersection traveling west and the radar reading is −85 mph?
7. An F-22 aircraft is flying at 530 mph with an elevation of 6600 ft on a straight-line path that will take it directly
over an anti-aircraft gun.
6600 ft
θ
x
How fast (in radians per second) must the gun be able to turn to accurately track the aircraft when the plane
is:
8. An F-22 aircraft is flying at 500 mi/h with an elevation of 100 ft on a straight-line path that will take it directly
over an anti-aircraft gun as in Exercise 4.2.7 (note the lower elevation here).
How fast must the gun be able to turn to accurately track the aircraft when the plane is:
24
1 ft/s
ft
At what rate is the top of the ladder sliding down the side of the house when the base is:
(a) 1 foot from the house?
(c) 22 feet?
(d) How long will the tank take to fill when starting at empty?
12. A rope, attached 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 connection point between rope and weight.
CHAPTER 4. APPLICATIONS OF THE DERIVATIVE 186
30 ft
2 ft/s
Suppose the man stands directly next to the weight (i.e., a total rope length of 60 feet) and begins to walk
away at a rate of 2 ft/s. How fast is the weight rising when the man has walked:
(a) 10 feet?
(b) 40 feet?
(c) How far must the man walk to raise the weight all the way to the pulley?
13. Consider the situation described in Exercise 4.2.12. Suppose the man starts 40 ft from the weight and begins to
walk away at a rate of 2 fts .
(c) How fast is the weight rising after the man has walked 30 feet?
(d) How far must the man walk to raise the weight all the way to the pulley?
14. A hot air balloon lifts off from ground rising vertically. From 90 feet away, a 6 ft tall woman tracks the path of
the balloon. When her sightline with the balloon makes a 45◦ angle with the horizontal, she notes the angle is
increasing at about 3◦ per minute.
(a) What is the elevation of the balloon?
(b) How fast is it rising?
15. A company that produces landscaping materials is dumping sand into a conical pile. The sand is being poured
3
at a rate of 5 fts . The physical properties of the sand, in conjunction with gravity, ensure that the cone’s height
is roughly 47 the length of the diameter of the circular base.
How fast is the cone rising when it has a height of 30 feet?
CHAPTER 4. APPLICATIONS OF THE DERIVATIVE 187
4.3 Optimization
In Section 3.1 we learned about extreme values — the largest and smallest
values a function attains on an interval. We motivated 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 section we apply [Link]/watch?v=nlLf3CcgblI
the concepts of extreme values to solve “word problems,” i.e., problems stated Figure 4.3.1 A simple optimization prob-
in terms of situations that require us to create the appropriate mathematical lem
framework in which to solve the problem.
We start with a classic example which is followed by a discussion of the topic
of optimization.
A man has 100 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?
Solution. One can likely guess the correct answer — that is great. We
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 situation. Our enclosure is sketched
twice in Figure 4.3.3, either with treetop 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 function — after all, we are trying to maximize the area.
y
y
x
x
Figure 4.3.3 A sketch of the enclosure in Example 4.3.2.
We let x and y denote the lengths of the sides of the rectangle. Clearly,
Area = xy.
We now have two equations and two unknowns. In the latter equa-
tion, we solve for y:
y = 50 − x.
Now substitute this expression for y in the area equation:
Note we now have an equation of one variable; we can truly call the
Area a function of x.
CHAPTER 4. APPLICATIONS OF THE DERIVATIVE 188
A(x) = x(50 − x)
= 50x − x2
A′ (x) = 50 − 2x
This example is very simplistic and a bit contrived. (After 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 equations that de-
scribe a situation, reduce an equation to a single variable, then find the needed
extreme value.
“In real life,” problems are much more complex. The equations are often
not reducible to a single variable (hence multi-variable calculus is needed) and
the equations themselves may be difficult to form. Understanding the princi-
ples here will provide a good foundation for the mathematics you will likely en-
counter later.
We outline here the basic process of solving these optimization problems.
Here is another classic calculus problem: A woman has a 100 feet of fenc-
ing, 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 won’t swim away.) Video Solution
What dimensions provide the maximal area?
Solution. We will follow the steps outlined by Key Idea 4.3.4.
x [Link]/watch?v=wIs5N5HOCrc
y
y
x
Figure 4.3.6 A sketch of the enclosure in Example 4.3.5
2. We want to maximize the area; as in the example before,
Area = xy.
Find the critical points. We have A′ (x) = 50−x; setting this equal
to 0 and solving for x returns x = 50. This gives an area of
A power line needs to be run from a power station located on the beach 1000 ft
to an offshore facility. Figure 4.3.8 shows the distances between the
power station to the facility.
5000 ft
It costs $50/ ft to run a power line along the land, and $130/ ft to
run a power line under water. How much of the power line should be Figure 4.3.8 Running a power line from
run along the land to minimize the overall cost? What is the minimal the power station to an offshore facil-
cost? ity with minimal cost in Example 4.3.7
Solution. We will follow the strategy of Key Idea 4.3.4 implicitly, with-
out specifically numbering steps.
There are two immediate solutions that we could consider, each of
which we will reject through “common sense.” First, we could minimize
the distance by directly connecting the two locations with a straight line.
However, this requires that all the wire be laid underwater, the most Video Solution
costly option. Second, we could minimize the underwater length by run-
ning a wire all 5000 ft along the beach, directly across from the offshore
facility. This has the undesired effect of having the longest distance of
all, probably ensuring a non-minimal cost.
The optimal solution likely has the line being run along the ground
for a while, then underwater, as the figure 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
[Link]/watch?v=qDK9rqloKRs
the distances as shown in Figure 4.3.9.
By choosing x as we did (instead of letting x be the distance along
the land), we make the expression under the square root simple. We
now create the cost function. 2
000
√ x2 +1 1000 ft
Cost = land cost + water cost
$50 × land distance + $130 × water distance
p 5000 − x x
50(5000 − x) + 130 x2 + 10002 .
√ Figure 4.3.9 Labeling unknown distances
So we have c(x) = 50(5000 − x) + 130 x2 + 10002 . This function in Example 4.3.7
only makes sense on the interval [0, 5000]. While we are fairly certain
the endpoints will not give a minimal cost, we still evaluate c(x) at each
to verify.
(Notice that if x = 0, the line is run the full 5000 ft along land and a full
1000 ft under water. If x = 5000, the line is run the maximum distance
underwater.)
We now find the critical values of c(x). We compute c′ (x) as
130x
c′ (x) = −50 + √ .
x2 + 10002
Recognize that this is never undefined. Setting c′ (x) = 0 and solving
for x, we have:
130x
−50 + √ =0
x2 + 10002
130x
√ = 50
x2 + 10002
1302 x2
= 502
x + 10002
2
In the exercises you will see a variety of situations that require you to com-
bine problem-solving skills with calculus. Focus on the process; learn how to
form equations from situations that can be manipulated into what you need. Es-
chew memorizing how to do “this kind of problem” as opposed to “that kind
of problem.” Learning a process will benefit one far longer than memorizing a
specific technique.
Before you begin the exercises, here is one more example, presented in video
form in Figure 4.3.10.
Section 4.4 introduces our final application of the derivative: differentials.
Given y = f (x), they offer a method of approximating the change in y after x
changes by a small amount.
[Link]/watch?v=XJYDMZe8JUk
Figure 4.3.10 Optimizing construction
of a box with no top
CHAPTER 4. APPLICATIONS OF THE DERIVATIVE 192
4.3.1 Exercises
Terms and Concepts
1. (□ True □ False) An “optimization problem” is essentially an “extreme values” problem in a “story prob-
lem” setting.
2. (□ True □ False) This section teaches one to find the extreme values of a function that has more than one
variable.
Problems
3. Find the maximum product of two numbers (not necessarily integers) that have a sum of 150.
4. Find the minimum sum of two positive numbers whose product is 560.
5. Find the maximum sum of two positive numbers whose product is 580.
6. Find the maximum sum of two numbers, each of which is less than or equal to 290, whose product is 400.
7. Find the maximal area of a right triangle with hypotenuse of length 2.
8. A rancher has 900 feet of fencing in which to construct adjacent, equally sized rectangular pens. What dimen-
sions should these pens have to maximize the enclosed area?
9. A standard soda can is roughly cylindrical and holds 355 cm3 of liquid. What dimensions should the cylinder
have 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.
Discuss whether or not your calculation suggests that a real world soda can is designed to minimize the
materials cost.
10. Find the dimensions of a cylindrical can with a volume of 206 in3 that minimizes the surface area.
The “#10 can”is a standard sized can used by the restaurant industry that holds about 206 in3 with a diameter
3
of 6 16 in and height of 7 in. Does it seem these dimensions where chosen with minimization in mind?
Discuss whether or not your calculation suggests that a #10 can is designed to minimize the materials cost.
11. A standard soda can is roughly cylindrical and holds 355 cm3 of liquid. A real-world soda can has material on
the top and bottom that is thicker than the material around the side. Assume that the top/bottom material
is twice as thick as the material around the side. What dimensions should the cylinder have to minimize the
material needed to produce the can? Based on your dimensions and the assumption about material thickness,
determine whether or not the standard can is produced to minimize the material costs.
Discuss whether or not your calculation suggests that a real world soda can is designed to minimize the
materials cost.
12. The United States Postal Service charges more for boxes whose combined length and girth exceeds 108 inches.
(The “length” of a package is the length of its longest side; the girth is the perimeter of the cross section, i.e.,
2w + 2h).
What is the maximum volume of a package with a square cross section (w = h) that does not exceed the
108 inch standard?
CHAPTER 4. APPLICATIONS OF THE DERIVATIVE 193
13. The strength S of a wooden beam is directly proportional to its cross sectional width w and the square of its
height h. that is, S = kwh2 for some constant k.
12 h
Given a circular log with diameter of 18 inches, what sized beam can be cut from the log with maximum
strength?
14. A power line is to be run to an offshore facility in the manner described in Example 4.3.7. The offshore facility
is 6 miles at sea and 4 miles along the shoreline from the power plant. It costs $35,000 per mile to lay a power
line underground and $70,000 to run the line underwater.
How much of the power line should be run underground? What is the minimum overall cost?
15. A power line is to be run to an offshore facility in the manner described in Example 4.3.7. The offshore facility
is 6 miles at sea and 2 miles along the shoreline from the power plant. It costs $45,000 per mile to lay a power
line underground and $75,000 to run the line underwater.
How much of the power line should be run underground? What is the minimum overall cost?
16. A woman throws a stick into a lake for her dog to fetch; the stick is 35 feet down the shore line and 13 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 stick before swimming. The dog runs about 19 fts and swims about 2 fts .
How far along the shore should the dog run to minimize the time it takes to get to the stick? (Hint: the figure
from Example 4.3.7 can be useful.)
17. A woman throws a stick into a lake for her dog to fetch; the stick is 25 feet down the shore line and 16 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 stick before swimming. The dog runs about 22 fts and swims about 1.7 fts .
How far along the shore should the dog run to minimize the time it takes to get to the stick? (Google “calculus
dog” to learn more about a dog’s ability to minimize times.)
18. What are the dimensions of the rectangle with largest area that can be drawn inside the unit circle?
CHAPTER 4. APPLICATIONS OF THE DERIVATIVE 194
4.4 Differentials
In Section 2.2 we explored the meaning and use of the derivative. This sec-
tion starts by revisiting some of those ideas.
Recall that the derivative of a function f can be used to find the slopes of
lines tangent to the graph of f . At x = c, the tangent line to the graph of f has [Link]/watch?v=YmODT2PolKY
equation Figure 4.4.1 Video introduction to Sec-
y = f ′ (c)(x − c) + f (c). tion 4.4
The tangent line can be used to find good approximations of f (x) for values
of x near c.
For instance, we can approximate sin(1.1) using the tangent line
√ to the graph
of f (x) = sin(x) at x = π/3 ≈ 1.05. Recall that sin(π/3) = 3/2 ≈ 0.866,
and f ′ (π/3) = cos(π/3) = 1/2. Thus the tangent line to f (x) = sin(x) at
x = π/3 is:
1
ℓ(x) = (x − π/3) + 0.866.
2
y y
1
√
3 ℓ(1.1) ≈ sin(1.1)
2 √ sin(1.1)
(π/3, 3/3) 0.89
0.88
0.5
0.87
√
3
2 ( √ )
x π/3, 3/3
x
π
π 1.1
3 3
(a) (b)
Figure 4.4.2 Graphing f (x) = sin(x) and its tangent line at x = π/3 in order to
estimate sin(1.1)
In Figure 4.4.2(a), we see a graph of f (x) = sin(x) graphed along with its
tangent line at x = π/3. The small rectangle shows the region that is displayed
in Figure 4.4.2(b). In this figure, we see how we are approximating sin(1.1) with
the tangent line, evaluated at 1.1. Together, the two figures show how close
these values are.
Using this line to approximate sin(1.1), we have:
1
ℓ(1.1) = (1.1 − π/3) + 0.866
2
1
= (0.053) + 0.866 = 0.8925.
2
(We leave it to the reader to see how good of an approximation this is.)
We now generalize this concept. Given f (x) and an x-value c, the tangent
line is y = ℓ(x), where ℓ(x) = f ′ (c)(x − c) + f (c). Clearly, f (c) = ℓ(c). Let ∆x
be a small number, representing a small change in the x-value. We assert that:
[Link]/watch?v=mQRelmurD-w
f (c + ∆x) ≈ ℓ(c + ∆x),
Figure 4.4.3 Approximating the value
since the tangent line to a function approximates well the values of that func- of sin(1.1)
tion near x = c. This tangent line approximation is used frequently enough in
applications that we give it a name.
CHAPTER 4. APPLICATIONS OF THE DERIVATIVE 195
Definition 4.4.4
The function ℓ(x) is often referred to as the linearization, or linear ap-
proximation of f at c. It is the linear function that best approximates
the value of f (x) when x is close to 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).
This final equation is important; it becomes the basis of Definition 4.4.5 and
Key Idea 4.4.7. In short, it says that when the x-value changes from c to c + ∆x,
the y value of a function f changes by about f ′ (c)∆x.
We introduce two new variables, dx and dy in the context of a formal defin-
ition.
dy = f ′ (x)dx.
We can solve for f ′ (x) in the above equation: f ′ (x) = dy/dx. This states Differentials and linearization. The
that the derivative of f with respect to x is the differential of y divided by the relationship between the differ-
dy
differential of x; this is not the alternate notation for the derivative, dx . This ential and the linearization given
latter notation was chosen because of the fraction-like qualities of the derivative, in Definition 4.4.4 is as follows:
but again, it is one symbol and not a fraction.
It is helpful to organize our new concepts and notations in one place. ℓ(x) = f (c) + dy,
∆y = f (x + ∆x) − f (x).
When students first encounter differentials, they are often left wondering
why dy and ∆y are different, while dx and ∆x are the same. The video in Fig-
ure 4.4.8 attempts to offer an explanation.
[Link]/watch?v=XxpcZw702nA
Figure 4.4.8 Why is it that dx = ∆x?
CHAPTER 4. APPLICATIONS OF THE DERIVATIVE 196
∆y ≈ dy
= f ′ (3)dx
= 2 · 3 · 0.1 = 0.6. [Link]/watch?v=KCDezzvfDKA
Of course, it is easy to compute the actual answer (by hand or with a calcula-
tor): 3.12 = 9.61. (Before we get too cynical and say “Then why bother?”, note
our approximation is really good!)
So why bother?
In “most” real life situations, we do not know the function that describes
a particular behavior. Instead, we can only take measurements of how things
change — measurements of the derivative.
Imagine water flowing down a winding channel. It is easy to measure the
speed and direction (i.e., the velocity) of water at any location. It is very hard
to create a function that describes the overall flow, hence it is hard to predict
where a floating object placed at the beginning of the channel will end up. How-
ever, we can approximate the path of an object using differentials. Over small
PID controllers. Another place
intervals, the path taken by a floating object is essentially linear. Differentials
differentials are used is in a PID
allow us to approximate the true path by piecing together lots of short, linear
controller, which stands for “Pro-
paths. This technique is called Euler’s Method, studied in introductory Differen-
portional Integral Derivative”. A
tial Equations courses.
PID controller uses concepts of
We use differentials once more to approximate the value of a function. Even
both derivative and integral cal-
though calculators are very accessible, it is neat to see how these techniques
culus to very accurately control
can sometimes be used to easily compute something that looks rather hard.
a process (such as maintaining a
Example 4.4.10 Using differentials to approximate a function value. stable temperature on an espresso
machine).
√
Approximate 4.5.
√ √
Solution. We expect 4.5 ≈ 2, yet we can do better. Let f (x)√= x,
and let c = 4. Thus f (4) = 2. We can compute f ′ (x) = 1/(2 x), so
f ′ (4) = 1/4.
We approximate the difference between f (4.5) and f (4) using dif- Video Solution
ferentials, with dx = 0.5:
f (4.5) − f (4) = ∆y ≈ dy
= f ′ (4) · dx
= 1/4 · 1/2
[Link]/watch?v=nFaq1O_wWso
CHAPTER 4. APPLICATIONS OF THE DERIVATIVE 197
= 1/8
= 0.125.
The approximate
√ change in f from x = 4 to x = 4.5 is 0.125, so we
approximate 4.5 ≈ 2.125.
quite often. While we don’t discuss here what all of that notation means, note
the existence of the differential dx. Proper handling of integrals comes with
proper handling of differentials.
In light of that, we practice finding differentials in general.
1. y = sin(x) 2. y = 3. yp =
ex x2 + 2 x2 + 3x − 1
Solution.
1. y = sin(x): As f (x) = sin(x), f ′ (x) = cos(x). Thus
dy = cos(x)dx.
2. y = ex x2 + 2 : Let f (x) = ex x2 + 2 . We need f ′ (x), requir-
ing the Theorem 2.4.2.
We have f ′ (x) = ex x2 + 2 + 2xex , so
dy = ex x2 + 2 + 2xex dx.
√ √
3. y = x2 + 3x − 1: Let f (x) = x2 + 3x − 1; we need f ′ (x),
requiring the Theorem 2.5.4.
− 1
We have f ′ (x) = 12 x2 + 3x − 1 2 (2x + 3) = 2√x2x+3
2 +3x−1
.
Thus
(2x + 3)dx
dy = √ .
2 x2 + 3x − 1
How close are f (x) and f (x + ∆x)? This is a difference in “y” values:
or ±1.5%.
We leave it to the reader to confirm this, but if the diameter of the
ball was supposed to be 10 cm, the same manufacturing tolerance would
give a propagated error in mass of ±12.33g, which corresponds to a
percent error of ±0.188%. While the amount of error is much greater
(12.33 > 0.493), the percent error is much lower.
CHAPTER 4. APPLICATIONS OF THE DERIVATIVE 199
4.4.1 Exercises
Terms and Concepts
1. (□ True □ False) Given a differentiable function y = f (x), we are generally free to choose a value for dx,
which then determines the value of dy.
2. (□ True □ False) The symbols “dx” and “∆x” represent the same concept.
3. (□ True □ False) The symbols “dy” and “∆y” represent the same concept.
4. (□ True □ False) Differentials are important in the study of integration.
5. How are differentials and tangent lines related?
6. (□ True □ False) In real life, differentials are used to approximate function values when the function itself
is not known.
Problems
31. A set of plastic spheres are to be made with a diameter of 4 cm. If the manufacturing process is accurate to
2 mm, what is the propagated error in volume of the spheres?
32. The distance, in feet, a stone drops in t seconds is given by d(t) = 16t2 . The depth of a hole is to be approximated
by dropping a rock and listening for it to hit the bottom. What is the propagated error if the time measurement
is accurate to 4/10 of a second and the measured time is:
(a) 4 seconds?
(b) 6 seconds?
33. What is the propagated error in the measurement of the cross sectional area of a circular log if the diameter is
measured at 20′′ , accurate to 1/8′′ ?
34. A wall is to be painted that is 8′ high and is measured to be 13′ , 2′′ long. Find the propagated error in the
measurement of the wall’s surface area if the measurement is accurate to 1/ − 2′′ .
CHAPTER 4. APPLICATIONS OF THE DERIVATIVE 200
Exercise Group. The following 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 computing.)
35. The length L of a long wall is to be 36. The length L of a long wall is to be
approximated. The angle θ, as shown in the approximated. The angle θ, as shown in the
diagram (not to scale), is measured to be 85.2◦ , diagram (not to scale), is measured to be 71.5◦ ,
◦
accurate to 1 . Assume that the triangle formed accurate to 1◦ . Assume that the triangle formed
is a right triangle. is a right triangle.
l=?
l=?
θ
25′ θ
(a) What is the measured length L of the 25′
wall?
(a) What is the measured length L of the
(b) What is the propagated error? wall?
(c) What is the percent error? (b) What is the propagated error?
θ
25′
(a) What is the measured length L of the
wall?
(b) What is the propagated error?
l =?
θ
25′
What is the approximate percent error?
CHAPTER 4. APPLICATIONS OF THE DERIVATIVE 202
x
−4 −2 2 4
y = p1 (x)
f (0) = 2 f ′′′ (0) = −1
f ′ (0) = 1 f (4) (0) = −12
−5 f ′′ (0) = 2 f (5) (0) = −19
Figure 4.5.2 A graph of f (x) and its Figure 4.5.3 Derivatives of f evalu-
tangent line at 0 ated at 0
One shortcoming of this approximation is that the tangent line only matches
the slope of f ; it does not, for instance, match the concavity of f . We can find a
polynomial, p2 (x), that does match the concavity near 0 without much difficulty,
though. The table in Figure 4.5.3 gives the following information:
p(x) = a0 + a1 x + a2 x2 .
p2 (x) = a0 + a1 x + a2 x2 p2 (0) = a0
p2 ′ (x) = a1 + 2a2 x p2 ′ (0) = a1
p2 ′′ (x) = 2a2 p2 ′′ (0) = 2a2 .
p2 (x) = 2 + x + x2 .
match those of f .
How do we ensure that the derivatives of our polynomial match those of f ? x
We simply begin with a polynomial of the desired degree, compute its deriva- −4 −2 2 4
tives, and compare them to those of f ! Recall that each term in a polynomial y = p4 (x)
y = f (x)
consists of a power of x, and a coefficient, like so: an xn . Our goal is to deter-
mine the value for each coefficient an so that the derivatives of our polynomial −5
match those of our function f . If we take k derivatives of the term an xn , with Figure 4.5.4 Plotting f , p2 and p4
k ≤ n, we obtain
dk
(an xn ) = n(n − 1) · · · (n − k + 1)an xn−k .
dxk
For k < n, the expression above vanishes when we set x = 0. However, for
n = k, we obtain the constant value
dk
(ak xk ) = k · (k − 1) · · · 2 · 1ak . (4.5.1)
dxk
Consider a polynomial
pn (x) = a0 + a1 x + · · · + ak xk + · · · + an xn
of degree n. If we take k derivatives, all of the terms involving powers of x less The notation k! is read as “k fac-
than k disappear, and when we set x = 0, all of the terms involving powers torial”. By convention, we also
of x larger than k disappear, leaving us with the single constant given in Equa- define 0! = 1, mostly because
tion (4.5.1). it makes our formulas look a lot
Recalling the notation k! = 1·2·3 · · · k for the product of the first k integers, nicer.
we have shown that
p(k)
n (0) = k!ak .
f ′′ (c)
pn (x) = f (c) + f ′ (c)(x − c) + (x − c)2
2!
f ′′′ (c) f (n) (c)
+ (x − c)3 + · · · + (x − c)n .
3! n!
Solution.
1. We start with creating a table of the derivatives of ex evaluated at
x = 0. In this particular case, this is relatively simple, as shown in [Link]/watch?v=ENf-Z2pLrJg
Figure 4.5.10.
By the definition of the Maclaurin polynomial, we have
f ′′ (c)
pn (x) = f (c) + f ′ (c)(x − c) + (x − c)2 + . . .
2!
f ′′′ (c) f (n) (c)
... (x − c)3 + · · · + (x − c)n
3! n!
0! 1!
= 0 + (x − 1) − (x − 1)2 + . . .
1! 2!
2! (−1)n+1 · (n − 1)!
. . . (x − 1)3 + · · · + (x − 1)n
3! n!
1 1
= (x − 1) − (x − 1)2 + (x − 1)3 − . . . f (x) = ln(x) ⇒ f (1) = 0
2 3
1 (−1) n+1 f ′ (x) = x1 ⇒ f ′ (1) = 1
. . . (x − 1)4 + · · · + (x − 1)n . f ′′ (x) = − x12 ⇒ f ′′ (1) = −1
4 n
f ′′′ (x) = x23 ⇒ f ′′′ (1) = 2
Note how the coefficients of the (x − 1) terms turn out to be
f (4) (x) = − x64 ⇒ f (4) (1) = −6
“nice.”
.. ..
. .
2. We can compute p6 (x) using our work above:
f (n) (x) = ⇒ f (n) (1) =
1 1 (−1) n+1
(n−1)!
p6 (x) = (x − 1) − (x − 1)2 + (x − 1)3 xn (−1)n+1 (n − 1)!
2 3
1 1 1 Figure 4.5.13 Derivatives of ln(x) eval-
− (x − 1) + (x − 1)5 − (x − 1)6 .
4
4 5 6 uated at x = 1
Since p6 (x) approximates ln(x) well near x = 1, we approximate
ln(1.5) ≈ p6 (1.5):
1 1
p6 (1.5) = (1.5 − 1) − (1.5 − 1)2 + (1.5 − 1)3 + . . .
2 3
1 1 1
· · · − (1.5 − 1) + (1.5 − 1)5 − (1.5 − 1)6
4
4 5 6
259
=
640
≈ 0.404688.
This is a good approximation as a calculator shows that ln(1.5) ≈
0.4055. Figure 4.5.14 below plots y = ln(x) with y = p6 (x). We
can see that ln(1.5) ≈ p6 (1.5).
CHAPTER 4. APPLICATIONS OF THE DERIVATIVE 206
−2
y = p6 (x) −4
−6
−4
1. When f (x) is known, but perhaps “hard” to compute directly. For in-
stance, we can define the cosine of an angle as either the ratio of sides
of a right triangle (“adjacent over hypotenuse”) or using the definition in
terms of the unit circle. However, neither of these provides a convenient
way of computing cos(2). A Taylor polynomial of sufficiently high degree
can provide a reasonable method of computing such values using only op-
erations usually hard-wired into a computer (+, −, × and ÷). Even though Taylor polynomials
could be used in calculators and
2. When f (x) is not known, but information about its derivatives is known. computers to calculate values of
This occurs more often than one might think, especially in the study of trigonometric functions, in prac-
differential equations. tice they generally aren’t. Other
In both situations, a critical piece of information to have is “How good is my more efficient and accurate meth-
approximation?” If we use a Taylor polynomial to compute cos(2), how do we ods have been developed, such
know how accurate the approximation is? as the CORDIC algorithm. How-
Although much of the content presented in Calculus concerns the search for ever, understanding how Taylor
exact answers to problems such as integration and differentiation, many practi- polynomials could be used is im-
cal applications of calculus involve attempts to find approximations; for example, portant to developing an under-
using Newton’s Method to approximate the zeros of a function, or numerical in- standing of various approximat-
tegration to approximate the value of an integral that cannot be solved exactly. ing techniques.
CHAPTER 4. APPLICATIONS OF THE DERIVATIVE 207
The first part of Taylor’s 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 approximation. The second part gives bounds on how big that error
can be. If the (n + 1)th derivative is large on I, the error may be large; if x
is far from c, the error may also be large. However, the (n + 1)! term in the
denominator tends to ensure that the error gets smaller as n increases.
The following example computes error estimates for the approximations of
ln(1.5) and ln(2) made in Example 4.5.12.
[Link]/watch?v=2IHECY8dFN0
Example 4.5.18 Finding error bounds of a Taylor polynomial. Figure 4.5.17 Video presentation of The-
orem 4.5.16
Use Theorem 4.5.16 to find error bounds when approximating ln(1.5)
and ln(2) with p6 (x), the Taylor polynomial of degree 6 of f (x) = ln(x)
at x = 1, as calculated in Example 4.5.12.
Solution. Video Solution
1. We start with the approximation of ln(1.5) with p6 (1.5). The the-
orem references an open interval I that contains both x and c.
The smaller the interval we use the better; it will give us a more
accurate (and smaller!) approximation of the error. We let I =
(0.9, 1.6), as this interval contains both c = 1 and x = 1.5. The
theorem references max f (n+1) (z) . In our situation, this is ask-
ing “How big can the 7th derivative of y = ln(x) be on the inter-
val (0.9, 1.6)?” The seventh derivative is y = −6!/x7 . The largest
absolute value it attains on I is about 1506. (There are no criti- [Link]/watch?v=TBV4-X7HoHk
cal numbers of f (7) in the interval so we evaluate the endpoints:
f (7) (0.9) ≈ 1506 and f (7) (1.6) ≈ 27.) In particular, we are eval-
uating at x = 1.5, so we let x = 1.5. Thus we can bound the error
as:
max f (7) (z)
|R6 (1.5)| ≤ (1.5 − 1)7
7!
1506 1
≤ ·
5040 27
CHAPTER 4. APPLICATIONS OF THE DERIVATIVE 208
≈ 0.0023.
We practice again. This time, we use Taylor’s theorem to find n that guaran-
tees our approximation is within a certain amount.
29+1
≈ 0.000282 < 0.001. Thus we want to approximate cos(2)
(9 + 1)!
with p9 (2).
f (x) = cos(x) ⇒ f (0) = 1
We now set out to compute p9 (x). We again need a table of the
derivatives of f (x) = cos(x) evaluated at x = 0. A table of these values f ′ (x) = − sin(x) ⇒ f ′ (0) = 0
is given in Figure 4.5.20. f ′′ (x) = − cos(x) ⇒ f ′′ (0) = −1
Notice how the derivatives, evaluated at x = 0, follow a certain pat- f ′′′ (x) = sin(x) ⇒ f ′′′ (0) = 0
tern. All the odd powers of x in the Taylor polynomial will disappear as f (4) (x) = cos(x) ⇒ f (4) (0) = 1
their coefficient is 0. While our error bounds state that we need p9 (x), f (5) (x) = − sin(x) ⇒ f (5) (0) = 0
our work shows that this will be the same as p8 (x). f (6) (x) = − cos(x) ⇒ f (6) (0) = −1
Since we are forming our polynomial at x = 0, we are creating a f (7) (x) = sin(x) ⇒ f (7) (0) = 0
Maclaurin polynomial, and: f (8) (x) = cos(x) ⇒ f (8) (0) = 1
f (9) (x) = − sin(x) ⇒ f (9) (0) = 0
f ′′ (0) 2 f ′′′ (0) 3 f (8) (0) 8
p8 (x) = f (0) + f ′ (0)x + x + x + ··· + x
2! 3! 8! Figure 4.5.20 A table of the derivatives
1 1 1 1 of f (x) = cos(x) evaluated at x = 0
= 1 − x2 + x4 − x6 + x8 .
2! 4! 6! 8!
We finally approximate cos(2):
y
131
cos(2) ≈ p8 (2) = − ≈ −0.41587. y = p8 (x) 1
315
y = cos(x)
Our error bound guarantee that this approximation is within 0.001
of the correct answer. Technology shows us that our approximation is x
actually within about 0.0003 of the correct answer. −5 −4 −3 −2 −1 1 2 3 4 5
Figure 4.5.21 shows a graph of y = p8 (x) and y = cos(x). Note how
well the two functions agree on about (−π, π).
−1
√
the fifth derivative of f (x) = x takes on this interval is near
x = 2.9, at about 0.0273. (We often graph the (n + 1)th deriva-
tive to find its extrema. In this case is f (5) (x) = 105/(32x9/2 ) is
always decreasing, so the maximum occurs at 2.9.) Thus
0.0273
|R4 (3)| ≤ (3 − 4)5 ≈ 0.00023.
5! y √
y= x
This shows our approximation is accurate to at least the first 2 3
places after the decimal. (It turns out that our approximation is ac- y = p4 (x)
tually accurate to 4 places after the decimal.) A graph of f (x) =
√
x and p4 (x) is given in Figure 4.5.24. Note how the two func- 2
tions are nearly indistinguishable on (2, 7).
1
Our final example gives a brief introduction to using Taylor polynomials to
solve differential equations.
x
Example 4.5.25 Approximating an unknown function. 2 4 6 8 10
√
Figure 4.5.24 A graph of f (x) = x
A function y = f (x) is unknown save for the following two facts.
and its degree 4 Taylor polynomial at
1. y(0) = f (0) = 1, and x=4
2. y ′ = y 2
(This second fact says that amazingly, the derivative of the function
is actually the function squared!)
Find the degree 3 Maclaurin polynomial p3 (x) of y = f (x).
Solution. One might initially think that not enough information is given
to find p3 (x). However, note how the second fact above actually lets us
know what y ′ (0) is:
y ′ = y 2 ⇒ y ′ (0) = y 2 (0).
y′ = y2
d ′ d 2
y = y
dx dx
y ′′ = 2y · y ′ .
We repeat this once more to find y ′′′ (0). We again use implicit dif-
ferentiation; this time the Product Rule is also required.
d ′′ d
y = 2yy ′
dx dx
y ′′′ = 2y ′ · y ′ + 2y · y ′′ .
CHAPTER 4. APPLICATIONS OF THE DERIVATIVE 211
In summary, we have:
Figure 4.5.26 shows this function plotted with p3 (x). Note how similar Figure 4.5.26 A graph of y = −1/(x−
they are near x = 0. 1) and y = p3 (x) from Example 4.5.25
It is beyond the scope of this text to pursue error analysis when using Tay-
lor polynomials to approximate solutions to differential equations. This topic is
often broached in introductory Differential Equations courses and usually cov-
ered in depth in Numerical Analysis courses. Such an analysis is very important;
one needs to know how good their approximation is. We explored this example
simply to demonstrate the usefulness of Taylor polynomials.
We first learned of the derivative in the context of instantaneous rates of
change and slopes of tangent lines. We furthered our understanding of the
power of the derivative by studying how it relates to the graph of a function
(leading to ideas of increasing/decreasing and concavity).
CHAPTER 4. APPLICATIONS OF THE DERIVATIVE 212
4.5.1 Exercises
Terms and Concepts
Problems
Exercise Group. In the following exercises, find the Maclaurin polynomial of degree n for the given function.
5. Find the Maclaurin polynomial of degree n = 3 6. Find the Maclaurin polynomial of degree n = 8
for f (x) = e−x . for f (x) = sin(x).
p3 (x) = p8 (x) =
7. Find the Maclaurin polynomial of degree n = 5 8. Find the Maclaurin polynomial of degree n = 6
for f (x) = x · ex . for f (x) = tan(x).
p5 (x) = p6 (x) =
9. Find the Maclaurin polynomial of degree n = 4 10. Find the Maclaurin polynomial of degree n = 4
for f (x) = e2x . 1
for f (x) = .
p4 (x) = 1−x
p4 (x) =
11. Find the Maclaurin polynomial of degree n = 4 12. Find the Maclaurin polynomial of degree n = 7
1 1
for f (x) = . for f (x) = .
1+x 1+x
p4 (x) = p7 (x) =
Exercise Group. In the following exercises, find the Taylor polynomial of degree n, at x = c, for the given function.
√
13. Find the Taylor polynomial for f (x) = x of 14. Find the degree n = 4 Taylor polynomial for
degree n = 4, at c = 1. f (x) = ln(x + 1), at c = 1.
p4 (x) = p4 (x) =
15. Find the degree n = 6 Taylor polynomial for 16. Find the degree n = 5 Taylor poplynomial for
f (x) = cos(x), at c = π/4. f (x) = sin(x), at c = π/6.
p6 (x) = p5 (x) =
17. Find the degree n = 5 Taylor poplynomial for 18. Find the degree n = 8 Taylor poplynomial for
f (x) = x1 , at c = 2. 1
f (x) = 2 , at c = 1.
p5 (x) = x
p8 (x) =
19. Find the degree n = 3 Taylor poplynomial for 20. Find the degree n = 2 Taylor polynomial for
1 f (x) = x2 cos(x), at c = π.
f (x) = 2 , at c = −1.
x +1 p2 (x) =
p3 (x) =
Exercise Group. In the following exercises, approximate the function value with the indicated Taylor polynomial and
give approximate bounds on the error.
21. Approximate sin(0.1) with the Maclaurin 22. Approximate cos(1) with the Maclaurin
polynomial of degree 3. polynomial of degree 4.
CHAPTER 4. APPLICATIONS OF THE DERIVATIVE 213
√
23. Approximate 10 with the Taylor polynomial of 24. Approximate ln(1.5) with the Taylor polynomial
degree 2 centered at x = 9. of degree 3 centered at x = 1.
Exercise Group. The following exercises ask for an n to be found such that pn (x) approximates f (x) within a certain
bound of accuracy.
25. Find n such that the Maclaurin polynomial of 26. Find n such that the √ Taylor polynomial of
degree n of f (x) = ex approximates e within degree n of f (x)
√ = x, centered at x = 4,
0.0001 of the actual value. approximates 3 within 0.0001 of the actual
value.
27. Find n such that the Maclaurin polynomial of 28. Find n such that the Maclaurin polynomial of
degree n of f (x) = cos(x) approximates degree n of f (x) = sin(x) approximates cos(π)
cos(π/3) within 0.0001 of the actual value. within 0.0001 of the actual value.
Exercise Group. In the following exercises, find the nth term of the indicated Taylor polynomial.
29. Find a formula for the nth term of the 30. Find a formula for the nth term of the
Maclaurin polynomial for f (x) = ex . Maclaurin polynomial for f (x) = cos(x).
31. Find a formula for the nth term of the 32. Find a formula for the nth term of the
Maclaurin polynomial for f (x) = sin x. 1
Maclaurin polynomial for f (x) = .
1−x
33. Find a formula for the nth term of the 34. Find a formula for the nth term of the Taylor
1 polynomial for f (x) = ln(x) centered at x = 1.
Maclaurin polynomial for f (x) = .
1+x
Exercise Group. In the following exercises, approximate the solution to the given differential equation with a degree
4 Maclaurin polynomial.
35. y ′ = y, y(0) = 1 36. y ′ = 5y, y(0) = 3
2
37. y ′ = , y(0) = 1
y
CHAPTER 4. APPLICATIONS OF THE DERIVATIVE 214
Solution.
[Link]/watch?v=Y2O3RD9tt34
CHAPTER 4. APPLICATIONS OF THE DERIVATIVE 215
2. √
x+3−2 by LHR
1
2 (x + 3)−1/2 1
lim = lim =− .
x→1 1−x x→1 −1 4
3.
x2 by LHR 2x
lim = lim .
x→0 1 − cos(x) x→0 sin(x)
4. We already know how to evaluate this limit; first factor the numer-
ator and denominator. We then have:
x2 + x − 6 (x − 2)(x + 3) x+3
lim = lim = lim = 5.
x→2 x − 3x + 2
2 x→2 (x − 2)(x − 1) x→2 x − 1
x2 + x − 6 by LHR 2x + 1
lim = lim = 5.
x→2 x2 − 3x + 2 x→2 2x − 3
Note that at each step where l’Hospital’s Rule was applied, it was needed:
the initial limit returned the indeterminate form of “0/0.” If the initial limit re-
turns, for example, 1/2, then l’Hospital’s Rule does not apply.
The following theorem extends our initial version of l’Hospital’s Rule in two
ways. It allows the technique to be applied to the indeterminate form ∞/∞
and to limits where x approaches ±∞.
1. Let lim f (x) = ±∞ and lim g(x) = ±∞, where f and g are
x→a x→a
differentiable on an open interval I containing a. If
f ′ (x)
lim = L,
x→a g ′ (x)
then
f (x)
lim = L,
x→a g(x)
f ′ (x)
lim = L,
x→∞ g ′ (x)
then
f (x)
lim = L,
x→∞ g(x)
where L is a real number, or L = ±∞. A similar statement can
be made for limits where x approaches −∞.
Recall that this means that the limit does not exist; as x approaches
∞, the expression ex /x3 grows without bound. We can infer from
this that ex grows “faster” than x3 ; as x gets large, ex is far larger
than x3 . (This has important implications in computing when con-
sidering efficiency of algorithms.)
Solution.
[Link]/watch?v=wJzKupOv8cg
CHAPTER 4. APPLICATIONS OF THE DERIVATIVE 217
lim x · e1/x = 0.
x→0−
Thus
x+1
lim ln(x + 1) − ln(x) = lim ln = 0.
x→∞ x→∞ x
Solution.
1. This is equivalent to a special limit given in Theorem 1.3.17; these
limits have important applications within mathematics and finance.
Note that the exponent approaches ∞ while the base approaches
1, leading to the indeterminate form 1∞ . Let f (x) = (1 + 1/x)x ;
the problem asks to evaluate lim f (x). Let’s first evaluate lim ln f (x) .
x→∞ x→∞
x [Link]/watch?v=wHCd7Wsxzug
1
lim ln f (x) = lim ln 1 +
x→∞ x→∞ x
1
= lim x ln 1 +
x→∞ x
1
ln 1 + x
= lim
x→∞ 1/x
= 1.
Thus lim ln f (x) = 1. We return to the original limit and apply
x→∞
Key Idea 4.6.7.
x
1
lim 1 + = lim f (x) = lim eln(f (x)) = e1 = e.
x→∞ x x→∞ x→∞
CHAPTER 4. APPLICATIONS OF THE DERIVATIVE 219
1
Our brief revisit of limits will be rewarded in the next section where we con- f (x) = xx
sider improper integration. So far, we have only Z considered definite integrals x
1
0.5 1 1.5 2
where the bounds are finite numbers, such as f (x) dx. Improper integra-
0
tion considers integrals where one, or both, of the bounds are “infinity.” Such Figure 4.6.9 A graph of f (x) = xx
integrals have many uses and applications, in addition to generating ideas that supporting the fact that as x → 0+ ,
are enlightening. f (x) → 1
CHAPTER 4. APPLICATIONS OF THE DERIVATIVE 220
4.6.4 Exercises
Terms and Concepts
Problems
17. 18.
ex − 1 ex − x − 1
lim lim
x→0+ x2 x→0+ x2
19. 20.
x − sin(x) x4
lim lim
x→0+ x3 − x2 x→∞ ex
21. √ 22.
x ex
lim lim
x→∞ ex x→∞ x2
23. 24.
ex ex
lim √ lim
x→∞ x x→∞ 2x
25. 26.
ex x3 − 5x2 + 3x + 9
lim x lim
x→∞ 3 x→3 x3 − 7x2 + 15x − 9
27. 28.
x3 + 4x2 + 4x ln(x)
lim 3 lim
x→−2 x + 7x2 + 16x + 12 x→∞ x
29. 30.
ln x2 ln2 (x)
lim lim
x→∞ x x→∞ x
CHAPTER 4. APPLICATIONS OF THE DERIVATIVE 221
31. 32. √
lim+ x · ln(x) lim x · ln(x)
x→0 x→0+
33. 34.
1
lim+ x · e x lim x3 − x2
x→0 x→∞
35. √ 36.
lim x − ln(x) lim x · ex
x→∞ x→−∞
37. 38.
1 −1 1
lim 2 · e x lim (1 + x) x
x→0 + x x→0+
39. 40.
lim+ (2x)x 2
x→0
lim ( )x
x→0+ x
41. 42.
lim (sin(x))x lim (1 − x)1−x
x→0+ x→1−
49. 50.
lim tan(x)sin(2x) 1 1
lim+ −
x→π/2 x→1 ln(x) x − 1
51. 52.
5 x 1
lim − lim xtan
x→3+ x2 − 9 x−3 x→∞ x
53. 54.
ln3 (x) x2 + x − 2
lim lim
x→∞ x x→1 ln(x)
CHAPTER 4. APPLICATIONS OF THE DERIVATIVE 222
Integration
G(x) = F (x) + C.
[Link]/watch?v=W-FUL0ApGL8
CHAPTER 5. INTEGRATION 225
integration. So: Z
sin(x) dx = − cos(x) + C.
This is asking: “What functions have a differential of the form dy?” The an-
swer is “Functions of the form y+C, where C is a constant.” What is y? We have
lots of choices, all differing by a constant; the simplest choice is y = − cos(x).
Understanding all of this is more important later as we try to find antideriv-
atives of more complicated functions. In this section, we will simply explore
the rules of indefinite integration, and one can succeed for now with answering
“What happened to the dx?” with “It went away.”
Let’s practice once more before stating integration rules.
This final step of “verifying our answer” is important both practically and
theoretically. In general, taking derivatives is easier than finding antiderivatives
so checking our work is easy and vital as we learn.
We also see that taking the derivative of our answer returns the function in
the integrand. Thus we can say that:
Z
d
f (x) dx = f (x).
dx
Z Z
d
(cf (x)) = c · f ′ (x) c · f (x) dx = c ·f (x) dx
dx
Z Z Z
d
(f (x) ± g(x)) = f ′ (x) ± g ′ (x) f (x) ± g(x) dx = f (x) dx ± g(x) dx
dx
Z
d
(C) = 0 0 dx = C
dx
Z Z
d
(x) = 1 1 dx = dx = x + C
dx
Z
d n 1
(x ) = n · xn−1 xn dx = xn+1 + C (n ̸= −1)
dx n+1
Z
d
(sin(x)) = cos(x) cos(x) dx = sin(x) + C
dx
Z
d
(cos(x)) = − sin(x) sin(x) dx = − cos(x) + C
dx
Z
d
(tan(x)) = sec2 (x) sec2 (x) dx = tan(x) + C
dx
Z
d
(csc(x)) = − csc(x) cot(x) csc(x) cot(x) dx = − csc(x) + C
dx
Z
d
(sec(x)) = sec(x) tan(x) sec(x) tan(x) dx = sec(x) + C
dx
Z
d
(cot(x)) = − csc2 (x) csc2 (x) dx = − cot(x) + C
dx
Z
d x
(e ) = ex ex dx = ex + C
dx
Z
d x 1
(a ) = ln(a) · ax ax dx = · ax + C
dx ln(a)
Z
d 1 1
(ln(x)) = , x > 0 dx = ln |x| + C
dx x x
• Z Z
c · f (x) dx = c · f (x) dx
In the last step we can consider the constant as also being multiplied by 5,
but “5 times a constant” is still a constant, so we just write “C”.
• Z Z Z
f (x) ± g(x) dx = f (x) dx ± g(x) dx
This is the Sum/Difference Rule: we can split integrals apart when the
integrand contains terms that are added/subtracted, as we did in Exam-
ple 5.1.7. So:
Z Z Z Z
2 2
(3x + 4x + 5) dx = 3x dx + 4x dx + 5 dx
Z Z Z
= 3 x2 dx + 4 x dx + 5 dx
1 1
= 3 · x3 + 4 · x2 + 5x + C
3 2
3 2
= x + 2x + 5x + C
In practice we generally do not write out all these steps, but we demon-
strate them here for completeness.
• Z
1
xn dx = xn+1 + C (n ̸= −1)
n+1
This is the Power Rule of indefinite integration. There are two important
things to keep in mind:
R
1. Notice the restriction that n ̸= −1. This is important: x1 dx ̸=
“ 01 x0 + C”; rather, see the last rule from the list.
2. We are presenting antidifferentiation as the “inverse operation” of
differentiation. Here is a useful quote to remember:
“Inverse operations do the opposite things in the opposite
order.”
When taking a derivative using the Power Rule, we first multiply by
the power, then second subtract 1 from the power. To find the anti-
derivative, do the opposite things in the opposite order: first add 1
to the power, then second divide by the power.
• Z
1
dx = ln |x| + C
x
Note that this rule uses 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.
CHAPTER 5. INTEGRATION 228
v(t) = −32t + C
v(3) = −10
−32(3) + C = −10
C = 86
Thus v(t) = −32t + 86. We can use this equation to understand the
motion of the object: when t = 0, the object had a velocity of v(0) =
86 fts . Since the velocity is positive, the object was moving upward.
When did the object begin moving down? Immediately after v(t) =
0:
43
−32t + 86 = 0 =⇒ t = ≈ 2.69s.
16
Recognize that we are able to determine quite a bit about the path of
the object knowing just its acceleration and its velocity at a single point
in time.
[Link]/watch?v=MB1dLY4lOew
CHAPTER 5. INTEGRATION 229
= sin(t) + C
= f ′ (t).
sin(0) + C = 3
C = 3.
Using the initial value, we have found f ′ (t) = sin(t)+3. We now find
f (t) by integrating again. We will use a different integration constant
since we have already defined C to equal 3 above.
Z Z
f (t) = f ′ (t) dt = (sin(t) + 3) dt = − cos(t) + 3t + D.
− cos(0) + 3(0) + D = 5
−1 + C = 5
C=6
5.1.1 Exercises
Terms and Concepts
Problems
Exercise Group. Evaluate the indefinite integral. Don’t forget your constant of integration!
R 7 R 9
9. 8x dx 10. x dx
R 2
R
11. 5x + 3 dx 12. dt
R R 1
13. 1 ds 14. 5t9 dt
R 6 R 1
15. 16. √ dx
t7 dt x
R R
17. sec(θ) tan(θ) dθ 18. sin(θ) dθ
R R θ
19. (sec(x) tan(x) − csc(x) cot(x)) dx 20. 2e dθ
R t R 4t
21. 3 dt 22. 3 dt
R R 4
23.
2
(5t + 6) dt 24. t − 1 t5 − 8t dt
R 6 2 R
25. x x dx 26. 1.41421e dx
R
27. r dx
Z Z
28. Consider the two integrals, s ds and sn dn.
n
Exercise Group. Find the function determined by the given initial value problem.
30. f ′ (x) = sin(x) and f (0) = −8
31. f ′ (x) = 2ex and f (0) = 10
32. f ′ (x) = 3x3 − 2x2 and f (2) = 5
33. f ′ (x) = sec(x) tan(x) and f π3 = −8
34. f ′ (x) = 5x and f (2) = 7
35. f ′′ (x) = 5 and f ′ (0) = 4, f (0) = 9
36. f ′′ (x) = 3x and f ′ (1) = −10, f (1) = −5
37. f ′′ (x) = 7ex and f ′ (0) = 4, f (0) = 5
38. f ′′ (θ) = cos(θ) and f ′ (0) = 2, f (0) = 9
39. f ′′ (x) = 29x4 + 6x + cos(x) and f ′ (0) = −3, f (0) = −1
40. f ′′ (x) = 0 and f ′ (−5) = 3, f (−5) = 8
CHAPTER 5. INTEGRATION 232
v (ft/s)
5
Now consider a slightly harder situation (and not particularly realistic): an
object travels in a straight line with a constant velocity of 5 fts for 10 seconds,
then instantly reverses course at a rate of 2 fts for 4 seconds. (Since the object is
traveling in the opposite direction when reversing course, we say the velocity is
a constant −2 fts .) How far away from the starting point is the object — what is
its displacement?
Here we use “Distance = Rate1 × Time1 + Rate2 × Time2 ,” which is
2 4 6 8 10
Distance = 5 · 10 + (−2) · 4 = 42 ft. t (s)
Hence the object is 42 feet from its starting location. Figure 5.2.2 The area under a constant
We can again depict this situation graphically. In Figure 5.2.3 we have the velocity function corresponds to dis-
velocities graphed as straight lines on [0, 10] and [10, 14], respectively. The dis- tance traveled
placement of the object is
“Area above the t-axis −Area below the t-axis,”
which is easy to calculate as 50 − 8 = 42 feet.
v (ft/s)
5
Now consider a more difficult problem.
v(0) = −32 · 0 + 48
= 48 Video Solution
Z
= (−32t + 48) dt
= −16t2 + 48t + C.
0 = −32t + 48
t = 48/32
= 1.5 s .
(Notice how we ended up just finding when the velocity was 0ft/s!)
The first derivative test shows this is a maximum, so the maximum height
of the object is found at
The above example does not prove a relationship between area under a ve-
locity function and displacement, but it does imply a relationship exists. Sec-
tion 5.4 will fully establish fact that the area under a velocity function is dis-
placement.
Given a graph of a function y = f (x), we will find that there is great use
in computing the area between the curve y = f (x) and the x-axis. Because of
this, we need to define some terms.
Let y = f (x) be defined on a closed interval [a, b]. The total signed area
from x = a to x = b under f is:
(area under y = f (x) and above the x-axis on [a, b]) − (area above
y = f (x) and under the x-axis on [a, b]).
The definite integral of f on [a, b] is the total signed area of f on
[a, b], denoted
Z b
f (x) dx,
a
By our definition, the definite integral gives the “signed area under f .” We
usually drop the word “signed” when talking about the definite integral, and
simply say the definite integral gives “the area under f ” or, more commonly,
“the area under the curve.” [Link]/watch?v=1kJUMKdjumQ
The previous section introduced the indefinite integral, which related to an- Figure 5.2.7 Video presentation of De-
tiderivatives. We have now defined the definite integral, which relates to areas finition 5.2.6
under a function. The two are very much related, as we’ll see when we learn
the Fundamental
R Theorem of Calculus in Section 5.4. Recall that earlier we said
that the “ ” symbol was an “elongated S” that represented finding a “sum.” In
the context of the definite integral, this notation makes a bit more sense, as we
are adding up areas under the function f . 1 y
−1
Z 3 Z 3
1. f (x) dx 4. 5f (x) dx
0 0
Z 5 Z 1
2. f (x) dx 5. f (x) dx
3 1
Z 5
3. f (x) dx
0
Video Solution
Solution.
R3
1. 0 f (x) dx is the area under f on the interval [0, 3]. This region is
R3
a triangle, so the area is 0 f (x) dx = 12 (3)(1) = 1.5.
R5
2. 3 f (x) dx represents the area of the triangle found under the x-
axis on [3, 5]. The area is 21 (2)(1) = 1; since it is found under the
R5
x-axis, this is “negative area.” Therefore 3 f (x) dx = −1. [Link]/watch?v=jrjjVT1j9uw
R5
3. 0 f (x) dx is the total signed area under f on [0, 5]. This is 1.5 +
(−1) = 0.5.
R3
4. 0 5f (x) dx is the area under 5f on [0, 3]. This is sketched in Fig-
ure 5.2.10. Again, the region is a triangle, with height R 35 times that y
of the height of the original triangle. Thus the area is 0 5f (x) dx =
1 4
2 (15)(1) = 7.5.
R1 2
5. 1 f (x) dx is the area under f on the “interval” [1, 1]. This de-
scribes a line segment, not a region; it has no width. Therefore x
the area is 0. 1 2 3 4 5
−2
This example illustrates some of the properties of the definite integral, given
−4
here.
Theorem 5.2.11 Properties of the Definite Integral. Figure 5.2.10 A graph of 5f in Exam-
Let f and g be defined on a closed interval I that contains the values a, ple 5.2.8. (Yes, it looks just like the
b and c, and let k be a constant. The following hold: graph of f in Figure 5.2.9, just with a
different y-scale.)
Z a
1. f (x) dx = 0
a
Z b Z c Z c
2. f (x) dx + f (x) dx = f (x) dx
a b a
Z b Z a
3. f (x) dx = − f (x) dx
a b
Z Z Z
b b b
4. f (x) ± g(x) dx = f (x) dx ± g(x) dx
a a a
Z b Z b
5. k · f (x) dx = k · f (x) dx
a a
[Link]/watch?v=sK5vZ_QrkNk
Figure 5.2.12 Video presentation of The-
orem 5.2.11
CHAPTER 5. INTEGRATION 236
Solution.
Rb
1. a f (x) dx has a positive value (since the area is above the x-axis)
Rc Rb
whereas b f (x) dx has a negative value. Hence a f (x) dx is big-
ger.
Rc
2. a f (x) dx is the total signed area under f between x = a and
x = c. Since the region below the x-axis looks to be larger than
the region above, we conclude that the definite integral has a value
less than 0.
3. Note how the second integral has the bounds “reversed.” There-
Rb Rc
fore c f (x) dx = − b f (x) dx represents a positive number,
greater than the area described by the first definite integral. Hence
Rb
c
f (x) dx is greater.
The area definition of the definite integral allows us to use geometry to com-
pute the definite integral of some simple functions.
area is: Z 3 p 1 9
9 − x2 dx = πr2 = π.
−3 2 2
10 y y
5
(5, 6)
5
R2
x
−2 2 4
R1
−5
(−2, −8) x
−10 −3 3
v (ft/s)
Figure 5.2.17 f (x) =
10
38
Example 5.2.18 Understanding motion given velocity. 5
line, given in Figure 5.2.19, where the numbers in the given regions gives a b c
the area of that region. Assume that the definite integral of a velocity 11 11
−5
function gives displacement. Find the maximum speed of the object and
its maximum displacement from its starting position.
Figure 5.2.19 A graph of a velocity in
Solution. Since the graph gives velocity, finding the maximum speed is
Example 5.2.18
simple: it looks to be 15ft/s.
At time t = 0, the displacement is 0; the object is at its starting po-
sition. At time t = a, the object has moved backward 11 feet. Between Video Solution
times t = a and t = b, the object moves forward 38 feet, bringing it into
a position 27 feet forward of its starting position. From t = b to t = c
the object is moving backwards again, hence its maximum displacement
is 27 feet from its starting position.
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 5.2.20, where a region below y = x2 is shaded.
What is its area? The function y = x2 is relatively simple, yet the shape it defines [Link]/watch?v=2zJzbg0hNXE
has an area that is not simple to find geometrically.
In Section 5.3 we will explore how to find the areas of such regions. y
10
x
1 2 3
5.2.1 Exercises
Terms and Concepts
Problems
Exercise Group. A graph of a function f (x) is given. Using the geometry of the graph, evaluate the definite integrals.
5. 6.
4 y 2 y
y = −2x + 4
2 1
x x
1 2 3 4 1 2 3 4 5
−2 −1
y = f (x)
−4 −2
R1 R2
(a) 0
(−2x + 4) dx (a) 0
f (x) dx
R2 R3
(b) 0
(−2x + 4) dx (b) 0
f (x) dx
R3 R5
(c) 0
(−2x + 4) dx (c) 0
f (x) dx
R3 R5
(d) 1
(−2x + 4) dx (d) 2
f (x) dx
R4 R3
(e) 2
(−2x + 4) dx (e) 5
f (x) dx
R1 R3
(f) 0
(−6x + 12) d (f) 0
−2f (x) dx
CHAPTER 5. INTEGRATION 240
7. 8.
y y
4 3
y =x−1
3 2
y = f (x)
2 1
x
1
1 2 3 4
x
−1
1 2 3 4
R2 R1
(a) 0
f (x) dx (a) 0
(x − 1) dx
R4 R2
(b) 2
f (x) dx (b) 0
(x − 1) dx
R4 R3
(c) 2
2f (x) dx (c) 0
(x − 1) dx
R1 R3
(d) 0
4x dx (d) 2
(x − 1) dx
R3 R4
(e) 2
(2x − 4) dx (e) 1
(x − 1) dx
R3 R4
(f) 2
(4x − 8) dx (f) 1
(x − 1) + 1 dx
9. 10.
y y
3
p
f (x) = 4 − (x − 2)2 f (x) = 3
2 3
2
1
1
x x
1 2 3 4 2 4 6 8 10
R2 R5
(a) f (x) dx (a) 0
f (x) dx
0
R4 R7
(b) f (x) dx (b) 3
f (x) dx
2
R4 R0
(c) f (x) dx (c) 0
f (x) dx
0
R4 Z b
(d) 0
5f (x) dx (d) f (x) dx, where 0 ≤ a ≤ b ≤ 10
a
Exercise Group. A graph of a function f (x) is given; the numbers inside the shaded regions give the area of that
region. Evaluate the definite integrals using this area information.
CHAPTER 5. INTEGRATION 241
11. 12.
50 y y
y = f (x) 1
f (x) = sin(πx/2)
11 21 x 4/π
59 1 2 3
x
1 2 3 4
−50
4/π
−1
−100
R1 R2
(a) f (x) dx (a) 0
f (x) dx
0
R2 R4
(b) f (x) dx (b) 2
f (x) dx
0
R3 R4
(c) f (x) dx (c) 0
f (x) dx
0
R2 R1
(d) −3f (x) dx (d) 0
f (x) dx
1
13. 14.
10
y f (x) = 3x2 − 3 4 y
3 f (x) = x2
5
2
4 4 x
−2 −1 −4 1 2 1
1/3 7/3 x
−5
1 2
R −1
(a) −2
f (x) dx R2
(a) 5x2 dx
R2 0
(b) 1
f (x) dx R2
(b) (x2 + 3) dx
R1 0
(c) −1
f (x) dx R3
(c) (x − 1)2 dx
R1 1
(d) f (x) dx R4
0
(d) 2
(x − 2)2 + 5 dx
Exercise Group. A graph is given of the velocity function of an object moving in a straight line. Answer the questions
based on the graph.
CHAPTER 5. INTEGRATION 242
15. 16.
3 4
y (ft/s) y (ft/s)
2 3
2
1
t (s) 1
1 2 3
t (s)
−1 1 2 3 4 5
(a) What is the object’s maximum velocity? (a) What is the object’s maximum velocity?
(b) What is the object’s maximum (b) What is the object’s maximum
displacement? displacement?
(c) What is the object’s total displacement (c) What is the object’s total displacement
on [0, 3]? on [0, 5]?
17. An object is thrown straight up with a velocity, in ft/s, given by v(t) = −32t + 64, where t is in seconds, from a
height of 48 feet.
(a) What is the object’s maximum velocity?
Use these values and properties of definite integrals to evaluate the indicated definite integral.
R2 R3
19. 0
f (x) + g(x) dx 20. 0
f (x) − g(x) dx
R3 22. Find
21. 3f (x) + 2g(x) dx Z 3 a formula for a in terms of b such that
2
af (x) + bg(x) dx = 0.
0
Exercise Group. The values of several definite integrals are given as follows:
Z 3 Z 5 Z 5 Z 5
s(t) dt = 10 s(t) dt = 8 r(t) dt = −1 r(t) dt = 11
0 3 3 0
CHAPTER 5. INTEGRATION 243
Use these values and properties of definite integrals to evaluate the indicated definite integral.
R3 R0
23. 0
s(t) + r(t) dt 24. 5
s(t) − r(t) dt
R3 26. Z Find a formula for a in terms of b such that
25. πs(t) − 7r(t) dt
3 5
ar(t) + bs(t) dt = 0.
0
CHAPTER 5. INTEGRATION 244
1
RHR MPR LHR other x
1 2 3 4
R4
Figure 5.3.4 Approximating 0 (4x −
x2 ) dx using rectangles. The heights
of the rectangles are determined us-
ing different rules.
CHAPTER 5. INTEGRATION 245
Example 5.3.5 Using the Left Hand, Right Hand and Midpoint Rules.
R4
Approximate the value of 0 (4x − x2 ) dx using the Left Hand Rule, the
Right Hand Rule, and the Midpoint Rule, using 4 equally spaced subin- Video Solution
tervals.
Solution. We break the interval [0, 4] into four subintervals as before.
In Figure 5.3.6(a) we see 4 rectangles drawn on f (x) = 4x − x2 using
the Left Hand Rule. (The areas of the rectangles are given in each figure.)
Note how in the first subinterval, [0, 1], the rectangle has height f (0) =
0. We add up the areas of each rectangle (height× width) for our Left
Hand Rule approximation:
y y y
4 4 4
3 3 3
2 4 2 4 2 3.75 3.75
3 3 3 3
1 1 1 1.75 1.75
0 x 0 x x
1 2 3 4 1 2 3 4 1 2 3 4
(a) using the Left Hand (b) using the Right (c) using the Midpoint
Rule Hand Rule Rule
R4
Figure 5.3.6 Approximating 0 (4x − x2 ) dx in Example 5.3.5
CHAPTER 5. INTEGRATION 246
X9
ai (5.3.1)
|{z}
i| =
{z 1} summand
i-index of summation
X
7
2. (3ai − 4)
i=3
X
4
3. (ai )2 Video Solution
i=1
Solution.
1.
X
6
ai = a1 + a2 + a3 + a4 + a5 + a6
i=1
= 1 + 3 + 5 + 7 + 9 + 11
= 36. [Link]/watch?v=GKaRI_a96-Q
X
7
(3ai − 4) = (3a3 − 4) + (3a4 − 4) + (3a5 − 4) + (3a6 − 4) + (3a7 − 4)
i=3
= 11 + 17 + 23 + 29 + 35
= 115.
CHAPTER 5. INTEGRATION 247
3.
X
4
(ai )2 = (a1 )2 + (a2 )2 + (a3 )2 + (a4 )2
i=1
= 1 2 + 32 + 52 + 72
= 84.
It might seem odd to stress a new, concise way of writing summations only
to write each term out as we add them up. It is. The following theorem gives
some of the properties of summations that allow us to work with them without
writing individual terms. Examples will follow.
X
n
X
j X
n X
n
1. c = c · n, where c is a 4. ai + ai = ai
i=1 i=m i=j+1 i=m
constant.
X
n
n(n + 1)
X
n X
n
5. i=
2. (ai ± bi ) = ai ± 2
i=1
i=m i=m
Xn
X
n
n(n + 1)(2n + 1)
bi 6. i2 =
i=m i=1
6
X
n X
n X
n 2
n(n + 1)
3. c · ai = c · ai 7. 3
i =
i=m i=m i=1
2
Solution.
X
6 X
6 X
6
(2i − 1) = 2i − (1)
i=1 i=1 i=1
!
X
6
= 2 i −6
i=1
6(6 + 1)
=2 −6
2
= 42 − 6 = 36
We obtained the same answer without writing out all six terms. When
dealing with small sizes of n, it may be faster to write the terms out
by hand. However, Theorem 5.3.9 is incredibly important when dealing
with large sums as we’ll soon see.
CHAPTER 5. INTEGRATION 248
xi = x0 + i∆x
starting subinterval
value size
So x9 = x0 + 9(4/16) = 2.25.
If we had partitioned [0, 4] into 100 equally spaced subintervals, each subin-
terval would have length ∆x = 4/100 = 0.04. We could compute x31 as
We use these formulas in the next two examples. The following example lets
us practice using the Right Hand Rule and the summation formulas introduced
in Theorem 5.3.9.
[Link]/watch?v=urkFFBmu9uQ
CHAPTER 5. INTEGRATION 249
gral as
X
16
f (xi )∆x.
i=1
We have ∆x = 4/16 = 0.25. Since xi = 0 + i∆x, we have
xi = 0 + i∆x = i∆x.
Using the summation formulas, consider:
Z 4 X
16
(4x − x2 ) dx ≈ f (xi )∆x
0 i=1
X
16
= f (i∆x)∆x
i=1
X
16
= 4i∆x − (i∆x)2 ∆x
i=1
X
16
= (4i∆x2 − i2 ∆x3 )
i=1
X
16 X
16
= (4∆x2 ) i − ∆x3 i2 (5.3.2)
i=1 i=1
16 · 17 16(17)(33)
= (4∆x2 ) − ∆x3
2 6
= 4 · 0.252 · 136 − 0.253 · 1496
= 10.625
We were able to sum up the areas of 16 rectangles with very lit-
tle computation. In Figure 5.3.14 the function and the 16 rectangles y
are graphed. While some rectangles over-approximate the area, other
4
under-approximate the area (by about the same amount). Thus our ap-
proximate area of 10.625 is likely a fairly good approximation. 3
Notice Equation (5.3.2); by replacing 16 by 1,000 (and appropriately
changing the value of ∆x), we can use that equation to sum up 1000 2
rectangles!
We do so here, skipping from the original summand to the equivalent 1
of Equation (5.3.2) to save space. Note that ∆x = 4/1000 = 0.004. x
Z 4 X
1000 1 2 3 4
RUsing
4
many, many rectangles, we have a likely good approximation
of 0
(4x − x2 )∆x. That is,
Z 4
(4x − x2 ) dx ≈ 10.666656.
0
CHAPTER 5. INTEGRATION 250
Before the above example, we stated what the summations for the Left Hand,
Right Hand and Midpoint Rules looked like. Each had the same basic structure,
which was:
1. each rectangle has the same width, which we referred to as ∆x, and
2. each rectangle’s height is determined by evaluating f at a particular point
in each subinterval. For instance, the Left Hand Rule states that each rec-
tangle’s height is determined by evaluating f at the left hand endpoint of
the subinterval the rectangle lives on.
One could partition an interval [a, b] with subintervals that do not have the
same size. We refer to the length of the ith subinterval as ∆xi . Also, one could
determine each rectangle’s height by evaluating f at any point ci in the ith subin-
terval. Thus the height of the ith subinterval would be f (ci ), and the area of the
ith rectangle would be f (ci )∆xi . These ideas are formally defined below.
Summations of rectangles with area f (ci )∆xi are named after mathemati-
cian Georg Friedrich Bernhard Riemann, as given in the following definition.
Key Idea 5.3.20 Riemann Sum Concepts. Figure 5.3.19 An example of a Rgeneral
4
Riemann sum to approximate 0 (4x−
Z b X
n 2
x ) dx
Consider f (x) dx ≈ f (ci )∆xi .
a i=1
CHAPTER 5. INTEGRATION 251
b−a
1. When the n subintervals have equal length, ∆xi = ∆x = .
n
2. The ith term of an equally spaced partition is xi = a + i∆x. (Thus
x0 = a and xn = b.)
X
n
3. The Left Hand Rule summation is: f (xi−1 )∆x.
i=1
X
n
4. The Right Hand Rule summation is: f (xi )∆x.
i=1
Xn
xi−1 + xi
5. The Midpoint Rule summation is: f ∆x.
i=1
2
Note the graph of f (x) = 5x+2 in Figure 5.3.22. The regions whose
area is computed by the definite integral are triangles, meaning we can
find the exact answer without summation techniques. We find that the
exact answer is indeed 22.5. One of the strengths of the Midpoint Rule is
that often each rectangle includes area that should not be counted, but
misses other area that should. When the partition size is small, these
two amounts are about equal and these errors almost “cancel each other
out.” In this example, since our function 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 function is negative, the rectangles have a
“negative” height. When we compute the area of the rectangle, we use
f (ci )∆x; when f is negative, the area is counted as negative.
Notice in the previous example that while we used 10 equally spaced inter-
vals, the number “10” didn’t play a big role in the calculations until the very end.
Mathematicians love to abstract ideas; let’s approximate the area of another re-
gion using n subintervals, where we do not specify a value of n until the very
end.
X
n
= f (−1 + i∆x)∆x
i=1
Xn
= (−1 + i∆x)3 ∆x
i=1
X
n
= (i∆x)3 − 3(i∆x)2 + 3i∆x − 1 ∆x (now distribute ∆x)
i=1
X
n
= i3 ∆x4 − 3i2 ∆x3 + 3i∆x2 − ∆x (now split up summation)
i=1
X
n X
n X
n X
n
= ∆x4 i3 − 3∆x3 i2 + 3∆x2 i− ∆x
i=1 i=1 i=1 i=1
2
n(n + 1) n(n + 1)(2n + 1) n(n + 1)
= ∆x 4
− 3∆x3 + 3∆x2 − n∆x
2 6 2
(use ∆x = 6/n)
X
n
• SL (n) = f (xi−1 )∆x, the sum of equally spaced rectangles formed
i=1
using the Left Hand Rule,
X
n
• SR (n) = f (xi )∆x, the sum of equally spaced rectangles formed us-
i=1
ing the Right Hand Rule, and
Xn
xi−1 + xi
• SM (n) = f ∆x, the sum of equally spaced rectangles
i=1
2
formed using the Midpoint Rule.
The following theorem states that we can use any of our three rules to find
Rb
the exact value of a definite integral a f (x) dx. It also goes two steps further.
The theorem states that the height of each rectangle doesn’t have to be de-
termined following a specific rule, but could be f (ci ), where ci is any point in
the ith subinterval, as discussed before Riemann Sums were defined in Defini-
tion 5.3.17.
The theorem goes on to state that the rectangles do not need to be of the
same width. Using the notation of Definition 5.3.15, let ∆xi denote the length
of the ith subinterval in a partition of [a, b] and let ∥∆x∥ represent the length
of the largest subinterval in the partition: that is, ∥∆x∥ is the largest of all the
∆xi . If ∥∆x∥ is small, then [a, b] must be partitioned 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 partition is growing to
infinity, as the largest subinterval length is becoming arbitrarily small. We then
interpret the expression
X
n
lim f (ci )∆xi
∥∆x∥→0
i=1
as “the limit of the sum of the areas of rectangles, where the width of each
rectangle can be different but getting small, and the height of each rectangle is
not necessarily determined by a particular rule.” The theorem states that this
Riemann Sum also gives the value of the definite integral of f over [a, b].
CHAPTER 5. INTEGRATION 256
Let f be continuous on the closed interval [a, b] and let SL (n), SR (n),
SM (n), ∆x, ∆xi and ci be defined as before. Then:
X
n Z b
2. lim f (ci )∆x = f (x) dx
n→∞ a
i=1
Z [Link]/watch?v=A-WLvclVMC0
X
n b
3. lim f (ci )∆xi = f (x) dx Figure 5.3.27 Video presentation of The-
∥∆x∥→0 a
i=1 orem 5.3.26
We summarize what we have learned over the past few sections here.
• Knowing the “area under the curve” can be useful. One common example:
the area under a velocity curve is displacement.
Rb
• We have defined the definite integral, a f (x) dx, to be the signed area
under f on the interval [a, b]. One of the things Theorem 5.3.26
tells us is that if f is continuous
• While we can approximate a definite integral many ways, we have focused on [a, b], then the definite inte-
Rb
on using rectangles whose heights can be determined using the Left Hand gral a f (x) dx is guaranteed to
Rule, the Right Hand Rule and the Midpoint Rule. exist.
Knowing that every continu-
• Sums of rectangles of this type are called Riemann sums.
ous function can be integrated
• The exact value of the definite integral can be computed using the limit of is useful, since most of the func-
a Riemann sum. We generally use one of the above methods as it makes tions we work with are continu-
the algebra simpler. ous. However, it turns out that a
function can be integrated even
We first learned of derivatives through limits then learned rules that made if it has a finite number of dis-
the process simpler. We know of a way to evaluate a definite integral using limits; continuities, as long as these are
in the next section we will see how the Fundamental Theorem of Calculus makes removable or jump discontinuities.
the process simpler. The key feature of this theorem is its connection between
the indefinite integral and the definite integral.
CHAPTER 5. INTEGRATION 257
5.3.4 Exercises
Terms and Concepts
Problems
Exercise Group. Write out each term of the summation and compute the sum.
P5 P3
5. i2 6. (5i + 2)
i=3 i=−2
P
2
πi
P
8
7. sin 2 8. 5
i=−2 i=1
P
6
1
P
8
i
9. i 10. (−1) i
i=1 i=1
3
P P
6
i
11. 1
i − 1
i+1 12. (−1) cos(πi)
i=1 i=0
P
n P
k P
n P
n P
n P
k
Exercise Group. Theorem 5.3.9 states ai = ai + ai , so ai = ai − ai . Use this fact, along
i=1 i=1 i=k+1 i=k+1 i=1 i=1
with other parts of Theorem 5.3.9, to evaluate the summation.
P
20 P
29
25. i 26. i3
i=11 i=17
P
15 P
13
27. 6 28. 3i3
i=8 i=8
Z b
Exercise Group. In the following exercises, a definite integral f (x) dx is given.
a
Exercise Group. A definite integral is given below. As demonstrated in Examples 5.3.23 and Example 5.3.24, do the
following:
(a) Find a formula to approximate the definite integral using n subintervals and the provided rule.
(b) Evaluate the formula using n = 10, 100, and 1000.
(c) Find the limit of the formula, as n → ∞, to find the exact value of the definite integral.
Z 1 Z 2
35. x3 dx, using the Left Hand Rule. 36. 2x2 dx, using the Left Hand Rule.
0 −1
Z Z
1 6
37. (3x + 2) dx, using the Midpoint Rule. 38. 4x2 + 1 dx, using the Left Hand Rule.
−2 2
Z Z
10 1
39. (3 − x) dx, using the Left Hand Rule. 40. x3 − x2 dx, using the Right Hand Rule.
−10 0
CHAPTER 5. INTEGRATION 259
8 y
6
t
a x b 4
Rx 2
Figure 5.4.2 The area of the shaded region is F (x) = a
f (t) dt
t
Example 5.4.3 Exploring the “Area so far” function. 1 x
−2
Consider f (t) = 2t pictured in R xFigure Figure 5.4.4 and its associated
“area so far” function, F (x) = 1 2t dt. Using the graph of f and geom- Figure 5.4.4 The Rarea of the shaded
etry, find an explicit formula for F . x
region is F (x) = 1 2t dt
Solution. We can see from Figure 5.4.5 that for x ≥ 1, the area under
the curve can be found by subtracting the area of two triangles. The Video Solution
larger triangle will have a base of x and a height of f (x) = 2x, while the
smaller triangle will have a base of 1 and a height of 2. Therefore, the
area under the curve for x ≥ 1 is given by A(x) = 12 (x)(2x)− 12 (1)(2) =
x2 − 1.
Note that this same formula holds for x < 1. If x < 1, then F (x) =
Rx R1
1
2t dt = − x 2t dt. The areas to the left of x = 1 will have oppo-
site signs (since they areas areR accumulated before x = 1). For exam-
1
ple, when x = 0, F (0) = − 0 2t dt = − 12 (1)(2) = −1. This is the
same valueRwe get from evaluating
R1 x2 − 1 for x = 0. Also notice that
−1 [Link]/watch?v=zVyMghQRLcI
F (−1) = 1 2t dt = − −1 2t dt. This integral is clearly 0 since the
areas over [−1, 0] and [0, 1] will sum to zero. Again, this is the same
answer obtained by evaluating x2 − 1 for x = −1.
8 y
Therefore, we can reasonably say that F (x) = x2 − 1. A plot of
both f (x) = 2x and F (x) = x2 − 1 are given in Figure Figure 5.4.6. You
6
should notice a familiar relationship between these two functions. This
relationship is formally stated in Theorem 5.4.7.
4
8 y
2x
2
6
2
1 t
4
1 x
x
2
−2
x
−1 1 2 3 Figure 5.4.5 The Rarea of the shaded
x
region is F (x) = 1 2t dt
−2
F ′ (x) = f (x).
In other words: Z x
d
f (t) dt = f (x).
dx a
What we have done in Example 5.4.9 was more than finding a complicated
Video Solution
way of computing an antiderivative. Consider a function f defined on an open
Rb
interval containing a, b and c. Suppose we want to compute a f (t) dt. First, let
Z x
F (x) = f (t) dt. (5.4.1)
c
Using the properties of the definite integral found in Theorem 5.2.11, we know
Z b Z c Z b
f (t) dt = f (t) dt + f (t) dt
a a c
Z a Z b [Link]/watch?v=7tHmgPcUZG4
=− f (t) dt + f (t) dt
c c
Using Equation (5.4.1), let x = a in the first integral and x = b in the second
Ra Rb
integral so that c f (t) dt = F (a) and c f (t) dt = F (b). Therefore:
Z b
f (t) dt = −F (a) + F (b)
a
= F (b) − F (a).
We now see how indefinite integrals and definite integrals are related: we
can evaluate a definite integral using antiderivatives! In fact, this is exactly what
CHAPTER 5. INTEGRATION 261
we noticed in Example 5.4.3. The “area so far” function was indeed an anti-
derivative of the integrand. This is the second part of the Fundamental Theorem
of Calculus.
As its name suggests, the Fundamental Theorem of Calculus is an important Figure 5.4.11 Video presentation of The-
result. In fact, it’s sufficiently important that it’s worth taking a moment to un- orem 5.4.10
derstand why it’s true. A proof is given in Figure 5.4.12.
Z 2 Z 5 Z 5
3 t
1. x dx 3. e dt 5. 2 dx
−2 0 1
Z Z
π 9 √
2. sin(x) dx 4. u du Video Solution
0 4
Solution.
1.
Z 2 2
1 4
x3 dx = x
−2 4 −2
1 4 1
= 2 − (−2) 4
4 4
[Link]/watch?v=YxQyFln5UIQ
= 0.
2.
Z π π
sin(x) dx = − cos(x)
0
0
= − cos(π) − − cos(0)
= 1 + 1 = 2.
(This is interesting; it says that the area under one “hump” of a
sine curve is 2.)
3.
Z 5 5
et dt = et
0 0
= e − e0 5
= e5 − 1 ≈ 147.41.
4.
Z Z
9 √ 9
1
u du = u 2 du
4 4
2 3 9
= u2
3 4
2 3 3
= 9 − 42
2
3
2 38
= 27 − 8 = .
3 3
5.
Z 5 5
2 dx = 2x
1 1
= 2(5) − 2
= 2(5 − 1) = 8.
This integral is interesting; the integrand is a constant function,
hence we are finding the area of a rectangle with width (5−1) = 4
and height 2. Notice how the evaluation of the definite integral led
Rb
to 2(4) = 8. In general, if c is a constant, then a c dx = c(b − a).
CHAPTER 5. INTEGRATION 263
add these two areas, we get the displacement of 4 ft. But when we add
the absolute value of both of these areas (as in Figure 5.4.17), we get
the total distance of 9 ft.
y y
20 20
10 10
A1 A1
A3
t t
5/8 A2 1 5/8 1
−10 −10
Figure 5.4.16 The area between Figure 5.4.17 The area between
v(t) and the t-axis can be used to |v(t)| and the t-axis can be used
represent displacement to represent distance
Integrating a rate of change function gives total change. Velocity is the rate
of position change; integrating velocity gives the total change of position, i.e.,
displacement.
Integrating a speed function gives a similar, though different, result. Speed
is also the rate of position change, but does not account for direction. That is,
the speed an object is the absolute valueR 1 of its velocity. This is what we saw
in Example 5.4.15 when we evaluated 0 |v(t)| dt. So integrating a speed func-
tion gives total change of position, without the possibility of “negative position
change.” Hence the integral of a speed function gives distance traveled.
As acceleration is the rate of velocity change, integrating an acceleration
function gives total change in velocity. We do not have a simple term for this
analogous to displacement. If a(t) = 5miles/h2 and t is measured in hours,
then Z 3
a(t) dt = 15
0
means the velocity has increased by 15m/h from t = 0 to t = 3.
What is the derivative of such a function? 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.
[Link]/watch?v=nGS4ENM8arI
CHAPTER 5. INTEGRATION 266
y y
f (x) f (x)
g(x) g(x)
[Link]/watch?v=UufnFHBnv88
x x
Figure 5.4.21 Video introduction to Sub-
a b a b
section 5.4.4
(a) (b)
Figure 5.4.22 Finding the area bounded by two functions on an interval by sub-
tracting the area under g from the area under f
The area can be found by recognizing that this area is “the area under f −
the area under g.” Using mathematical notation, the area is
Z b Z b
f (x) dx − g(x) dx.
a a
15 y y = x2 + x − 5
Example 5.4.24 Finding area between curves.
10
Find the area of the region enclosed by y = x2 + x − 5 and y = 3x − 2.
Solution. It will help to sketch these two functions, as done in Figure 5.4.25. y = 3x − 2
5
The region whose area we seek is completely bounded by these two
functions; they seem to intersect at x = −1 and x = 3. To check, set x
x2 + x − 5 = 3x − 2 and solve for x: −2 −1 1 2 3 4
x2 + x − 5 = 3x − 2
(x2 + x − 5) − (3x − 2) = 0
x2 − 2x − 3 = 0 Figure 5.4.25 Sketching the region en-
closed by y = x2 + x − 5 and y =
(x − 3)(x + 1) = 0
3x − 2 in Example 5.4.24
CHAPTER 5. INTEGRATION 267
x = −1, 3.
One of the things we have to be careful about when finding the area between [Link]/watch?v=Bgji1b7Wdr4
curves is that the curves might cross, so that the distinction between “upper Figure 5.4.26 Finding the area between
curve” and “lower curve” can change. The video example in Figure 5.4.26 illus- curves that intersect multiple times
trates this phenomenon.
R 4 Consider the graph of a function f in Figure 5.4.27 and the area defined by
1
f (x) dx. Three rectangles are drawn in Figure 5.4.28; in Figure 5.4.28(a), the
height of the rectangle is greater than f on [1, 4], hence the area of this rectangle
R4
is is greater than 1 f (x) dx.
In Figure 5.4.28(b), the height of the rectangle is smaller than f on [1, 4],
R4
hence the area of this rectangle is less than 1 f (x) dx.
x
Finally, in Figure 5.4.28(c) the
R 4height of the rectangle is such that the area of 1 2 3 4
the rectangle is exactly that of 1 f (x) dx. Since rectangles that are “too big”,
as in R(a), and rectangles that are “too little,” as in (b), give areas greater/lesser
4
than 1 f (x) dx, it makes sense that there is a rectangle, whose top intersects
Figure 5.4.27 A graph of a function f
f (x) somewhere on [1, 4], whose area is exactly that of the definite integral.
to introduce the Mean Value Theorem
y y y
x x x
1 2 3 4 1 2 3 4 1 2 3 4
Let f be continuous on [a, b]. There exists a value c in [a, b] such that
Z b
f (x) dx = f (c)(b − a).
a
CHAPTER 5. INTEGRATION 268
axis on [a, b] is the same as the amount of area below the x-axis above f ; see
Figure 5.4.33 for an illustration of this. In this sense, we can say that f (c) is the Figure 5.4.32 A graph of y = sin(x)
average value of f on [a, b]. on [0, π] and the rectangle guaranteed
by the Mean Value Theorem
y y
y = f (x)
y = f (x) − f (c)
f (c) f (c)
x x
a c b a c b
Figure 5.4.33 On the left, a graph of y = f (x) and the rectangle guaranteed by
the Mean Value Theorem. On the right, y = f (x) is shifted down by f (c); the
resulting “area under the curve” is 0
The value f (c) is the average value in another sense. First, recognize that
CHAPTER 5. INTEGRATION 269
for some value of c in [a, b]. Replacing the integral with the limit of a Riemann
sum (as in Theorem 5.3.26):
Z b
1
f (c) = f (x) dx
b−a a
1 Xn
= lim f (ci ) ∆x Using Theorem 5.3.26
b − a n→∞ i=1
1 X
n
b−a b−a
= lim f (ci ) ∆x =
b − a n→∞ i=1 n n
X
n
1
= lim f (ci ) Cancelling the common factor of b − a.
n→∞
i=1
n
Pn
Examining this last line closely, the expression i=1 f (ci ) n1 represents adding
up n sample values of f (x)and then dividing by n. This is exactly what we do
when we calculate the average of a set Pof n numbers. Now when we consider
n
taking the limit as n goes to ∞, lim 1
i=1 f (ci ) n , we are adding up all of the
n→∞
function’s output values over [a, b] and dividing by the “number of numbers”. In
a sense, we are adding up an infinite number of output values and then dividing
by the number of terms we summed (which is again infinite).
This leads us to a definition.
An application of this definition is given in the following example. Figure 5.4.35 Video presentation of De-
finition 5.4.34
Example 5.4.36 Finding the average value of a function.
An object moves back and forth along a straight line with a velocity given
by v(t) = (t − 1)2 on [0, 3], where t is measured in seconds and v(t) is
measured in ft/s.
What is the average velocity of the object?
Solution. By our definition, the average velocity is:
Z 3 Z
1 1 3 2
(t − 1)2 dt = t − 2t + 1 dt
3−0 0 3 0
3
1 1 3
= t −t +t
2
3 3
0
1 1 3 1 3
= (3) − (3) + (3) −
2
(0) − (0) + (0)
2
3 3 3
= 1 ft/s .
CHAPTER 5. INTEGRATION 270
5.4.6 Exercises
Terms and Concepts
Problems
29.
Z 1
(a) Explain why xn dx = 0 when n is a positive, odd integer.
−1
Z 1 Z 1
n
(b) Explain why x dx = 2 xn dx when n is a positive, even integer.
−1 0
Z a+2π
30. Explain why sin t dt = 0 for all values of a.
a
CHAPTER 5. INTEGRATION 272
Rb
Exercise Group. Find all values c such that a
f (x) dx = f (c)(b − a), as guaranteed by the Theorem 5.4.29.
R 2 R 2
31. x dx 32. x dx
R x R √
33. e dx 34. x dx
Exercise Group. Find the average value of the function on the given interval.
35. f (x) = sin(x) on π2 , π 36. y = cos(x) on [0, π]
37. y = x on [0, 7] 38. y = x2 on [0, 8]
39. y = x3 on [0, 9] 1
40. y = on [1, e]
t
Exercise Group. A velocity function is given for an object moving along a straight line. Find the displacement of the
object over the given time interval.
ft ft
41. v(t) = −32t + 28 s on [0, 4] 42. v(t) = −32t + 200 s on [0, 9]
43. v(t) = 7 ft
on [0, 4] 44. v(t) = 8t mph on [−2, 3]
s
ft √
45. v(t) = cos(t) s on 0, 3π
2 46. v(t) = 5 t fts on [0, 32]
Exercise Group. An acceleration function of an object moving along a straight line is given. Find the change of the
object’s velocity over the given time interval.
ft ft
47. a(t) = −32 s2 on [0, 8] 48. a(t) = 11 on [0, 9]
s2
ft
49. a(t) = t s2 on [0, 1] 50. a(t) = sin(t) sft2 on π2 , 3π
2
Exercise Group. Sketch the given relations and find the area of the enclosed region.
51. y = 2x, y = 5x, and x = 3 52. y = −x + 1, y = 3x + 6, x = 2 and x = −1
53. y = x2 − 2x + 5, y = 5x − 5 54. y = 2x2 + 2x − 5, y = x2 + 3x + 7,
R −x2
e dx.
This section outlines three common methods of approximating the value of
definite integrals. We describe each as a systematic method of approximating
area under a curve. By approximating this area accurately, we find an accurate
approximation of the corresponding definite integral.
We will apply the methods we learn in this section to the following definite
integrals:
R 1 −x2 R π2 R 4π sin(x)
0
e dx, −π
sin(x3 ) dx, 0.5 x dx,
4
y y y
1
1
y = e−x
2
1 y = sin(x3 )
sin(x)
y=
x
0.5
0.5
0.5
x
−1 −0.5 0.5 1 1.5 x
x
5 10 15
−0.2 0.2 0.4 0.6 0.8 1
−0.5
Key Idea 5.3.20 states that to use the Left Hand Rule we use the summation
X
n Xn
f (xi−1 )∆x and to use the Right Hand Rule we use f (xi )∆x. We review
i=1 i=1
the use of these rules in the context of examples.
y y
−x2
y = e−x
2
1 y=e 1
0.5 0.5
x x
0.2 0.4 0.6 0.8 1 0.2 0.4 0.6 0.8 1
(a) Using the Left Hand Rule (b) Using the Right Hand Rule
R1
e−x dx in Example 5.5.2
2
Figure 5.5.3 Approximating 0
y y y = sin(x3 )
1 1
y = sin(x3 )
0.5 0.5
x x
−1 −0.5 0.5 1 1.5 −1 −0.5 0.5 1 1.5
−0.5 −0.5
(a) (b)
R π
Figure 5.5.6 Approximating 2
−π
sin(x3 ) dx in Example 5.5.4
4
approximated with 5 trapezoids of equal width; the top “corners” of each trape-
using 5 trapezoids of equal widths
zoid lies on the graph of f (x). It is clear from this figure that these trapezoids
more accurately approximate
R1 the area under f and hence should give a better
approximation of 0 e−x dx. (In fact, these trapezoids seem to give a great
2
e−xi
2
1 + 0.9608 xi
(0.2) = 0.1961.
2
Moving right, the next trapezoid has legs of length 0.9607 and 0.8521 0 1
and a height of 0.2. Thus its area is: 0.2 0.9608
0.4 0.8521
0.9608 + 0.8521 0.6 0.6977
(0.2) = 0.1813.
2 0.8 0.5273
The sum of the areas of all 5 trapezoids is: 1 0.3679
There are many things to observe in this example. Note how each term in
the final summation was multiplied by both 1/2 and by ∆x = 0.2. We can factor
these coefficients out, leaving a more concise summation as:
1 h
(0.2) (1 + 0.9608) + (0.9608 + 0.8521) + (0.8521 + 0.6977)
2 i
+ (0.6977 + 0.5273) + (0.5273 + 0.3679) .
Now notice that all numbers except for the first and the last are added twice.
Therefore we can write the summation even more concisely as
0.2 h i
1 + 2(0.9608 + 0.8521 + 0.6977 + 0.5273) + 0.3679 .
2
CHAPTER 5. INTEGRATION 277
Rb
This is the heart of the Trapezoidal Rule, wherein a definite integral a f (x) dx
is approximated by using trapezoids of equal widths to approximate the corre-
sponding area under f . Using n equally spaced subintervals with endpoints x0 ,
b−a
x1 , . . ., xn , we again have ∆x = . Thus:
n
Z b Xn
f (xi−1 ) + f (xi )
f (x) dx ≈ ∆x
a i=1
2
∆x X
n
= f (xi−1 ) + f (xi )
2 i=1
!
∆x h X i
n−1
= f (x0 ) + 2 f (xi ) + f (xn ) .
2 i=1
0.236 h i
≈ − 0.4657 + 2 − 0.1654 + (−0.031) + . . . + 0.68999 + (−0.67)
2
= 0.4258.
Notice how “quickly” the Trapezoidal Rule can be implemented once the ta-
ble of values is created. This is true for all the methods explored in this section;
the real work is creating a table of xi and f (xi ) values. Once this is completed,
approximating the definite integral is not difficult. Again, using technology is
wise. Spreadsheets can make quick work of these computations and make us-
ing lots of subintervals easy.
Also notice the approximations the Trapezoidal Rule gives. It is the average
of the approximations given by the Left and Right Hand Rules! This effectively
renders the Left and Right Hand Rules obsolete. They are useful when first learn-
ing about definite integrals, but if a real approximation is needed, one is gener-
ally better off using the Trapezoidal Rule instead of either the Left or Right Hand
Rule. However, there are two other methods that are also generally more accu-
rate than the Left or Right Hand Rule.
X
n
= f (xi ) ∆x
i=1
where xi is the midpoint of each subinterval,
1
xi = a + ∆x i −
2
Example 5.5.12 Using the Midpoint Rule.
Z 1
e−x dx.
2
Use the Midpoint Rule with n = 5 to approximate
0
Solution. We cannot use the table in Figure 5.5.10 that we used for the
Trapezoidal, Right and Left Hand Rules when using the Midpoint Rule.
The Trapezoidal rule averages the outputs of the function to obtain a
more accurate estimate of the definite integral. The Midpoint Rule av-
erages
the inputs
of each subinterval to create a rectangle with height
xi−1 +xi xi−1 +xi
f 2 . Generally f 2 ̸= f (xi−12)+f (xi ) .
So we will create a new table of values as shown in Figure 5.5.13. We
have ∆x = (1 − 0)/5 = 0.2. The midpoint of the first subinteval is at
e−xi
2
xi
0 + 0.2(1/2) = 0.1 and each successive midpoint is 0.2 from the last.
So we have
Z 1 0.1 0.9900
e−x dx ≈ 0.2(0.99 + 0.9139 + 0.7788 + 0.6126 + 0.4449)
2
0.3 0.9139
0 0.5 0.7788
≈ 0.7480 0.7 0.6126
Z 0.9 0.4449
1
e−x dx ≈ 0.7480.
2
We approximate
Figure 5.5.13 A table of values of e−x
2
The actual answer, accurate to 4 decimal places is 0.4609. So the Mid- Figure 5.5.15 Values used to approxi-
Rπ
point Rule with 10 subintervals is an overrapproximation by about 0.0183. mate −2 π sin(x3 ) dx in Example 5.5.14
4
Notice that this error is about half of the error in using the Trapezoidal
Rule.
CHAPTER 5. INTEGRATION 279
In many cases, the Midpoint Rule will more accurate than the Trapezoidal
Rule. You may wonder though, how can we improve on the Trapezoidal and
Midpoint Rules, apart from using more and more subintervals? The answer is
clear once we look back and consider what we have really done so far. The
Left Hand Rule, Right Hand Rule and Midpoint Rules are not really about using
rectangles to approximate area. Instead, they approximate a function f with
constant functions on small subintervals and then compute the definite integral
of these constant functions. The Trapezoidal Rule is really approximating a func-
tion f with a linear function on a small subinterval, then computing the definite
integral of this linear function. In all of these cases the definite integrals are easy
to compute in geometric terms.
So we have a progression: we start by approximating f with a constant func-
tion and then with a linear function. What is next? A quadratic function. By
approximating the curve of a function with lots of parabolas, we generally get
an even better approximation of the definite integral. We call this process Simp-
son’s Rule, named after Thomas Simpson (1710-1761), even though others had
used this rule as much as 100 years prior.
CHAPTER 5. INTEGRATION 280
e−xi
2
xi y
y = e−x
2
1
0 1
0.25 0.939
0.5 0.779
0.5
0.75 0.570
1 0.368
(a) x
0.25 0.5 0.75 1
(b)
R1
e−x dx, along with a
2
Figure 5.5.19 A table of values to approximate 0
graph of the function
Simpson’s Rule states that
Z 1
0.25 h i
e−x dx ≈
2
1+4(0.939)+2(0.779)+4(0.570)+0.368 = 0.74683.
0 3
≈ 0.4701
x
Recall that the actual value, accurate to 3 decimal places, is 0.4609. −1 −0.5 0.5 1 1.5
Our approximation is within one 1/100th of the correct value. The graph
in Figure 5.5.22 shows how closely the parabolas match the shape of the −0.5
graph.
Rπ
Figure 5.5.22 Approximating −2 π sin(x3 ) dx
4
in Example 5.5.20 with Simpson’s Rule
and 10 equally spaced intervals
CHAPTER 5. INTEGRATION 282
Z b
1. Let ET and EM be the error in approximating f (x) dx using
a
the Trapezoidal and Midpoint Rules respectively, with n subinter-
vals. If f has a continuous second derivative on [a, b] and K is any
upper bound of |f ′′ (x)| on [a, b], then
(b − a)3
ET ≤ K.
12n2
and
(b − a)3
EM ≤ K.
24n2
Z b
2. Let ES be the error in approximating f (x) dx using Simpson’s
a
Rule with n subintervals.. If f has a continuous 4th derivative on
[a, b] and K is any upper bound of f (4) (x) on [a, b], then
(b − a)5
ES ≤ K.
180n4
Figure 5.5.26 shows a graph of f ′′ (x) on [0, 1]. It is clear that the
largest value of f ′′ , in absolute value, is 2.
CHAPTER 5. INTEGRATION 284
Thus we let K = 2 and apply the error formula from Theorem 5.5.24. 0.5
y
(1 − 0)3 x
ET ≤ · 2 = 0.006. 0.2 0.4 0.6 0.8 1
12 · 52
−0.5
Since the maximum error in the Midpoint rule is half the error in the
Trapezoidal Rule, we can say: EM ≤ 0.003 −1
Our error estimation formula states that our approximation of 0.7444
found in Example 5.5.9 is within 0.0067 of the correct answer. Hence we −1.5 y = e−x (4x2 − 2)
2
But we can do better than this with the Midpoint Rule since its error is
at most half of the error of the Trapezoidal Rule. Our error estimate for-
mula state that our approximate of 0.7480 found in Example 5.5.12 is
within 0.0034 of the correct answer. Hence Hence we know that the ac-
tual value is within [0.7480−0.0034, 0.7480+0.0033] = [0.7447, 0.7513].
We had earlier stated the actual answer, correct to 4 decimal places,
to be 0.7468, affirming the validity of Theorem 5.5.24.
Simpson’s Rule with n = 4:
We start by computing the 4th derivative of f (x) = e−x :
2
y
Figure 5.5.27 shows a graph of f (4) (x) on [0, 1]. It is clear that the
y = e−x (16x4 − 48x2 + 12)
2
One of the authors drove his daughter home from school while she recorded Time Speed
their speed every 30 seconds. The data is given in Figure 5.5.29. Approx- (min) (mph)
imate the distance they traveled. 0 0
Solution. Recall that by integrating a speed function we get distance 1 25
traveled. We have information about v(t); we will use Simpson’s Rule to 2 22
Z b
3 19
approximate v(t) dt.
a 4 39
The most difficult aspect of this problem is converting the given data 5 0
into the form we need it to be in. The speed is measured in miles per 6 43
hour, whereas the time is measured in minutes. 7 59
We need to compute ∆x = (b−a)/n. With 25 data points collected, 8 54
there are n = 24 subintervals. What are a and b? Since we start at time
9 51
t = 0, we have a = 0. The final recorded time was t = 12 minutes,
10 43
which is 1/5 of an hour. Thus we have
11 35
b−a 1/5 − 0 1 ∆x 1 12 40
∆x = = = ; = .
n 24 120 3 360 13 43
Thus the distance traveled is approximately: 14 30
15 0
Z 0.2
1 h i 16 0
v(t) dt ≈ f (x0 ) + 4f (x1 ) + 2f (x2 ) + · · · + 4f (xn−1 ) + f (xn )
0 360 17 28
1 h i 18 40
= 0 + 4 · 25 + 2 · 22 + · · · + 2 · 40 + 4 · 23 + 0
360 19 42
≈ 6.2167 miles. 20 40
21 39
We approximate the author drove 6.2 miles. (Because we are sure 22 40
the reader wants to know, the author’s odometer recorded the distance 23 23
as about 6.05 miles.)
24 0
5.5.6 Exercises
Terms and Concepts
Problems
Exercise Group. In the following exercises, approximate the definite integral with the Trapezoidal Rule and Simpson’s
Rule, with n = 4. Then find the exact value.
R1 R 10
5. For the integral −1 x2 dx: 6. For the integral 0 5x dx:
(a) Approximate using the trapezoidal rule: (a) Approximate using the trapezoidal rule:
(b) Approximate using Simpson’s rule: (b) Approximate using Simpson’s rule:
(c) Find the exact value: (c) Find the exact value:
Rπ R4√
7. For the integral 0
sin(x) dx: 8. For the integral x dx:
0
(a) Approximate using the trapezoidal rule: (a) Approximate using the trapezoidal rule:
(b) Approximate using Simpson’s rule: (b) Approximate using Simpson’s rule:
(c) Find the exact value: (c) Find the exact value:
R3 R1
9. For the integral 0
(x3 + 2x2 − 5x + 7) dx: 10. For the integral 0
x4 dx:
(a) Approximate using the trapezoidal rule: (a) Approximate using the trapezoidal rule:
(b) Approximate using Simpson’s rule: (b) Approximate using Simpson’s rule:
(c) Find the exact value: (c) Find the exact value:
R 2π R3 √
11. For the integral 0
cos(x) dx: 12. For the integral −3
9 − x2 dx:
(a) Approximate using the trapezoidal rule: (a) Approximate using the trapezoidal rule:
(b) Approximate using Simpson’s rule: (b) Approximate using Simpson’s rule:
(c) Find the exact value: (c) Find the exact value:
Exercise Group. In the following exercises, approximate the definite integral with the Trapezoidal Rule and Simpson’s
Rule, with n = 6.
CHAPTER 5. INTEGRATION 287
R1 R1 2
13. For the integral 0
cos x2 dx: 14. For the integral −1
ex dx:
(a) Approximate using the trapezoidal rule: (a) Approximate using the trapezoidal rule:
(b) Approximate using Simpson’s rule: (b) Approximate using Simpson’s rule:
R5√ Rπ
15. For the integral x2 + 1 dx: 16. For the integral 0
x sin(x) dx:
0
(a) Approximate using the trapezoidal rule: (a) Approximate using the trapezoidal rule:
(b) Approximate using Simpson’s rule: (b) Approximate using Simpson’s rule:
R π/2 p R4
17. For the integral 0
cos(x) dx: 18. For the integral 1
ln(x) dx:
(a) Approximate using the trapezoidal rule: (a) Approximate using the trapezoidal rule:
(b) Approximate using Simpson’s rule: (b) Approximate using Simpson’s rule:
R1 1
R6 1
19. For the integral −1 sin(x)+2
dx: 20. For the integral 0 sin(x)+2
dx:
(a) Approximate using the trapezoidal rule: (a) Approximate using the trapezoidal rule:
(b) Approximate using Simpson’s rule: (b) Approximate using Simpson’s rule:
Exercise Group. In the following exercises, find n such that the error in approximating the given definite integral is
less than 0.0001 when using the Trapezoidal Rule and Simpson’s Rule.
Rπ R4
21. For the integral 0 sin(x) dx: 22. For the integral 1 √1x dx:
(a) Trapezoid rule: (a) Trapezoid rule:
n >= n >=
(b) Simpson’s rule: (b) Simpson’s rule:
n >= n >=
Rπ R5
23. For the integral 0 cos x2 dx: 24. For the integral 0 x4 dx:
(a) Trapezoid rule: (a) Trapezoid rule:
n >= n >=
(b) Simpson’s rule: (b) Simpson’s rule:
n >= n >=
Exercise Group. In the following exercises, a region is given. Find the area of the region using Simpson’s Rule:
25. 26.
3.6
4.7
6.3
6.6
6.9
3.6
4.5
6.6
5.6
5.1
CHAPTER 5. INTEGRATION 289
5.6 Substitution
We motivate this section with an example. Let f (x) = (x2 + 3x − 5)10 . We
can compute f ′ (x) using the Chain Rule. It is:
How would we have evaluated this indefinite integral without starting with
f (x) as we did?
This section explores integration by substitution. It allows us to “undo the
Chain Rule.” Substitution allows us to evaluate the above integral without know-
ing the original function first.
R The underlying principle is to rewrite a R“complicated” integral of the form
f (x) dx as a not-so-complicated integral h(u) du. We’ll formally establish
later
R how this is done. First, consider again our introductory indefinite integral,
(20x + 30)(x2 + 3x − 5)9 dx. Arguably the most “complicated” part of the
integrand is (x2 + 3x − 5)9 . We wish to make this simpler; we do so through a
substitution. Let u = x2 + 3x − 5. Thus
(x2 + 3x − 5)9 = u9 .
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 section contains numerous examples through which the reader will gain
understanding and mathematical maturity enabling them to regard substitution
as a natural tool when evaluating integrals.
We stated before that integration by substitution “undoes” the Chain Rule.
Specifically, let F (x) and g(x) be differentiable functions and consider the de-
rivative of their composition:
d
F g(x) = F ′ (g(x))g ′ (x).
dx
CHAPTER 5. INTEGRATION 290
Thus Z
F ′ (g(x))g ′ (x) dx = F (g(x)) + C.
The
R point of substitution is to make the integration step easy. Indeed, the
step F ′ (u) du = F (u) + C looks easy, as the antiderivative of the deriva-
tive of F is just F , plus a constant. The “work” involved is making the proper
substitution. There is not a step-by-step process that one can memorize; rather,
experience will be one’s guide. To gain experience, we now embark on many
examples.
1
= − cos(u) + C (now replace u with x2 + 5)
2
1
= − cos(x2 + 5) + C.
2
R
Thus x sin(x2 + 5) dx = − 12 cos(x2 + 5) + C. We can check our
work by evaluating the derivative of the right hand side.
substitution.
Z Z
7 7 du
dx =
−3x + 1 u −3
Z
−7 du
=
3 u
−7
= ln |u| + C
3
7
= − ln |−3x + 1| + C.
3
Using Key Idea 5.6.5 is faster, recognizing that u is linear and a =
−3. One may want to continue writing out all the steps until they are
comfortable with this particular shortcut. [Link]/watch?v=-6CFSvtMCDU
Figure 5.6.7 Video presentation of Ex-
Not all integrals that benefit from substitution have a clear “inside” function.
amples 5.6.3–5.6.6
Several of the following examples will demonstrate ways in which this occurs.
Our examples so far have required “basic substitution.” The next example
demonstrates how substitutions can be made that often strike the new learner
as being “nonstandard.”
The final answer is interesting; the natural log of the natural log. Take
the derivative to confirm this answer is indeed correct.
The next three examples will help fill in some missing pieces of our antideriv-
ative knowledge. We know the antiderivatives of the sine and cosine functions;
what about the other standard functions tangent, cotangent, secant and cose-
cant? We discover these next.
− ln |cos(x)| + C = ln (cos(x))−1 + C
1
= ln +C
cos(x)
= ln |sec(x)| + C.
R
Thus the result they give is tan(x) dx = ln |sec(x)| + C. These
two answers are equivalent.
Z
sec2 (x) + sec(x) tan(x)
= dx.
sec(x) + tan(x)
We can use similar techniques to those used in Examples 5.6.12 and Exam-
ple 5.6.13 to find antiderivatives of cot(x) and csc(x) (which the reader can
explore in the exercises.) We summarize our results here.
R
1. sin(x) dx = − cos(x) + C,
R
2. cos(x) dx = sin(x) + C,
R
3. tan(x) dx = − ln |cos(x)| + C,
R
4. csc(x) dx = − ln |csc(x) + cot(x)| + C,
R
5. sec(x) dx = ln |sec(x) + tan(x)| + C,
R
6. cot(x) dx = ln |sin(x)| + C,
1 + cos(2x)
cos2 (x) = .
2
The right hand side of this equation is not difficult to integrate. We
have:
Z Z
1 + cos(2x)
cos2 (x) dx = dx
2
Z
1 1
= + cos(2x) dx
2 2
1 1 sin(2x)
= x+ +C
2 2 2
1 sin(2x)
= x+ + C,
2 4
CHAPTER 5. INTEGRATION 296
x2 + 2x + 3
= x 2 + 2x 2 + 3x− 2 .
3 1 1
x1/2
We can now integrate using the Power Rule:
Z 2 Z
x + 2x + 3
2 + 2x 2 + 3x− 2
3 1 1
dx = x dx
x1/2
2 5 4 3 1
= x 2 + x 2 + 6x 2 + C
5 3
This is a perfectly fine approach. We demonstrate how this can also
be solved using
√ substitution as its implementation is rather clever.
1
Let u = x = x 2 ; therefore
1 1
du = √ dx ⇒ 2du = √ dx.
2 x x
Z 2 Z
x + 2x + 3
This gives us √ dx = (x2 + 2x + 3) · 2 du. What are
x
1
we to do with the other x terms? Since u = x 2 , u2 = x, etc. We can
then replace x2 and x with appropriate powers of u. We thus have
Z 2 Z
x + 2x + 3
√ dx = (x2 + 2x + 3) · 2 du
x
Z
= 2(u4 + 2u2 + 3) du
2 5 4 3
= u + u + 6u + C
5 3
2 5 4 3 1
= x 2 + x 2 + 6x 2 + C,
5 3
which is obviously the same answer we obtained before. In this situation,
substitution is arguably more work than our other method. The fantastic
thing is that it works. It demonstrates how flexible integration is.
Thus Z Z
1 1 1
2
dx =
x 2
dx.
25 + x 25 1+ 5
Z Z Z
1 1 1
1. dx 2. √ dx 3. q dx
9 + x2 5 − x2 x x −
2 1
100
1 x−2
= tan−1 + C.
3 3
As with all definite integrals, you can check your work by differentiation.
This workflow works fine, but substitution offers an alternative that is pow-
erful and amazing (and a little time saving).
CHAPTER 5. INTEGRATION 301
In effect, Theorem 5.6.25 states that once you convert to integrating with
respect to u, you do not need to switch back to evaluating with respect to x. A
few examples will help one understand.
1 y 1 y
y = cos(3x − 1)
0.5 0.5
1
x y= 3 cos(u) u
−1 1 2 3 4 5 −1 1 2 3 4 5
−0.5 −0.5
−1 −1
(a) (b)
Figure 5.6.27 Graphing the areas defined by the definite integrals of Ex-
ample 5.6.26
CHAPTER 5. INTEGRATION 302
The graphs in Figure 5.6.27 tell more of the story. In Figure 5.6.27(a)
the area defined by the original integrand is shaded, whereas in Fig-
ure 5.6.27(b) the area defined by the new integrand is shaded. In this
particular situation, the areas look very similar; the new region is “shorter”
but “wider,” giving the same area.
1 y 1 y y=u
y = sin(x) cos(x)
0.5 0.5
x u
π 1 π
2 2
−0.5 −0.5
(a) (b)
Figure 5.6.29 Graphing the areas defined by the definite integrals of Ex-
ample 5.6.28
5.6.5 Exercises
Terms and Concepts
Problems
Exercise Group. Use Substitution to evaluate the indefinite integral involving trigonometric functions.
Z Z
15. sin3 (x) cos(x) dx 16. cos4 (x) sin(x) dx
Z Z
17. sin(1 − 6x) dx 18. sec2 (7 − 5x) dx
Z Z
19. sec(7x) dx 20. tan8 (x) sec2 (x) dx
Z Z
8 9
21. x cos x dx 22. tan2 (x)dx
Z Z
23. cot(x) dx 24. csc(x) dx
Do not just refer to Theorem 5.6.14 for the Do not just refer to Theorem 5.6.14 for the
answer; justify it through Substitution. answer; justify it through Substitution.
Exercise Group. Use Substitution to evaluate the indefinite integral involving exponential functions.
Z Z
5
25. e4x−5 dx 26. ex x4 dx
Z Z x
2 e +3
27. ex +2x+1 (x + 1) dx 28. dx
ex
Z Z x
ex e + e−x
29. dx 30. dx
ex + 7 e4x
Z Z
31. 99x dx 32. 28x dx
Exercise Group. Use Substitution to evaluate the indefinite integral involving logarithmic functions.
CHAPTER 5. INTEGRATION 304
Z Z 4
ln(x) (ln(x))
33. dx 34. dx
x x
Z Z
ln x5 1
35. dx 36. dx
x x ln(x6 )
Exercise Group. Use Substitution to evaluate the indefinite integral involving rational functions.
Z 2 Z 3
x + 3x − 8 x + x2 + x + 1
37. dx 38. dx
x x
Z 3 Z 2
x −2 x + 9x − 9
39. dx 40. dx
x+1 x+8
Z Z
4x − 8x2 − 2 x2 − 4x
41. dx 42. dx
x−6 x − 6x2 + 9
3
Exercise Group. Use Substitution to evaluate the indefinite integral involving inverse trigonometric functions.
Z Z
6 5
43. dx 44. √ dx
x2 + 6 25 − x2
Z Z
7 9
45. √ dx 46. √ dx
10 − x 2 x x − 49
2
Z Z
6x x
47. √ dx 48. √ dx
x − 64x
6 4 1 − x4
Z Z
1 3
49. dx 50. √ dx
x2 + 18x + 96 −x + 16x − 28
2
Z Z
6 8
51. √ dx 52. 2 − 8x + 80
dx
−x + 12x − 27
2 x
x2 + y 2 = 1 1 y x2 − y 2 = 1 y
(cos(θ),sin(θ)) (cosh(θ),sinh(θ))
2 [Link]/watch?v=-6y0xCwCy4s
θ Figure 5.7.1 Video introduction to Sec-
θ tion 5.7
2 x x
2
−1 −0.5 0.5 1 −2 2
−2
−1
(a) (b)
Figure 5.7.2 Using trigonometric functions to define points on a circle and hyper-
bolic functions to define points on a hyperbola. The area of the shaded regions
are included in them.
ex + e−x 1
1. cosh(x) = 4. sech(x) =
2 cosh(x)
ex − e−x 1
2. sinh(x) = 5. csch(x) =
2 sinh(x)
sinh(x) cosh(x)
3. tanh(x) = 6. coth(x) = Pronunciation Note:
cosh(x) sinh(x)
“cosh” rhymes with “gosh,”
These hyperbolic functions are graphed in Figure 5.7.4 and Figure 5.7.6. “sinh” rhymes with “pinch,”
In the graph of cosh(x) in Figure 5.7.4(a), the graphs of ex /2 and e−x /2 are and
included with dashed lines. In the graph of sinh(x) in Figure 5.7.4(b), the graphs “tanh” rhymes with “ranch.”
of ex /2 and −e−x /2 are included with dashed lines. As x gets “large,” cosh(x)
and sinh(x) each act like ex /2; when x is a large negative number, cosh(x) acts
like e−x /2 whereas sinh(x) acts like −e−x /2.
CHAPTER 5. INTEGRATION 307
5 5
ex /2 e−x /2 x ex /2 −e−x /2 x
−3 −2 −1 1 2 3 −3 −2 −1 1 2 3
−5 −5
−10 −10
(a) (b)
Figure 5.7.4 Graphs of sinh(x) and cosh(x)
In Figure Figure 5.7.6, notice the domains of tanh(x) and sech(x) are (−∞, ∞),
whereas both coth(x) and csch(x) have vertical asymptotes at x = 0. Also note
the ranges of these functions, especially tanh(x): as x → ∞, both sinh(x) and [Link]/watch?v=0YP4mVrroVk
cosh(x) approach e−x /2, hence tanh(x) approaches 1.
Figure 5.7.5 Video presentation of graphs
y y
and basic properties of hyperbolic func-
tions
2 coth(x) sech(x) 2
csch(x)
tanh(x)
x x
−3 −2 −1 1 2 3 −3 −2 −1 1 2 3
−2 −2
(a) (b)
Figure 5.7.6 Graphs of tanh(x), coth(x), csch(x) and cosh(x)
The following example explores some of the properties of these functions
that bear remarkable resemblance to the properties of their trigonometric coun-
terparts.
[Link]/watch?v=VunyFD8keVg
CHAPTER 5. INTEGRATION 308
sinh2 (x) 1
tanh2 (x) + sech2 (x) = +
cosh2 (x) cosh2 (x)
sinh2 (x) + 1
= Now use identity from Part 1
cosh2 (x)
cosh2 (x)
= = 1.
cosh2 (x)
The following Key Idea summarizes many of the important identities relat-
ing to hyperbolic functions. Each can be verified by referring back to Defini-
tion 5.7.3.
d
1. cosh(x) = sinh(x)
dx
d
2. sinh(x) = cosh(x)
dx
d
3. tanh(x) = sech2 (x)
dx
d
4. sech(x) = − sech(x) tanh(x)
dx
d
5. csch(x) = − csch(x) coth(x)
dx
d
6. coth(x) = − csch2 (x)
dx
CHAPTER 5. INTEGRATION 310
Z
1. cosh(x) dx = sinh(x) + C
Z
2. sinh(x) dx = cosh(x) + C
Z
3. tanh(x) dx = ln(cosh(x)) + C
Z
4. coth(x) dx = ln |sinh(x) | + C
Solution.
d
1. Using the Chain Rule directly, we have dx cosh(2x) = 2 sinh(2x).
Just to demonstrate that it works, let’s also use the Basic Identity
found in Key Idea 5.7.8: cosh(2x) = cosh2 (x) + sinh2 (x).
d d
cosh(2x) = cosh2 (x) + sinh2 (x)
dx dx
= 2 cosh(x) sinh(x) + 2 sinh(x) cosh(x) [Link]/watch?v=MwYZHh9UaRo
= 4 cosh(x) sinh(x).
Using another Basic Identity, we can see that 4 cosh(x) sinh(x) =
2 sinh(2x). We get the same answer either way.
2. We employ substitution, with u = 7t − 3 and du = 7dt. Applying
Key Ideas 5.6.5 and Key Idea 5.7.8 we have:
Z
1
sech2 (7t − 3) dt = tanh(7t − 3) + C.
7
3.
Z ln(2) ln(2)
cosh(x) dx = sinh(x)
0 0
= sinh(ln(2)) − sinh(0)
= sinh(ln(2)).
We can simplify this last expression as sinh(x) is based on expo-
nentials:
eln(2) − e− ln(2)
sinh(ln(2)) =
2
CHAPTER 5. INTEGRATION 311
2 − 1/2
=
2
3
= .
4
y 10 y
10 y = cosh(x)
y = sinh(x)
8 5
6
x
4 y = cosh −1
(x) −10 −5 5 10
2 −5 y = sinh−1 (x)
x
2 4 6 8 10 −10
(a) (b)
y y
3
2
2
y = coth−1 (x)
1
y = sech−1 (x)
x x
−3 −2 −1 1 2 3 −3 −2 −1 1 2 3
−1
y = tanh−1 (x) y = csch−1 (x)
−2 −2
−3
(c) (d)
Figure 5.7.15 Graphs of the hyperbolic functions (with restricted domains) and
their inverses
Key Idea 5.7.16 Logarithmic definitions of Inverse Hyperbolic Func-
tions.
p
1. cosh−1 (x) = ln x + x2 − 1 ; x ≥ 1
1 1+x
2. tanh−1 (x) = ln ; |x| < 1
2 1−x
√ !
−1 1 + 1 − x2
3. sech (x) = ln ;0<x≤1
x
p
4. sinh−1 (x) = ln x + x2 + 1
1 x+1
5. coth−1 (x) = ln ; |x| > 1
2 x−1
√ !
−1 1 1 + x2
6. csch (x) = ln + ; x ̸= 0
x |x|
The following Key Ideas give the derivatives and integrals relating to the in-
verse hyperbolic functions. In Key Idea 5.7.18, both the inverse hyperbolic and
logarithmic function representations of the antiderivative are given, based on
Key Idea 5.7.16. Again, these latter functions are often more useful than the
former. Note how inverse hyperbolic functions can be used to solve integrals
we used Trigonometric Substitution to solve in Section 6.3.
CHAPTER 5. INTEGRATION 313
d 1
1. cosh−1 (x) = √ ;
dx x −1
2
x>1
d 1
2. sinh−1 (x) = √
dx 2
x +1
d 1
3. tanh−1 (x) = ;
dx 1 − x2
|x| < 1
d −1
4. sech−1 (x) = √ ;
dx x 1 − x2
0<x<1
d −1
5. csch−1 (x) = √ ;
dx |x| 1 + x2
x ̸= 0
d 1
6. coth−1 (x) = ;
dx 1 − x2
|x| > 1
1.
Z p
1
√ dx = ln x + x2 − a2 + C
−a
x2 2
x
(for 0 < x < a) = cosh−1 +C
a
2.
Z p
1
√ dx = ln x + x2 + a2 + C
x2 + a2
x
= sinh−1 +C
a
3.
Z
1 1 a+x
dx = ln +C
a2 −x 2 2a a−x
(
1
tanh−1 xa + C x2 < a2
= 1 a
−1 x
a coth a +C a2 < x2
4.
Z
1 1 x
√ dx = ln √ +C
x a2 − x2 a a + a2 − x2
1 x
(for 0 < x < a) = − sech−1 +C
a a
CHAPTER 5. INTEGRATION 314
5.
Z
1 1 x
√ dx = ln √ +C
x x2 + a2 a a + a2 + x2
1 x
= − csch−1 +C
a a
√
Note a2 = 10, hence a = 10. Now apply the integral rule.
1 3x
= sinh−1 √ +C
3 10
1 p
= ln 3x + 9x2 + 10 + C.
3
This section covers a lot of ground. New functions were introduced, along
with some of their fundamental identities, their derivatives and antiderivatives,
their inverses, and the derivatives and antiderivatives of these inverses. Four
Key Ideas were presented, each including quite a bit of information.
Do not view this section as containing a source of information to be mem-
orized, but rather as a reference for future problem solving. Key Idea 5.7.18
contains perhaps the most useful information. Know the integration forms it
helps evaluate and understand how to use the inverse hyperbolic answer and
the logarithmic answer.
The next section takes a brief break from demonstrating new integration
techniques. It instead demonstrates a technique of evaluating limits that return
indeterminate forms. This technique will be useful in Section 6.5, where limits
will arise in the evaluation of certain definite integrals.
CHAPTER 5. INTEGRATION 316
5.7.3 Exercises
Terms and Concepts
Z
1. In Key Idea 5.7.8, the equation tanh(x) dx = ln(cosh(x)) + C is given. Why is “ln |cosh(x)|” not used — i.e.,
why are absolute values not necessary?
2. The hyperbolic functions are used to define points on the right hand portion of the hyperbola x2 − y 2 = 1, as
shown in Figure 5.7.2. How can we use the hyperbolic functions to define points on the left hand portion of the
hyperbola?
Problems
Exercise Group. In the following exercises, verify the given identity using Definition 5.7.3, as done in Example 5.7.7.
3. Verify the identity coth2 (x) − csch2 (x) = 1 using the definitions of the hyperbolic functions.
4. Verify the identity cosh(2x) = cosh2 (x) + sinh2 (x) using the definitions of the hyperbolic functions.
cosh(2x) + 1
5. Verify the identity cosh2 (x) = using the definitions of the hyperbolic functions.
2
cosh(2x) − 1
6. Verify the identity sinh2 (x) = using the definitions of the hyperbolic functions.
2
d
7. Verify the identity [sech(x)] = − sech(x) tanh(x) using the definitions of the hyperbolic functions.
dx
d
8. Verify the identity [coth(x)] = − csch2 (x) using the definitions of the hyperbolic functions.
dx
Z
9. Verify the identity tanh(x) dx = ln(cosh(x)) + C using the definitions of the hyperbolic functions.
Z
10. Verify the identity coth(x) dx = ln |sinh(x)| + C using the definitions of the hyperbolic functions.
Exercise Group. In the following exercises, find the derivative of the given function.
11. Find the derivative of f (x) = sinh(2x). 12. Find the derivative of f (x) = cosh2 x.
13. Find the derivative of f (x) = tanh(x2 ). 14. Find the derivative of f (x) = ln(sinh(x)).
15. Find the derivative of f (x) = sinh(x) cosh(x). 16. Find the derivative of
f (x) = x sinh(x) − cosh(x).
17. Find the derivative of f (x) = sech−1 (x2 ). 18. Find the derivative of f (x) = sinh−1 (3x).
19. Find the derivative of f (x) = cosh−1 (2x2 ). 20. Find the derivative of f (x) = tanh−1 (x + 5).
21. Find the derivative of f (x) = tanh−1 (cos(x)). 22. Find the derivative of f (x) = cosh−1 (sec(x)).
Exercise Group. In the following exercises, find the equation of the line tangent to the function at the given x-value.
23. Find the equation of the tangent line to 24. Find the equation of the tangent line to
y = f (x) at x = 0, where f (x) = sinh(x). y = f (x) at x = ln(2), where f (x) = cosh(x).
y= y=
25. Find the equation of the tangent line to 26. Find the equation of the tangent line to
y = f (x) at x = − ln(3), where y = f (x) at x = ln(3), where f (x) = sech2 (x).
f (x) = tanh(x). y=
y=
27. Find the equation of the tangent line to 28. Find the equation√of the tangent line to
y = f (x) at x = 0, where f (x) = sinh−1 (x). y = f (x) at x = 2, where f (x) = cosh−1 (x).
y= y=
CHAPTER 5. INTEGRATION 317
Exercise Group. In the following exercises, evaluate the given indefinite integral.
Z
30. Evaluate
Z the indefinite integral
29. Evaluate the indefinite integral tanh(2x) dx.
cosh(3x − 7) dx.
Z
31. Evaluate
Z the indefinite integral
32. Evaluate the indefinite integral x cosh(x) dx.
sinh(x) cosh(x) dx.
Z Z
1
33. Evaluate the indefinite integral x sinh(x) dx. 34. Evaluate the indefinite integral √ dx.
x2 +1
Z Z
1 1
35. Evaluate the indefinite integral √ dx. 36. Evaluate the indefinite integral dx.
−9 x2 9 − x2
Z Z √
2x x
37. Evaluate the indefinite integral √ dx. 38. Evaluate the indefinite integral √ dx.
x4 − 4 1 + x3
Z Z
1 1
39. Evaluate the indefinite integral dx. 40. Evaluate the indefinite integral dx.
x4 − 16 x2 + x
Z Z
ex
41. Evaluate the indefinite integral dx. 42. Evaluate the indefinite integral sinh−1 (x) dx.
e2x + 1
Z
43. Evaluate
Z the indefinite integral
44. Evaluate the indefinite integral sech(x) dx.
tanh−1 (x) dx.
cosh(x)
(Hint: mutiply by cosh(x) ; set u = sinh(x).)
Exercise Group. In the following exercises, evaluate the given definite integral.
Z 1 46. Evaluate the definite integral
45. Evaluate the definite integral sinh(x) dx. Z ln(2)
−1 cosh(x) dx.
− ln(2)
Z 1 Z 2
1
47. Evaluate the definite integral sech2 (x) dx. 48. Evaluate the definite integral √ dx.
0 0 x2 + 1
INTEGRATION 318
1001
Appendix A
Problems
1.1.7. 5 1.1.8. 3
1.1.9. DNE 1.1.10. 23
1.1.11. −4 1.1.12. DNE or ∞
1.1.13. DNE 1.1.14. 6
1.1.15. 1 1.1.16. DNE
1.1.17. 1 1.1.18. DNE
1.1.19. DNE 1.1.20. 1
1.1.21. −7 1.1.22. 9
1.1.23. 5 1.1.24. −0.111111
1.1.25. 29 1.1.26. 0.2
1.1.27. −1 1.1.28. 0
1002
APPENDIX A. ANSWERS TO SELECTED EXERCISES 1003
1.2.2. y-tolerance
1.2.3. True
1.2.4. True
1.3.6. True
Problems
1.3.7. 9 1.3.8. 6
1.3.9. 0 1.3.10. DNE
1.3.11. 3 1.3.12. not possible to know
1.3.13. 3 1.3.14. −45
1.3.19. 23 4
π−5
1.3.20. π−8
√
1.3.21. 3 1.3.22. − 16
5
4
1.3.23. DNE 1.3.24. 256
√
1.3.25. 2 3 1.3.26. ln(4)
3
π 2 −4π−2 1.3.28. 2π−4
1.3.27. 2π 2 −2π+1 5π−5
1.3.29. 1
4 1.3.30. − 72
17 13
1.3.31. 4 1.3.32. 3
4 5
1.3.33. 9 1.3.34. 4
1.3.35. 0 1.3.36. 0
1.3.37. 1 1.3.38. 9
1.3.39. 8 1.3.40. 9
8
π
1.3.41. 1 1.3.42. 180
1.4.2. False
1.4.3. False
APPENDIX A. ANSWERS TO SELECTED EXERCISES 1004
1.4.4. True
Problems
1.4.5. 1.4.6.
(a) 2 (a) 0
(b) 2 (b) 4
(c) 2 (c) DNE
(d) 1 (d) 4
(e) DNE (e) DNE
(f) 4 (f) 1
1.4.7. 1.4.8.
(a) DNE or ∞ (a) 2
(b) DNE or ∞ (b) 3
(c) DNE or ∞ (c) DNE
(d) DNE (d) 4
(e) 5
(f) 4
1.4.9. 1.4.10.
(a) 1 (a) −5
(b) 1 (b) 1
(c) 1 (c) DNE
(d) 1 (d) 3
1.4.11. 1.4.12.
(a) 2 (a) a − 1
(b) 2 (b) a
(c) 2 (c) DNE
(d) 0 (d) a
(e) 2
(f) 2
(g) 2
(h) DNE
APPENDIX A. ANSWERS TO SELECTED EXERCISES 1005
1.4.13. 1.4.14.
(b) 6 (b) 0
(d) 2 (d) 0
1.4.15. 1.4.16.
(a) 9 (a) −1
(b) 9 (b) 0
(c) 9 (c) DNE
(d) 9 (d) 0
(e) 126
(f) 126
(g) 126
(h) 126
1.4.17. 1.4.18.
(a) 1 − cos2 (a) (a) 0
(b) sin2 (a) (b) 1
(c) 1 − cos2 (a) or sin2 (a) (c) DNE
(d) sin2 (a) (d) −2
1.4.19. 1.4.20.
(a) −4 (a) c
(b) −4 (b) c
(c) −4 (c) c
(d) −2 (d) c
1.4.21.
(a) −1
(b) 1
(c) DNE
(d) 0
1.5 · Continuity
1.5 · Exercises
Terms and Concepts
1.5.5. False
1.5.6. True
APPENDIX A. ANSWERS TO SELECTED EXERCISES 1006
1.5.7. True
1.5.8. False
1.5.9. False
1.5.10. True
Problems
1.5.19. 1.5.20.
(a) Yes. (a) Yes.
1.5.39. 1.23633
1.5.40. 0.523633
1.5.41. 0.693164
1.5.42. 0.785547
1.6.1. False
1.6.2. True
1.6.3. False
1.6.4. True
1.6.5. True
APPENDIX A. ANSWERS TO SELECTED EXERCISES 1007
Problems
1.6.9. 1.6.10.
(a) −∞ (a) −∞
(b) ∞ (b) ∞
(c) DNE
(d) ∞
(e) ∞
(f) ∞
1.6.11. 1.6.12.
1.6.15. 1.6.16.
(a) −∞ (a) −∞
(b) ∞ (b) −∞
(c) DNE (c) −∞
1.6.17. 1.6.18.
(a) ∞ (a) 1.8
(b) ∞ (b) 1.8
(c) ∞ (c) 1.8
1.6.25. ∞ 1.6.26. ∞
1.6.27. ∞ 1.6.28. ∞
2 · Derivatives
2 · Derivatives
2.1 · Instantaneous Rates of Change: The Derivative
2.1 · Exercises
APPENDIX A. ANSWERS TO SELECTED EXERCISES 1008
2.1.1. True
2.1.2. True
Problems
2.1.7. 0 2.1.8. 2
2.1.9. −3 2.1.10. 2x
2.1.11. 3x2 2.1.12. 6x − 1
2.1.13. −1
x2 2.1.14. −1
(s−2)2
(a) −2, 0, 4
(b) 2x
(c) Undefined
2.1.28.
(c) Undefined
2.1.33. (a) (−2, 0) ∪ (2, ∞) 2.1.34. (a) (−2, 2)
(b) (−∞, −2) ∪ (0, 2) (b) (−∞, −2) ∪ (2, ∞)
(c) {−2, 0, 2} (c) {−2, 2}
(d) (−1, 1) (d) (−1, 0) ∪ (1, ∞)
(e) (−∞, −1) ∪ (1, ∞) (e) (−∞, −1) ∪ (0, 1)
(f) {−1, 1} (f) {−1, 0, 1}
2.1.35. no
2.1.36. yes
2.2.1. velocity
2.2.3. linear functions
Problems
2.2.4. 20
2.2.5. −89
2.2.6. 91
2.2.7. f(10.1)
2.2.8. −2
2.2.9. 7
2.2.10. decibels per customer
2.2.11. foot per second squared
2.2.12. foot per hour
2.2.15. Choice 1 2.2.16. Choice 2
2.2.17. Choice 2 2.2.18. Choice 2
Problems
√
2.3.37. (a) y = 2·1
x− π
+−2 3 2.3.38. (a) 9 − 9x
2 6 2
√
(b) y = −1
− (−9)) + 90
(b) y = − 1
· 2 x − π6 + −2 3 −9 (x
2 2
2.4.1. False
2.4.2. False
2.4.3. True
2.4.4. the quotient rule
2.4.5. False
Problems
2.4.41. 17 2.4.42. 0
2
2.4.43. NONE 2.4.44. 0, 4
2.5.1. True
2.5.2. False
2.5.3. False
2.5.4. True
2.5.5. True
2.5.6. True
Problems
9 4
2.5.7. 10 4x3 − x 12x2 − 1 2.5.8. 15(3t − 2)
2 2
2.5.9. 3(sin(θ) + cos(θ)) (cos(θ) − sin(θ)) 2.5.10. (6t + 1) e3t +t−1
3 1 5
2.5.11. 4 ln(x) − x4 x − 4x
3 2.5.12. 0.693147 · 2q +4q
5q 4 + 4
4
2.5.14. −5 sin(5t)
2.5.13. 5 y + y1 1 − y12
2.5.15. 2 sec2 (2q) 2.5.16. − csc2 θ2 + 3 · 2θ
2
2.5.17. 6t5 − (t3t3 )2 cos t6 + 1 2.5.18. −5 cos4 (7q) · 7 sin(7q)
t3
1
2.5.41. x
k
2.5.42. x
Problems
1
√ 1 1 5
1
2.6.5. 1
√ + 2 w 2.6.6. √ 5 + y 0.166667
2 w √ 2 6 ( 6 y) 6
( w)
√
2.6.7. √1
2 9+t2
· 2t 2.6.8. 1
√
2 w
tan(w) + sec2 (w) w
2.6.9. 1.2y 0.2 2.6.10. πrπ−1 + 3.8r2.8
√ √
1 1
(cos(x) + ex ) + (ex − sin(x))
1
√
2.6.11.
w−(w−8) 2 w 2.6.12. √ 5
6
x
√ 2 6 ( 6 x)
( w)
−4x3 −y 0.6
2.6.13. 2y+1 2.6.14. x0.6
y
2.6.15. sin(x) sec(y) 2.6.16. x
y
2.6.17. −(ex x(x+2)·2−y )
x 2.6.18. ln(2)
2.6.27. 2.6.28.
(a) y = 0 (a) x = 1
√
−3 3
√ √
(b) y = −1.859(x − 0.1) + 0.2811 (b) y = 8 x− 0.6 + 0.8
2.6.29. 2.6.30.
(a) y = 4 (a) y = −x + 1
1 √
(b) y = 3
1 (x − 2) − 108 4 (b) y = 3 3
4
108 4
2.6.31. 2.6.32.
−1
√
(a) y = √
3
x− 7
2 + 6+3 3
2 (a) y = 1
√ √ −2
√
3(x−(4+3 3)) 3
(b) y = √
5
(x + 1) + 1
2 −1 + 5
(b) y = +
2 2
√
(c) y = √2 (x
5
+ 1) + 1
2 −1 − 5
! 0.6 0.6 ·3
( ( )) x0.6 ·3 −0.4 −y
2 − 4x3 − y −y x−0.4
− (2y+1)·12x2 −4x3 5 x0.6 5
2y+1 2.6.34. x1.2
2.6.33. (2y+1)2
1
ln(1+x) x2
2.6.37. (a) (1 + x) x 1
x(x+1) − x2
2.6.38. (a) (2x) (2x ln(2x) + x)
(b) y = (1 − 2 ln(2)) (x − 1) + 2 (b) y = (2 + 4 ln(2)) (x − 1) + 2
xx
2.6.39. (a) x+1 ln(x) + 1 − x+1
1
2.6.40. (a) xsin(x)+2 cos(x) ln(x) + sin(x)+2
x
(b) y = 14 (x − 1) + 12 3π 2
(b) y = 4 x − 2 + 2 π π 3
x+2 x+1 −
2.6.41. (a) x+1 1 1
x+2 2.6.42. (a) (x+1)(x+2) 1
(x+3)(x+4) x+1 + 1
x+2 − 1
x+3 − 1
x+4
(b) y = 19 (x − 1) + 2
3 (b) y = 11
72 x + 1
6
2.7.1. False
Problems
1
2.7.9. 7
2.7.10. − 14
1
2.7.11. −0.5
1
2.7.12. 132
2.7.13. − 25
4
1
2.7.14. 12
2.7.15. − √ 1
·4 2.7.16. − √1 ·7
1−(4w)2 |7x| (7x)2 −1
APPENDIX A. ANSWERS TO SELECTED EXERCISES 1014
2.7.17. 1
1+(2r)2
·2 2.7.18. cos−1 (w) − w √1−w
1
2
√ √
− 3
2.7.29. y = 2 x − 2 + − π3 2.7.30. y = −4 x − 4
3
+ π
6
Problems
3.3.3. Answers will vary; graphs should be steeper near x = 0 than near x = 2.
3.3.5. False
Problems
Problems
3.5.3. True
3.5.4. True
3.5.5. True
4.1.1. False
4.1.2. False
Problems
4.2.1. True
4.2.2. False
Problems
4.2.3.
cm
(a) 0.198944 s
cm
(b) 0.0198944 s
cm
(c) 0.00198944 s
4.2.4.
cm
(a) 0.397887 s
cm
(b) 0.00397887 s
4.2.6.
mi
(a) 68.75 h
mi
(b) 75 h
4.2.7.
rad
(a) 258.537 hr
rad
(b) 413.417 hr
rad
(c) 424 hr
4.2.8.
rad
(a) 0.0225641 s
rad
(b) 0.553459 s
rad
(c) 7.33333 s
4.2.9.
ft
(a) 0.0417029 s
ft
(b) 0.458349 s
APPENDIX A. ANSWERS TO SELECTED EXERCISES 1020
ft
(c) 3.35489 s
(d) ∞
4.2.10.
ft
(a) 30.5941 min
ft
(b) 36.0555 min
ft
(c) 301.496 min
4.2.11.
ft
(a) 19.1658 s
ft
(b) 0.191658 s
ft
(c) 0.0395988 s
(d) 381.791 s
4.2.12.
ft
(a) 0.632456 s
ft
(b) 1.6 s
(c) 51.9615 ft
4.2.13.
(a) 80 ft
ft
(b) 1.71499 s
ft
(c) 1.83829 s
(d) 74.162 ft
4.2.14.
(a) 96 ft
ft
(b) 9.42478 s
ft
4.2.15. 0.00230973 s
4.3 · Optimization
4.3 · Exercises
Terms and Concepts
4.3.1. True
4.3.2. False
Problems
4.3.3. 5625
APPENDIX A. ANSWERS TO SELECTED EXERCISES 1021
√
4.3.4. 2 560
4.3.5. DNE
8450
4.3.6. 29
4.3.7. 1
225
4.3.8. 150 ft; 2 ft
4.3.9. (a) 3.83722 cm
(b) 7.67443 cm
4.3.10. (a) 3.20058 in
(b) 6.40117 in
4.3.11. (a) 3.0456 cm
(b) 12.1824 cm
4.3.12. 11664 in3
4.3.13. 10.3923 in; 14.6969 in
4.3.14. (a) 0.535898 mi
(b) $503,730.67
4.3.15. (a) 0 mi
(b) $474,341.65
4.3.16. 33.6239 ft
4.3.17. 23.7599 ft
√ √
4.3.18. 2; 2
4.4 · Differentials
4.4 · Exercises
Terms and Concepts
4.4.1. True
4.4.2. True
4.4.3. False
4.4.4. True
4.4.6. True
Problems
(b) 76.8
4.4.33. 3.92699
4.4.34. −4 ft2
4.4.35. 4.4.36.
(a) 297.717 ft (a) 298.868 ft
(b) 62.3155 ft (b) 17.335 ft
(c) 20.9% (c) 5.8%
4.4.37. 4.4.38. Isosceles ... feet
(a) 298.868 ft
(b) 8.66751 ft
(c) 2.9%
4.4.39. 1%
4.5.2. True
4.5.3. 6 + 3x − 4x2
4.5.4. 30
Problems
2 2
4.5.13. 1 + 0.5(x − 1) − 0.125(x − 1) + 4.5.14. 0.693147 + 0.5(x − 1) − 0.125(x − 1) +
3 4 3 4
0.0625(x − 1) − 0.0390625(x − 1) 0.0416667(x − 1) − 0.015625(x − 1)
4.5.15. 2
2 4.5.16. 0.5 + 0.866025 x − π6 − 0.25 x − π6 −
0.707107 − 0.707107 x − π4 − 0.353553 x − π4 + 3 4
3 4 0.144338 x − π6 + 0.0208333 x − π6 +
0.117851 x − π4 + 0.0294628 x − π4 − 5
5 6 0.00721688 x − π6
0.00589256 x − π4 − 0.000982093 x − π4
APPENDIX A. ANSWERS TO SELECTED EXERCISES 1023
4.5.17. 4.5.18.
2 3
0.5 − 0.25(x − 2) + 0.125(x − 2) − 0.0625(x − 2) + 2 3 4
1 − 2(x − 1) + 3(x − 1) − 4(x − 1) + 5(x − 1) −
4 5
0.03125(x − 2) + 0.015625(x − 2) 5 6 7
6(x − 1) + 7(x − 1) − 8(x − 1) + 9(x − 1)
8
2 π 2 −2 2
4.5.19. 0.5 + 0.5(x + 1) + 0.25(x + 1) 4.5.20. −π 2 − 2π(x − π) + 2 (x − π)
4.6.2. False
4.6.3. False
Problems
5 · Integration
5 · Integration
5.1 · Antiderivatives and Indefinite Integration
5.1 · Exercises
APPENDIX A. ANSWERS TO SELECTED EXERCISES 1024
5.1.2. an antiderivative
5.1.4. (a) opposite
(b) opposite
5.1.6. velocity
5.1.7. velocity
5.1.8. F (x) + G(x)
Problems
5.1.30. − (cos(x) + 7)
5.1.31. 2ex + 8
4 3
5.1.32. 3 x4 − 2 x3 − 5
3
5.1.38. 2θ − cos(θ) + 10
29x6 6x
5.1.39. 30 + 3.2104 − cos(x) − 3.55811x − 0.311487
5.1.40. 3x + 23
5.2.3. 0
R
5.2.4. 02 (2x + 3) dx
APPENDIX A. ANSWERS TO SELECTED EXERCISES 1025
Problems
5.2.5. 5.2.6.
(a) 3 (a) −4
(b) 4 (b) −5
(c) 3 (c) −3
(d) 0 (d) 1
(e) −4 (e) −2
(f) 9 (f) 10
5.2.7. 5.2.8.
(a) 4 (a) − 12
(b) 2 (b) 0
3
(c) 4 (c) 2
3
(d) 2 (d) 2
9
(e) 1 (e) 2
15
(f) 2 (f) 2
5.2.9. 5.2.10.
(a) π (a) 15
(b) π (b) 12
(c) 2π (c) 0
5.2.11. 5.2.12.
(c) −4 (c) 8
3
(d) −2 (d) 38
3
5.2.15. 5.2.16.
ft ft
(a) 2 s (a) 3 s
5.2.17.
ft
(a) 64 s
(b) 64 ft
(c) 2 s
(d) 4.64575 s
5.2.18.
ft
(a) 96 s
(b) 6 s
(c) 6 s
(d) 208 ft
5.2.19. 2 5.2.20. 5
5.2.21. 16 5.2.22. a = − 27 b
5.2.23. 22 5.2.24. −7
5.2.25. 0 5.2.26. a = − 18
11 b
5.3.1. limits
5.3.2. 14
5.3.3. rectangles
5.3.4. True
Problems
5.3.13. 1; 4; 5i 5.3.14. 0; 6; i2 − 2
i i
5.3.15. 1; 5; i+4 5.3.16. 1; 5; − (−e)
(n−1)2 −9 9
5.3.35. (a) 5.3.36. (a) 6 + 1n + 1n2
4n2
(b) 0.2025 (b) 5.19
(c) 0.245025 (c) 5.9109
(d) 0.2495 (d) 5.99101
(e) 1 (e) 6
4
3 844
−256 128
5.3.37. (a) 2 5.3.38. (a) 3 + 1n + 3n2
3
(b) 2 (b) 256.16
3
(c) 2 (c) 278.778
3
(d) 2 (d) 281.077
3 844
(e) 2 (e) 3
5.3.39. (a) 60 − 200
n 5.3.40. (a) − 12
1
+ 1
12n2
(b) 40 (b) −0.0825
(c) 58 (c) −0.083325
(d) 59.8 (d) −0.0833332
(e) 60 (e) − 12
1
5.4.2. 0
5.4.3. True
Problems
5.4.5. 40 5.4.6. 98
3
5.4.7. 0 5.4.8. 1
√
5.4.9. 2 − 2 5.4.10. 7
( 4095
512 )
5.4.12. −14
5.4.11. ln(8)
5.4.13. 6 5.4.14. e2 − e1
5.4.15. 248 5.4.16. 4
3
968 5.4.18. ln(6)
5.4.17. 5
6 242
5.4.19. 7 5.4.20. 1215
1 1
5.4.21. 2 5.4.22. 3
1 1
5.4.23. 4 5.4.24. 90
5.4.25. 14 5.4.26. 16
5.4.27. 0 5.4.28. 1
APPENDIX A. ANSWERS TO SELECTED EXERCISES 1028
5.4.35. 2 5.4.36. π
π
7 64
5.4.37. 2 5.4.38. 3
729 1
5.4.39. 4 5.4.40. e1 −1
ft ft
5.4.47. −256 s 5.4.48. 99 s
1 ft ft
5.4.49. 2 s 5.4.50. 0 s
5.5.1. False
5.5.4. A quadratic function (i.e., parabola)
Problems
5.5.5. 5.5.6.
(a) 0.75 (a) 250
(b) 0.666667 (b) 250
(c) 0.666667 (c) 250
5.5.7. 5.5.8.
(a) 1.89612 (a) 5.14626
(b) 2.00456 (b) 5.25221
(c) 2 (c) 5.33333
5.5.9. 5.5.10.
5.5.11. 5.5.12.
(a) 0 (a) 12.2942
(b) 0 (b) 13.3923
(c) 0 (c) 14.1372
5.5.13. 5.5.14.
5.5.21. 5.5.22.
(a) 161 (a) 130
(b) 12 (b) 18
5.5.23. 5.5.24.
(a) 994 (a) 5591
(b) 62 (b) 46
5.6 · Substitution
5.6 · Exercises
Terms and Concepts
Problems
1
8 1
10
5.6.3. 8 x4 + 8 +C 5.6.4. x2 + 9x + 1
10 +C
5 10
10 x − 7 7x − 6x2 + 5
1 2 3
5.6.5. +C 5.6.6. 10 +C
1 2
√
5.6.7. 4 ln(|4x + 6|) + C 5.6.8. 5 5x + 3 + C
APPENDIX A. ANSWERS TO SELECTED EXERCISES 1030
√
5.6.9. 23 (x − 2) x+1+C 5.6.10. x( 2 ) 29 x3 + 2 + C
3
√ √
5.6.11. 2e x
+C 5.6.12. 27 x7 + 7 + C
2 ln2 (x)
5.6.13. C − 1 1
4 x2 +9 5.6.14. 2 +C
(sin(x))4 (cos(x))5
5.6.15. 4 +C 5.6.16. C − 5
cos(1−6x) tan(7−5x)
5.6.17. 6 +C 5.6.18. C − 5
1 9
5.6.19. 7 ln(|sec(7x) + tan(7x)|) + C 5.6.20. 19 (tan(x)) + C
5.6.21. 1
9 sin x 9
+C 5.6.22. tan(x) − x + C
5.6.23. ln(|sin(x)|) + C 5.6.24. − ln(|csc(x) + cot(x)|) + C
5.6.25. 14 e4x−5 + C
5
5.6.26. 15 ex + C
5.6.27. 12 e(x+1) + C
2
5.6.28. x − 3e−x + C
1 −3x
5.6.29. ln(ex + 7) + C 5.6.30. C − 3e + 51 e−5x
99x 28x
5.6.31. 19.775 +C 5.6.32. 5.54518 +C
x2 x3 x2
5.6.37. 2 + 3x − 8 ln(|x|) + C 5.6.38. 3 + 2 + x + ln(|x|) + C
5.6.39. 5.6.40. (x+8)2
− 7(x + 8) − 17 ln(|x + 8|) + C
1 3 3 2 2
3 (x + 1) + 2 (x + 1) +3(x + 1)−ln(|x + 1|)+C
5.6.41. 5.6.42. 1
3 ln x3 − 6x2 + 9 +C
2
C − 4(x − 6) + 92(x − 6) + 266 ln(|x − 6|)
5.6.43. 2.44949 tan−1 x
2.44949 +C 5.6.44. 5 sin−1 x5 + C
5.6.45. 7 sin−1 x
3.16228 +C 5.6.46. 97 sec−1 |x| +C
7
5.6.47. 3
sec−1 |x|
+C 5.6.48. 0.5 sin−1 x2 + C
4 8
5.6.49. 0.258199 tan−1 x+9
15 +C 5.6.50. 3 sin−1 x−86 +C
−1 x−6
−1 x−4
5.6.51. 6 sin 3 +C 5.6.52. tan 8 +C
6
5.6.53. C − 5(x51−2) 5.6.54. 5x6 − 6x4 + 6
1
+C
36
√
5.6.55. 16 6 + 6x2 + C 5.6.56. tan x8 + 3 + C
5.6.57. C − 23 (cos(x))( 2 )
3
5.6.58. C − 1
9 cos(9x + 5)
5.6.59. ln(|x − 8|) + C 5.6.60. 2 ln(|x + 5|) + C
5.6.61. 2x + 3x + ln x − 4x + 5
2 2
+C 5.6.62. ln x2 − 2x − 3 + C
5.6.63. 2 ln x2 − 9x + 3 + C 5.6.64. − 12 x2 + 9x + ln x2 + 3x + 4 + C
2
1
5.6.65. 14 tan−1 x7 + C 5.6.66. tan−1 (8x) + C
5.6.67. sec−1 (|9x|) + C 5.6.68. 1
5 sin−1 5 x2 + C
5.6.69.
5.6.70.
2 ln x − 12x + 117 −
5 2 8
9 tan−1 x−6
9 +C 5
3 tan
−1 x−4
6 + 3
2 ln x2 − 8x + 52 +C
APPENDIX A. ANSWERS TO SELECTED EXERCISES 1031
5.6.71. x2
5.6.72. − 18 ln x2 + 36 +C
x+20.1246 tan−1
2
+9 ln x2 − 4x + 9 +C
x−2
2.23607
5.6.73. 12 x2 − 6x + 4 ln x2 + 6x + 19 + 5.6.74. − tan−1 (cos(x)) + C
11.7004 tan−1 3.16228
x+3
+C
5.6.75. tan−1 (sin(x)) + C 5.6.76. C − ln(|csc(x) + cot(x)|)
√ √
5.6.77. 5 x2 + 16x + 56 + C 5.6.78. x2 + 12x + 34 + C
5
10190
5.6.79. ln 8
5.6.80. 3
1 1
5.6.81. 3 5.6.82. 8
π
5.6.83. 1
2 e4 − e 4 5.6.84. 2
π
5.6.85. 2 5.6.86. 1
6 π
5.7.19. √ 1 2 · 2 · 2x 5.7.20. 1
(2x2 ) −1 1−(x+5)2
Quick Reference
d d
1. (cx) = c 5. (u(v)) = u′ (v)v ′
dx dx
d
2. (u ± v) = u′ ± v ′
dx d
6. (c) = 0
d dx
3. (u · v) = uv ′ + u′ v
dx
d u vu′ − uv ′ d
4. ( )= 7. (x) = 1
dx v v2 dx
d n d
1. (x ) = nxn−1 10. (tan x) = sec2 x
dx dx
d x d
2. (e ) = ex 11. (cot x) = − csc2 x
dx dx
d x
3. (a ) = ln a · ax d
dx 12. (cosh x) = sinh x
dx
d 1
4. (ln x) =
dx x d
13. (sinh x) = cosh x
d 1 1 dx
5. (loga x) = ·
dx ln a x d
14. (sech x) = − sech x tanh x
d dx
6. (sin x) = cos x
dx
d
d 15. (tanh x) = sech2 x
7. (cos x) = − sin x dx
dx
d d
8. (csc x) = − csc x cot x 16. (csch x) = − csch x coth x
dx dx
d d
9. (sec x) = sec x tan x 17. (coth x) = − csch2 x
dx dx
1087
APPENDIX B. QUICK REFERENCE 1088
d 1 d 1
1. (sin−1 x) = √ 7. (cosh−1 x) = √
dx 1 − x2 dx x −1
2
d −1 d 1
2. (cos−1 x) = √ 8. (sinh−1 x) = √
dx 1 − x2 dx 2
x +1
d −1 d −1
3. (csc−1 x) = √ 9. (sech−1 x) = √
dx |x| x2 − 1 dx x 1 − x2
d 1 d −1
4. (sec−1 x) = √ 10. (csch−1 x) = √
dx |x| x2 − 1 dx |x| 1 + x2
d 1 d 1
5. (tan−1 x) = 11. (tanh−1 x) =
dx 1 + x2 dx 1 − x2
d −1 d 1
6. (cot−1 x) = 12. (coth−1 x) =
dx 1 + x2 dx 1 − x2
Z Z Z
1. c · f (x) dx = c f (x) dx 3. 0 dx = C
Z Z Z Z
2. f (x) ± g(x) dx = f (x) dx ± g(x) dx 4. 1 dx = x + C
Z Z
1
1. x x
e dx = e + C 4. dx = ln |x| + C
x
Z Z
1
2. ln x dx = x ln x − x + C 5. xn dx = xn+1 + C, n ̸= −1
n+1
Z
1
3. ax dx = · ax + C
ln a
Z
1. cos x dx = sin x + C
Z
2. sin x dx = − cos x + C
Z
3. tan x dx = − ln |cos x| + C
Z
4. sec x dx = ln |sec x + tan x| + C
APPENDIX B. QUICK REFERENCE 1089
Z
5. csc x dx = − ln |csc x + cot x| + C
Z
6. cot x dx = ln |sin x| + C
Z
7. sec2 x dx = tan x + C
Z
8. csc2 x dx = − cot x + C
Z
9. sec x tan x dx = sec x + C
Z
10. csc x cot x dx = − csc x + C
Z
1 1
11. cos2 x dx = x + sin 2x + C
2 4
Z
1 1
12. x − sin 2x + C
sin2 x dx =
2 4
Z
1 1 −1 x
13. dx = tan +C
x2 + a2 a a
Z x
1
14. √ = sin−1 +C
a2 − x2 a
Z
1 1 |x|
15. √ = sec−1 +C
x x −a
2 2 a a
( √ ) (0, 1) ( √ )
−1, 3 1
, 3
( √ √ 2) 2 2 ( 2√ √ )
− 2
2
, 22 π/2
2
2
, 2
2
( √ ) 2π/3 π/3 (√ )
− 23 , 12 3π/4 90◦ π/4
3 1
,2
120◦ 60 ◦ 2
◦ 45◦ π/6
5π/6 135
150◦ 30◦
210◦ 330◦
7π/6 225◦ 315◦ 11π/6
( √ ) ◦ (√ )
5π/4 240 300◦ 7π/4
− 23 , − 12 270◦ 2
3
, − 21
( √ √ ) 4π/3 5π/3 (√ √ )
− 22 , − 22 3π/2
2
, − 2
( √ ) ( √2 ) 2
− 12 , − 23 1
2
, − 23
(0, −1)
y sin θ = y cos θ = x
1 1
(x, y) csc θ =
y
sec θ =
x
y x
y θ tan θ =
x
cot θ =
y
x
x
APPENDIX B. QUICK REFERENCE 1091
O H
sin θ = csc θ =
se H O
Opposite
u
ten A H
o cos θ = sec θ =
H yp H A
O A
θ tan θ =
A
cot θ =
O
Adjacent
2. tan2 x + 1 = sec2 x 2.
2 tan x
3. tan 2x =
1 − tan2 x
π
1. sin − x = cos x 1. sin(−x) = − sin x
2
π 2. cos(−x) = cos x
2. cos − x = sin x
2 3. tan(−x) = − tan x
π
3. tan − x = cot x 4. csc(−x) = − csc x
2
π
5. sec(−x) = sec x
4. csc − x = sec x
2
π 6. cot(−x) = − cot x
5. sec − x = csc x
2 List B.3.4 Even/Odd Identities
π
6. cot − x = tan x
2
2 1 − cos 2x x+y x−y
1. sin x = 1. sin x + sin y = 2 sin cos
2 2 2
1 + cos 2x
2. cos2 x = x−y x+y
2 2. sin x − sin y = 2 sin cos
2 2
1 − cos 2x
3. tan2 x = 3. cos x + cos y =
1 + cos 2x x+y x−y
2 cos cos
List B.3.5 Power-Reducing Formulas 2 2
4. cos x − cos y =
x+y x−y
−2 sin sin
2 2
1
1. sin x sin y = cos(x − y) − cos(x + y)
2
1
2. cos x cos y = cos(x − y) + cos(x + y)
2
1
3. sin x cos y = sin(x + y) + sin(x − y)
2
Area = 12 bh
θ
Surface
√ Area = h
b πr r2 + h2 + πr2
Law of Cosines: r
c2 = a2 +b2 −2ab cos θ
APPENDIX B. QUICK REFERENCE 1093
Trapezoids a Sphere
Area = 12 (a + b)h Volume = 43 πr3
h
Surface Area =4πr2
r
b
B.5 Algebra
Let p(x) = an xn +an−1 xn−1 +· · ·+a1 x+a0 be a polynomial. If p(a) = 0, then a is a zero of the polynomial
and a solution of the equation p(x) = 0. Furthermore, (x − a) is a f actor of the polynomial.
An nth degree polynomial has n (not necessarily distinct) zeros. Although all of these zeros may be imaginary,
a real polynomial of odd degree must have at least one real zero.
Quadratic Formula.
√
If p(x) = ax2 + bx + c, and 0 ≤ b2 − 4ac, then the real zeros of p are x = (−b ± b2 − 4ac)/2a
APPENDIX B. QUICK REFERENCE 1094
Special Factors.
x2 − a2 = (x − a)(x + a)
x3 − a3 = (x − a)(x2 + ax + a2 )
x3 + a3 = (x + a)(x2 − ax + a2 )
x4 − a4 = (x2 − a2 )(x2 + a2 )
n(n − 1) n−2 2
(x + y)n = xn + nxn−1 y + x y + · · · + nxy n−1 + y n
2!
n(n − 1) n−2 2
(x − y)n = xn − nxn−1 y + x y − · · · ± nxy n−1 ∓ y n
2!
Binomial Theorem.
(x + y)2 = x2 + 2xy + y 2
(x − y)2 = x2 − 2xy + y 2
(x + y)3 = x3 + 3x2 y + 3xy 2 + y 3
(x − y)3 = x3 − 3x2 y + 3xy 2 − y 3
(x + y)4 = x4 + 4x3 y + 6x2 y 2 + 4xy 3 + y 4
(x − y)4 = x4 − 4x3 y + 6x2 y 2 − 4xy 3 + y 4
Factoring by Grouping.
Arithmetic Operations.
a c ad + bc a+b a b
ab + ac = a(b + c) + = = +
a b d bd c c c
a
a d ad
cb = = b = a a
=
ac
b c bc c bc b b
d c
b ab a−b b−a ab + ac
a = = =b+c
c c c−d d−c a
√
a0 = 1, a ̸= 0 (ab)x = ax bx ax ay = ax+y a = a1/2
ax √ a x ax √
= ax−y n
a = a1/n = x n
am = am/n
ay b b r
√ √ √
1 √ a n
a
a−x = x = √
n n
ab = n
a b (ax )y = axy n
n
a b b
APPENDIX B. QUICK REFERENCE 1095
Summation Formulas:.
X
n X
n
n(n + 1)
c = cn i=
i=1 i=1
2
X
n Xn 2
n(n + 1)(2n + 1) n(n + 1)
i2 = i3 =
i=1
6 i=1
2
Trapezoidal Rule:.
Z b
∆x
f (x) dx ≈ f (x0 ) + 2f (x1 ) + 2f (x2 ) + · · · + 2f (xn−1 ) + f (xn )
a 2
(b − a)3
with Error ≤ 2
max |f ′′ (x)|
12n
Simpson’s Rule:.
Z b
∆x
f (x) dx ≈ f (x0 ) + 4f (x1 ) + 2f (x2 ) + 4f (x3 ) + · · · + 2f (xn−2 ) + 4f (xn−1 ) + f (xn )
a 3
(b − a)5
with Error ≤ max f (4) (x)
180n4
Arc Length:.
Z b p
L= 1 + f ′ (x)2 dx
a
Surface of Revolution:.
Z b p
2π f (x) 1 + f ′ (x)2 dx
a
(where f (x) ≥ 0)
Z b p
S = 2π x 1 + f ′ (x)2 dx
a
(where a, b ≥ 0)
1097
INDEX 1098